url
stringlengths
31
38
title
stringlengths
7
229
abstract
stringlengths
39
2.87k
text
stringlengths
1
3.74M
meta
dict
https://arxiv.org/abs/1912.00224
Almost sharp bounds on the number of discrete chains in the plane
The following generalisation of the Erdős unit distance problem was recently suggested by Palsson, Senger and Sheffer. Given $k$ positive real numbers $\delta_1,\dots,\delta_k$, a $(k+1)$-tuple $(p_1,\dots,p_{k+1})$ in $\mathbb{R}^d$ is called a $(\delta,k)$-chain if $\|p_j-p_{j+1}\| = \delta_j$ for every $1\leq j \leq k$. What is the maximum number $C_k^d(n)$ of $(k,\delta)$-chains in a set of $n$ points in $\mathbb{R}^d$, where the maximum is taken over all $\delta$? Improving the results of Palsson, Senger and Sheffer, we essentially determine this maximum for all $k$ in the planar case. error term It is only for $k\equiv 1$ (mod) $3$ that the answer depends on the maximum number of unit distances in a set of $n$ points. We also obtain almost sharp results for even $k$ in $3$ dimension.
\section{Introduction}\label{sec:introduction} Determining the maximum possible number of pairs $u_d(n)$ at distance $1$ apart in a set of $n$ points in $\mathbb{R}^d$ for $d=2$ is one of the central questions in combinatorial geometry, known as the Erdős Unit Distance problem. The question dates back to 1946, and despite much effort, the best known upper and lower bounds are still very far apart. For some constants $C,c>0$, we have \[n^{1+c/\log\log n}\leq u_2(n)\le Cn^{4/3},\] where the lower bound is due to Erdős \cite{erdos} and the upper bound is due to Spencer, Szemerédi and Trotter \cite{SST}. Recently, there has been great progress in a closely related problem of determining the minimum number of distinct distances between $n$ points on the plane due to Guth and Katz \cite{GK}, but the powerful algebraic machinery they used has not yet given any improvement for the unit distance question. As in the planar case, the best known upper and lower bounds in the $3$-dimensional case are also far apart. For every $\varepsilon>0$ there are constants $c,C>0$ such that we have \begin{equation}\label{eqzahl} cn^{4/3}\log\log n\le u_3(n)\le Cn^{295/197+\varepsilon}, \end{equation} where the lower bound is due to Erdős \cite{erdos3}, and the upper bound is due to Zahl \cite{Za2}. The latter is a recent improvement upon the upper bound $O(n^{3/2})$ by Kaplan, Matou\v sek, Safernová, and Sharir \cite{KMSS}, and Zahl \cite{Za}. This paper can be seen as an effort to find generalisations of the Unit Distance problem that are within the reach of our current methods. In what follows, we describe the generalisation that we work with. Palsson, Senger and Sheffer \cite{Shef} suggested the following question. Let $\bm\delta=(\delta_1,\dots,\delta_k)$ be a fixed sequence of $k$ positive reals. A $(k+1)$-tuple $(p_1,\dots,p_{k+1})$ of distinct points in $\mathbb{R}^d$ is called a \emph{$k$-chain} if $\|p_i-p_{i+1}\|=\delta_i$ for all $i=1,\dots,k$. For every fixed $k$ determine $C^d_k(n)$, the maximum number of $k$-chains that can be spanned by a set of $n$ points in $\mathbb{R}^d$. We do not include $\bm\delta$ in the notation, since our results, with the exception of Proposition \ref{3dbound} do not depend on $\bm\delta$ up to the order of magnitude. The authors of \cite{Shef} give the following lower bound on $C^2_k(n)$: \[C^2_k(n)=\Omega\left (n^{\lfloor (k+1)/3 \rfloor+1}\right ). \] They also provided upper bounds in terms of the maximum number of unit distances. \begin{prop}[Palsson, Senger, and Sheffer \cite{Shef}] \begin{equation*} C^2_k(n) = \begin{cases} O\left (n\cdot u_2(n)^{k/3} \right ) & \text{\rm if $k\equiv 0$ (mod $3$),}\\ O\left (u_2(n)^{(k+2)/3}\right ) & \text{\rm if $k\equiv 1$ (mod $3$),}\\ O\left (n^2\cdot u_2(n)^{(k-2)/3}\right ) & \text{\rm if $k\equiv 2$ (mod $3$).} \end{cases} \end{equation*} \end{prop} If $u_2(n) = O(n^{1+\varepsilon})$ for any $ {\varepsilon}>0$, which is conjectured to hold, then the upper bounds in the proposition above almost match the lower bound given above. However, as we have already mentioned, determining the order of magnitude of $u_2(n)$ has proved to be a very hard problem and is very far from its resolution. Thus, it is interesting to obtain ``unconditional'' bounds, that depend on the value of $u_2(n)$ as little as possible. In \cite{Shef}, the following ``unconditional'' upper bounds were proved in the planar case. \begin{thm}[Palsson, Senger, and Sheffer \cite{Shef}]\label{PSS2} $C_2^2(n)=\Theta(n^2)$, and for every $k\geq 3$ we have \[C^2_k(n)=O\left (n^{2k/5+1+\gamma_k}\right ), \] where $\gamma_k\leq \frac{1}{12}$, and $\gamma_k \to \frac{4}{75}$ as $k\to \infty$. \end{thm} In our main result, in two-third of the cases we almost determine the value of $C^2_k(n),$ no matter what the value of $u_2(n)$ is, by matching the lower bounds given in Theorem~\ref{PSS2}. Further, we show that in the remaining cases determining $C^2_k(n)$ essentially reduces to determining the maximum number of unit distances. \begin{thm}\label{main}For every integer $k\geq 1$ we have\,\footnote{In what follows $f(n)= \tilde O(g(n))$ means that there exist positive constants $c, C$ such that $ f(n)/g(n) \le C\log^{c} n$ for every sufficiently large $n.$ We write $f(n) = \tilde{\Omega}(g(n))$ if $g(n) = \tilde O(f(n))$, and $f(n) = \tilde{\Theta}(g(n))$ if $f(n) = \tilde O(g(n))$ and $g(n) = \tilde O(f(n)).$} \[C^2_k(n)=\tilde{\Theta}\left (n^{\lfloor (k+1)/3 \rfloor+1}\right ) \textrm{ if } k\equiv 0,2 \text{ \rm (mod } 3 \text{\rm )},\] and for any $\varepsilon> 0$ we have \[C^2_k(n)=\Omega\left (n^{(k-1)/3}u_2(n)\right) \textrm{ and } C^2_k(n)=O\left(n^{(k-1)/3+\varepsilon}u_2(n)\right) \textrm { if } k\equiv 1 \text{ \rm(mod $3$)}. \] \end{thm} Let us turn our attention to the $3$-dimensional case. The following was proved in \cite{Shef}. \begin{thm}[Palsson, Senger, and Sheffer \cite{Shef}]\label{thmshefr3} For any integer $k\geq 2$, we have \[C^3_k(n)=\Omega \left (n^{\lfloor k/2 \rfloor +1}\right ),\] and \begin{equation*} C^3_k(n) = \begin{cases} O\left (n^{2k/3+1}\right ) & \text{\rm if $k\equiv 0$ (mod $3$),}\\ O\left (n^{2k/3+23/33+\varepsilon}\right ) & \text{\rm if $k\equiv 1$ (mod $3$),}\\ O\left (n^{2k/3+2/3}\right ) & \text{\rm if $k\equiv 2$ (mod $3$).} \end{cases} \end{equation*} \end{thm} We improve their upper bound and essentially settle the problem for even $k$. \begin{thm}\label{3d}For any integer $k\geq 2$ we have \[C^3_k(n)=\tilde O\left (n^{k/2+1}\right).\] Furthermore, for even $k$ we have \[C^3_k(n)=\tilde{\Theta}\left (n^{k/2+1}\right).\] \end{thm} We also improve the lower bound from Theorem~\ref{thmshefr3} for odd $k$ and $\bm\delta=(1,\dots,1)$. Let $us_3(n)$ be the maximum possible number of pairs at unit distance apart in $X\times Y$, where $X$ is a set of $n$ points in $\mathbb{R}^3$ and $Y$ is a set of $n$ points on a sphere in $\mathbb{R}^3$. \begin{prop}\label{oddk}Let $k\geq 3$ odd. Then for $\bm\delta=(1,\dots,1)$ we have \[C^3_k(n)=\Omega\left (\max\left \{\frac{u_3(n)^k}{n^{k-1}},us_3(n)n^{(k-1)/2}\right \}\right ).\] \end{prop} By using stereographic projection we obtain that $us_3(n)$ equals the maximum number of incidences between a set of $n$ points and a set of $n$ circles (not necessarily of the same radii) in the plane. Thus we have \[cn^{4/3}\leq us_3(n)=\tilde{O}\left ( n^{15/11} \right )\] (For the lower bound see \cite{us1}, and for the upper bound see \cite{us2,AS,us3}).) Therefore, in general we cannot tell which of the two bounds in Proposition~\ref{oddk} is better. However, for large $k$ the second term is larger than the first due to \eqref{eqzahl}. Finally, we note that for $d\geq 4$ we have $C_k^d(n)=\Theta(n^{k+1})$. Indeed, we clearly have $C_k^d(n)=O(n^{k+1})$. To see that $C_k^d(n)=\Omega(n^{k+1})$, take two orthogonal circles of radius $1/\sqrt{2}$ centred at the origin and choose $n/2$ points on each of them. Then any sequence of $k+1$ points that alternate between the two circles forms a path in which all edges have unit length. The exact value of $u_d(n)$ for large $n$ and even $d\geq 4$ was determined by Brass \cite{brass} ($d=4)$ and Swanepoel \cite{Sw} ($d\geq 6$), by using stability results form extremal graph theory. \section{Preliminaries} We denote by $u_d(m,n)$ the maximum number of incidences between a set of $m$ points and $n$ spheres\footnote{circles, if $d=2$} of fixed radius in $\mathbb{R}^d$. In other words, $u_d(m,n)$ is the maximum number of red-blue pairs spanning a given distance in a set of $m$ red and $n$ blue points in $\mathbb{R}^d$. By the result of Spencer, Szemerédi and Trotter \cite{SST}, we have \begin{equation}\label{2drich} u_2(m,n)=O\left (m^{\frac{2}{3}}n^{\frac{2}{3}}+m+n\right). \end{equation} For given $r$ and $\delta$ we say that a point $p$ is \emph{$r$-rich} with respect to a set $P\subseteq \mathbb{R}^d$ and to a distance $\delta$, if the sphere of radius $\delta$ around $p$ contains at least $r$ points of $P$. If $P\subseteq \mathbb{R}^2$ and $|P|=n^x$, then \eqref{2drich} implies that the number of points that are $n^{\alpha}$-rich with respect to $P$ and to a given distance $\delta$ is \begin{equation}\label{2drichness} O\left (n^{2x-3\alpha}+n^{x-\alpha}\right ). \end{equation} The bound \begin{equation}\label{eq3d} u_3(m,n)=O\left (m^{\frac3{4}}n^{\frac34}+m+n\right) \end{equation} is due to Zahl \cite{Za2} and Kaplan, Matou\v sek, Safernov\'a, and Sharir \cite{KMSS}. It implies that for $P\subseteq \mathbb{R}^3$ with $|P|=n^x$ the number of points that are $n^{\alpha}$-rich with respect to $P$ and to a given distance $\delta$ is \begin{equation}\label{richness} O\left (n^{3x-4\alpha}+n^{x-\alpha}\right ). \end{equation} \section{Bounds in \texorpdfstring{$\bm{\mathbb{R}^2}$}{Lg}}\label{sec3} For a fixed $\bm\delta=(\delta_1,\dots,\delta_k)$ and $P_1\dots,P_{k+1}\subseteq \mathbb{R}^2$ we denote by $\mathcal{C}_k(P_1,\dots,P_{k+1})$ the family of $(k+1)$-tuples $(p_1,\dots,p_{k+1})$ with $p_i\in P_i$ for all $i\in[k+1]$, $\|p_i-p_{i+1}\|=\delta_i$ for all $i\in[k]$ and with $p_i\neq p_j$ for $i\neq j$. Let $C_k(P_1,\dots,P_{k+1})=|\mathcal{C}_k(P_1,\dots,P_{k+1})|$ and \[C_k(n_1,\dots,n_{k+1})=\max C_k(P_1,\dots,P_{k+1}),\] where the maximum is taken over all sets sets $P_1,\ldots, P_{k+1}$ subject to $|P_i|\le n_i$ for all $i\in [k+1]$. We have $C^2_k(n)\leq C_k(n,\dots,n)\leq C^2_k\left((k+1)n\right)$. Indeed, for the lower bound choose $P_i=P$ for every $1\leq i \leq k+1$, and for the upper bound note that $|P_1\cup \dots \cup P_{k+1}|\leq (k+1)n$. Since we are only interested in the order of magnitude of $C^2_k(n)$ for fixed $k$, we are going to bound $C_k(n,\dots,n)$ instead of $C^2_k(n)$. In Section~\ref{sec31}, we are going to prove the lower bounds from Theorem~\ref{main}. In Section~\ref{sec32}, we are going to prove an upper bound on $C_k(n,\dots,n)$, which is almost tight for $k\equiv 0,2$ (mod $3$). The case $k\equiv 1$ (mod $3$) is significantly more complicated. We will explain the case $k=4$ case separately in Section~\ref{sec33}, and then the general case in Section~\ref{sec34}. \subsection{Lower bounds}\label{sec31} For completeness, we present constructions for all congruence classes modulo $3$. For $k\equiv 0,2$ they were described in \cite{Shef}. \begin{prop}\label{proplow} For any fixed distance-vector $\bm\delta = (\delta_1,\ldots,\delta_k)$ of positive reals we have $$C_k(n,\ldots,n)= \begin{cases} \Omega(n^{\lfloor (k+1)/3\rfloor +1}),& \mbox{if } k \equiv 0,2({\rm mod\ } 3),\\ \Omega(n^{(k-1)/3}\cdot u_2(n)), & \mbox{if } k \equiv 1({\rm mod\ } 3). \end{cases}$$ \end{prop} In the proof of the proposition, we shall need the following proposition, using which was suggested to us by D\"om\"ot\"or P\'alv\"olgyi. \begin{prop}\label{Propprob}Fix $\varepsilon>0$. Then there exists $\gamma = \gamma(\varepsilon)>0$, such that for any $n$ there exist two sets $X_1,X_2$ of points on the plane, $|X_1|,|X_2|\le n$, such that: \begin{itemize} \item[(i)] The diameter of $X_2$ is at most $\varepsilon;$ \item[(ii)] The number of unit distances between $X_1$ and $X_2$ is at least $\gamma u_2(n,n).$ \end{itemize} \end{prop} \begin{proof} Take sets $Z_1,Z_2$ of $n$ points on the plane each, such that the number of unit distances between them is $u_2(n,n)$. Consider a bipartite graph $G = (Z_1\cup Z_2,E)$, where edges connect vertices at unit distance apart. Take an infinite grid formed by lines $y = 10i+\eta_1, x = 10j+\eta_2,$ where $i,j\in \mathbb Z$ and $\eta_1,\eta_2$ are chosen from $[0,10)$ uniformly at random. Then the expected number of edges of $G$ that are `cut' by a line in the grid is at most $\frac 12 u_2(n,n)$ and, therefore, putting $G' = G'(\eta_1,\eta_2) = (V, E')$ to be a subgraph of $G$ formed by all edges that are not `cut' by the grid (i.e., which endpoints lie inside the same square of the grid), there exist a choice of $\eta_1$, $\eta_2$ such that $|E'|\ge \frac 12 |E|$ and, moreover, no vertex of $G$ lies on a line of the grid. For a square $S$ of the grid, let $G'[S]$ be a subgraph of $G'$ induced on the vertices lying in $S$. Note that $G'$ is a disjoint union of $G'[S]$ over all possible $S$. Now translate all vertices in $G'[S]$ by an appropriate vector, so that a) all of them lie within a square $[0,11]\times [0,11]$ and b) no vertices from different translates coincide. Denote the new graph $G'' = (Y_1\cup Y_2,E'')$, where $Y_i$ is the union of translates of vertices from $Z_i$, $i=1,2$. It is clear that $|V(G'')| = |V(G')|$ and $|E''| = |E(G')|\ge \frac 12 u_2(n,n)$. Moreover, $V(G'')$ lies inside the square $[0,11]\times [0,11]$. Put $X_1:=Y_1$. Next, partition $[0,11]$ into $O(\varepsilon^2)$ squares of side at most $\varepsilon/2$ and choose a square $C$ out of them such that the number of edges from $G''$ emanating from vertices in $X_2:=Y_2\cap C$ is maximal. We claim that $X_1,X_2$ as above satisfy the conditions of the proposition. First, it is easy to see that the diameter of $X_2$ is at most $\varepsilon.$ Second, by the choice of $C,$ the number of edges from $G''$ between $X_1$ and $X_2$ is at least $\Omega(\varepsilon^2 |E''|) = \Omega(\varepsilon^2 u_2(n,n)),$ as desired. \end{proof} \begin{proof}[Proof of Proposition~\ref{proplow}] We prove the following slightly stronger statement by induction on $k$. For any fixed $\varepsilon>0$ the lower bound claimed in the proposition can be achieved on sets $P_1,\ldots, P_{k+1}$, such that the diameter of $P_{k+1}$ is at most $\varepsilon$ (with $\Omega$ depending on $\varepsilon$). First, we show the base of induction ($k=0,1,2$). Note that $C_0(n) = n$ and the set $P_1$ can be chosen to have diameter at most $\varepsilon.$ For $k=1$, we can use the construction from Proposition~\ref{proplow}. For $k=2$, let $P_2 = \{x\}$ for some point $x$, and let $P_1$, $P_3$ be disjoint sets of $n$ points on arcs of length $\varepsilon$ lying on the circles around $x$ of radii $\delta_1$ and $\delta_3$, respectively. Then we have that $C_2(P_1,P_2,P_3)=n^2$. Next, fix a vector $\bm\delta = (\delta_1,\ldots,\delta_{k+3}),$ put $\varepsilon = \min \{ \delta_{k+1}/3,\delta_{k+2}/3\}$ and let $\bm\delta' = (\delta_1,\ldots,\delta_k)$. We apply the inductive statement with $\varepsilon$ and $\bm \delta '$, and obtain the corresponding sets $P_1,\ldots, P_{k+1}.$ Put $P_{k+3} = \{x\},$ where $x$ is an arbitrary point on the plane that is at distance $\delta_{k+1}+\delta_{k+2}-2\varepsilon$ from some point $y$ in $P_{k+1}$. Put $P_{k+2}$ to be a set of $|P_{k+1}|$ points on the circle $C$ of radius $\delta_{k+2}$ around $x,$ such that for each point in $P_{k+1}$ there is at least one point in $P_{k+2}$ at distance $\delta_{k+1}$. For each point $z$ in $P_{k+1}$ it is possible to choose the corresponding point for $P_{k+2}$ since the minimum of the distance between a point on the circle $C$ and $z$ is at most $\|x-y\|-\delta_{k+2}+{\rm diam}(P_{k+1}) = \delta_{k+1}-\varepsilon<\delta_{k+1}$, while the maximum distance is at least $\|x-y\|+\delta_{k+2}-{\rm diam}(P_{k+1}) =\delta_{k+1}+2\delta_{k+2}-3\varepsilon>\delta_{k+1}$. (Note that for bounding both the maximum and the minimum distances we used triangle inequality and the fact that $P_{k+1}$ has diameter at most $\varepsilon.$) We use flexibility in the choice of $x$ to assure that all additional points are different from the points in $P_1,\ldots, P_{k+1}$. Finally, put $P_{k+4}$ to be a set of $n$ points on a sufficiently small arc on the circle of radius $\delta_{k+3}$ around $x.$ Since by construction every $k$-chain from $P_1\times \dots \times P_{k+1}$ can be extended to a $k+3$ chain in $P_1\times \dots \times P_{k+1}$ in at least $n$ different ways, we obtain that $C_2(P_1,\ldots, P_{k+4}) \ge n C_2(P_1,\ldots, P_{k+1})$. Further, $P_{k+4}$ can be chosen to satisfy any fixed requirement on the diameter. \end{proof} \subsection{Upper bound for the \texorpdfstring{$\bm{{k\equiv} 0,2}$}{Lg} (mod \texorpdfstring{$\bm{3}$}{Lg}) cases}\label{sec32} We fix $\bm\delta=(\delta_1,\dots,\delta_k)$ throughout the remainder of Section~\ref{sec3}. All logs are base $2$. \begin{thm}\label{firstbound} For any fixed integer $k\geq 0$ and $x,y\in [0,1]$, we have \begin{equation*}\label{eqmain} C_k(n^{x},n,\ldots,n,n^{y})=\tilde O\left ( n^{\frac {f(k)+x+y}{3}} \right ), \end{equation*} where $f(k) = k+2$ if $k\equiv 2$ \emph{(mod $3$)} and $f(k) = k+1$ otherwise. \end{thm} Theorem \ref{firstbound} implies the upper bounds in Theorem \ref{main} for $k\equiv 0,2$ (mod $3$) by taking $x=y=1$. It is easier, however, to prove this more general statement than the upper bounds in Theorem~\ref{main} directly. Having varied sizes of the first and the last groups of points allows for a seamless use of induction. \begin{proof}[Proof of Theorem \ref{firstbound}] The proof is by induction on $k$. Let us first verify the statement for $k\le 2.$ (Note that, for $k=0$, we should have $x=y$.) We have \begin{align} C_0(n^x)&\le n^x = O\left(n^{\frac {1+x+y}3}\right), \nonumber\\ \label{k=1} C_1(n^x,n^y)&\leq u_2(n^x,n^y)=O\left (n^{\frac{2}{3}(x+y)}+n^x+n^y\right )=O\left (n^{\frac{2+x+y}3}\right),\\ \label{ktwo} C_2(n^x,n,n^y)&\leq 2 n^xn^y= O\left (n^{\frac{4+x+y}{3}}\right ), \end{align} where \eqref{k=1} follows from \eqref{2drich} and \eqref{ktwo} follows from the fact that each pair $(p_1,p_3)$ can be extended to a $2$-chain $(p_1,p_2,p_3)$ in at most $2$ different ways. Next, let $k\geq 3$. Take $P_1,\dots,P_{k+1}\subseteq \mathbb{R}^2$ with $|P_1|=n^x$, $|P_{k+1}|=n^y$, and $|P_i|=n$ for $2\leq i \leq k$. Denote by $P_2^{\alpha}\subseteq P_2$ the set of those points in $P_2$ that are at least $n^{\alpha}$-rich but at most $2n^{\alpha}$-rich with respect to $P_1$ and $\delta_1$. Similarly, we denote by $P_k^{\beta}\subseteq P_k$ the set of those points in $P_k$ that are at least $n^{\beta}$-rich but at most $2n^{\beta}$-rich with respect to $P_{k+1}$ and $\delta_k$. Applying a standard dyadic decomposition argument twice implies that \[\mathcal{C}_k(P_1,P_2\dots,P_k,P_{k+1})= \bigcup_{\alpha,\beta} \mathcal{C}_k(P_1,P_2^{\alpha},P_3,\dots, P_{k-1},P_k^{\beta},P_{k+1}),\] where the union is taken over all $\alpha,\beta\in \left \{\frac{i}{\log n}: i=0,\ldots, \lceil\log n\rceil\right\}$. Since the cardinality of the latter set is at most $\log n+2$, it is sufficient to prove that for every $\alpha$ and $\beta$ we have \begin{equation}\label{alphabeta} C_k(P_1,P_2^{\alpha},P_3,\dots,P_{k-1},P_k^{\beta},P_{k+1})= \tilde O\left ( n^{\frac{f(k)+x+y}{3}}\right ). \end{equation} To prove this, we consider three cases. \bigskip {\bf Case 1: $\bm{\alpha\geq \frac{x}{2}}$.} By \eqref{2drichness} we have $|P_2^{\alpha}|=O(n^{x-\alpha})$. Therefore the number of pairs \mbox{$(p_1,p_2)\in P_1\times P_2^{\alpha}$} with $\|p_1-p_2\|=\delta_1$ is at most $O(n^x)$. Since every pair $(p_1,p_2)\in P_1\times P_2^{\alpha}$ and every $(k-3)$-chain $(p_4,\dots,p_{k+1})\in P_4\times\dots \times P_{k}^{\beta}\times P_{k+1}$ can be extended to a $k$-chain $(p_1,\dots,p_{k+1})\in P_1\times \dots\times P_{k+1}$ in at most two different ways, we obtain \[C_k(P_1,P_2^{\alpha},\dots,P_k^{\beta},P_{k+1})\leq 4O(n^x)C_{k-3}(P_4,\dots,P_k^{\beta},P_{k+1}).\] By induction we have \[C_{k-3}(P_4,\dots,P_k^{\beta},P_{k+1})= \tilde O\left ( n^{\frac{f(k-3)+1+y}{3}}\right ).\] These two displayed formulas and the fact that $f(k-3)=f(k)-3$ imply \eqref{alphabeta}. \bigskip {\bf Case 2: $\bm{\beta \geq \frac{y}{2}}$.} By symmetry, this case can be treated in the same way as Case 1. \bigskip {\bf Case 3: $\bm{\alpha \leq \frac{x}{2}}$ and $\bm{\beta\leq \frac{y}{2}}$.} By \eqref{2drichness} we have $|P_2^{\alpha}|= O\left (n^{2x-3\alpha}\right )$ and $|P_k^{\beta}|= O\left (n^{2y-3\beta}\right)$. The number of $(k-2)$-chains in $P_2^{\alpha}\times P_3\times\dots\times P_{k-1}\times P_k^{\beta}$ is $C_{k-2}(P_2^{\alpha},P_3,\dots,P_{k-1}, P_k^{\beta})$, and every $(k-2)$-chain $(p_2,\dots,p_k)\in P_2^{\alpha}\times P_3\times\dots\times P_{k-1}\times P_k^{\beta}$ can be extended at most $4n^{\alpha+\beta}$ ways to a $k$-chain in $P_1\times P_2^{\alpha}\times \dots\times P_k^{\beta}\times P_{k+1}$. Thus \[C_{k}(P_1,P_2^{\alpha},\dots,P_k^{\beta},P_{k+1})\leq 4n^{\alpha+\beta}C_{k-2}(P_2^{\alpha},\dots,P_k^{\beta}).\] By induction we have \[C_{k-2}(P_2^{\alpha},\dots,P_k^{\beta})= \tilde O\left (n^{\frac{f(k-2)+2x-3\alpha+2y-3\beta}{3}}\right ).\] For $k \equiv 0,2$ (mod $3$) we have $f(k)\ge f(k-2)+2,$ and thus \begin{multline*} C_{k}(P_1,P_2^{\alpha},\dots,P_k^{\beta},P_{k+1})= \tilde O\left(n^{\alpha+\beta}n^{\frac{f(k-2)+2x-3\alpha+2y-3\beta}{3}}\right)\\[4pt] = \tilde O\left(n^{\frac{f(k)-2+2x+2y}{3}}\right) = \tilde O\left(n^{\frac{f(k)+x+y}{3}}\right). \end{multline*} If $k\equiv 1$ (mod $3$) then $f(k)<f(k-2)+2$, and thus the argument above does not work. However, we then have $f(k)= f(k-1)+1$, and we can use the bound \[C_{k}(P_1,P_2^{\alpha},\dots,P_k^{\beta},P_{k+1})\leq 2n^{\alpha}C_{k-1}(P_2^{\alpha},P_3,\dots,P_{k+1}),\] obtained in an analogous way. This gives \[C_{k}(P_1,P_2^{\alpha},P_3,\dots,P_{k+1})= \tilde O\left(n^{\alpha}n^{\frac{f(k-1)+2x-3\alpha+y}{3}}\right) = \tilde O\left(n^{\frac{f(k)-1+2x+y}{3}}\right) = \tilde O\left(n^{\frac{f(k)+x+y}{3}}\right).\] \end{proof} \begin{remark} The proof above is not sufficient to obtain an almost sharp bound in the $k \equiv 1$ (mod $3$) case for two reasons. First, for these $k$ any analogue of Theorem \ref{firstbound} would involve taking maximums of two expressions, where one contains $u_2(n^x,n)$ and the other contains $u_2(n^y,n)$. However, due to our lack of good understanding of how $u_2(n^x,n)$ changes as $x$ is increasing, this is difficult to work with. Second, on a more technical side, while Case 1 and Case 2 in the above proof would go through with any reasonable inductive statement, Case 3 would fail. The main reason for this is that $C_k$ as a function of $k$ makes jumps at every third value of $k$, and remains essentially the same, or changes by $u_2(n,n)/n$ for the other values of $k$. Thus one would need to remove three vertices from the path to make the induction work. However, the path has only two ends, and removing vertices other than the endpoints turns out to be intractable. \end{remark} \subsection{Upper bound for \texorpdfstring{$\bm{k=4}$}{Lg}}\label{sec33} In this section we prove the upper bound in Theorem~\ref{main} for $k=4$. Let $P_1,\dots, P_5$ be five sets of $n$ points. We will show that $C_4(P_1,\dots,P_5)=\tilde O(u_2(n)n)$, which is slightly stronger than what is stated in Theorem \ref{main}. Instead of \eqref{2drichness} we need the following more general bound on the number of rich points. \begin{obs}[Richness bound]\label{richnessbound} Let $n^y$ be the maximum possible number of points that are $n^{\alpha}$-rich with respect to a set of $n^{x}$ points and some distance $\delta$. Then we have \begin{equation}\label{rich1} n^{y+\alpha}\leq u_2(n^x,n^y), \end{equation} or, equivalently \begin{equation*} n^{\alpha}\leq \frac{u_2(n^x,n^y)}{n^y}. \end{equation*} \end{obs} The proof of \eqref{rich1} follows immediately from the definition of $n^{\alpha}$-richness and $u_2(n^x,n^y)$. \medskip Let $\Lambda:=\big\{\frac i{\log n}: i = 0,\ldots, \lceil\log n\rceil \big\}^4$. For any $\bm{\alpha}=(\alpha_2,\alpha_3,\alpha_4,\alpha_5) \in \Lambda$ let $Q_1^{\bm{\alpha}} = P_1$ and for $i=2,\ldots, 5$ define recursively $Q_i^{\bm\alpha}$ to be the set of those points in $P_i$ that are at least $n^{\alpha_i}$-rich but at most $2n^{\alpha_i}$-rich with respect to $Q_{i-1}$ and $\delta_i$. It is not difficult to see that \[\mathcal C_4(P_1,\ldots, P_5)= \bigcup_{\bm\alpha\in \Lambda} \mathcal{C}_4\left(Q_1^{\bm\alpha},\ldots,Q_5^{\bm\alpha}\right).\] We have $|\Lambda| = \tilde O(1)$ and thus, in order to prove the theorem, it is sufficient to show that for every $\bm{\alpha}\in \Lambda$ we have \begin{equation*} C_4\left(Q_1^{\bm{\alpha}},\dots,Q_5^{\bm{\alpha}}\right)= O\left(n \cdot u_2(n,n)\right). \end{equation*} From now on, fix $\bm{\alpha}=(\alpha_2,\dots,\alpha_5)$, and denote $Q_i=Q_i^{\bm\alpha}$. Choose $x_i\in [0,1]$ so that $|Q_i|=n^{x_i}$. Then we have \begin{equation}\label{eq51} C_4(Q_1,\dots,Q_5)= O\left(n^{x_5+\alpha_5+\alpha_4+\alpha_3+\alpha_2}\right).\end{equation} Indeed, each chain $(p_1,\dots,p_5)$ with $p_i\in Q_i$ can be obtained in the following five steps. \vspace{0.2cm} \begin{itemize} \item \bf{Step 1: } \rm Pick $p_5\in Q_5$. \vspace{0.1cm} \item \bf{Step i ($2\le i\le 5$): } \rm Pick a point $p_{6-i}\in Q_{6-i}$ at distance $\delta_{6-i}$ from $p_{7-i}$. \end{itemize} \vspace{0.2cm} In the first step we have $n^{x_5}$ choices, and for $i\geq 2$ in the $i$-th step we have at most $2n^{\alpha_{6-i}}$ choices. Further, by Observation~\ref{richnessbound}, for each $i\geq 2$ we have \begin{equation}\label{eq52}n^{\alpha_i}\leq \frac{u_2(n^{x_{i-1}},n^{x_i})}{n^{x_i}}.\end{equation} Combining \eqref{eq51} and \eqref{eq52}, we obtain \begin{equation}\label{eq53} C_4(Q_1,\dots,Q_5)=O\left( u_2(n^{x_4},n^{x_5})\frac{u_2(n^{x_3},n^{x_4})}{n^{x_4}}\frac{u_2(n^{x_2},n^{x_3})}{n^{x_3}} \frac{u_2(n^{x_1},n^{x_2})}{n^{x_2}}\right). \end{equation} By \eqref{2drich} we have \[u_2(n^{x_{i-1}},n^{x_i})= O\left (\max \big\{n^{\frac{2}{3}(x_i+x_{i-1})},n^{x_i},n^{x_{i-1}}\big\}\right ).\] Note that the maximum is attained on the second (third) term iff $x_{i-1}\le \frac {x_i}2$ ($x_i\le \frac{x_{i-1}}2$). To bound $C_4(Q_1,\dots,Q_5)$ we consider several cases depending on which of these three terms the maximum above is attained on for different $i$. \bigskip \bf{Case 1: }\rm For all $2\leq i \leq 5$ we have $u_2(n^{x_{i-1}},n^{x_i})= O\left( n^{\frac{2}{3}(x_i+x_{i-1})}\right)$. Then \begin{equation*} \frac{u_2(n^{x_4},n^{x_5})u_2(n^{x_3},n^{x_4})u_2(n^{x_2},n^{x_3})}{n^{x_2+x_3+x_4}}=O\left( n^{\frac{2}{3}x_5+\frac{1}{3}x_4+\frac{1}{3}x_3-\frac{1}{3}x_2}\right) \end{equation*} and \begin{equation*} \frac{u_2(n^{x_3},n^{x_4})u_2(n^{x_2},n^{x_3})u_2(n^{x_1},n^{x_2})}{n^{x_2+x_3+x_4}}=O\left( n^{-\frac{1}{3}x_4+\frac{1}{3}x_3+\frac{1}{3}x_2+\frac{2}{3}x_1}\right). \end{equation*} Substituting each of these two displayed formulas into \eqref{eq53} and taking their product, we obtain \[C_4(Q_1,\dots,Q_5)^2 = O\left( u_2(n^{x_1},n^{x_2})u_2(n^{x_4},n^{x_5})\cdot n^{\frac{2}{3}x_1+\frac{2}{3}x_3+\frac{2}{3}x_5}\right)= O\left( u_2(n,n)^2\cdot n^{2}\right),\] which concludes the proof in this case. \bigskip \bf{Case 2: } \rm There is an $2\leq i \leq 5$ such that \begin{equation}\label{eq55} \min \{x_{i-1},x_i\}\le \frac 12\max\{x_{i-1},x_i\}\ \ \ \text{and thus}\ \ \ \ u_2(n^{x_{i-1}},n^{x_i})= O\left (\max \{n^{x_{i-1}},n^{x_{i}}\}\right).\end{equation} We distinguish three cases based on for which $i$ holds. \bigskip {\bf Case 2.1:} \eqref{eq55} holds for $i=2$ or $5$. In particular, this implies that $u_2(n^{x_{1}},n^{x_2})= O(n)$ or $u_2(n^{x_{4}},n^{x_5})= O(n)$. The following lemma finishes the proof in this case. \begin{lem}\label{smalln} Let $R_1,\ldots, R_5\subseteq \mathbb{R}^2$ such that $|R_i|\leq n$ for every $i\in [5]$. If $u_2(R_1,R_2)= O(n)$ or $u_2(R_4,R_5)= O(n)$ holds, then $C_4(R_1,\dots,R_5)= O\left(n\cdot u_2(n,n)\right)$. \end{lem} \begin{proof} We have \[C_4(R_1,\dots,R_5)\leq 2u_2(R_1,R_2)u_2(R_4,R_5)=O\left(n\cdot u_2(n,n)\right).\] Indeed, every $4$-tuple $(r_1,r_2,r_4,r_5)$ with $r_i\in R_i$ can be extended in at most two different ways to a $4$-chain $(r_1,\dots,r_5)\in R_1\times\dots\times R_5$. At the same time, the number of $4$-tuples with $\|r_1-r_2\|=\delta_1$, $\|r_4-r_5\|=\delta_4$ is at most $u_2(R_1,R_2)u_2(R_4,R_5)$. \end{proof} \bigskip \bf{Case 2.2: }\rm \eqref{eq55} holds for $i=4.$ Note that if $x_4\leq \frac{x_3}{2}\le \frac 12$, then $u_2(n^{x_5},n^{x_4})=O(n)$, and we can apply Lemma~\ref{smalln} to conclude the proof in this case. Thus we may assume that $x_3\le \frac{x_4}2$, and hence $u_2(n^{x_4},n^{x_3})=O(n^{x_4})$. This means that $n^{\alpha_4}=O(1)$ by Observation \ref{richnessbound}. Thus to finish the proof of this case, it is sufficient to prove the following claim. \begin{cla} Let $R_1,\dots,R_5\subseteq \mathbb{R}^2$ such that $|R_i|\leq n$ for all $i\in [5]$ and every point of $R_4$ is $O(1)$ rich with respect to $R_3$ and $\delta_3$. Then $C_4(R_1,\dots,R_5) = O\left(n\cdot u_2(n,n)\right)$. \end{cla} \begin{proof} Every $4$-chain $(r_1,\dots,r_5)$ can be obtained in the following steps. \vspace{0.2cm} \begin{itemize} \item Pick a pair $(r_4,r_5)\in R_4\times R_5$ with $\|r_4-r_5\|=\delta_4$. \vspace{0.1cm} \item Choose $r_3\in R_3$ at distance $\delta_3$ from $r_4$. \vspace{0.1cm} \item Pick a point $r_1\in R_1$. \vspace{0.1cm} \item Extend $(r_1,r_3,r_4,r_5)$ to a $4$-chain. \end{itemize} In the first step, we have at most $u_2(n,n)$ choices, in the third at most $n$ choices, and in the other two steps at most $O(1)$. \end{proof} {\bf Case 2.3 } \eqref{eq55} holds for $i=3$ {\it only}. Arguing as in Case 2.2, we may assume that $u_2(n^{x_3},n^{x_2})=O(n^{x_2})$. Then we have \begin{multline*} C_4(Q_1,\dots,Q_5)= O\left( u_2(n^{x_4},n^{x_5})\frac{u_2(n^{x_3},n^{x_4})}{n^{x_4}}\frac{u_2(n^{x_2},n^{x_3})}{n^{x_3}} \frac{u_2(n^{x_1},n^{x_2})}{n^{x_2}}\right)\\[10pt] = O\left(u_2(n^{x_1},n^{x_2})\cdot n^{\frac{2}{3}(x_4+x_5)+\frac{2}{3}(x_3+x_4)-x_4-x_3}\right)= O\left(u_2(n,n)\cdot n\right), \end{multline*} which finishes the proof. \subsection{Upper bound for the \texorpdfstring{$\bm{k\equiv 1}$}{Lg} (mod \texorpdfstring{$\bm{3}$)}{Lg} case}\label{sec34} We will prove the upper bound in Theorem \ref{main} for $k\equiv 1$ by induction. The $k=1$ case follows from the definition of $u_2(n,n)$, thus we may assume that $k\geq 4$. For the rest of the section fix ${\varepsilon}'>0$, and sets $P_1,\ldots, P_{k+1}\subseteq \mathbb{R}^2$ of size $n$, further let $\varepsilon=\frac{\varepsilon'}{4k}$. We are going to show that $C_k(P_1,\dots,P_{k+1})=O(n^{(k-1)/3+\varepsilon'}u_2(n))$. The first step of the proof is to find a certain covering of $P_1\times \dots \times P_{k+1}$, which resembles the one used for the $k=4$ case, although is more elaborate.\footnote{This covering brings in the $\varepsilon$-error term in the exponent, that we could avoid in the $k=4$ case.} (The goal of this covering is to make the corresponding graph between each of the two consecutive parts `regular in both directions' in a certain sense.) Let \[\Lambda=\Big\{i{\varepsilon}: i=0,\ldots, \Big\lfloor\frac 1{\varepsilon}\Big\rfloor\Big\}^{k+1}.\] We cover the product ${\bf P}=P_1\times\dots\times P_{k+1}$ by fine-grained classes $P_1^{\bm{\gamma}}\times \ldots\times P_{k+1}^{\bm{\gamma}}$ encoded by the sequence $\bm{\gamma} = (\bm{\gamma^1},\bm{\gamma^2},\ldots )$ of length at most $(k+1){\varepsilon}^{-1}+1$ with $\bm{\gamma^j}\in \Lambda$ for each $j=1,2,\dots$. One property that we shall have is \begin{equation*} P_1\times\dots \times P_{k+1} = \bigcup_{\bm{\gamma}}P_1^{\bm\gamma}\times \ldots\times P_{k+1}^{\bm\gamma}. \end{equation*} To find the covering, first we define a function $D$ that receives a parity digit $j\in \{0,1\}$, a product set ${\bf R}:= R_1\times\ldots \times R_{k+1}$ and an $\bm{\alpha}\in \Lambda$, and outputs a product set $D(j,\bm{R},\bm{\alpha})={\bf R(\bm \alpha)}= R_1(\bm\alpha)\times\ldots \times R_{k+1}(\bm\alpha)$. \medskip {\bf Definition of $\mathbf{D}$} \begin{itemize} \item If $j=1$ then let $R_1(\bm{\alpha}):=R_1$ and for $i=2,\ldots, k+1$ define $R_i(\bm{\alpha})$ iteratively to be the set of points in $R_i$ that are at least $n^{\alpha_i}$, but at most $n^{\alpha_i+{\varepsilon}}$-rich with respect to $R_{i-1}(\bm{\alpha})$ and $\delta_{i-1}.$ \item If $j=0$ then apply the same procedure, but in reverse order. That is, let $R_{k+1}(\bm{\alpha})=R_{k+1}$ and for $i=k,k-1,\dots,1$ define $R_{i}(\bm{\alpha})$ iteratively to be the set of points in $R_i$ that are at least $n^{\alpha_i}$ but at most $n^{\alpha_i+{\varepsilon}}$-rich with respect to $R_{i+1}(\bm{\alpha})$ and $\delta_i$. \end{itemize} \vspace{0.2cm} Note that \begin{equation}\label{equnion}{\bf R}= \bigcup_{\bm{\alpha}\in \Lambda} {\bf R}(\bm{\alpha}).\end{equation} For a sequence $\bm{\gamma} = (\bm{\gamma^1},\bm{\gamma^2}, \ldots)$ with $\bm{\gamma^j}\in \Lambda,$ we define ${\bf P}^{\bm\gamma}$ recursively as follows. Let ${\bf P}^{\emptyset} :={\bf P}$, and for each $j\geq 1$ let \[{\bf{P}}^{(\bm{\gamma^1},\ldots,\bm{\gamma^{j}})}=D(j \text{ (mod }2),{\bf{P}}^{(\bm{\gamma^1},\ldots,\bm{\gamma^{j-1}})},\bm{\gamma^j}).\] We say that a sequence $\bm{\gamma}$ is {\it stable at $j$} if \begin{equation*}\label{eqdecomp} \big|{\bf P}^{(\bm{\gamma^1},\ldots,\bm{\gamma^{j}})}\big|\ge \big|{\bf P}^{(\bm{\gamma^1},\ldots,\bm{\gamma^{j-1}})}\big|\cdot n^{-{\varepsilon}}.\end{equation*} Otherwise $\bm{\gamma}$ is \emph{unstable at $j$}. \begin{defi}\label{defups} Let $\Upsilon$ be the set of those sequences $\bm{\gamma}$ that are stable at their last coordinate, but are not stable for any previous coordinate, and for which ${\bf P}^{\bm\gamma}$ is non-empty. \end{defi} The set $\Upsilon$ has several useful properties, some of which are summarised in the following lemma. \begin{lem}\label{decompprop} 1. Any $\bm{\gamma}\in \Upsilon$ has length at most $(k+1) {\varepsilon}^{-1}+1$. \\ \vspace{0.1cm} 2. $|\Upsilon| = O_{{\varepsilon}}(1).$ \\ \vspace{0.1cm} 3. ${\bf P} = \bigcup_{\bm{\gamma}\in \Upsilon}{\bf P}^{\bm\gamma}.$\\ \end{lem} \begin{proof} \begin{enumerate} \item If $\bm{\gamma}$ is unstable at $j$ then \[|{\bf P}^{(\bm\gamma^1,\ldots,\bm\gamma^{j})}|\le |{\bf P}^{(\bm\gamma^1,\ldots,\bm\gamma^{j-1})}|\cdot n^{-{\varepsilon}}.\] Since $|{\bf P}| = n^{k+1}$ and $|{\bf P}^{\bm\gamma}|\ge 1,$ we conclude that $\bm{\bm\gamma}$ is unstable at at most $(k+1) {\varepsilon}^{-1}$ indices $j$. \vspace{0.1cm} \item It follows from part 1 by counting all possible sequences of length at most $(k+1) {\varepsilon}^{-1}+1$ of elements from the set $\Lambda$. (Note that $|\Lambda| = O_{{\varepsilon}}(1).$) \vspace{0.1cm} \item For a nonnegative integer $j$ let $\Lambda^{\le j}$ be the set of all sequences of length at most $j$ of elements from $\Lambda$. Let \[\Upsilon_j:= \left(\Upsilon\cap \Lambda^{\le j}\right)\cup\Psi_j, \textrm{ where } \ \Psi_j:=\big\{\bm{\gamma}\in \Lambda^j: \bm{\gamma} \text{ is not stable for any }\ell \le j\big\}.\] By part 1 of the lemma, $\Upsilon_j = \Upsilon$ for $j>(k+1){\varepsilon}^{-1}.$ We prove by induction on $j$ that ${\bf P}= \bigcup_{\bm{\gamma}\in \Upsilon_j}{\bf P}^{\bm\gamma}$. $\Upsilon_0$ consists of an empty sequence, thus the statement is clear for $j=0$. Next, assume that the statement holds for $j$. We have \[{\bf P}=\bigcup_{\bm{\gamma}\in \Upsilon_j}{\bf P}^{\bm\gamma}=\bigcup_{\bm{\gamma}\in \Lambda^{\leq j}}{\bf P}^{\bm\gamma}\cup\bigcup_{\bm{\gamma}\in \Psi_j}{\bf P}^{\bm\gamma}.\] By $\eqref{equnion}$ we have that ${\bf P}^{\bm\gamma} = \bigcup_{\bm{\gamma'}}{\bf P}^{\bm\gamma'}$ holds for any $\bm\gamma\in \Psi_j$, where the union is taken over the sequences from $\Lambda^{j+1}$ that coincide with $\bm{\gamma}$ on the first $j$ entries. This, together with $\bm\gamma'\in \left(\Upsilon\cap \Lambda^{j+1}\right)\cup \Psi_{j+1}$ when ${\bf P}^{\bm\gamma'}$ is nonempty finishes the proof. \end{enumerate} \end{proof} Parts 2 and 3 of Lemma~\ref{decompprop} imply that in order to complete the proof of the \mbox{$k\equiv 1$ (mod $3$)} case, it is sufficient to show that for any $\bm{\gamma}\in \Upsilon$ we have \begin{equation}\label{eq61} C_k(P_1^{\bm\gamma},\ldots, P_{k+1}^{\bm\gamma})= O\left(u_2(n)\cdot n^{\frac{k-1}3+4k{\varepsilon}}\right). \end{equation} From now on fix $\bm{\gamma}\in \Upsilon$. For each $i=1,\ldots, k+1$ let $R_i:=P_i^{\bm\gamma}$ and $Q_i:= P_i^{\bm\gamma'}$, where $\bm{\gamma'}$ is obtained from $\bm{\gamma}$ by removing the last element of the sequence. Without loss of generality, assume that the length $\ell$ of $\bm{\gamma}$ is even. For each $i=1,\ldots, k+1,$ choose $x_i,y_i$ such that \[|Q_i| = n^{x_i}, \ \ \ \ |R_i| = n^{y_i}.\] Let $\alpha_i:= \bm\gamma^{\ell-1}_i$ and $\beta_i:= \bm\gamma^{\ell}_i$. By the definition of $\bf{P}^{\bm{\gamma}}$ we have that each point in $Q_i$ is at least $n^{\alpha_i}$-rich but at most $n^{\alpha_i+\varepsilon}$-rich with respect to $Q_{i-1}$ and $\delta_{i-1}$, and each point in $R_i$ is at least $n^{\beta_i}$-rich but at most $n^{\beta_i+\varepsilon}$-rich with respect to $R_{i+1}$ and $\delta_i$. By Observation~\ref{richnessbound}, we have \begin{equation}\label{eqkey} n^{\alpha_i}\leq \frac{u_2(n^{x_{i-1}},n^{x_i})}{n^{x_i}}\ \ \ \ \text{and} \ \ \ \ n^{\beta_i}\leq \frac{u_2(n^{y_{i}},n^{y_{i+1}})}{n^{y_i}}\le \frac{u_2(n^{x_{i}},n^{x_{i+1}})}{n^{x_i-{\varepsilon}}}. \end{equation} The last inequality follows from two facts: first $u_2(n^{y_{i}},n^{y_{i+1}})\le u_2(n^{x_{i}},n^{x_{i+1}})$ and, second, since $\bm\gamma$ is stable at its last coordinate\footnote{This is the only place where we use the stability of $\bm\gamma$ directly.}, we have $n^{y_i} = |R_i|\ge |Q_i|\cdot n^{-{\varepsilon}} = n^{x_i-{\varepsilon}}.$ In the same fashion as in the beginning of Section~\ref{sec33}, we can show that \begin{align*} C_k(R_1,\dots,R_{k+1})\leq& n^{y_{1}}n^{\beta_{1}+\dots+\beta_k+k\varepsilon}, \textrm{ and }\\[5pt] C_k(R_1,\dots,R_{k+1})\leq C_k(Q_1,\dots,Q_{k+1})\leq& n^{x_{k+1}}n^{\alpha_{k+1}+\alpha_{k}+\dots+\alpha_2+k\varepsilon}. \end{align*} Combining the first of these displayed inequalities with \eqref{eqkey}, we have \begin{equation*} C_{k}(R_1,\dots,R_{k+1}) \leq u_2(n^{x_1},n^{x_{2}})\prod_{2\leq i \leq k}\frac{u_2\left (n^{x_{i}},n^{x_{i+1}}\right)}{n^{x_i}}n^{2k\varepsilon} . \end{equation*} Recall that \begin{equation}\label{eq66} u_2(n^{x_{i}},n^{x_{i+1}})=O\left (\max \{n^{\frac{2}{3}(x_i+x_{i+1})},n^{x_i},n^{x_{i+1}}\}\right ). \end{equation} To bound $C_k(R_1,\dots,R_{k+1})$, we consider several cases based on which of these three terms can be used to bound $u_2(n^{x_{i}},n^{x_{i+1}})$ for different values of $i$. \bigskip {\bf Case 1: } Either $u_2(n^{x_1},n^{x_2})= O(n)$ or $u_2(n^{x_k},n^{x_{k+1}})= O(n)$ holds. As in the proof of Lemma~\ref{smalln}, we have \begin{multline*} C_k(R_1,\dots,R_{k+1})\\[5pt] \leq \min\big\{ 2u_2(n^{y_1},n^{y_{2}})C_{k-3}(R_4,\dots,R_{k+1}),2u_2(n^{y_k},n^{y_{k+1}})C_{k-3}(R_1,\dots,R_{k-2})\big \}. \end{multline*} By induction we obtain $C_{k-3}(R_4,\dots,R_{k+1}),C_{k-3}(R_1,\dots,R_{k-2})= O\left(n^{\frac{k-4}{3}+\varepsilon}\cdot u_2(n)\right )$. Together with the assumption of Case 1, and the fact that $u_2(n^{y_1},n^{y_{2}})\leq u_2(n^{x_1},n^{x_{2}})$ and \mbox{$u_2(n^{y_k},n^{y_{k+1}})\leq u_2(n^{x_k},n^{x_{k+1}})$}, this implies \eqref{eq61} and finishes the proof.\\ {\bf Case 2: } For some $i=1,\ldots, (k-1)/3,$ one of the following holds: \vspace{0.3cm} \begin{itemize} \item $u_2(n^{x_{3i+1}},n^{x_{3i+2}})= O(\max\{n^{x_{3i+1}},n^{x_{3i+2}}\})$; \vspace{0.1cm} \item $u_2(n^{x_{3i-1}},n^{x_{3i}})= O(n^{x_{3i-1}})$; \vspace{0.1cm} \item $u_2(n^{x_{3i}},n^{x_{3i+1}})= O(n^{x_{3i+1}})$. \end{itemize} We will show how to conclude in the first case. The other cases are very similar and we omit the details of their proofs. If $u_2(n^{x_{3i+1}},n^{x_{3i+2}})= O(n^{x_{3i+2}})$ then $n^{\alpha_{3i+2}}=O(1)$ by \eqref{eqkey}. Every chain $(r_1,\dots,r_{k+1})\in \mathcal C_k(Q_1,\ldots, Q_{k+1})$ can be obtained as follows. \vspace{0.3cm} \begin{enumerate} \item Pick a $(3i-2)$-chain $(r_1,\dots,r_{3i-1})$ with $r_j\in Q_j$ for every $j$. \vspace{0.2cm} \item Pick a $(k-3i-1)$-chain $(r_{3i+2},r_{3i+3},\dots,r_{k+1})$ with $r_j\in Q_j$ for every $j$. \vspace{0.2cm} \item Extend $(r_{3i+2},r_{3i+3},\dots,r_{k+1})$ to a $(k-3i-2)$ chain $(r_{3i+1},r_{3i+2},\dots,r_{k+1})$. \vspace{0.2cm} \item Connect $(r_1,\dots,r_{3i-1})$ and $(r_{3i+1},r_{3i+2},\dots,r_{k+1})$ to obtain a $k$-chain. \end{enumerate} In the first step, we have $O\left(n^{\frac{3i-3}3+\varepsilon}\cdot u_2(n)\right)$ choices by induction on $k$. In the second step, we have $\tilde O\left(n^{\frac{k-3i+2}{3}}\right)$ choices by the $k\equiv 0$ (mod $3$) case of Theorem~\ref{main}. In the third step, we have at most $n^{\alpha_{3i+2}+\varepsilon}= O(n^{\varepsilon})$ choices. Finally, in the fourth step we have at most $2$ choices. Thus the number of $k$-chains is at most \[O\left(n^{\frac{3i-3}{3}+\varepsilon}\cdot u_2(n)\right)\cdot \tilde O\left (n^{\frac{k-3i+2}{3}}\right )\cdot O\left(n^{\varepsilon}\right)\cdot 2=O\left(n^{\frac{k-1}{3}+3\varepsilon}\cdot u_2(n)\right),\] finishing the proof of the first case. \smallskip If $u_2(n^{x_{3i+1}},n^{x_{3i+2}})= O(n^{x_{3i+1}})$ then $n^{\beta_{3i+1}}= O( n^{\varepsilon})$ by \eqref{eqkey}.\footnote{This is the key application of \eqref{eqkey}, and the reason why we needed a decomposition with regularity in both directions between the consecutive parts.} We proceed similarly in this case, but we count the $k$-chains now in $R_1\times\ldots\times R_{k+1}$ instead in $Q_1\times \ldots \times Q_{k+1}$ (and get an extra factor of $n^{{\varepsilon}}$ in the bound). In all cases, we obtain \eqref{eq61}. \bigskip {\bf Case 3: } Neither the assumptions of Case 1 nor that of Case 2 hold. We define four sets $S'$, $S'_+$, $S'_{++}$, and $S'_-$ of indices in $\{2,\ldots, k\}$ as follows. Let \begin{flalign*} S'& := \setbuilder{i}{u_2(n^{x_i},n^{x_{i-1}})= O(n^{\frac{2}{3}(x_i+x_{i-1})}) \textrm{ and } u_2(n^{x_{i+1}},n^{x_{i}})= O(n^{\frac{2}{3}(x_{i+1}+x_{i})}) }, \\[7pt] S'_+& := \Big\{i : u_2(n^{x_i},n^{x_{i-1}})= O(n^{\frac{2}{3}(x_i+x_{i-1})}) \textrm{ and } u_2(n^{x_{i+1}},n^{x_{i}})= O(n^{x_{i}}), \textrm{ or } \\ & \phantomrel{=}{} \phantomrel{=}{} \phantomrel{=}{} \phantomrel{=}{} u_2(n^{x_i},n^{x_{i-1}})= O(n^{x_i}) \textrm{ and } u_2(n^{x_{i+1}},n^{x_{i}})= O(n^{\frac{2}{3}(x_{i+1}+x_{i})}) \Big \}, \\[7pt] S'_{++}& :=\Big\{i : u_2(n^{x_i},n^{x_{i-1}})= O(n^{x_i}) \textrm{ and } u_2(n^{x_{i+1}},n^{x_{i}})= O(n^{x_{i}}) \Big \}, \textrm{ and} \\[7pt] S'_-& :=\Big\{i : u_2(n^{x_i},n^{x_{i-1}})= O(n^{\frac{2}{3}(x_i+x_{i-1})}) \textrm{ and } u_2(n^{x_{i+1}},n^{x_{i}})= O(n^{x_{i+1}}), \textrm{ or } \\ & \phantomrel{=}{} \phantomrel{=}{} \phantomrel{=}{} \phantomrel{=}{} u_2(n^{x_i},n^{x_{i-1}})= O(n^{x_{i-1}}) \textrm{ and } u_2(n^{x_{i+1}},n^{x_{i}})= O(n^{\frac{2}{3}(x_{i+1}+x_{i})}) \Big \}. \end{flalign*} Since the conditions of Case 2 are not satisfied, we have \[\{2,\dots,k\}\subseteq S'\cup S'_+\cup S'_{++}\cup S'_-.\] Indeed, for each $i\in\{2,\ldots, k\},$ there are $9$ possible pairs of maxima in \eqref{eq66} with $i,i+1.$ The four sets above encompass $6$ possibilities. In total, there are $4$ possible pairs of maxima with only the two last terms from \eqref{eq66} used. For $i \equiv 1,2$ (mod $3$), any of those $4$ are excluded due to the first condition in Case 2 (in fact, then $i\in S'\cup S'_-$). If $i \equiv 0$ (mod $3$), then the second and the third condition in Case 2 rule out all possibilities but the one defining $S'_{++}$. From these it directly follows that if $i\in S'_{++}$, then $i-1,i+1\in S'_-$, while if $i\in S'_+$ then one of $i-1,i+1$ is in $S'_-$. (Recall that $i\in S'_+\cup S'_{++}$ only if $i\equiv 0$ (mod $3$).) These together imply \begin{equation}\label{eq67} |S'_+|+2|S'_{++}|\le |S'_-|. \end{equation} We partition $\{2,\dots,k\}$ using these sets as follows: let $S_- = S'_-, S = S'\setminus S'_-, S_+ = S'_+\setminus (S'_-\cup S')$ and $S_{++}=\{2,\dots,k\}\setminus S'_-\cup S'\cup S'_{+}$. Note that the analogue of \eqref{eq67} holds for the new sets. That is, we have \begin{equation*}\label{eq68} |S_+|+2|S_{++}|\le |S_-|. \end{equation*} Recall that \begin{equation}\label{24} C_{k}(R_1,\dots,R_{k+1}) \leq u_2(n^{x_1},n^{x_{2}})\prod_{2\leq i \leq k}\frac{u_2\left (n^{x_{i}},n^{x_{i+1}}\right)}{n^{x_i}}n^{2k\varepsilon}. \end{equation} Since the assumptions of Case 1 and 2 do not hold, we have $2,k\in S$. Indeed, $2,k\ne 0$ \mbox{(mod $3$)} and thus $2,k\notin S_+,S_{++}$. Further, if say $k\in S_-=S'_-$ then by the definition of $S'_-$ we either have $u_2(n^{x_{k+1}},n^{x_k}) = O(n)$, or $u_2(n^{x_k},n^{x_{k-1}})=O(n^{x_{k-1}})$. The first case cannot hold since the assumption of Case 1 does not hold. Further, the second case cannot hold either, since it would imply $x_k\le \frac {x_{k-1}}2\le \frac 12$, meaning $u_2(n^{x_{k+1}},n^{x_k}) = O(n)$. Using $2,k\in S$ and expanding \eqref{24}, we obtain {\small \begin{equation}\label{first} C_{k}(R_1,\dots,R_{k+1})\leq n^{2k\varepsilon}u_2(n^{x_1},n^{x_2})n^{-\frac{1}{3}x_2}n^{\frac{2}{3}x_{k+1}}\prod_{\substack{i\in S,\\i\neq 2}}n^{\frac{1}{3}x_i}\prod_{i\in S_+}n^{\frac{2}{3}x_i}\prod_{i\in S_{++}}n^{x_i}\prod_{i\in S_-}n^{-\frac{1}{3}x_i}, \end{equation}} and {\small \begin{equation}\label{second} C_{k}(R_1,\dots,R_{k+1})\leq n^{2k\varepsilon}u_2(n^{x_k},n^{x_{k+1}})n^{-\frac{1}{3}x_k}n^{\frac{2}{3}x_1}\prod_{\substack{i\in S,\\ i\neq k}}n^{\frac{1}{3}x_i}\prod_{i\in S_+}n^{\frac{2}{3}x_i}\prod_{i\in S_{++}}n^{x_i}\prod_{i\in S_-}n^{-\frac{1}{3}x_i}. \end{equation}} Taking the product of \eqref{first} and \eqref{second} we obtain \begin{multline*} C_{k}(R_1,\dots,R_{k+1})^2 \leq \\ n^{4k\varepsilon}\cdot u_2(n^{x_1},n^{x_2})u_2(n^{x_k},n^{x_{k+1}})n^{\frac{2}{3}(x_1+x_{k+1})}\left (\prod_{ \substack{i\in S,\\ i\neq 2,k}} n^{\frac{1}{3}x_i}\prod_{i\in S_+}n^{\frac{2}{3}x_i}\prod_{i\in S_{++}}n^{x_i}\prod_{i\in S_-}n^{-\frac{1}{3}x_i} \right)^2\\[10pt] \leq n^{4k\varepsilon}\cdot u_2(n,n)^2\cdot n^{2\left(\frac{2}{3}+ \frac{1}{3}|S\setminus\{2,k\}|+\frac{2}{3}|S_+|+|S_{++}|\right ) }= u_2(n,n)^2\cdot n^{\frac{2(k-1)}{3}+4k{\varepsilon}}. \end{multline*} The last equality follows from $|S_+|+2|S_{++}|\le |S_-|$, which is equivalent to $\frac 23|S_+|+|S_{++}|\le \frac 13(|S_+|+|S_{++}|+|S_-|)$, and from the fact that $S$, $S_+$,$S_{++}$, and $S_-$ partition $\{2,\dots,k\}$. This finishes the proof. \section{Bounds in \texorpdfstring{\bm{$\mathbb{R}^3$}}{Lg}} Similarly as in the planar case, for a fixed $\bm\delta=(\delta_1,\dots,\delta_k)$ and $P_1\dots,P_{k+1}\subseteq \mathbb{R}^3$ we denote by $\mathcal{C}_k^3(P_1,\dots,P_k)$ the family of $(k+1)$-tuples $(p_1,\dots,p_{k+1})$ with $p_i\in P_i$ for all $i\in[k+1]$ and with $\|p_i-p_{i+1}\|=\delta_i$ for all $i\in[k]$. Let $C_k^3(P_1,\dots,P_{k+1})=|\mathcal{C}_k^{3}(P_1,\dots,P_{k+1})|$ and \[C^3_k(n_1,\dots,n_{k+1})=\max C_k^{3}(P_1,\dots,P_{k+1}),\] where the maximum is taken over all choices of $P_1,\ldots, P_{k+1}$ subject to $|P_i|\le n_i$ for all $i\in [k+1]$. Similarly to the planar case it follows that $C^3_k(n)\leq C^3_k(n,\dots,n)\leq C^3_k\left((k+1)n\right)$. Since we are only interested in the order of magnitude of $C^3_k(n)$ for fixed $k$, sometimes we are going to work with $C^3_k(n,\dots,n)$ instead of $C^3_k(n)$. \subsection{Lower bounds} For completeness, we recall the constructions from \cite{Shef} for even $k\geq 2$. Let $\bm{\delta}=(\delta_1,\dots,\delta_k)$ be any given sequence. For every even $2\leq i\leq k$, let $P_i=\{p_i\}$ be a single point such that the sphere of radius $\delta_i$ centred at $p_i$ and the sphere of radius $\delta_{i+1}$ centred at $p_{i+2}$ intersect in a circle. Further, let $P_1$ be a set of $n$ points contained in the sphere of radius $\delta_1$ centred at and $p_2$, and $P_{k+1}$ be a set of $n$ points contained in the sphere of radius $\delta_k$ centred at and $p_2$. Finally, for every odd $3\leq i \leq k-1$, let $P_i$ be a set of $n$ points contained in the intersection of the sphere of radius $\delta_{i-1}$ centred at $p_{i-1}$ and of the sphere of radius $\delta_{i}$ centred at $p_{i+1}$. Then $P_1\times\dots\times P_{k+1}$ contains $n^{\frac{k}{2}+1}$ many $k$-chains, since every element of $P_1\times\dots\times P_{k+1}$ is a $k$-chain, and $|P_1\times\dots\times P_{k+1}|=n^{\frac{k}{2}+1}$. Next, we prove the lower bounds for odd $k\geq 3$ and $\bm\delta=(1,\dots,1)$ given in Proposition \ref{oddk}. \begin{proof}[Proof of Proposition \ref{oddk}] First we show that $C_k^3(n)=\Omega\left (\frac{u_3(n)^k}{n^{k-1}}\right )$. Take a set $P'\subset {\mathbb R}^3$ of size $n$ that contains $u_3(n)$ point pairs at unit distance apart. It is a standard exercise in graph theory to show that since $u_3(n)$ is superlinear, there is $P\subset P'$ such that $\frac{n}{2}\leq |P|\leq n$ and for every $p\in P$ there are at least $\frac{u_3(n)}{4n}$ points $p'\in P$ at distance $1$ from $p$. Then $P$ contains $\Omega\left (\frac{u_3(n)^k}{n^{k-1}}\right )$ many $k$-chains with $\bm\delta=(1,\dots,1)$. To prove $C_k^3(n)=\Omega\left (us_3(n)n^{(k-1)/2}\right)$, we modify and extend the construction used for $k-1$ as follows. Let $P_1,\dots,P_{k-1}$ be as in the construction for $(k-1)$-chains with $\bm\delta=(1,\dots,1)$ (from the even case). Further, let $P_{k}$ be a set of $n$ points on the unit sphere around $p_{k-1}$, and $P_{k+1}$ be a set of $n$ points such that $u_3(P_k,P_{k+1})=us_3(n)$. Since every $(p_1,\dots,p_{k+1})\in P_1\times \dots \times P_{k+1}$ with $\|p_k-p_{k+1}\|=1$ is a $k$-chain, we obtain that $P_1\times\dots\times P_{k+1}$ contains $\Omega\left (us_3(n)n^{(k-1)/2}\right)$ many $k$-chains with $\bm\delta=(1,\dots,1)$. \end{proof} \subsection{Upper bound} We again fix $\bm\delta=(\delta_1,\dots,\delta_k)$ throughout the section. The following result with $x=1$ implies the upper bound in Theorem~\ref{3d}. \begin{thm}\label{3dbound} For any fixed integer $k\geq 0$ and $x\in [0,1]$, we have \begin{equation*} C^3_k(n^{x},n,\ldots,n)=\tilde O\left ( n^{\frac{k+1+x}{2}}\right ). \end{equation*} \end{thm} \begin{proof} The proof is by induction on $k$. For $k=0$ the bound is trivial, and for $k=1$ it follows from \eqref{eq3d}. For $k\geq 2$ let $P_1,\dots,P_{k+1}\subseteq \mathbb{R}^3$ be sets of points satisfying $|P_1|=n^x$, and $|P_i|=n$ for $2\leq n \leq k+1$. Denote by $P_2^{\alpha}\subseteq P_2$ the set of those points in $P_2$ that are at least $n^{\alpha}$-rich but at most $2n^{\alpha}$-rich with respect to $P_1$ and $\delta_1$. It is not hard to see that \[\mathcal{C}^3_k(P_1,P_2\dots,P_{k+1})\subseteq \bigcup_{\alpha\in \Lambda} \mathcal{C}^3_k(P_1,P_2^{\alpha},P_3,\ldots,P_{k+1}),\] where $\Lambda:= \{\frac{i}{\log n}: i= 0,1,\ldots, \lfloor \log n\rfloor\}$. Since $|\Lambda| = \tilde O(1)$, it is sufficient to prove that, for every $\alpha\in \Lambda,$ we have \begin{equation*}\label{alpha3} C^3_k(P_1,P_2^{\alpha},P_3,\ldots, P_{k+1})= \tilde O\left ( n^{\frac{k+1+x}2}\right ). \end{equation*} Assume that $|P^{\alpha}_2| = n^y.$ The number of $(k-1)$-chains in $P_2^{\alpha}\times P_3\times \dots\times P_{k+1}$ is at most $C^3_{k-1}(n^{y},n,\dots,n)$, and each of them may be extended in $2n^{\alpha}$ ways. By induction, we get \[C^3_k(P_1,P_2^{\alpha},P_3,\ldots, P_{k+1})= \tilde O\left(n^{\alpha}\cdot n^{\frac{k+y}{2}}\right),\] and we are done as long as \begin{equation}\label{eqtocheck}2\alpha+k+y\le k+1+x.\end{equation} To show this, we need to consider several cases depending on the value of $\alpha.$ Note that $\alpha\le x$. \vspace{0.1cm} \begin{itemize} \item If $\alpha\geq \frac{2x}{3}$, then by \eqref{richness} we have $y\leq x-\alpha$, and the LHS of \eqref{eqtocheck} is at most $\alpha+k+x\le 1+k+x$. \vspace{0.1cm} \item If $\frac x2 \le \alpha\le \frac{2x}{3}$ then by \eqref{richness} we have $y\leq 3x-4\alpha.$ The LHS of \eqref{eqtocheck} is at most $k+3x-2\alpha\le k+2x\le k+1+x$. \vspace{0.1cm} \item If $\alpha\le \frac x2$ then we use a trivial bound $y\leq 1$. The LHS of \eqref{eqtocheck} is at most $2\alpha+k+1\le x+k+1.$ \end{itemize} \end{proof} \section{Concluding remarks} If $\bf{\delta}=(1,\dots,1)$, then the results of this paper are about the finding the maximum number of $k$-long paths that can be determined by a unit-distance graph on $n$ vertices. As a generalisation, one can study the maximum number of isomorphic copies of a given tree in unit distance graphs. More generally, we propose the following problem. Let $T=(V,E)$ be a tree with $V=\{v_1,\dots,v_{k+1}\}$ and $E=\{(v_{i_1},v_{j_1}),\dots,(v_{i_k},v_{j_k})\}$. For a fixed sequence $\bm{\delta}=\{\delta_1,\dots,\delta_k\},$ a $(k+1)$-tuple of distinct points $(p_1,\dots,p_{k+1})$ in $\mathbb{R}^d$ is a \emph{$T$-tree}, if $\|p_{i_\ell}-p_{j_\ell}\|=\delta_{\ell}$ for every $\ell=1,\ldots, k$. What is the maximum possible number $C_T^d(n)$ of $T$-trees in a set of $n$ points in ${\mathbb R}^d$? For $d=2$ we write $C_T^2(n)=C_T(n)$. Note that, as in the case of chains, $C_T^d(n) = \Omega\big(n^{|V(T)|}\big)=\Omega(n^{k+1})$ for $d\ge 4$. However, for $d=2$, $C_T(n)$ depends on $T$, not only on the number of vertices of $T$. Indeed, we saw that if $T$ is a path, then $C_T(n)$ is roughly $n^{k/3}$. At the same time, if $T$ is a star with $k$ leaves then $C_T(n)=\Theta(n^k)$. To see this, fix the centre of the star and distribute the remaining points equally on concentric circles of radii $\delta_1,\ldots, \delta_k$ around the centre of the star. One can similarly find examples in for $d=3$ where $C_T^3(n)$ depends on the tree itself. For some trees determining $C_T(n)$ trivially reduces to determining $C_k(n)$ for some $k$. However, for many other trees the problem seems challenging, and new ideas are needed to tackle it. Subdivisions of stars show that it in some cases it might not be possible to determine $C_T(n)$ without knowing $u_2(n)$, even in terms of $u_2(n)$. To see this, let $T_{\ell,3}$ be a star-shaped tree on $3\ell+1$ vertices, with one (central) vertex of degree $\ell$, and $\ell$ paths on $3$ vertices joined to the central vertex. (This tree for $\ell =3$ is depicted on Figure \ref{fig2}, right.) Generalising the lower bound constructions we used for the chains in two different ways, we can obtain $C_{T_{\ell,3}}(n)=\Omega(u_2(n)^{\ell})$ (by fixing the central vertex of the tree) and $C_{T_{\ell,3}}(n)=\Omega(n^{\ell+1})$ (by fixing all vertices that are neighbours of the leaves). This is illustrated on Figure \ref{fig1} for $\ell=3$. It would be interesting to prove that $C_{T_{\ell,3}(n)}$ is the maximum of these two lower bounds. \begin{figure}[h] \centering {\includegraphics[width=10cm]{TwoTree.pdf}} \caption{Star-shaped trees.}\label{fig2} \end{figure} \begin{figure}[h] \centering {\includegraphics[width=16.5cm]{2new.pdf}} \caption{Examples providing lower bounds for the number of copies of $T_{\ell,3}$ with $\ell=3$. Left: $\Omega(n^{\ell+1})$ copies of $T_{\ell,3}$. Right: $\Omega(u_2^{\ell})$ copies of $T_{\ell,3}$. Vertices of red colour indicate parts (corresponding to the vertices of the tree and) that consist of a single vertex.}\label{fig1} \end{figure} \begin{pro}\label{pro1}Is it true that $C_{T_{\ell,3}}(n)=\Theta(\max\{n^{\ell+1},u_2(n)^{\ell}\})$? \end{pro} Motivated by the constructions described before, we propose the following more general question. \begin{pro}\label{treeconj}Is it true that for every tree $T$ there are integers $m,\ell$ such that $C_T(n)=\Theta(n^mu_2(n)^\ell)$? \end{pro} The smallest tree $T$, for which we cannot determine the order of magnitude of $C_T(n)$, is a star-shaped tree on $7$ vertices with one central vertex of degree $3$ and three and three paths on $2$ vertices joined to the central vertex (see the left tree of Figure \ref{fig2}). \begin{pro}\label{small} Is it true that for the star-shaped tree $T$ on $7$ vertices described above we have $C_T(n)=\Theta(n^3)$? \end{pro} Note that it may be easier (and also very interesting) to obtain upper bounds in Problems \ref{pro1}-\ref{small} with a poly-logarithmic or $n^{\varepsilon}$ error term, as we did in Theorem \ref{main}. In a follow-up paper, we find almost sharp bounds for $C_T(n)$ for some `non-trivial' trees (i.e., those that cannot be reduced to the chains case). Further, we will describe a general lower bound construction that motivates Problem \ref{treeconj}. We will also provide some partial results for Problem \ref{small} and connect it to an interesting incidence problem. Finally, it would be interesting to decide if the lower bounds in Proposition \ref{oddk} are sharp for $\bm\delta=(1,\dots,1)$. The first open case is $k=3$. \begin{pro} Is it true that $C_3^3(n)=\Theta\left (\max\left\{ \frac{u_3(n)^3}{n^2},us_3(n)n\right \}\right )$ for $\bm\delta=(1,\dots,1)$? \end{pro} \subsection*{Acknowledgements.} We thank Konrad Swanepoel and Peter Allen for helpful comments on the manuscript. We also thank Dömötör Pálvölgyi for suggesting Proposition \ref{Propprob}. The authors acknowledge the financial support from the Russian Government in the framework of MegaGrant no 075-15-2019-1926. \bibliographystyle{amsplain}
{ "timestamp": "2020-10-19T02:13:27", "yymm": "1912", "arxiv_id": "1912.00224", "language": "en", "url": "https://arxiv.org/abs/1912.00224", "abstract": "The following generalisation of the Erdős unit distance problem was recently suggested by Palsson, Senger and Sheffer. Given $k$ positive real numbers $\\delta_1,\\dots,\\delta_k$, a $(k+1)$-tuple $(p_1,\\dots,p_{k+1})$ in $\\mathbb{R}^d$ is called a $(\\delta,k)$-chain if $\\|p_j-p_{j+1}\\| = \\delta_j$ for every $1\\leq j \\leq k$. What is the maximum number $C_k^d(n)$ of $(k,\\delta)$-chains in a set of $n$ points in $\\mathbb{R}^d$, where the maximum is taken over all $\\delta$? Improving the results of Palsson, Senger and Sheffer, we essentially determine this maximum for all $k$ in the planar case. error term It is only for $k\\equiv 1$ (mod) $3$ that the answer depends on the maximum number of unit distances in a set of $n$ points. We also obtain almost sharp results for even $k$ in $3$ dimension.", "subjects": "Combinatorics (math.CO); Metric Geometry (math.MG)", "title": "Almost sharp bounds on the number of discrete chains in the plane", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9840936096877063, "lm_q2_score": 0.8152324938410783, "lm_q1q2_score": 0.8022650875987776 }
https://arxiv.org/abs/2112.05094
Weak limits of consecutive projections and of greedy steps
Let $H$ be a Hilbert space. We investigate the properties of weak limit points of iterates of random projections onto $K\geq 2$ closed convex sets in $H$ and the parallel properties of weak limit points of residuals of random greedy approximation with respect to $K$ dictionaries. In the case of convex sets these properties imply weak convergence in all the cases known so far. In particular, we give a short proof of the theorem of Amemiya and Ando on weak convergence when the convex sets are subspaces. The question of the weak convergence in general remains open.
\section*{Introduction} In what follows $H$ is a real Hilbert space with scalar product $\langle\,\cdot\,,\cdot\,\rangle$ and norm $|\,\cdot\,|$. Let $A_1,\dots,A_K$ be closed and convex sets in $H$, $K\geq 2$, so that $A_1\cap\dots\cap A_K=\{0\}$. Let $P_i$ denote the metric projection onto $A_i$. Let $i(n)\in \{1,\dots, K\}$ be a fixed sequence containing each $k\in \{1,\dots, K\}$ infinitely often. For $x_0\in H$, we consider the sequence \begin{equation}\label{pr} x_n=P_{i(n)} x_{n-1}, \qquad n=1,2,\dots \end{equation} In the case when $A_i$ are closed subspaces of $H$ the convergence properties of the sequence $\{x_n\}$ are well understood. If the sequence of the indices $\{i_n\}$ is periodic then the sequence $\{x_n\}$ converges in norm \cite{N}, \cite{Ha}. The rate of the convergence depending on the position of the subspaces and of the initial point is known \cite{BaGM1}, \cite{BaGM2}, \cite{BDH}, \cite{BK}. In this context an interplay with the convergence properties of the greedy algorithm was discovered recently \cite{BK}. If no extra information about the indices, or about the position of the subspaces is known already for $K=3$ divergence might occur \cite{P}, \cite{KM}, \cite{KP}, \cite{K20}. The sequence $\{x_n\}$, however, always converges weakly to zero \cite{AA}. In the lack of linearity, when the sets $A_i$ are just closed and convex, the situation is different. Already for $K=2$ the sequence $\{x_n\}$ might diverge in norm, although the sequence of indices is inevitably periodic \cite{H},\cite{K08},\cite{MR}. Weak convergence is known only under additional conditions: when $K\leq 3$ \cite{DR}, or when the indices are periodic \cite{Bre}, or when the sets are ``somewhat symmetric" \cite{DR}. We denote by $W=W(x_0)$ the set of all partial weak limits of the sequence (\ref{pr}) and face \begin{problem}\label{problem1} Is it true that $W=\{0\}$? \end{problem} We investigate the structure of the set $W$ and give new short proofs of the weak convergence in all of the cases mentioned above. In particular, we give a short proof of the theorem of Amemiya and Ando on weak convergence when the convex sets are subspaces. The general case remains, however, open. In the spirit of \cite{BK}, we establish an interplay with the weak convergence problem of greedy approximation with respect to $K$ dictionaries. The structural properties of the set of weak partial limits of this greedy approximation turn out to be the same. We have hit the same bounds of knowledge while seeking weak convergence. \section{Projections on convex sets}\label{sec:convex} Let $A_1,\dots,A_K$ be closed and convex sets in $H$, $K\geq 2$ so that $A_1\cap\dots\cap A_K=\{0\}$. Let $i(n)\in \{1,\dots, K\}$ be a fixed sequence containing each $k\in \{1,\dots, K\}$ infinitely often and let the sequence $\{x_n\}$ be defined by (\ref{pr}). We assume $i(n)\not=i(n+1)$ without loss of generality. We study the structure of the set $W=W(x_0)$ of all partial weak limits of the sequence $\{x_n\}$. Since the nearest point projection onto a convex set is a $1$-Lipschitz mapping, the norms $|x_n|$ decrease and the set $W$ is always nonempty. We may assume that $|x_n|\searrow R>0$, as $R=0$ implies convergence in norm and hence $W=\{0\}$. For $w\in W$, we denote by $J(w)$ the maximal subset of $\{1,\dots, K\}$ such that $w\in A_{J(w)}$. Here we use the notation $A_J=\cap_{j\in J}A_j$. Since $|x_n-x_{n-1}|^2\leq |x_{n-1}|^2-|x_n|^2$, for $x_n\in A_{i(n)}$ we have \begin{equation}\label{dist} {\rm dist\,} (x_n,A_{i(n\pm m)})\to 0 \qquad (n\to \infty) \end{equation} for any fixed $m$. Therefore $|J(w)|\geq 2$ for each $w\in W$, and $W$ is a weakly closed subset of $\cup_{|J|\geq 2}A_J \cap B(0,R)$, where $B(0,R)$ is the ball centered at 0 of radius R. If $w\not=0$, then $|J(w)|<K$, since $\bigcap A_i=\{0\}$. It also follows from (\ref{dist}) that in case $i(n)\equiv n ({\rm mod}\, K)$ of alternating projections we have $J(w)=\{1,\dots,K\}$ for each $w$, and hence $W=\{0\}$. In particular, if we have just two convex sets then the sequence $\{x_n\}$ converges weakly. Next we show that if $W$ contains an element of maximal norm, then $W=\{0\}$. \begin{theorem}\label{theorem1} For each $w\in W$, $w\not=0$, one can find another element $w'\in W$ with the following properties: \begin{itemize} \item[(i)] $|J(w')\setminus J(w)|\geq 1$; \item[(ii)] $|J(w')\cap J(w)|\geq 2$ \item[(iii)] $|J(w')|\geq 3$; \item[(iv)] $\langle w'-w,a\rangle\geq 0$ for every $a\in A_{J(w)}$. \end{itemize} In particular $|w'|>|w|$ in view of (i), since $\langle w'-w,w\rangle\geq 0$, hence also $|w'|^2\geq |w|^2+|w-w'|^2$. \end{theorem} \begin{proof} (i) Let $$ x_{n_k}\rightharpoonup w, \qquad i(n_k)\in J(w). $$ Taking a subsequence of $k$'s if needed, we can choose $q\notin J(w)$ with the following property: for any $k$ there is a number $m_k\in (n_k, n_{k+1})$ with $i(m_k)=q$, so that for any $n\in [n_k, m_k)$ we have $i(n)\in J(w)$, and hence $i(n)\not= q$. Again taking a subsequence of $k$'s if needed, we get $x_{m_k}\rightharpoonup w'$, and this is the definition of $w'$. Clearly $w'\in A_q$, hence $J(w')\ni q $ and (i) holds. (ii) The numbers $i(m_k-1)$ and $i(m_k-2)$ belong to $J(w)$ and are distinct. We choose two different numbers $i,j\in J(w)$ so that $i(m_k-1)=i$ and $i(m_k-2)=j$ for infinitely many $k$'s. In view of (\ref{dist}) this implies $w'\in A_i\cap A_j$, hence $i,j \in J(w')\cap J(w) $ and (ii) holds. The property (iii) follows from (i) and (ii). (iv) For any $a\in A_{J(w)}$, we have $$ \langle w'-w,a\rangle = \lim_{k\to \infty} \langle x_{m_k}-x_{n_k},a\rangle $$ $$ = \lim_{k\to \infty} \sum_{n=n_k+1}^{m_k}\langle x_{n}-x_{n-1},a\rangle= \lim_{k\to \infty} \sum_{n=n_k+1}^{m_k-1}\langle x_{n}-x_{n-1},a\rangle $$ $$ = \lim_{k\to \infty}\frac{1}{2} \sum_{n=n_k+1}^{m_k-1}(| x_{n-1}-a|^2-|x_n-a|^2 + |x_n|^2 -|x_{n-1}|^2) $$ $$ =\frac{1}{2} \lim_{k\to \infty} \left(\sum_{n=n_k+1}^{m_k-1}(| x_{n-1}-a|^2-|P_{i(n)}x_{n-1}-P_{i(n)}a|^2 ) + |x_{m_k-1}|^2 -|x_{n_k}|^2\right)\geq 0, $$ since every term in the sum is non-negative and $\lim_{k\to \infty}|x_{m_k-1}|=\lim_{k\to \infty}|x_{n_k}|= R$. \end{proof} \begin{remark}\label{remark1} The inequality (iv) holds for $a\in A_{J(w,w')}$, where $J(w,w')=\{i(n): n\in [n_k,m_k-1], k=1,2,\dots\}$. Since $J(w,w')\subset J(w)$, $A_{J(w,w')}$ can be strictly larger than $A_{J(w)}$. \end{remark} The following corollary is a special case of Theorem~2 of \cite{DR}; our proof is different. \begin{corollary}\label{cor1.1} If $K\leq 3$, then $W=\{0\}$. \end{corollary} \begin{proof} The case $K=2$ we have explained above Theorem~\ref{theorem1}. Assume that $K=3$ and that there is $w\in W\setminus \{0\}$. By Theorem \ref{theorem1} there is $w'\in W$ with $|w'|>|w|$ and $J(w')=\{1,2,3\}$ Hence $w'=0$ which is a contradiction. \end{proof} Assume all the convex sets $A_i$ are cones. Assume, moreover, that the intersection of any triple of these cones with the unit sphere has a positive distance to the intersection of any other triple. Then $W=\{0\}$ according to the next corollary. \begin{corollary}\label{cor1.2} Suppose for every $r>0$ there exists $\delta(r)>0$ so that for any two different triples $\{i,j,k\}$ and $\{i,j,l\}$ and elements $u\in A_{\{i,j,k\}}\cap S(0,r)$, $v\in A_{\{i,j,l\}}\cap S(0,r)$ we have $|u-v|>\delta(r)$. Then $W=\{0\}$. \end{corollary} \begin{proof} Suppose $W\not=\{0\}$. Using Theorem \ref{theorem1}, we construct a sequence $w_n\in W$ so that $w_1\not=0$, $w_{n+1}=w_n'$, $|J(w_n)|\ge 3$, $J(w_n)\not= J(w_{n+1})$ and $|J(w_n)\cap J(w_{n+1})|\ge 2$ for each $n$. So we get $w_n\in A_{\{i,j,k\}} $ and $w_{n+1}\in A_{\{i,j,l\}} $ for some $i$, $j$ and $k\not= l$ depending on $n$. Since the sequence $|w_n|$ is bounded and increasing, let $r=\lim_{n\to \infty} |w_n|$. Hence, $u_n=rw_n/(2|w_n|)\in A_{\{i,j,k\}}$ for all sufficiently large $n$. Denoting $w_n=(1+t_n)u_n$, $t_n>0$, for those $n$ we have $$ \begin{array}{rcl} |w_{n+1}|^2&\ge & |w_n|^2+|w_{n+1}-w_n|^2 =\\ &=& |w_n|^2+|(1+t_{n+1})u_{n+1}-(1+t_n)u_n|^2 \\ &\ge & |w_n|^2+|u_{n+1}-u_n|^2+2\langle u_{n+1}-u_n, t_{n+1}u_{n+1}-t_nu_n \rangle\\ &\ge &|w_n|^2+|u_{n+1}-u_n|^2+ 2(r/2)^2(t_{n+1}+t_n-(t_{n+1}+t_n))\\ &=&|w_n|^2+|u_{n+1}-u_n|^2>|w_n|^2+\delta(r/2)^2. \end{array} $$ That means, however, that $|w_n|$ is unbounded. \end{proof} \begin{theorem}\label{theorem2} If $W\not=\{0\}$, then one can find two different elements $w,w'\in W$ so that $$ w={\rm weak}\lim_{k\to \infty} x_{n_k}, \qquad w'={\rm weak}\lim_{k\to \infty} x_{m_k}, $$ where $n_1<m_1<n_2<m_2<\dots$, and $i(n)\in J(w)\cap J(w')$ for any $n\in \cup_k (n_k,m_k)$. Consequently, $(i)$ $\langle w'-w,a\rangle\geq 0$ for every $a\in A_{J(w)}$, $(ii)$ $\langle w'-w,b\rangle\geq 0$ for every $b\in A_{J(w')}$. \end{theorem} \begin{proof} Both inequalities (i) and (ii) follow from the first statement of the Theorem. The proof follows that of (iv) of Theorem \ref{theorem1}: all projections between $n_k$ and $m_k$ have indices from $J(w)\cap J(w')$. To prove the first statement we take $w$ and $w'$ from Theorem \ref{theorem1}. All indices in $J=\{i(n): n\in \cup_k (n_k,m_k)\}$ belong to $J(w)$ by the proof of Theorem \ref{theorem1}. If $J\subset J(w')$, we are done. Otherwise we define $\nu_k$ as the largest numbers $n\in (n_k, m_k)$ such that $i(n)\notin J(w')$. By taking a subsequence of $k$'s so that all these $i(n)$ are the same, we get $x_{\nu_k}\rightharpoonup v\not= w'$. Then we redefine $w:=v$, $n_k:=\nu_k$. The renewed set $J=\{i(n): n\in \cup_k (n_k,m_k)\}$ is now a subset of $J(w')$, and the number of elements in it has decreased by at least one. If this new $J$ is also included in the new $J(w)$, we stop. Otherwise we this time choose the numbers $\nu_k$ as the least numbers $n\in (n_k, m_k)$ such that $i(n)\notin J(w)$. Then we redefine $w'$ and $m_k$'s. Since $|J|$ is decreasing, this oscillation process stops in a finite number of steps: $|J|$ cannot become less than 2. In case $|J|=2$ obviously $J\subset J(w)\cap J(w')$. \end{proof} Dye and Reich used in \cite{DR} the so-called weak internal points (WIP) of a convex set to prove a nonlinear result that properly contains the original linear theorem of Amemiya and Ando: if all the $K$ closed convex sets are linear subspaces then the sequence $\{x_n\}$ converges weakly \cite{AA}. In our version of the theorem we assume that zero is a WIP in each of the convex sets $A_k$. Again, the result is a special case of Theorem~5 of \cite{DR}; our proof is different. \begin{corollary}\label{cor2.1} Assume that zero is a weak internal point of each of the $K$ convex sets $A_k$: if $a\in A_k$ then $-\lambda a\in A_k$ for some $\lambda=\lambda(a,k)>0$. Then $W=\{0\}$. In particular, if all $A_k$ are closed linear subspaces of $H$ then the sequence (\ref{pr}) converges weakly. \end{corollary} \begin{proof} Assuming $W\neq \{0\}$ we take the two different elements $w,w'\in W$ from Theorem \ref{theorem2}. Using (i) of Theorem \ref{theorem2} for $a=w$ gives $\langle w'-w,w\rangle\geq 0$ and $|w'|^2\geq |w|^2+|w-w'|^2$. Using (ii) of Theorem \ref{theorem2} for $b=-\lambda w'$ gives $\langle w'-w,-\lambda w'\rangle\geq 0$ and $|w|^2\geq |w'|^2+|w-w'|^2$. Hence $w=w'$, which is a contradiction. \end{proof} \section{Parallels between projecting onto convex sets and greedy approximation} A subset $D$ of the the unit sphere $S(H)$ of the Hilbert space $H$ is called a {\em dictionary} if its span is dense in $H$. Assume, moreover, that $D$ does not lie in a half-space: for any nonzero $v\in H$, there exists $g\in D$ such that $\langle v,g\rangle>0$. The greedy approximation algorithm then generates for $D$ and for any element $x=x_0\in H$ the sequence \begin{equation} \label{eq1} x_{n+1}=x_n-\langle x_n,g_{n+1}\rangle g_{n+1}, \qquad n=0,1,\dots, \end{equation} where the element $g_{n+1}\in D$ is such that $$ \langle x_n,g_{n+1}\rangle=\max\{\langle x_n,g\rangle\colon g\in D\}. $$ The existence of $\max\{\langle x,g\rangle\colon g\in D\}$ for every $x\in H$ is an additional condition on $D$. If the maximum is attained on several elements of $D$, any of them is selected as $g_{n+1}$. More precisely, this algorithm is called the \textit {pure greedy} algorithm, in contrast to other approximation algorithms whose names contain the word ``greedy'', see~\cite{T}. For any symmetric dictionary $D$ the pure greedy algorithm converges in norm, see~\cite[Ch.~2]{T}. That is, $x_n\to 0$ for any initial element $x=x_0$, and $x$ is represented as a norm-convergent series $\sum_{n=0}^\infty \langle x_n,g_{n+1}\rangle g_{n+1}$. If $D$ is not symmetric, the greedy algorithm may diverge in norm~\cite{B21}, although it always converges weakly to zero~\cite{B20}. Several details of the divergence construction in~\cite{B21} occur to be similar to that of~\cite{K08}. The ``bridge" between this two seemingly different examples is the theorem of Moreau \cite{M}: \begin{equation}\label{moreau} P_A(x)=x-P_{A^*}(x) \end{equation} for any $x\in H$, any convex cone $A\subset H$ and its polar cone $$ A^*=\{y\in H: \langle y, z\rangle\le 0 \quad \forall z\in A\}. $$ Recall that both papers~\cite{H} and \cite{K08} provide examples of convex cones $A_1,A_2\subset H$ so that $A_1\cap A_2=\{0\}$ and alternating projections on those cones diverge in norm for certain starting elements. The formula (\ref{moreau}) allows us to interpret this result as an example of a divergent greedy algorithm with respect to the dictionary $D=(A_1^*\cup A_2^*)\cap S(H)$. Indeed, $D$ does not lie in a half-space as $A_1\cap A_2=\{0\}$, and for any greedy residual $x_n$ lying in, say, $A_1$, we have $$ \max\{\langle x_n,g\rangle\colon g\in D\}=\max\{\langle x_n,g\rangle\colon g\in A_2^*\cap S(H)\}, $$ so that $x_{n+1}=x_n-P_{A_2^*}(x_n)=P_{A_2}(x_n)\in A_2$. Thus the author of~\cite{B21} didn't have to reinvent the wheel:~\cite{H} and \cite{K08} both provided the example he needed. However, the example in \cite{B21} is simpler than those of \cite{H} and \cite{K08}: it uses a discrete dictionary without the extra care needed to build it of convex cones. The above parallels between projecting onto convex sets and greedy approximation have already been noticed in \cite{BK} in the special case of subspaces. In the context of this paper, these parallels bring up the question of weak divergence of random greedy steps with respect to several dictionaries. This problem is considered in the next section. It turns to have the same ``bounds of knowledge" as the problem of the weak divergence of random projections onto several convex sets. \section{Greedy approximation with respect to several dictionaries} Let $K\ge 2$, $D_1,\dots,D_K$ be subsets of $S(H)$ so that their union $\bigcup_{i=1}^K D_i$ is contained in no half-space: for any nonzero $v\in H$, there exists $g\in \bigcup_{i=1}^K D_i$ such that $\langle v,g\rangle>0$. This implies that the set $\bigcup_{i=1}^K D_i$ is spanning; we will call here the sets $D_i$ dictionaries. Assume that for each $x\in H$ and each $i\in \{1,\dots,K\}$ the following condition holds: if $\sup_{g\in D_i} \langle x, g \rangle> 0$, then the supremum is attained on some element $g_i(x)\in D_i$. If it is attained at several elements of $D_i$, then we denote by $g_i(x)$ any one of them. If $\sup_{g\in D_i} \langle x, g \rangle \le 0$, we put $g_i(x)=0$. Clearly, our assumption means that the set $\Lambda(D_i)=\{\lambda g: \lambda\ge 0, g\in D_i\} $ is proximal, and the element $\langle x, g_i(x)\rangle g_i(x)$ belongs to the metric projection $P_{\Lambda(D_i)}(x)$. Let $G_i$ denote the mapping corresponding to one step of the greedy algorithm with respect to the dictionary $D_i$: $$ G_i(x)=x-\langle x, g_i(x)\rangle g_i(x). $$ Note that \begin{equation}\label{Pyth1} |G_i(x)|^2= |x|^2- |x-G_i(x)|^2. \end{equation} Let $i(n)\in \{1,\dots, K\}$ be a fixed sequence containing each $k\in \{1,\dots, K\}$ infinitely often and such that $i(n)\neq i(n+1)$ for all $n\in \mathbb N$. For $x_0\in H$, we consider the sequence $$ x_n=G_{i(n)} x_{n-1}, \qquad n=1,2,\dots. $$ As we have already mentioned above, this sequence may diverge in norm even in case of one dictionary. Both examples in \cite{H} and \cite{K08} can be interpreted as norm divergence examples of residuals $x_n$ for alternating greedy steps with respect to two dictionaries. So we are interested in weak convergence, just as in case of projections. Denoting $W=W(x_0)$ the set of all partial weak limits of the sequence $\{x_n\}$, we face \begin{problem}\label{problem2} Is it true that $W=\{0\}$? \end{problem} We may assume $x_n\not= x_{n-1}$ for all $n$, that is, $\sup_{g\in D_{i(n)}} \langle x, g \rangle> 0$. According to (\ref{Pyth1}), \begin{equation}\label{pyth} |x_{n+1}|^2=|x_{n}|^2-|x_n-x_{n+1}|^2, \end{equation} hence the norms $|x_n|$ are decreasing. We may assume that $|x_n|\searrow R>0$, since $R=0$ implies $W=\{0\}$. We define the closed convex cones $$ A_i=\{y\in H: \langle y, g \rangle\le 0 \mbox{ for all } g\in D_i\}, \qquad i=1,\dots, K. $$ Notice, that $A_i$ is the polar cone of $\overline{{\rm conv\,}}\Lambda(D_i)$. As in Section~\ref{sec:convex}, for $w\in W$, we denote by $J(w)$ the maximal subset of $\{1,\dots, K\}$ such that $w\in A_{J(w)}$, and again use the notation $A_J=\cap_{j\in J}A_j$. Let us nevertheless stress, that the set $W$ is here the result of greedy approximation with respect to the dictionaries $D_1, \dots, D_K$. Let us prove that $|J(w)|\geq 2$ for each $w\in W$. The convergence $x_{n_j}\rightharpoonup w$ implies the convergence $x_{n_j+m}\rightharpoonup w$ for any fixed $m$, since $\lim_{i\to \infty}|x_i-x_{i+m}|=0$ by (\ref{pyth}). Suppose the sequence $i(n_j+m)$ contains some $k$ infinitely often. If $w\notin A_k$, then $ \langle w, g \rangle > \delta> 0$ for some $g\in D_k$, which yields $ \langle x_{n_j+m}, g \rangle > \delta$ for all sufficiently large $j$, so that $|x_{n_j+m+1}|^2\leq |x_{n_j+m}|^2- \delta^2$ for such $j$ with $i(n_j+m)=k$, and a contradiction with $|x_n|\searrow R>0$. So we get $w\in A_k$, and since one can find at least two such $k$'s using different $m$'s, we arrive at $|J(w)|\geq 2$. The same argument shows that in case $i(n)\equiv n ({\rm mod}\, K)$ of alternating greedy algorithm we have $J(w)=\{1,\dots,K\}$ for each $w$, and hence $W=\{0\}$. Thus, $W$ is a weakly closed subset of $\cup_{2\le |J|}A_J \cap B(0,R)$. \begin{theorem}\label{theorem3} For each $w\in W$, $w\not=0$, one can find another element $w'\in W$ with the following properties: \begin{itemize} \item[(i)] $|J(w')\setminus J(w)|\geq 1$; \item[(ii)] $|J(w')\cap J(w)|\geq 2$ \item[(iii)] $|J(w')|\geq 3$; \item[(iv)] $\langle w'-w,a\rangle\geq 0$ for every $a\in A_{J(w)}$. \end{itemize} In particular $|w'|>|w|$ in view of (i), since $\langle w'-w,w\rangle\geq 0$, hence also $|w'|^2\geq |w|^2+|w-w'|^2$. \end{theorem} \begin{proof} Theorem \ref{theorem3} is formally identical to Theorem \ref{theorem1}, and the proofs of (i)-(iii) follow the same reasoning. The proof of (iv) is slightly different. As in the proof of Theorem \ref{theorem1}, we have two alternating sequences $n_1<m_1<n_2<m_2<\dots$ so that $$ x_{n_k}\rightharpoonup w, \qquad x_{m_k}\rightharpoonup w', $$ and $i(n)\in J(w)$ for all $n\in \cup_k[n_k,m_k)$. For any $a\in A_{J(w)}$, we have \begin{equation}\notag \begin{split} \langle w'-w,a\rangle &= \lim_{k\to \infty} \langle x_{m_k}-x_{n_k},a\rangle \\ &= \lim_{k\to \infty} \sum_{n=n_k+1}^{m_k}\langle x_{n}-x_{n-1},a\rangle= \lim_{k\to \infty} \sum_{n=n_k+1}^{m_k-1}\langle x_{n}-x_{n-1},a\rangle \\ &= \lim_{k\to \infty}\sum_{n=n_k+1}^{m_k-1} (-1)\langle x_{n-1}, g_{i(n)}(x_{n-1})\rangle \langle g_{i(n)}(x_{n-1}),a\rangle \geq 0. \end{split} \end{equation} The last inequality holds since each of the summands is non-negative: $\langle x, g_i(x)\rangle\ge 0$ for any $x$ and $i$ by the definition of $g_i$, and $\langle g_{i(n)}(x_{n-1}),a\rangle\le 0$ since $i(n)\in J(w)$ and $a\in A_{J(w)}$. \end{proof} \begin{remark}\label{remark2} The inequality (iv) holds for $a\in A_{J(w,w')}$, where $J(w,w')=\{i(n): n\in [n_k,m_k-1], k=1,2,\dots\}$. Since $J(w,w')\subset J(w)$, $A_{J(w,w')}$ can be strictly larger than $A_{J(w)}$. \end{remark} \begin{corollary}\label{corol3.1.} If $K\leq 3$, then $W=\{0\}$. \end{corollary} \begin{proof} If $K=2$ we have an alternating greedy algorithm, hence convergence as we have explained above Theorem~\ref{theorem3}. Assume that $K=3$ and that there is $w\in W\setminus \{0\}$. By Theorem \ref{theorem3} there is $w'\in W$ with $|w'|>|w|$ and $J(w')=\{1,2,3\}$ Hence $w'=0$ which is a contradiction. \end{proof} \begin{corollary}\label{cor3.2} Suppose for any four indices $i,j,k,l\in \{1,\dots,K\}$ the inequality \begin{equation}\label{ijkl} \inf_{s\in S(H)}\sup_{g\in D_i\cup D_j\cup D_k\cup D_l} \langle s, g\rangle>0 \end{equation} holds. Then $W=\{0\}$. \end{corollary} \begin{proof} The inequalities (\ref{ijkl}) provide $\delta>0$ so that for any distinct $i,j,k,l$ and $u\in A_{\{i,j,k\}}\cap S(H)$ there exists $g\in D_l$ such that $\langle u,g \rangle>\delta$. Hence, for any two different triples $\{i,j,k\}$ and $\{i,j,l\}$ and unit elements $u\in A_{\{i,j,k\}}$, $v\in A_{\{i,j,l\}}$ we have $|u-v|>\delta$: $$ |u-v|\ge \langle u-v, g \rangle \ge \langle u,g \rangle>\delta. $$ Further we repeat the proof of Corollary~\ref{cor1.2}. Suppose $W\not=\{0\}$. By Theorem \ref{theorem3}, we can produce a sequence $w_n\in W$ so that $w_{n+1}=w_n'$, $|J(w_n)|\ge 3$, $J(w_n)\not= J(w_{n+1})$ and $|J(w_n)\cap J(w_{n+1})|\ge 2$ for each $n$. So we get $w_n\in A_{\{i,j,k\}} $ and $w_n\in A_{\{i,j,l\}} $ for some $i$, $j$ and $k\not= l$ depending on $n$. Therefore, using that the sets $A$ are cones, we can refine the inequality from Theorem \ref{theorem3}: $$ |w_{n+1}|^2\ge |w_n|^2+|w_{n+1}-w_n|^2\ge |w_n|^2+\delta^2|w_n|^2/2. $$ That, however, means that $|w_n|$ is unbounded. \end{proof} \begin{theorem}\label{theorem4} If $W\not=\{0\}$, then one can find two different elements $w,w'\in W$ so that $$ w={\rm weak}\lim_{k\to \infty} x_{n_k}, \qquad w'={\rm weak}\lim_{k\to \infty} x_{m_k}, $$ where $n_1<m_1<n_2<m_2<\dots$, and $i(n)\in J(w)\cap J(w')$ for any $n\in \cup_k (n_k,m_k)$. Consequently, $(i)$ $\langle w'-w,a\rangle\geq 0$ for every $a\in A_{J(w)}$, $(ii)$ $\langle w'-w,b\rangle\geq 0$ for every $b\in A_{J(w')}$. \end{theorem} \begin{proof} We repeat the proof of Theorem~\ref{theorem2}; it is purely combinatorial. The inequalities follow from the first statement as in the proof of part (iv) of Theorem~\ref{theorem3}. \end{proof} \begin{corollary}\label{cor4.1} If all $D_k$ are symmetric, then $W=\{0\}$. \end{corollary} \begin{proof} Assume that $W\neq \{0\}$. We take the two different elements $w,w'\in W$ from Theorem~\ref{theorem4}. Using (i) of Theorem~\ref{theorem4} for $a=w$ gives $\langle w'-w,w\rangle\geq 0$, hence $|w'|^2\geq |w|^2+|w-w'|^2$. Using (ii) of Theorem~\ref{theorem4} for $b=-\lambda w'$ gives $\langle w'-w,-\lambda w'\rangle\geq 0$, hence $|w|^2\geq |w'|^2+|w-w'|^2$. Thus we get $w=w'$, which is a contradiction. \end{proof}
{ "timestamp": "2021-12-10T02:26:30", "yymm": "2112", "arxiv_id": "2112.05094", "language": "en", "url": "https://arxiv.org/abs/2112.05094", "abstract": "Let $H$ be a Hilbert space. We investigate the properties of weak limit points of iterates of random projections onto $K\\geq 2$ closed convex sets in $H$ and the parallel properties of weak limit points of residuals of random greedy approximation with respect to $K$ dictionaries. In the case of convex sets these properties imply weak convergence in all the cases known so far. In particular, we give a short proof of the theorem of Amemiya and Ando on weak convergence when the convex sets are subspaces. The question of the weak convergence in general remains open.", "subjects": "Functional Analysis (math.FA)", "title": "Weak limits of consecutive projections and of greedy steps", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9840936082881853, "lm_q2_score": 0.8152324826183822, "lm_q1q2_score": 0.802265075413659 }
https://arxiv.org/abs/1007.4508
Finite Size Percolation in Regular Trees
In the context of percolation in a regular tree, we study the size of the largest cluster and the length of the longest run starting within the first d generations. As d tends to infinity, we prove almost sure and weak convergence results.
\section{Introduction} Fix a positive integer $r$ and let $\mathbb{T}$ be the infinite $r$-ary tree, rooted at $\rho_0$. We consider a Bernoulli percolation on $\mathbb{T}$. Formally, to each node $v \in \mathbb{T}$, we associate a random variable $X_v$, where the variables $\{X_v:v \in \mathbb{T}\}$ are i.i.d.~Bernoulli with $\pr{X_v = 1} = 1 -\pr{X_v = 0} = p \in (0,1)$. For a subset $A \subset \mathbb{T}$, let $X_A = \prod_{v \in A} X_v$. We say that $A$ is {\it open} if $X_A = 1$. \subsection{The size of the largest cluster.} We use the term {\it cluster} to denote a connected component (i.e.~subtree) of $\mathbb{T}$ when undirected. Let $\mathcal{K}$ denote the set of clusters in $\mathbb{T}$. For a node $v \in \mathbb{T}$, let ${\rm gen}(v)$ be its generation, i.e.~the number of nodes in the shortest path from the root $\rho_0$ to $v$, not counting $\rho_0$. Note that ${\rm gen}(\rho_0) = 0$. Let $\mathbb{T}_d$ be the set of nodes with generation not exceeding $d$, namely $\mathbb{T}_d = \{v \in \mathbb{T}: {\rm gen}(v) \leq d\}$. For a cluster $A \in \mathcal{K}$, we let $|A|$ denote its size (i.e.~number of nodes) and $\rho(A)$ its root, namely $\rho(A) = \argmin\{{\rm gen}(v): v \in A\}$. For $d \in \mathbb{N}$, define $K_d$ to be the size of the largest open cluster with root of generation not exceeding $d$: $$K_d = \max\{|A|: A \in \mathcal{K},\, \rho(A) \in \mathbb{T}_d,\, X_A = 1\}.$$ In particular, $K_0$ is the size of the largest open cluster containing the root $\rho_0$. In this paper we study the limit behavior of $K_d$, as $d \to \infty$. In the context of the one-dimensional lattice $\mathbb{Z}$, the corresponding results are often referred to as the Erd\"os-R\'enyi Law~\cite{MR0272026} and, in that context, our approach follows that of Arratia, Goldstein and Gordon~\cite{MR972770}. In higher dimensions, the problem is much more intricate and many questions remain without answer. For a sample of sophisticated results, see e.g.~\cite{MR1868996,MR1372330,MR1880230}. The book by Grimmett~\cite{MR1707339} is a standard reference on percolation. For references more specific to trees, we refer the reader to a survey paper by Pemantle~\cite{MR1368099} and the book of Lyons and Peres~\cite{tree-book}. Though the literature on percolation is vast, most of it focuses on the existence of an infinite cluster and its characteristics when it exists. On the applications side, Patil and Taillie~\cite{MR2109372} identify regions of interest in a network by thresholding the response from each site in the network and computing connected components, which amounts to extracting the open clusters. One imagines that the largest cluster might receive the most attention. In particular, they mention monitoring water quality in a network of freshwater streams, where each stream may be modeled as a tree, though an irregular one. It is well-known that, in the supercritical setting where $p > 1/r$, the cluster at the origin has positive probability of being infinite, and, in fact, $\pr{K_d = \infty}$ tends to 1 as $d$ increases. We restrict our attention to the subcritical and critical cases, i.e~$p < 1/r$ and $p = 1/r$ respectively, where the cluster at the origin is finite with probability one. We start with the critical case, where we show that $K_d$ behaves like the maximum of $r^d$ independent random variables with distribution the total progeny of a Galton-Watson process with offspring distribution $\text{Bin}(r, 1/r)$. Let $\log_r$ denote the logarithm in base $r$. \begin{thm} \label{thm:cluster-cri} Assume $p=1/r$. Then with probability one, $$\frac{\log_r K_d}{d} \rightarrow 2, \quad d \to \infty.$$ Moreover, \[ \pr{\log_r K_d \leq 2d +x} \to \to \exp(- \Cl{cluster-cri}\, r^{-x/2}), \quad d \to \infty, \quad \Cr{cluster-cri} := \frac{2\, n^{-1/2}}{\sqrt{2 \pi r (r-1)}} \] \end{thm} For the subcritical case, we obtain similar results without transformation. Here, a Poisson approximation applies showing that $K_d$ behaves like the maximum of $|\mathbb{T}_d| = r^{d+1}/(r-1)$ independent random variables with distribution the total progeny of a Galton-Watson process with offspring distribution $\text{Bin}(r, p)$. Define \begin{equation} \label{kappa} \kappa = p (1-p)^{r-1}\ \frac{r^r}{(r-1)^{r-1}}. \end{equation} Note that $\kappa < 1$ for all $p < 1/r$. Let $[x]$ denote the entire part of $x \in \mathbb{R}$. \begin{thm} \label{thm:cluster-sub} Assume $p<1/r$. Then with probability one, $$\frac{K_d}{d} \rightarrow \frac{1}{\log_r(1/ \kappa)}, \quad d \to \infty.$$ Moreover, the sequence of random variables $(K_d -\mu_d: d \geq 0)$ is tight, where $\mu_d := \frac{d -\frac{3}{2} \log_r d}{\log_r(1/ \kappa)}$. In addition, a subsequence $K_d -\mu_d$ converges weakly if, and only if, $a := \lim_{d \to \infty} (\mu_d -[\mu_d])$ exists, in which case the weak limit is $[Z+a] -a$, where \[ \pr{Z \leq z} = \exp\left(-\Cl{cluster}\, \kappa^{z}\right), \] for an explicit constant $\Cr{cluster} > 0$ depending only on $(p,r)$ \end{thm} The behavior of the size of the largest open cluster in the subcritical regime is therefore similar in the context of the regular tree and in the context of the one-dimensional lattice, the latter corresponding to the length of the longest perfect head run in a sequence of coin tosses~\cite[Ex.~3]{MR972770}. \subsection{The length of the longest run.} We use the term {\it run} for a path in $\mathbb{T}$ when directed away from the root $\rho_0$. Note that runs are special clusters. Let $\mathcal{R}$ denote the set of runs and define $R_d$ to be the length of the longest open run with root of generation not exceeding $d$: $$R_d = \max\{|A|: A \in \mathcal{R},\, \rho(A) \in \mathbb{T}_d,\, X_A = 1\}.$$ Of course, runs and clusters coincide in the one-dimensional lattice $\mathbb{Z}$. For a general reference on runs in dimension one, see~\cite{MR1882476}. Using the Chen-Stein method, Chen and Huo~\cite{MR2268049} proved results on the longest left-right run in a thin two-dimensional lattice of the form $([0,d]\times[0,a]) \cap \mathbb{Z}^2$, with the width $a$ remaining constant. We also mention the work of Arias-Castro, Donoho and Huo~\cite{MR2275244} who used a statistic based on the longest run in a particular, non-planar graph to detect filaments in point-clouds. The results we obtain for runs are parallel to those we obtain for clusters. In the critical case, we show that $R_d$ behaves like the maximum of $r^d$ independent random variables with distribution the height of a Galton-Watson process with offspring distribution $\text{Bin}(r, 1/r)$. \begin{thm} \label{thm:run-cri} Assume $p=1/r$. Then with probability one, $$\frac{\log_r R_d}{d} \rightarrow 1, \quad d \to \infty.$$ Moreover, for any $x \in \mathbb{R}$, \[ \pr{\log_r R_d \leq d + x} \to \exp(- \Cl{run-cri}\, r^{-x}), \quad d \to \infty, \quad \Cr{run-cri} := \frac{2 rp}{r-1}. \] \end{thm} In the subcritical case, we show that $R_d$ behaves like the maximum of $|\mathbb{T}_d|$ independent random variables with distribution the height of a Galton-Watson process with offspring distribution $\text{Bin}(r, p)$. Again, a Poisson approximation applies. The constant that appears in the exponent is only defined implicitly. \begin{thm} \label{thm:run-sub} Assume $p<1/r$. Then with probability one, $$\frac{R_d}{d} \rightarrow \frac{1}{\log_r (1/p) -1}, \quad {\rm as} \ d \rightarrow \infty.$$ Moreover, the sequence of random variables $(K_d -\nu_d: d \geq 0)$ is tight, where $\nu_d := \frac{d}{\log_r (1/p) -1}$. In addition, a subsequence $K_d -\nu_d$ converges weakly if, and only if, $a := \lim_{d \to \infty} (\nu_d -[\nu_d])$ exists, in which case the weak limit is $[Z+a] -a$, where \[ \pr{Z \leq z} = \exp\left(-\Cl{run}\, (rp)^{z}\right), \] for an explicit constant $\Cr{run} > 0$ depending only on $(p,r)$ \end{thm} \subsection{Contents.} The rest of the paper is devoted to proving our results. In \secref{proof-cluster} we prove \thmref{cluster-cri} and \thmref{cluster-sub}. In \secref{proof-run} we prove \thmref{run-cri} and \thmref{run-sub}. \subsection{Additional Notation.} Let $\partial \mathbb{T}_d = \{v \in \mathbb{T}: {\rm gen}(v) = d\}$. For a cluster $A$, let $\underline{A}$ denote the set of nodes not in $A$ whose parents belong to $A$, and if $\rho(A) \neq \rho_0$, let $\mathring{A}$ denote the parent of $\rho(A)$. Also, define $(1 -X)_A = \prod_{v \in A} (1 - X_v)$. For two sequences of real numbers $(a_n)$ and $(b_n)$, we use the notation $a_n \sim b_n$ to indicate that $a_n/b_n \to 1$ and $a_n \asymp b_n$ to indicate that the ratio $a_n/b_n$ is bounded away from zero and infinity, both understood as $n \to \infty$. Throughout the paper $C$ denotes a finite, positive constant depending only on $r$ and $p$, whose value may change with each appearance. \section{The size of the largest open cluster} \label{sec:proof-cluster} In this section, we prove \thmref{cluster-cri} and \thmref{cluster-sub}. We start with some notation. For a vertex $v \in \mathbb{T}$, let $K(v)$ be the size of the largest open cluster with root $v$, $$K(v) = \max\{|A|: A \in \mathcal{K},\, \rho(A) = v,\, X_A = 1\}.$$ In particular, $$K_d = \max\{K(v): v \in \mathbb{T}_d\}.$$ The distribution of $K(v)$ does not depend on $v \in \mathbb{T}$, and, in fact, given $X_v = 1$, coincides with that of the total progeny of a Galton-Watson tree starting with one individual and with offspring distribution Bin$(r,p)$. Define \[ \psi_n = \pr{K(v) = n}, \quad \Psi_n = \pr{K(v) > n}. \] Applying a well-known identity by Dwass~\cite{MR0253433} (called the Otter-Dwass formula in~\cite{tree-book}), we get \begin{eqnarray*} \psi_n &=& \frac{p}{n} \pr{\xi_1 + \cdots + \xi_n = n-1}, \text{ where } \xi_1, \dots, \xi_n \stackrel{\rm i.i.d.}{\sim} \text{Bin}(r, p) \\ &=& \frac{p}{n} \pr{\text{Bin}(n r, p) = n-1} \\ &=& {\rm Cat}_{n}\, p^{n} (1-p)^{n (r-1) + 1}, \end{eqnarray*} where \[ {\rm Cat}_{n} := \frac{1}{n} {n r \choose n-1} = \frac{1}{(r-1) n + 1} {r n \choose n} \] is the {\it $n$th generalized Catalan number}~\cite{hil91}, which among other interpretations, is the number of subtrees of $\mathbb{T}$ of size $n$ rooted at the origin, i.e.~ \[ {\rm Cat}_{n} = |\{A \in \mathcal{K}: \rho(A) = \rho_0,\, |A| = n\}|. \] We could have obtained the expression for $\psi_n$ using this definition of ${\rm Cat}_n$. Indeed, for $n > 0$, $K(v) = n$ if, and only if, there is a (unique) subtree $A$ with $|A| = n$ , $\rho(A) = v$ and $X_A (1 -X)_{\underline{A}} = 1$, so that $A$ cannot be extended and still be an open cluster. We then use the fact that a subtree of size $n$ has exactly $(r-1)n+1$ children. With the use of Stirling's formula, we arrive at the following conclusions; see also~\cite{MR0386042, ney}. \begin{lem} \label{lem:Kv} In the critical case $p =1/r$, \[ \Psi_n \sim \frac{\Cr{cluster-cri}}{\sqrt{n}}. \] In the subcritical case $p < 1/r$, \[ \Psi_n \sim \Cl{cluster-aux}\, \frac{\kappa^{n+1}}{n^{3/2}}, \quad \Cr{cluster-aux} := \frac{1}{\sqrt{2 \pi}(1-\kappa)} \frac{(1-p) r^{1/2}}{(r-1)^{3/2}}. \] \end{lem} \subsection{Proof of \thmref{cluster-cri}} \label{sec:proof-cluster-cri} Define \[ K^\partial_d := \max\{K(v): v \in \partial \mathbb{T}_d\}. \] We first prove that the conclusions of \thmref{cluster-cri} hold for $K^\partial_d$. For $x \in \mathbb{R}$, let $n_d(x) = [r^{2 d + x}].$ As $K^\partial_d$ only involves independent random variables, we have \begin{equation} \nonumber \pr{\log_r K^\partial_d \leq 2d +x} = \pr{K^\partial_d \leq n_d(x)} = (1 -\Psi_{n_d(x)})^{r^d} = \exp(- \Cr{cluster-cri}\, r^{-x/2} + O(r^{-d -x})). \end{equation} Letting $d \to \infty$, we obtain the weak convergence, and by choosing $x = \varepsilon d$, with $\varepsilon > -2$ fixed, and applying the Borel-Cantelli Lemma, we obtain the almost sure convergence. It therefore suffices to show that $K_d = (1 + o_P(1)) K^\partial_d$. Clearly, $K_d \geq K^\partial_d$, so we focus on the upper bound. Define \[ B_d = \{v \in \partial \mathbb{T}_d: K(v) > r^d/d\}, \quad B^2_d = \{v \in \partial \mathbb{T}_d: K(v) > r^{2d}/d\}. \] For any open cluster $A$ with $\rho(A) \in \mathbb{T}_d$, we have \[ |A| = |A \cap \mathbb{T}_{d-1}| + \sum_{v \in A \cap \partial \mathbb{T}_d} K(v) \leq r^d + r^{2d}/d + \sum_{v \in A \cap B_d} K(v). \] We turn to bounding the sum. We first show that, with probability tending to one, there is no open cluster $A$ containing three or more nodes in $B_d$. Indeed, take $v_1, v_2, v_3 \in \partial \mathbb{T}_d$ distinct. Let $w$ denote their most recent common ancestor and let $k = d -{\rm gen}(w)$. Either the paths $v_j \to \rho_0$ meet at $w$ for the first time or two of the paths meet at a node $u$ with ${\rm gen}(u) > {\rm gen}(w)$, in which case we let $\ell = d -{\rm gen}(u)$. Now, the nodes $v_1, v_2, v_3$ belong to the same open cluster if, and only if, the smallest subtree containing $w$ and $v_1, v_2, v_3$ is open, and this subtree is of size $\ell + 2 k + 1$, and therefore, the probability that they belong to the same open cluster is $p^{\ell + 2k + 1}$. In addition, the number of such triplets is bounded by \[ \left(r^{d-k} {r^k \choose 3}\right) \cdot \left(r^k r^{k -\ell} {r^\ell \choose 2}\right) {r^k \choose 3}^{-1} \asymp r^{d +k +\ell}. \] The first factor comes from the fact that the three nodes are leaves of a subtree with root at generation $d-k$. Given that, the second factor comes from the fact that two of them belong to a subtree of that subtree with root at (relative) generation $k -\ell$. Hence, remembering that $p = 1/r$ and using \lemref{Kv}, we have \begin{eqnarray*} \pr{\exists A \in \mathcal{K}: X_A = 1,\, |A \cap B_d| \geq 3} &\leq& C\, \pr{K(v) > r^d/d}^3 \cdot \sum_{k=0}^d \sum_{\ell=0}^k r^{d +k +\ell} p^{\ell + 2k + 1} \\ &\leq& C\, (r^d/d)^{-3/2} r^{d} \asymp d^{3/2} r^{-d/2}. \end{eqnarray*} By the same token, with probability tending to one (in fact of order at most $d/r^d$), there is no open cluster $A$ containing two or more nodes in $B^2_d$. Now, when $|A \cap B_d| \leq 2$ and $|A \cap B^2_d| \leq 1$, we have \[ \sum_{v \in A \cap B_d} K(v) \leq \max_{v \in A \cap B_d} K(v) + r^{2d}/d \leq K^\partial_d + r^{2d}/d. \] In the end, with probability tending to one, \[ |A| \leq r^d + 2\, r^{2d}/d + K^\partial_d, \] for any open cluster $A$ with $\rho(A) \in \mathbb{T}_d$. Hence, \[ K_d \leq K^\partial_d + O_P(r^{2d}/d), \] and we conclude by the fact that $K^\partial_d$ is of order exceeding $r^{2d}/d$ with probability tending to one. \subsection{Proof of \thmref{cluster-sub}} \label{sec:proof-cluster-sub} The proof of the almost sure convergence may be obtained following the arguments provided in \secref{proof-cluster-cri} or using the bounds we are about to prove below. We omit details. The proof of the weak convergence is based on the Chen-Stein method for Poisson approximation as formulated by Arratia, Goldstein and Gordon~\cite{MR972770}. Define $$Y_A = \left\{\begin{array}{ll} X_A (1 -X)_{\underline{A}}, & \rho(A) = \rho_0,\\[.05in] X_A (1-X)_{\mathring{A}} (1 -X)_{\underline{A}}, & \rho(A) \neq \rho_0; \end{array}\right.$$ Also, let $\mathcal{K}_{d,n}$ be the set of clusters of size exceeding $n$ with root in $\mathbb{T}_d$, and define $$W_{d,n} = \sum_{A \in \mathcal{K}_{d,n}} Y_A.$$ \vspace{-.1in} By definition, $$\{K_d \leq n\} = \{Y_A = 0,\, \forall A \in \mathcal{K}_{d,n}\} = \{W_{d,n} = 0\}.$$ We approximate the law of $W_{d,n}$ by the Poisson distribution with same mean $\lambda_{d,n} = \expect{W_{d,n}}$. We start by estimating $\lambda_{d,n}$ using \lemref{Kv}, obtaining \begin{eqnarray*} \lambda_{d,n} & = & \sum_{A \in \mathcal{K}_{d,n}} \pr{Y_A = 1} \\ & = & \pr{K(\rho_0) > n} + (1-p) \sum_{v \in \mathbb{T}_d, v \neq \rho_0} \pr{K(v) > n} \\ & = & \Psi_n + (1-p) (|\mathbb{T}_d| -1) \Psi_n. \end{eqnarray*} In particular, as $n, d \to \infty$, \[ \lambda_{d,n} \sim \Cr{cluster}\ r^d n^{-3/2} \kappa^{n+1}, \quad \Cr{cluster} := \frac{\Cr{cluster-aux} (1-p) r}{r-1}. \] For a cluster $A \in \mathcal{K}_{d,n}$, define its neighborhood $\mathcal{B}(A)$ as the set of clusters $B \in \mathcal{K}_{d,n}$ such that $$(\mathring{B} \cup B \cup \underline{B}) \ \cap \ (\mathring{A} \cup A \cup \underline{A}) \neq \emptyset.$$ Define the following sums \begin{eqnarray*} F_{d,n} & = & \sum_{A \in \mathcal{K}_{d,n}} \sum_{B \in \mathcal{B}(A)} \pr{Y_A = 1} \pr{Y_B = 1},\\ G_{d,n} & = & \sum_{A \in \mathcal{K}_{d,n}} \sum_{B \in \mathcal{B}(A), B \neq A} \pr{Y_A = Y_B = 1},\\ H_{d,n} & = & \sum_{A \in \mathcal{K}_{d,n}} \expect{\left|\expect{Y_A - \expect{Y_A}|Y_B, B \notin \mathcal{B}(A)}\right|}. \end{eqnarray*} Then by the second part of~\cite[Th.~1]{MR972770}, $$\left|\pr{W_{d,n} = 0} - \exp(-\lambda_{d,n})\right| \leq F_{d,n} + G_{d,n} + H_{d,n}.$$ For $x \in \mathbb{R}$, define $n_d(x) = \left[\mu_d +x\right]$. When $x$ is fixed and $d \to \infty$, $\lambda_{d,n_d(x)} \asymp 1$, with $$\lambda_{d,n_d(x)} \to \Cr{cluster}\, \kappa^{[a+x] -a +1}, \ \text{ when } \mu_d -[\mu_d] \to a, \ x -[x] \neq 1-a,$$ with $$\pr{[Z+a] -a \leq x} = \exp(-\Cr{cluster}\, \kappa^{[a+x] -a +1}).$$ Therefore, to conclude it suffices to prove that $F_{d,n}, G_{d,n}, H_{d,n} \to 0$ when $d,n \to \infty$ in such a way that $\lambda_{d,n} \asymp |\mathbb{T}_d| \Psi_n \asymp 1$. First, $H_{d,n} = 0$ by independence of $Y_A$ and $Y_B, B \notin \mathcal{B}(A)$. For $G_{d,n}$, the only pairs $A,B \in \mathcal{K}_{d,n}$ that contribute to the sum satisfy either $\mathring{B} \in \underline{A}$ or $\mathring{A} \in \underline{B}$, and in both cases $$\pr{Y_A = Y_B = 1} = (1-p)^{-1} \pr{Y_A = 1} \pr{Y_B = 1}.$$ Hence, using the fact that there are ${\rm Cat}_{m}$ subtrees of size $m$ with a given root, each with $(r-1)m + 1$ children, and then \lemref{Kv}, we have \begin{eqnarray*} G_{d,n} & \leq & 2 (1-p)^{-1} |\mathbb{T}_d| \sum_{m > n}\ {\rm Cat}_{m}\, p^m (1-p)^{(r-1)m+1} ((r-1)m + 1) \cdot \Psi_n \\ & \leq & C\, \lambda\, \sum_{m > n} m \psi_{m} = C\, \lambda\, \left((n+1) \Psi_n + \sum_{m > n} \Psi_{m}\right) \asymp n^{-1/2} \kappa^n \to 0, \quad n \to \infty. \end{eqnarray*} For $F_{d,n}$, the only pairs $A,B \in \mathcal{K}_{d,n}$ that contribute to the sum satisfy either $\mathring{B} \in \mathring{A} \cup A \cup \underline{A}$ or $\mathring{A} \in \mathring{B} \cup B \cup \underline{B}$. The computations are then similar. \section{The length of the longest open run} \label{sec:proof-run} The arguments are parallel to those provided in \secref{proof-cluster}. For $A \subset \mathbb{T}$, define its height as $\tau(A) = \sup\{{\rm gen}(v): v \in A\} -{\rm gen}(\rho(A))$. For a vertex $v \in \mathbb{T}$, let $R(v)$ be the length of the longest run with root $v$, \[ R(v) = 1 + \max\{\tau(A): A \in \mathcal{K}, \rho(A) = v\}. \] In particular, \[ R_d = \max\{R(v): v \in \mathbb{T}_d\}. \] The distribution of $R(v)$ does not depend on $v \in \mathbb{T}$, and, in fact, given $X_v = 1$, coincides with that of the height (plus one), i.e.~extinction time, of a Galton-Watson tree with offspring distribution Bin$(r,p)$. Define \[ \phi_h = \pr{R(v) = h}, \quad \Phi_h = \pr{R(v) > h}. \] We have the following results on the asymptotic behavior of $\Phi_h$~\cite{ney}. \begin{lem} \label{lem:Rv} In the critical case $p =1/r$, \[ \Phi_h \sim \frac{\Cr{run-cri}}{h}. \] In the subcritical case $p < 1/r$, there is an implicit constant $\Cl{run-aux} > 0$ such that \[ \Phi_h \sim \Cr{run-aux}\, (rp)^h. \] \end{lem} Let ${\rm Cat}_{n,h}$ denote the number of subtrees rooted at the origin, of size $n$ and height $h$. See~\cite{MR1249127} for some results on ${\rm Cat}_{n,h}$. As in \secref{proof-cluster}, we can argue that \[ \phi_h = \sum_{n > h} {\rm Cat}_{n,h}\, p^n (1-p)^{(r-1)n +1}, \ \text{ implying } \ \Phi_h = \sum_{\ell > h} \sum_{n > \ell} {\rm Cat}_{n,\ell}\, p^n (1-p)^{(r-1)n +1}. \] \subsection{Proof of \thmref{run-cri}} The proof is based on the following observation \[ R^\partial_d \leq R_d \leq R^\partial_d + d, \quad R^\partial_d := \max\{R(v): v \in \partial \mathbb{T}_d\}, \] where the $d$ term bounds the length of any run in $\mathbb{T}_{d-1}$. As $R^\partial_d$ only involves independent random variables, \begin{equation} \label{eq:partial-Rd} \pr{R^\partial_d \leq h} = (1 -\Phi_{h})^{r^d}. \end{equation} Choosing $h = r^{[d/2]}$ and using \lemref{Rv}, we obtain \[ \pr{R^\partial_d \leq r^{[d/2]}} \leq \exp(- C\, r^{-d/2}), \ \text{ for some } C > 0, \] so that, applying the Borel-Cantelli Lemma, $R^\partial_d \geq r^{d/2}$ eventually, with probability one. Hence, $\log_r R_d = (1 +o_P(1)) \log_r R^\partial_d$, and it is therefore enough to prove the results for $R^\partial_d$ in place of $R_d$. The almost sure convergence is obtained in a similar way by choosing $h = r^{(1+\varepsilon) d}$ with $\varepsilon$ fixed, either positive or negative. For the weak convergence, fix $x$ and let $h_d(x) = [r^{d+x}]$. By \lemref{Rv} and \eqref{eq:partial-Rd}, we have \[ \pr{\log_r R^\partial_d \leq d + x} \to \exp(- \Cr{run-cri}\, r^{-x}), \quad d \to \infty. \] \subsection{Proof of \thmref{run-sub}} We again omit the details of the proof of the almost sure convergence and focus on proving the weak convergence. Let $\mathcal{K}_{d,h}$ denote the set of clusters with root in $\mathbb{T}_d$ and height exceeding $h$. We use the notation introduced in \secref{proof-cluster-sub}, with $\mathcal{K}_{d,h}$ in place of $\mathcal{K}_{d,n}$. By definition, $$\{R_d \leq h\} = \{W_{d,h} = 0\}.$$ Using \lemref{Rv}, we obtain \begin{eqnarray*} \lambda_{d,h} & = & \sum_{A \in \mathcal{K}_{d,h}} \pr{Y_A = 1}\\ & = & \pr{R(\rho_0) > h} + (1-p) \sum_{v \in \mathbb{T}_d} \pr{R(v) > h}\\ & = & \Phi_h + (1-p) (|\mathbb{T}_d| -1) \Phi_h. \end{eqnarray*} In particular, as $h, d \to \infty$, \[ \lambda_{d,h} \sim \Cr{run}\, r^d (rp)^{h+1}, \quad \Cr{run} := \frac{\Cr{run-aux}\, (1-p)}{p(r-1)}. \] For $x \in \mathbb{R}$, define $h_d(x) = [\nu_d + x]$. When $x$ is fixed and $d \to \infty$, we have $\lambda_{d,h_d(x)} \asymp 1$, with $$\lambda_{d,h_d(x)} \to \Cr{run} (rp)^{[a+x] -a +1}, \ \text{ when } \nu_d -[\nu_d] \to a, \ x -[x] \neq 1-a.$$ It then suffices to show that $F_{d,h}, G_{d,h}, H_{d,h} \to 0$ when $d,h \to \infty$ in such a way that $\lambda_{d,h} \asymp |\mathbb{T}_d| \Phi_h \asymp 1$, and the computations are parallel to those in \secref{proof-cluster-sub}. We focus on $G_{d,h}$. Fix $\tilde{p} \in (p, 1/r)$ and let $\tilde{\Phi}_h$ be defined as $\Phi_h$, with $\tilde{p}$ in place of $p$. For $h$ large enough, we then have \begin{eqnarray*} G_{d,h} &\leq& 2 \sum_{A \in \mathcal{K}_{d,h}}\ \sum_{\mathcal{B}(A), \mathring{B} \in \underline{A}} (1-p)^{-1} \pr{Y_A = 1} \pr{Y_B = 1} \\ & \leq & C\, |\mathbb{T}_d| \sum_{\ell > h} \sum_{n > \ell} {\rm Cat}_{n,\ell}\ p^n (1-p)^{(r-1)n + 1} ((r-1)n+1) \cdot \Phi_h \\ & \leq & C\, \lambda\, \sum_{\ell > h} \sum_{n > \ell} {\rm Cat}_{n,\ell}\ \tilde{p}^n (1-\tilde{p})^{(r-1)n + 1} \\ & = & C\, \lambda\, \tilde{\Phi}_h \asymp (r \tilde{p})^h \to 0, \quad d \to \infty. \end{eqnarray*} \subsection*{Acknowledgements} The author would like to thank Philippe Flajolet for fruitful conversations and Jason Schweinsberg for reading an early version of the manuscript, pointing out some errors and helping with the proof of \thmref{cluster-cri}. This work was partially supported by a grant from the National Science Foundation (DMS-0603890) and a grant from the Office of Naval Research (N00014-09-1-0258). {\small \bibliographystyle{abbrv}
{ "timestamp": "2010-07-27T02:03:30", "yymm": "1007", "arxiv_id": "1007.4508", "language": "en", "url": "https://arxiv.org/abs/1007.4508", "abstract": "In the context of percolation in a regular tree, we study the size of the largest cluster and the length of the longest run starting within the first d generations. As d tends to infinity, we prove almost sure and weak convergence results.", "subjects": "Probability (math.PR)", "title": "Finite Size Percolation in Regular Trees", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9898303413461358, "lm_q2_score": 0.8104789155369048, "lm_q1q2_score": 0.8022366216197404 }
https://arxiv.org/abs/2209.03325
Pancyclicity of Hamiltonian graphs
An $n$-vertex graph is Hamiltonian if it contains a cycle that covers all of its vertices, and it is pancyclic if it contains cycles of all lengths from $3$ up to $n$. In 1972, Erdős conjectured that every Hamiltonian graph with independence number at most $k$ and at least $n = \Omega(k^2)$ vertices is pancyclic. In this paper we prove this old conjecture in a strong form by showing that if such a graph has $n = (2+o(1))k^2$ vertices, it is already pancyclic, and this bound is asymptotically best possible.
\section{Introduction} Hamiltonicity is one of the most central notions in graph theory, and it has been extensively studied by numerous researchers. The problem of deciding Hamiltonicity of a graph is NP-complete and therefore, a central theme in Combinatorics is to derive sufficient conditions for this property. The most classical one is Dirac’s theorem \cite{dirac1952some} which dates back to 1952 and states that every $n$-vertex graph with minimum degree at least $n/2$ contains a Hamilton cycle. Since then, many other interesting results about various aspects of Hamiltonicity have been obtained, see e.g. \cite{ajtai1985first,chvatal1972note,kuhn2013hamilton,krivelevich2011critical,krivelevich2014robust,MR3545109,ferber2018counting, cuckler2009hamiltonian, posa1976hamiltonian}, and the surveys \cite{gould2014recent, MR3727617}. A related notion to Hamiltonicity is that of pancyclicity. An $n$-vertex graph is said to be \emph{pancyclic} if it contains all cycles of length from $3$ up to $n$. Trivially, pancyclicity is a stronger property than Hamiltonicity, and one might ask how much stronger it really is. In 1973, Bondy \cite{bondy10pancyclic} stated his celebrated meta-conjecture, indicating that the first property should be only slightly stronger than the latter. Indeed, he claimed that any non-trivial condition which implies that a graph is Hamiltonian should also imply that it is pancyclic (up to a certain collection of simple exceptional graphs). As an example, Bondy \cite{bondy1971pancyclic} himself first showed that every $n$-vertex graph with minimum degree of at least $n/2$ is either pancyclic or isomorphic to the complete bipartite graph $K_{n/2,n/2}$, thus strengthening Dirac's theorem. His meta-conjecture sparked a lot of research which in turn has led to various appealing results and methods. For example, Bauer and Schmeichel \cite{bauer1990hamiltonian}, relying on previous results of Schmeichel and Hakimi \cite{schmeichel1988cycle}, showed that the sufficient conditions for Hamiltonicity given by Bondy \cite{bondy1980longest}, Chvátal \cite{chvatal1972hamilton} and Fan \cite{fan1984new} all imply pancyclicity, up to a certain small family of exceptional graphs. Furthermore, much like with Dirac's theorem, the classical result of Chvátal and Erdős \cite{chvatal1972hamilton} that a graph with connectivity number $\kappa(G)$ at least as a large as its independence number $\alpha(G)$ is Hamiltonian, has also been addressed. Namely, in 1990, Jackson and Ordaz \cite{jackson1990chvatal}, conjectured that if $\kappa(G) > \alpha(G)$, then $G$ must be pancyclic and an approximate form of this was proven by Keevash and Sudakov \cite{keevash2010pancyclicity}, who showed that $\kappa(G) \geq 600\alpha(G)$ is already sufficient. Bondy's meta-conjecture is about conditions for Hamiltonicity which imply pancyclicity. A natural and closely related question in a similar direction was first studied by Erd\H{o}s in the 1970s. Let $G$ be a Hamiltonian graph; under which assumptions can we guarantee that $G$ is also pancyclic or more generally, that it has many cycle lengths? One example of such a problem was suggested by Jacobson and Lehel at the 1999 conference “Paul Erd\H{o}s and His Mathematics”. They asked for the minimal number of cycle lengths in a $k$-regular $n$-vertex Hamiltonian graph. They conjectured (see Verstraëte \cite{verstraete2016extremal} for a stronger conjecture) that already when $k \geq 3$, there are $\Omega(n)$ many lengths. Improving on the previously best known lower bound of $\Omega(\sqrt{n})$ by Milans et al. \cite{milans2012cycle}, recently Buci\'c, Gishboliner and Sudakov \cite{bucic2021cycles} showed that any Hamiltonian graph with minimum degree at least $3$ has $n^{1-o(1)}$ different cycle lengths. As we already mentioned above, the earliest question of this flavor was studied by Erd\H{o}s about 50 years ago. In 1972 he asked the following in \cite{erdos1972some}. Given an $n$-vertex Hamiltonian graph with independence number $\alpha(G)\leq k$, how large does $n$ have to be in terms of $k$ in order to guarantee that $G$ is pancyclic? Erd\H{o}s \cite{erdos1972some} proved that it is enough to have $n=\Omega(k^4)$ and conjectured that already $n=\Omega(k^2)$ should be enough. A simple construction shows that this is best possible. Let $C_1,\ldots, C_k$ be disjoint cliques of size $2k-2$, and let each $C_i$ have two distinguished vertices $a_i$ and $b_i$. Let $G$ be the graph obtained by connecting $a_i$ and $b_{i+1}$ by an edge for each $i$ (taking addition modulo $k$). Notice that this graph has $n = 2k^2- 2k$ vertices, is Hamiltonian and its independence number is $k$. On the other hand, it is easy to check that it does not contain a cycle of length $2k - 1$, and thus it is not pancyclic. Indeed, observe that every cycle must be either a subgraph of one of the cliques $C_i$, or contain all the vertices $a_i,b_i$ for each $i$. The first type of cycles all have length at most $2k - 2$ and the latter have length at least $2k$. In the last 50 years, there have been several improvements upon Erd\H{o}s's initial result. Firstly, Keevash and Sudakov \cite{keevash2010pancyclicity} proved that $n=\Omega(k^{3})$ vertices are enough to guarantee pancyclicity. Then, Lee and Sudakov \cite{lee2012hamiltonicity} improved this to $n=\Omega(k^{7/3})$, and more recently Dankovics \cite{dankovics2020low} showed that $n=\Omega(k^{11/5})$ vertices suffice. In this paper we completely resolve the conjecture of Erd\H{o}s in the following strong form. \begin{thm}\label{thm:main} Every Hamiltonian graph $G$ with $\alpha(G)\leq k$ and at least $2k^2+o(k^2)$ vertices is pancyclic. \end{thm} \noindent As shown by the previous construction, our result is tight up to the $o(k^2)$ error term. The rest of this paper is organized as follows. In \Cref{sec:preliminarieas}, we state some well-known tools, and introduce some useful definitions. There we also prove the two key propositions that are used in the proof of Theorem~\ref{thm:main} , which is given in \Cref{sec:proof}. Finally, in \Cref{sec:concludingrem} we make some concluding remarks and mention some open questions; we also show how to get a short proof of the conjecture of Erd\H{o}s, that is, a proof of \Cref{thm:main} with a sufficiently large constant $C$ instead of the precise factor of 2. \section{Preliminaries}\label{sec:preliminarieas} \subsection{Notation and definitions} We mostly use standard graph theoretic notation. Let $G$ be a finite graph. Denote by $V(G)$ its vertex set, and let $S_1,S_2\subseteq V(G)$. We denote by $G[S_1]$ the subgraph of $G$ induced by $S_1$, and by $E[S_1,S_2]$ the set of edges with one endpoint in $S_1$ and the other in $S_2$. Let $H$ be a subgraph of $G$. We denote by $G[H]$ the graph $G[V(H)]$. A path $P=(x_0,x_1,\ldots,x_\ell)$ of length $\ell$ is a graph on vertex set $\{x_0,x_1,\ldots,x_\ell\}$ with an edge between $x_{i-1}$ and $x_{i}$ for all $i\in[\ell]$. We say that $x_0$ and $x_\ell$ are the endpoints of $P$, and we call $P$ an $x_0x_\ell$-path. If the vertices of the graph $G$ come with a given ordering, then we say that a path $P=(x_0,x_1,\ldots,x_\ell)$ contained in $G$ is increasing if $x_0<x_1<\ldots<x_\ell$. We denote by $\alpha(G)$ the independence number of $G$. Given a digraph $D$, its independence number $\alpha(D)$ is defined as the independence number of the underlying graph. Given sets $A_1,A_2\subset \mathbb N$, we denote by $A_1+A_2$ the set of integers $c$ such that $c=a_1+a_2$ for some $a_1\in A_1$ and $a_2\in A_2.$ Throughout the paper we omit floor and ceiling signs for clarity of presentation, whenever it does not impact the argument. \begin{defn} Let $a,b,p$ be positive real numbers. Given a graph $G$, and two vertices $x$ and $y$, we say that the pair $xy$ is $p$-dense in the interval $[a,b]$ if for every subinterval $[a',b']$ with $b'-a'\geq p$ there is an integer $\ell\in[a',b']$ and an $xy$-path in $G$ of length $\ell$. \end{defn} \subsection{Standard tools} \noindent Here we state and prove some standard facts which we use in our proof. We start with the following well-known result about directed graphs of Gallai and Milgram \cite{gallai1960verallgemeinerung}. A \emph{path cover} in a directed graph is a partition of its vertex set into directed paths, and its size is the number of such paths. \begin{lem}[\cite{gallai1960verallgemeinerung}]\label{lem:partition} Every directed graph $D$ has a path cover of size at most $\alpha(D)$. \end{lem} \noindent We also use the celebrated Ramsey's theorem. \begin{thm}\label{lem:ramsey} For every two positive integers $k,t$, there exists a large enough integer $n$, such that for any $k$-coloring of the edges of $K_n$, there is a monochromatic copy of $K_t$ in $K_n$.\end{thm} \noindent The next lemma shows that we can partition a large proportion of the vertex set of a graph into sets with small diameter, such that there are no edges between the parts. \begin{lem}\label{lem:BFSpartition} Let $G$ be an $n$-vertex graph and let $0<\gamma <\frac{1}{2}$. Then, there exists a collection of vertices $v_1, v_2, \ldots, v_r$ and disjoint sets $U_1, U_2, \ldots, U_r \subseteq V(G)$ such that the following hold. \begin{enumerate} \item $v_j \in U_j$ for all $j$, and $\left| \bigcup_j U_j \right| \geq (1-\gamma)n$. \item Every vertex $u \in U_j$ has $\text{dist}(v_j,u) \leq \log_{1+\gamma} n$. \item There is no edge between two sets $U_j, U_{j'}$ with $j \neq j'$. \end{enumerate} \end{lem} \begin{proof} We find the required sets and vertices with the following process. We start with an arbitrary vertex $v_1 \in V(G)$ and consider the breadth-first-search tree rooted at $v_1$, that is, consider the sets $V^{(1)}_0,V^{(1)}_1,V^{(1)}_2, \ldots \subseteq V(G)$ defined as $V^{(1)}_i := \{u \in G : \text{dist}(v_1,u) = i\}$. Now, define $i_1 \geq 0$ to be the minimal $i$ such that $|V^{(1)}_{i+1}| \leq \gamma \left|V^{(1)}_0 \cup \ldots \cup V^{(1)}_i \right|$ and let $U_1 := V^{(1)}_0 \cup \ldots \cup V^{(1)}_{i_1}$. We now continue the process and do the same on the graph $G' := G \setminus \left(U_1 \cup V^{(1)}_{i_1 + 1} \right)$. More precisely, take an arbitrary vertex $v_2 \in G'$ and consider again the breadth-first-search tree in $G'$ rooted at $v_2$. Like before, this gives an $i_2$ and sets $V^{(2)}_0, \ldots, V^{(2)}_{i_2+1}, U_2$. We then repeat this on the graph $G'' := G' \setminus \left(U_2 \cup V^{(2)}_{i_1 + 1} \right)$ and continue doing this until we have no vertices left. By construction, the desired properties hold. Indeed, the only vertices not contained in $\bigcup_j U_j$ are in some $ V^{(j)}_{i_j+1}$, hence there are at most $\gamma n$ of them. For the second property, observe that each $U_j$ is of size at least $(1+\gamma)^{i_j}$, so $i_j\leq \log_{1+\gamma}n$. The third condition holds by construction, since we deleted all the neighbors of $U_i$ before defining $U_{i+1}$ . \end{proof} \noindent Finally, we state a trivial observation used throughout the proof of Theorem \ref{thm:main}. It will be used to state that appropriate combinations of internally vertex-disjoint paths of different lengths result in the construction of cycles of many different lengths. \begin{obs}\label{obs:combining} Let $G$ be a graph whose vertex set contains $t$ disjoint sets $S_1,\ldots, S_t$ and another set of $t$ vertices $v_1,\ldots,v_t$ outside of $\bigcup_{i=1}^t S_i$. For each $i\in[t]$, let $A_i\subset \mathbb N$ and suppose that for every $i$ the induced subgraph $G[v_i\cup S_i\cup v_{i+1}]$ is such that it contains a $v_iv_{i+1}$-path of length $\ell$ for each $\ell\in A_i$ (with $v_{t+1}=v_1$). Then for every $\ell\in A_1+\ldots+A_t$, the graph $G$ contains a cycle of length $\ell$. \end{obs} \subsection{Finding consecutive path lengths} Next we show that in a graph with small independence number, we can find two vertices between which there exist paths of almost all `possible' lengths. We believe that this result is of independent interest and pose a problem related to it after its proof. \begin{prop}\label{lem:pathlengthsininterval} Let $G$ be a $n$-vertex graph with $\alpha(G) = k$ and let $0<\gamma <1/2$. Then there exist two vertices $u,v \in V(G)$ such that for every $\ell\in[\log_{1+\gamma}n, (1-\gamma)\frac{n}{k}]$ there is a $uv$-path of length $\ell$. \end{prop} \begin{proof} First, we apply Lemma \ref{lem:BFSpartition} to $G$ to get vertices $v_i$ and sets $U_i$ for all $i\leq r$. Fix the graph $H := G[U_1 \cup \ldots \cup U_r]$ and let us orient the edges of $H$ in the following manner. For an edge $xy$ in $H$ with $x,y \in U_j$ (recall property \textit{3} of the sets $U_j$) orient it as $x \rightarrow y$ if $\text{dist}(v_j,x) < \text{dist}(v_j,y)$ and as $y \rightarrow x$ if $\text{dist}(v_j,x) > \text{dist}(v_j,y)$. In the case that $\text{dist}(v_j,x) = \text{dist}(v_j,y)$, orient the edge arbitrarily. Now, since $\alpha(H) \leq \alpha(G) = k$ and $|H| \geq (1-\gamma)n$, by Lemma \ref{lem:partition} there must exist a directed path $\overrightarrow{P} = x_1 \rightarrow x_2 \rightarrow \ldots \rightarrow x_m$ in $H$ of length $m := \frac{|H|}{\alpha(H)} \geq (1-\gamma)\frac{n}{k}$. Let $j$ be such that $\overrightarrow{P} \subseteq U_j$, and for each $i\in[m]$ denote $d_i:=\text{dist}(v_j,x_i)$ and note that $d_i\leq \log_{1+\gamma}n$. Now we show that for all $\ell \in \left[d_m, m + d_1 \right] \supseteq \left[\log_{1+\gamma} n, m\right]$ there is path in $G$ of length $\ell$ between $v_j$ and $x_m$. For each $i\in[m]$ look at the $v_jx_m$-path $P_i$ obtained by concatenating the shortest $v_jx_i$-path with the path $x_{i+1}x_{i+2}\ldots x_m$. Since by definition, we have that $d_i \leq d_{\ell}$ for $i<\ell$, these two paths are vertex disjoint and their union is indeed a path as well. Moreover, for each $i$, we have that $|P_i|-1\leq|P_{i+1}|\leq |P_i|$ since $x_i \rightarrow x_{i+1}$ is an edge and thus, $d_i \leq d_{i+1} \leq d_i +1 $. Because $|P_1|=d_1+m$ and $|P_m|=d_m$, each path length in $\left[d_m, m + d_1 \right]$ is then attained by at least one of the constructed paths. Since $d_m\leq \log_{1+\gamma}n$ and $m \geq (1-\gamma)\frac{n}{k}$, this finishes the proof. \end{proof} Before moving on to the next section, it is worth noting that the above proposition is asymptotically tight. Indeed, an $n$-vertex graph $G$ with $\alpha(G) = k$ does not even necessarily need to contain a path of length larger than $\frac{n}{k}$, as we can see from a disjoint union of cliques of size $n/k$. Since the proof of this proposition uses a result about directed paths, we further ask if the following directed variant of it might be is true as well. \begin{prob} Does Proposition \ref{lem:pathlengthsininterval} generalize to directed graphs? If $G$ is a directed graph, how large an interval $I \subseteq [0,\frac{n}{k}]$ can we guarantee for which there are vertices $u,v$ with a directed $uv$-path of length $\ell$ for all $\ell \in I$? \end{prob} \subsection{Path shortening} In this section we show that if a graph has small independence number and contains a long path $P$, then we can find a slightly shorter path $P'$ with the same endpoints, which satisfies certain additional properties. In a graph with independence number $k$, a path can clearly be shortened by considering $2k+1$ consecutive vertices on the path, and observing that there must be an edge (not contained in the path) between those vertices. This is the statement of the next simple lemma. \begin{lem}\label{lem:easyjump} Let $G$ be a graph with independence number $k$, and $P$ a path in $G$ with endpoints $x,y$ such that $|P| > 2k$. Then, there is an $xy$-path $P'$ on the same vertex set $V(P)$ with $|P| - 2k \leq |P'| < |P|$. \end{lem} \noindent The following proposition is one of the central results of this paper. It shows that we can shorten a path only by a little while also preserving a pre-specified set of vertices in the newly obtained path. \begin{prop}\label{lem:jumpwithzigzag} Let $G$ be an $n$-vertex graph with independence number $k$, let $P$ be a path in $G$ with endpoints $x,y$, let $c \in \mathbb{N}$ and $U \subseteq V(P)$. Then there is an $xy$-path $P'$ with the following properties. \begin{enumerate} \item $U \subseteq V(P') \subseteq V(P)$. \item $|P| - (4c-3) \leq |P'| < |P|$ if for $c$ it holds that $c \left(\frac{|P|-(4c-1)|U|}{2k}-1 \right)>k$. \end{enumerate} \end{prop} \noindent Before we give a proof of Proposition~\ref{lem:jumpwithzigzag}, we introduce some useful concepts. The key idea to prove the proposition is to find a certain structure in our graph which can be used to shorten a path in a graph with low independence number. The properties of this structure are captured by the notion of a \emph{special edge set}, defined below. \begin{defn}\label{def:special} Given a graph $H$ with ordered vertex set $(1,2,\ldots, n)$, we say that a sequence of vertices $(v_1,\ldots,v_{\ell})$ is \emph{special} if the following hold (see Figure \ref{fig:special set}). \begin{itemize} \item $v_{i+1} > v_{i}$ for all $i\in[\ell-1]$. \item $(v_{i},v_{i+1}+1)$ is an edge in $G$ for all $i\in [\ell-1]$. We call the set formed by those edges a \emph{special edge set}. \end{itemize} \end{defn} \begin{figure}[ht] \centering \begin{tikzpicture}[scale=1.5,main node/.style={circle,draw,color=black,fill=black,inner sep=0pt,minimum width=3pt}] \tikzset{cross/.style={cross out, draw=black, fill=none, minimum size=2*(#1-\pgflinewidth), inner sep=0pt, outer sep=0pt}, cross/.default={2pt}} \tikzset{rectangle/.append style={draw=brown, ultra thick, fill=red!30}} \foreach \i in {1,...,46} { \node[main node, scale=0.4] (aux) at (\i*0.2-0.6,0){}; } \node[rectangle, color=red, scale=0.4] (a) at (0,0) [label=below:$v_1$]{}; \node[rectangle, color=red, scale=0.4] (a1) at (2,0)[label=below:$v_2$]{}; \node[rectangle, color=red, scale=0.4] (a2) at (3,0)[label=below:$v_3$]{}; \node[rectangle, color=red, scale=0.4] (a3) at (5.4,0)[label=below:$v_4$]{}; \node[rectangle, color=red, scale=0.4] (a4) at (8,0)[label=below:$v_5$]{}; \node[main node] (a4) at (8.2,0){}; \node[main node] (b1) at (2.2,0){}; \node[main node] (b2) at (3.2,0){}; \node[main node] (b3) at (5.6,0){}; \draw[line width= 1 pt] (a) to [bend left=60](b1); \draw[line width= 1 pt] (a1) to [bend left=60](b2); \draw[line width= 1 pt] (a2) to [bend left=60](b3); \draw[line width= 1 pt] (a3) to [bend left=60](a4); \end{tikzpicture} \caption{An illustration of a special edge set. The special vertices are colored red.} \label{fig:special set} \end{figure} \noindent We can now find special vertex sequences in graphs in the following manner. \begin{lem}\label{lem:zigzag} Let $H$ be a graph on the vertex set $[n]$, let $U' \subseteq [n]$ and suppose $H$ has independence number $k$. Then, there exist a special vertex sequence with at least $\frac{n-|U'|}{2k}$ vertices in $[n]\setminus U'$. \end{lem} \begin{proof} Let $H'$ be the graph obtained from $H$ by removing all edges of the form $\{i,i+1\}$ where $|u-v| = 1$. Since the removed edges form a graph with chromatic number at most 2, we must have that $\alpha (H') \leq 2 \alpha(G) \leq 2k$. Let us also direct the edges of $H'$ so that an edge $ij$ is oriented from $i$ to $j$ if $i < j$. Applying Lemma~\ref{lem:partition} to this directed graph, we obtain a partition of $[n]$ into at most $2k$ directed paths. Denote the digraph formed by the union of those paths as $F$. An important property of $F$ that we are going to use is that all outdegrees and indegreees of vertices in $F$ is at most one. Having obtained this path cover, we want to find another decomposition but now of the edges of $F$, and into a small number of special edge sets. Simply take $\mathcal{M}$ to be a smallest collection of edge disjoint special edge sets which decompose the edges of $F$ (this exists as one such collection is the set of all edges in $F$). If $(v_1,v_2,\ldots,v_\ell)$ is the special sequence corresponding to a special edge set $M$ in $\mathcal{M}$, then note that $v_\ell$ must be a vertex of out-degree $0$ in $F$. Indeed, if this is not the case, then there exists $u > v_\ell$ such that $v_\ell \rightarrow u$ is an edge of $F$. Let $M' \in \mathcal{M}$ be the special edge set which contains it and note that then $v_\ell$ is the first vertex of this special edge set, since otherwise it would contain the edge $\{v_{\ell -1}, v_{\ell}+1\} \in M$ and contradict the edge-disjointness of the special edge sets in $\mathcal{M}$. But now we can 'concatenate' $M$ and $M'$ to form a larger special edge set $M \cup M'$, which contradicts the minimality of $\mathcal{M}$. Notice that the special edge sets in $\mathcal{M}$ have disjoint corresponding special vertex sequences. Indeed, since these special edge sets are non-empty and edge-disjoint, the only possibility for a common vertex in two different special vertex sequences would be if some vertex $v$ is the first vertex of one sequence and the last vertex of another one. In turn, as shown in the previous paragraph, this would contradict the maximality of $\mathcal{M}$. Now, a consequence of having disjoint corresponding special vertex sequences in $\mathcal{M}$ and the previous paragraph is that each one of these has a unique vertex of out-degree $0$. Moreover, let $S \subseteq V(H)$ denote the set of vertices which do not belong to any special vertex sequence of $\mathcal{M}$ and notice that all $v \in S$ also have out-degree $0$ in $F$. In turn, since $F$ is a decomposition of $[n]$ with at most $2k$ paths, there are at most $2k$ vertices of outdegree zero and thus $|\mathcal{M}| + |S| \leq 2k$. Hence, there exists one special vertex sequence in $\mathcal{M} \cup S$ (allowing a single vertex to be a special vertex sequence) with at least $\frac{n - |U'|}{|\mathcal{M}|+|S|} \geq \frac{n-|U'|}{2k}$ vertices not in $U'$, which finishes the proof. \end{proof} \noindent We can now show the announced proof of our proposition. \begin{proof}[ of Proposition \ref{lem:jumpwithzigzag}] Let $(1,2,\ldots,|P|)$ be an ordering of the vertices of $P$, with the endpoints $x=1$ and $y=|P|$. Let $U_c$ be the set of vertices $x$ in $P$ such that there is a $u\in U$ with $|x-u|\leq 2c-1$. By applying Lemma~\ref{lem:zigzag} to $G[V(P)]$ and the set $U_c$, we obtain a special vertex sequence $(v_1,\ldots,v_\ell)$ corresponding to a special edge set $M$, with at least $\frac{|P|-|U_c|}{2k}$ vertices outside of $U_c$. First, notice that for each $v_{i+1}\notin U_c$ we may assume that $v_{i+1}-v_i\geq 2c$. Indeed, if that was not the case, then we obtain the desired path $P'$ from $P$ by replacing the interval $[v_i,v_{i+1}+1]$ by the edge $(v_i,v_{i+1}+1)$, noting that there are no vertices from $U$ inside of the removed interval, since otherwise $v_{i+1}$ would be in $U_c$, a contradiction. Now we define, for each special vertex $v = v_i$ which is not in $U_c\cup \{v_1\}$, the $c$-element set $S_v:=\{v-1, v-3,\ldots, v-(2c-1)\}$ contained in the $(2c-1)$-element interval $I_v:=[v-(2c-1),v-1]$, which is disjoint to $U$. Since $v_{i}-v_{i-1}\geq 2c$, all of these sets are disjoint and further, disjoint to $U$. Now, the union $S$ of those sets is of size at least $\left(\frac{|P|-|U_c|}{2k}-1 \right)c$. Therefore, since $|U_c|\leq (4c-1)|U|$, we get that $|S|\geq k+1$ by our assumption on $c$. Hence, there exists an edge in $G$ spanned by $S$, since the independence number of $G$ is $k$. We may assume that this edge does not lie inside some $S_v$, as otherwise we can again get the desired path $P'$ by using this edge instead of the interval which it bridges, avoiding at least one, but at most $2c-1$ vertices which are not in $U$. Hence, the found edge $ab$ with $a<b$ is between two distinct sets $S_{v_i}$ and $S_{v_j}$. Now we can find the required path $P'$ as shown in Figure~\ref{fig:n-1}, avoiding at most $4c-3$ vertices. \noindent More precisely, we obtain the required path $P'$ as the union of the following paths: \begin{itemize} \item The part of the path $P$ which connects $x$ to $a$, plus the edge $(a,b)$. \item The increasing path $P_2$ obtained by the following iterative procedure. First, initialize $P_2$ to be the edge $(v_i,v_i+1)$. Repeat the following. Let $r$ be the last vertex of $P_2$. If $r$ is a special vertex $r=v_t$, then add the edge $(r,v_{t+1}+1)$ to $P_2$, and update $r=v_{t+1}+1$. If $r$ is not a special vertex, add the edge $(r,r+1)$ to $P_2$, and update $r=r+1$. We stop when either $r=b$ or $r=v_j+1$. \item The increasing path $P_3$ obtained by the following iterative procedure. First we initialize $P_3$ to be the edge $(v_i,v_{i+1}+1)$. Repeat the following (exactly as for the previous path). Let $r$ be the last vertex of $P_3$. If $r$ is a special vertex $r=v_t$, then add the edge $(r,v_{t+1}+1)$ to $P_3$, and update $r=v_{t+1}+1$. If $r$ is not a special vertex, add the edge $(r,r+1)$ to $P_3$, and update $r=r+1$. We stop when either $r=b$ or $r=v_j+1$. \item The part of the path $P$ which connects $v_j+1$ to $y$. \end{itemize} It is easy to see that this path contains only vertices of $P$ and that it does not contain the vertex $v_j$. Furthermore, the only other vertices from $P$ which the new path $P'$ avoids are the vertices in $I_{v_i}$ which are strictly larger than $a$, and the vertices in $I_{v_j}$ which are strictly larger than $b$. So in total, the new path $P'$ avoids at most $1+(|I_{v_i}|-1)+(|I_{v_j}|-1)=4c-3$ vertices of $P$, which completes the proof. \end{proof} \begin{figure}[ht] \centering \begin{tikzpicture}[scale=1,main node/.style={circle,draw,color=black,fill=black,inner sep=0pt,minimum width=7pt}] \tikzset{rectangle/.append style={draw=brown, ultra thick, fill=red!30}} \node[rectangle, color=black, scale = 0.3] (s1) at (0,0) {$x$}; \node[rectangle, color=black, scale = 0.3] (s2) at (17,0) {$y$}; \node[rectangle, scale=0.6, color=red, opacity=1] (a1) at (1,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a2) at (4,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a3) at (6,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a4) at (9,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a5) at (11,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a6) at (13,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a7) at (16,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b0) at (1.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b1) at (4.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b2) at (6.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b3) at (9.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b4) at (11.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b5) at (13.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b6) at (16.3,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p1) at (3.7,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p11) at (3.4,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p111) at (3.1,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p3) at (8.7,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p4) at (8.4,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p5) at (8.1,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p6) at (12.7,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p6) at (12.4,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p6) at (12.1,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p7) at (15.7,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p5) at (15.4,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p9) at (15.1,0){}; \draw[line width= 1 pt] (a1) to [bend left=50](b1); \foreach \i in {2,...,6} { \draw[line width= 2 pt, color = blue] (a\i) to [bend left=50](b\i); } \draw[line width= 2 pt, color = blue] (p11) to [bend right=28](p9); \draw[line width= 2 pt, color = blue] (p11) to (s1); \draw[line width= 2 pt, color = blue] (a2) to (a3); \draw[line width= 2 pt, color = blue] (b2) to (a4); \draw[line width= 2 pt, color = blue] (b3) to (a5); \draw[line width= 2 pt, color = blue] (b4) to (a6); \draw[line width= 2 pt, color = blue] (b5) to (p9); \draw[line width= 2 pt, color = blue] (b6) to (s2); \node[rectangle, scale=0.6, color=red, opacity=1] (a1) at (1,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a2) at (4,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a3) at (6,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a4) at (9,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a5) at (11,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a6) at (13,0){}; \node[rectangle, scale=0.6, color=red, opacity=1] (a7) at (16,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b0) at (1.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b1) at (4.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b2) at (6.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b3) at (9.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b4) at (11.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b5) at (13.3,0){}; \node[main node, scale=0.6, color=black, opacity=1] (b6) at (16.3,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p1) at (3.7,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p11) at (3.4,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p111) at (3.1,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p3) at (8.7,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p4) at (8.4,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p5) at (8.1,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p6) at (12.7,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p6) at (12.4,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p6) at (12.1,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p7) at (15.7,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p5) at (15.4,0){}; \node[main node, scale=0.6, fill=white, opacity=1] (p9) at (15.1,0){}; \node[scale=1] (a) at (0, -0.3){$x$}; \node[scale=1] (a) at (17,-0.3){$y$}; \node[scale=1] (a) at (3.2,0.4){$S_{v_i}$}; \node[scale=1] (a) at (15.4,0.3){$S_{v_j}$}; \node[scale=1] (a) at (3.4,-0.2){$a$}; \node[scale=1] (a) at (15.1,-0.2){$b$}; \end{tikzpicture} \caption{ The shorter path $P'$ is drawn in blue, the special vertices are represented with red squares, while the vertices after them on $P$ are represented with black dots. The vertices in the sets $S_v$ are represented with white dots. Also note the following about the black dots and red squares which are between $a$ and $b$ in $P$. It could be that one of these black dots and the red square which is after it in our drawing, are the same vertex. However, the construction of the path $P'$ remains the same (for example, the fourth black dot and the fifth red square could be the same vertex). Notice also in the drawing that some of the red squares do not have a corresponding set of white dots. These are precisely those special vertices which are in $U_c \cup \{v_1\}$. } \label{fig:n-1} \end{figure} \section{Proof of Theorem \ref{thm:main}}\label{sec:proof} Let $\varepsilon>0$ be a small enough constant, and let $k$ be sufficiently large in terms of $\varepsilon$. Let $G$ be a Hamiltonian graph with $\alpha(G)\leq k$ on $n \geq (2+\varepsilon)k^2$ vertices. Our goal is to prove that $G$ is pancyclic. It will be convenient for us to consider different ranges of cycle lengths, and for each range we have a separate subsection which deals with it. \subsection{Lower range: from $3$ to $(2+\varepsilon)k$} Showing that $G$ contains all cycles of lengths between $3$ and $(2+\varepsilon)k$ only requires the fact that $G$ has no independent set of size $k+1$. Indeed, this boils down to the study of cycle-complete Ramsey numbers. Namely, the cycle-complete Ramsey number $r(C_\ell,K_s)$ is the smallest number $N$ such that every graph on $N$ vertices either contains a copy of $C_\ell$ or an independent set of size $s$. The following result of Erd\H{o}s, Faudree, Rousseau and Schelp \cite{erdos1978cycle}, along with a more recent result by Keevash, Long and Skokan \cite{keevash2021cycle} cover the mentioned range of cycle lengths we need. \begin{thm}[\cite{erdos1978cycle}]\label{thm:erdos cycle-complete} Let $\ell\geq 3$ and $s\geq 2$. Then $r(C_\ell, K_s) \leq\left((\ell-2)(s^{1/x}+2)+1\right)(s-1)$, where $x=\lfloor \frac{\ell-1}{2}\rfloor$. \end{thm} The next result by Keevash, Long and Skokan gives the precise behaviour of cycle-complete Ramsey numbers in a wide range of parameters, and proves a conjecture from \cite{erdos1978cycle}. \begin{thm}[\cite{keevash2021cycle}]\label{precise cycle-complete} There exists $C \geq 1$ so that $r(C_\ell , K_s) = (\ell - 1)(s- 1) + 1$ for $s \geq 3$ and $\ell\geq C \frac{\log s}{\log\log s}$. \end{thm} Now, note that since $G$ contains no independent set of size $k+1$, Theorem~\ref{thm:erdos cycle-complete} implies the existence of a cycle of length $\ell$ for every $\ell\in[3,\log k]$, while Theorem~\ref{precise cycle-complete} covers the range of $[\log k,(2+\varepsilon)k]$. \subsection{Upper range: from $\frac{1000}{\varepsilon^2}k$ to $n$}\label{sec:upperrange} First, note that all cycle lengths in $[2k^2+2k,n]$ can be obtained by iteratively applying Proposition~\ref{lem:jumpwithzigzag} with $c=1$, and $U=\emptyset$, and always shortening the cycle by one. Indeed, denote by $P$ any Hamilton path contained in a Hamilton cycle in $G$, and let $x$ and $y$ be its endpoints. As long as $n> 2k^2+2k$, by applying the mentioned proposition we get a path which is by $c=1$ shorter than $P$ and has the same endpoints, so adding the edge $xy$ to it creates a cycle of length $n-1$. We remove the remaining vertex and repeat. This gives all cycle lengths in $[2k^2+2k,n]$. Now we turn to the cycle lengths in $\left[\frac{1000}{\varepsilon^2}k,2k^2+2k\right]$. For this we need the following lemma. \begin{lem}\label{lem:partitionintomatchingcycle} Let $G$ be a Hamiltonian graph on $n$ vertices with independence number $k$. Then, there is a partition of the vertices of $G$ into a cycle $C$ and a set $S$ of size $|S|= \frac{\varepsilon n}{20}$, such that there is a matching $M \subseteq E[C,S]$ which covers $S$. \end{lem} \begin{proof} To show this claim, we apply Proposition~\ref{lem:jumpwithzigzag} iteratively $\varepsilon n/20$ times as follows. We always have $c=1$, and in the beginning we set $U_M=S=\emptyset$, and we set $M$ to be an empty matching and $C$ a Hamilton cycle in $G$. During the procedure $C$ and $S$ are always disjoint and partition the vertices of $G$, $M$ is a matching in $E[C,S]$ which covers $S$, and $U_M$ is the set of endpoints of $M$ in $C$. In the first step we apply Proposition~\ref{lem:jumpwithzigzag} with the mentioned values of $c=1$ and $U=\emptyset$, to get a cycle $C'$ of length $n-1$, and a vertex $v$ which is not on $C'$. Let $v'$ be a neighbor of $v$ in the cycle $C$. Now we set $C=C'$, $S = \{v\}$, $M=\{vv'\}$ and $U_M=\{v'\}$. In the $i$-th step of the procedure, we let $U$ denote the set of vertices which are either in $U_M$, or adjacent to a vertex in $U_M$ on the current cycle $C$. We apply Proposition~\ref{lem:jumpwithzigzag} with $c=1$ and $U$ to the graph $G[C]$, to get a cycle $C'$ of length $|C|-1$ and a vertex $v \in C \setminus C'$ which is not in $U$. Again, we denote by $v'$ the neighbor of $v$ in $C$, and we set $M:=M\cup \{vv'\}$, $S := S \cup \{v\}$, $U_M=U_M \cup\{v'\}$ and $C=C'$. Notice that since we only perform $\varepsilon n/20$ steps, at each point we have that $|U_M|\leq \varepsilon n/20$, and that $|C|\geq n-\varepsilon n/20$. Together with the fact that $|U|\leq 3|U_M|$, this gives that $\frac{|C|-3|U|}{2k}-1> k$, so we can always successfully apply Proposition~\ref{lem:jumpwithzigzag} with $c=1$. The resulting matching is then of size $\frac{\varepsilon n}{20}$, and evidently satisfies the given requirements. \end{proof} We are ready to show how to get the cycle lengths in $\left[\frac{1000}{\varepsilon^2}k,2k^2+2k\right]$. We first apply the above lemma to get a cycle $C$ of length $n-\frac{\varepsilon n}{20}\geq 2k^2+\frac{2\varepsilon k^2}{3}$ and outside of it a set $S$ of size $|S|=\frac{\varepsilon n}{20}$, together with a matching $M$ between them which covers $S$. Split the cycle $C$ into $4/\varepsilon$ intervals of (almost) equal size. By pigeonholing, at least one of those intervals contains at least $\frac{\varepsilon^2n}{80}$ endpoints of $M$. Let $S'$ be the subset of $S$ of vertices corresponding to those endpoints in $M$. We apply Lemma~\ref{lem:pathlengthsininterval} to the graph $G[S']$ with say $\gamma=1/100$, and conclude that there are two vertices $x'$ and $y'$ in $G[S']$, between which there exists a path of length $\ell$ in $G[S']$, for every $\ell\in [\frac{\varepsilon^2k}{100}, \frac{\varepsilon^2k}{50}]$. Let $x$ and $y$ be the vertices in $C$ corresponding to $x'$ and $y'$ in $M$. Let $P$ be the longer path in $C$ which connects $x$ and $y$, so that by the choice of $S'$ we have $|P|\geq |C|-\frac{\varepsilon n}{4}>2k^2+2k$. Now we use Proposition~\ref{lem:jumpwithzigzag} to show that in $G[P]$ the pair $(x,y)$ is $\frac{\varepsilon^2 k}{100}$-dense in $\left[\frac{900k}{\varepsilon^2},2k^2+2k\right]$. Denote $P_0:=P$, and we obtain the path $P_i$ from path $P_{i-1}$ by applying Proposition~\ref{lem:jumpwithzigzag} with $c=\frac{\varepsilon^2k}{400}$ and $U=\emptyset$. We do this until $|P_i|<\frac{900k}{\varepsilon^2}$ and then we stop our procedure. Notice we could do each step of the procedure as we always had $c \left(\frac{|P_i|}{2k}-1 \right)>k$. Hence we obtain a sequence of $xy$-paths of decreasing lengths, where $|P_i|\geq|P_{i-1}|-4c+3\geq |P_{i-1}|-\frac{\varepsilon^2k}{100}$. Since $|P_0|>2k^2+2k$ and the last path is of length at most $\frac{900k}{\varepsilon^2}$, we indeed get that in $G[P]$ the pair $(x,y)$ is $\frac{\varepsilon^2 k}{100}$-dense in $\left[\frac{900k}{\varepsilon^2},2k^2+2k\right]$. Now, applying \Cref{obs:combining} to the graph $G[P\cup S']$ with $v_1=x$ and $v_2=y$, gives all cycle lengths in $\left[\frac{1000k}{\varepsilon^2},2k^2+2k\right]$. \subsection{Middle range : from $(2+\varepsilon) k$ to $\frac{1000}{\varepsilon^2}k$}\label{sec:middlerange} By Lemma \ref{lem:partitionintomatchingcycle}, there exists a partition of the vertices of $G$ into a cycle $C$ and a set $S$ such that $|S| = \varepsilon n/20$, along with a matching $M \subseteq E[C,S]$ which covers $S$. Denote the vertices along the cycle $C$ with $(1,\ldots, N)$. For each vertex $x \in S$, we denote by $m(x)$ the vertex in $C$ matched to $x$ in $M$. We now remove from $S$ the at most $\frac{1000k}{\varepsilon^2}$ vertices $x$ which have $m(x)\in \{1,2,\ldots,\frac{1000k}{\varepsilon^2}\}$. Hence, there are at least $\varepsilon n/22$ vertices remaining in $S$. We let $B_1:=S$ and apply Proposition \ref{lem:pathlengthsininterval} with $\gamma=1/100$ to the graph $G[B_1]$, to find a pair of vertices $x_1,y_1$ such that for all $\ell \in \left[\frac{\varepsilon k}{100}, \frac{\varepsilon k}{50}\right]$, there is an $x_1y_1$ path of length $\ell$ in $G[S]$; we set $B_2:=B_1-\{x_1,y_1\}$. We repeat this $t= \varepsilon k^2/40$ times, i.e., we apply Proposition~\ref{lem:pathlengthsininterval} to $B_i$ to obtain vertices $x_i,y_i$ such that for all $\ell \in \left[\frac{\varepsilon k}{100}, \frac{\varepsilon k}{50}\right]$, there is an $x_iy_i$-path of length $\ell$ in $G[B_i]\subseteq G[S]$, and we set $B_{i+1}:=B_i-\{x_i,y_i\}$. Note that each $B_i$ is of size at least $|S|-2t\geq \frac{\varepsilon n}{22} - 2t \geq \frac{\varepsilon k^2}{11} - \frac{\varepsilon k^2}{20} \geq \frac{\varepsilon k^2}{30}$, so we always can successfully apply Proposition~\ref{lem:pathlengthsininterval}. Before we make our first crucial observation, we introduce some notation. First, with possible relabeling, let us suppose that for each $i$, we have that $m(x_i) < m(y_i)$. Now, for each $i$, let $P_i$ denote the subpath $\left(m(x_i)-\frac{1000k}{\varepsilon^2}, m(x_i)-\frac{1000k}{\varepsilon^2}-1,\ldots, m(x_i) -1,m(x_i)\right)$ of length $\frac{1000k}{\varepsilon^2}$. Notice that, since $m(x_i)>\frac{1000k}{\varepsilon^2}$, none of those paths contains vertex $1$. \begin{lem} \label{induced} If there is an $i$ such that $G[P_i]$ does not contain an increasing path of length $\varepsilon^3k$ as an induced subgraph with $m(x_i)$ as an endpoint \footnote{Recall that a path $P$ is an induced subgraph of $G[P_i]$ if its vertices belong to $V(P_i)$ and except for the edges of $P$, $G[P]$ does not contain any other edges.}, then $G$ contains all cycle lengths in $\left[(2+\varepsilon)k,\frac{1000k}{\varepsilon^2}\right]$. \end{lem} \begin{proof} Assume for sake of contradiction that $G[P_i]$ does not contain such a path. Then, the following holds. \begin{claim*} In the graph $G[P_i]$, the endpoints of $P_i$ are $\varepsilon^3k$-dense in $\left[0,\frac{1000k}{\varepsilon^2}\right]$. \end{claim*} \begin{proof}[ of Claim] Consider the following procedure. We begin with $P_i$; by assumption, there exists a chordal edge among the last $\varepsilon^3 k$ vertices of $P_i$ ending with $m(x_i)$. Otherwise, these vertices would induce an increasing path in $G[P_i]$. Thus we obtain a path $P_i'$ by adding this edge to $P_i$ instead of the interval between the endpoints of this edge. We repeat this procedure, each time finding a chordal edge in the newly obtained path, and we do this until our path has length at most $\varepsilon^3k$. Hence, we obtain a sequence of paths such that two consecutive paths lengths are at most $\varepsilon^3k$ apart, while the last path has length at most $\varepsilon^3k$. Since the endpoints always remain the same, this implies the statement of the claim. \end{proof} \noindent Consider now the path $P$ contained in the cycle $C$, and which is spanned by the vertices in the interval $\left[m(y_i),N]\cup[1,m(x_i)-\frac{1000k}{\varepsilon^2}\right]$. By Lemma \ref{lem:easyjump}, there exists a path $P'$ with the following properties: $V(P') \subseteq V(P)$, the endpoints of $P'$ are the same as those of $P$, and $|P'| \leq 2k$. Then, in order to finish, recall that $x_i,y_i$ are such that for all $\ell \in \left[\frac{\varepsilon k}{100}, \frac{\varepsilon k}{50}\right]$, there exists an $x_iy_i$-path of length $\ell$ in $G[S]$. Further, by the above claim we got that in the graph $G[P_i]$, the endpoints of $P_i$ are $\varepsilon^3k$-dense in $\left[0,\frac{1000k}{\varepsilon^2}\right]$. Hence we can use \Cref{obs:combining} on the graph $G$ with $v_1=m(x_i), v_2=m(x_i)-\frac{1000k}{\varepsilon^2}$ and $v_3=m(y_i)$, where $S_1$ are the internal vertices of $P_i$, $S_2$ the internal vertices of $P$ and $S_3=S$ to obtain all cycle lengths in $\left[(2+\varepsilon)k,\frac{1000}{\varepsilon^2}k\right]$. \end{proof} We have $t=\varepsilon k^2/40$ paths $P_i$ of length $1000k/\varepsilon^2$ such that each of them corresponds to an interval of vertices in $C$ and therefore intersects at most $2000k/\varepsilon^2$ other such paths. Thus, we can choose a collection of $r=\frac{t}{2000k/\varepsilon^2+1}\geq \sqrt{k}$ of those paths which are all disjoint. With possible renaming, w.l.o.g.\ we may assume that those paths are $P_1,\ldots,P_r$. Using Lemma \ref{induced}, we may also assume that there are induced increasing subpaths $Q_1, \dots,Q_r$ of $G[P_1],\ldots,G[P_r]$ with endpoints $m(x_1),\ldots,m(x_r)$ respectively and of length $\varepsilon^3 k$. Let us now define an auxiliary colored complete graph $H$ on $[r]$ in the following manner. For each $i \in [r]$, partition $Q_i$ into three consecutive subpaths $Q^{3}_i,Q^{2}_i,Q^{1}_i$ of size $|Q_i|/3=\varepsilon^3 k/3$, with $Q^{1}_i$ containing $m(x_i)$. Now, for $i,j \in [r]$, we color the edge $ij$ in $H$ \emph{red} if in $G$ both $E[Q^{1}_i,Q^{1}_j]$ and $ E[Q^{3}_i,Q^{3}_j]$ are non-empty. We color it \emph{blue} if $E[Q^{1}_i,Q^{1}_j] = \emptyset$, and in the remaining case, we color it \emph{green}. \begin{claim*} There are no blue or green cliques in $H$ of size larger than $6/\varepsilon^3$. \end{claim*} \begin{proof} Suppose there exists a blue clique $\{i_1, \ldots, i_\ell\}$ in $H$. Since each $Q^{1}_{i_j}$ is an induced path, its odd vertices form an independent set of size $|Q^{1}_{i_j} |/2$. Moreover, by assumption, there are no edges between two $Q^{1}_{i_j}$'s and therefore the set $\bigcup_{1 \leq j \leq \ell} V(Q^{1}_{i_j})$ must contain an independent set of size at least $$\sum_{1 \leq j \leq \ell} \frac{|Q^{1}_{i_j}|}{2} \geq \ell \cdot \left(\varepsilon^3 k/6 \right) .$$ Since $\alpha(G) \leq k$, this implies that $\ell \leq 6/\varepsilon^3$. An analogous argument deals with green cliques. \end{proof} Given the above claim and Theorem \ref{lem:ramsey}, and since $H$ has $r>\sqrt{k}$ vertices where we chose $k$ large enough in terms of $\varepsilon$, we have that there exists a red clique in $H$ of size at least $4\varepsilon^{-7}$. Denote by $I$ the vertices/indices contained in this clique, so that for all $i,j\in I$ we have that there is an edge $e^1_{ij}$ between $Q^{1}_i$ and $Q^{1}_j$ and an edge $e^3_{ij}$ between $ Q^{3}_i$ and $Q^{3}_j$. For simplicity of notation, w.l.o.g.\ we may assume that the indices in $I$ are $\{1, 2, 3, \ldots, |I|\}$ according to the ordering of the vertices $m(x_i)$ for $i \in I$ - that is, we now have $m(x_1) < m(x_2) < \ldots < m(x_{|I|})$. Denote by $z$ the endpoint of $Q_1$ which is not $m(x_1)$. In order to complete the proof, we will need the following lemma. \begin{lem}\label{lem:jump with Q} The path $Q$, defined by the interval $\left[z,m(x_{|I|})\right]$ is such that its endpoints are $3\varepsilon^3k$-dense in $\left[0,\frac{k}{\varepsilon^4}\right]$ in the graph $G[Q]$. \end{lem} \begin{proof} Denote by $R_1$ the path consisting only of vertex $z$. We now recursively define for each $i< |I|$ a path $R_i$ whose one endpoint is $z$ and the other endpoint $z_i$ lies in either $Q_i^1$ or $Q_i^3$ (see Figure~\ref{fig:jumpswithQ} for an illustration). First, let $z_1=z$. Suppose $z_i$ is in $Q_i^a$ for some $a\in\{1,3\}$, and let $b\in\{1,3\}\setminus\{a\}$; let $R_{i+1}$ be the path obtained from $R_{i}$ by concatenating to it the path contained in $Q_i$ which starts at $z_i$, goes through $Q_i^2$ and touches the edge $e^b_{i,i+1}$, and also add the edge $e^b_{i,i+1}$ itself. Now we also define paths $R_i'$ for each $i \leq |I|-1$, obtained from $R_i$ as follows. If $z_i\in Q^{a}_i$, again we let $b\neq a$ and $b\in\{1,3\}$. Let $R_i'$ be the path obtained by concatenating with $R_i$ the path starting at $z_i$, going through $Q_i$ until the edge $e^b_{i,|I|}$, then also adding this edge $e^b_{i,|I|}$ itself, together with the path in $Q_{|I|}$ which connects the endpoint of this edge with $m(x_{|I|})$. Note that the length of two paths $R_i'$ and $R_{i+1}'$ differs by at most $|Q_i|+|Q_{i+1}|+|Q_I|\leq 3\varepsilon^3k$, since the only vertices which belong to exactly one of these two paths are contained in $G[Q_i\cup Q_{i+1}\cup Q_I]$. Furthermore, the length of the first path $R'_1$ obtained by our procedure is at most $|Q_1| + |Q_{|I|}| \leq 2 \varepsilon^3 k$ and the length of the last path $R_{|I|-1}'$ is at least $(|I|-1)\varepsilon^3k/3\geq \frac{k}{\varepsilon^4}$, since it contains all paths $Q_i^2$ for all $i\leq |I|-1$. This implies that the path $Q$, defined as the path between $z=z_1$ and $m(x_{|I|})$ is such that its endpoints are $3\varepsilon^3k$-dense in $\left[0,\frac{k}{\varepsilon^4}\right]$ in the graph $G[Q]$. \end{proof} \begin{figure}[ht] \centering \begin{tikzpicture}[scale=0.62,main node/.style={circle,draw,color=black,fill=black,inner sep=0pt,minimum width=7pt}] \tikzset{rectangle/.append style={draw=brown, ultra thick, fill=blue!30}} \foreach \i in {1,...,4} { \node[main node, scale=0.5, color=black, opacity=1] (a\i) at (5.7*\i+1,0){}; \node[main node, scale=0.5, color=black, opacity=1] (b\i) at (5.7*\i+2,0){}; \node[main node, scale=0.5, color=black, opacity=1] (c\i) at (5.7*\i+3,0){}; \node[main node, scale=0.5, color=black, opacity=1] (d\i) at (5.7*\i+4,0){}; \draw[line width= 1 pt] (a\i) to (d\i); \node[main node, scale=0.5, color=blue, opacity=1] (z1) at (6.7,0){}; \node[main node, scale=0.5, color=blue, opacity=1] (t1) at (9.2,0){}; \draw[line width= 2 pt, color=blue] (z1) to (t1); \ifthenelse {\i>1 \and \i<5} { \pgfmathtruncatemacro\aa{(-1)^\i} \node[main node, scale=0.5, color=blue, opacity=1] (z\i) at (5.7*\i+2.5+\aa,0){}; \node[main node, scale=0.5, color=blue, opacity=1] (t\i) at (5.7*\i+2.5-\aa,0){}; \draw[line width= 2 pt, color=blue] (z\i) to (t\i); } {} \ifthenelse {\i<5 \and \i>1} { \pgfmathtruncatemacro\aa{\i-1} \draw[line width= 2 pt, color=blue] (z\i) to [bend right = 50] (t\aa); } {} } \foreach \i in{1,...,4} { \node[main node, scale=0.5, color=black, opacity=1] (a\i) at (5.7*\i+1,0){}; \node[main node, scale=0.5, color=black, opacity=1] (b\i) at (5.7*\i+2,0){}; \node[main node, scale=0.5, color=black, opacity=1] (c\i) at (5.7*\i+3,0){}; \node[main node, scale=0.5, color=black, opacity=1] (d\i) at (5.7*\i+4,0){}; \node[scale=0.8, color=black, opacity=1] (t) at (5.7*\i+1.5,-0.6){$Q_\i^3$}; \node[scale=0.8, color=black, opacity=1] (t) at (5.7*\i+2.5,-0.6){$Q_\i^2$}; \node[scale=0.8, color=black, opacity=1] (t) at (5.7*\i+3.5,-0.6){$Q_\i^1$}; } \node[main node, scale=0.5, color=black, opacity=1] (a) at (31,0){}; \node[main node, scale=0.5, color=black, opacity=1] (b) at (32,0){}; \node[main node, scale=0.5, color=black, opacity=1] (c) at (33,0){}; \node[main node, scale=0.5, color=black, opacity=1] (d) at (34,0){}; \node[scale=0.8, color=black, opacity=1] (t) at (31.5,-0.6){$Q_{|I|}^3$}; \node[scale=0.8, color=black, opacity=1] (t) at (32.5,-0.6){$Q_{|I|}^2$}; \node[scale=0.8, color=black, opacity=1] (t) at (33.5,-0.6){$Q_{|I|}^1$}; \draw[line width= 1 pt] (a) to (d); \draw[color=red,line width= 2 pt, opacity =0.8] (t4) to [ bend left = 70](31.5,0); \draw[color=red,line width= 2.5 pt, opacity=0.8] (34,0) to (31.5,0); \node[main node, scale=0.5, color=black, opacity=1] (a) at (31,0){}; \node[main node, scale=0.5, color=black, opacity=1] (b) at (32,0){}; \node[main node, scale=0.5, color=black, opacity=1] (c) at (33,0){}; \node[main node, scale=0.5, color=black, opacity=1] (d) at (34,0){}; \node[scale=0.8, color=black, opacity=1] (t) at (6.7,0.6){$z_1$}; \node[scale=0.8, color=black, opacity=1] (t) at (14.9,0.6){$z_2$}; \node[scale=0.8, color=black, opacity=1] (t) at (18.6,0.6){$z_3$}; \node[scale=0.8, color=black, opacity=1] (t) at (26.3,0.6){$z_4$}; \node[scale=0.8, color=black, opacity=1] (t) at (34.3,0.6){$m(x_{|I|})$}; \node[scale=1.2, color=black, opacity=1] (t) at (28.7,0){$\ldots$}; \node[scale=1, color=black, opacity=1] (t) at (28.3,2.85){$e_{4,|I|}^3$}; \node[scale=0.8, color=black, opacity=1] (t) at (10.7,0){$m(x_1)$}; \end{tikzpicture} \caption{ The thick blue path represents path $R_4$, and adding to it the red path creates $R_4'$. } \label{fig:jumpswithQ} \end{figure} Let $Q^*$ be the path in $C$ spanned by the interval $[1,z]\cup \left[m(y_{|I|}),N\right]$. By applying Lemma~\ref{lem:easyjump}, we get a path $Q'$ of length at most $2k$ in $G[Q^*]$ with the same endpoints as $Q^*$. Recalling that the pair of vertices $x_{|I|},y_{|I|}$ is connected by paths of all lengths in $\left[\frac{\varepsilon k}{100},\frac{\varepsilon k}{50}\right]$ in the subgraph $G[S]$, and Lemma~\ref{lem:jump with Q} above, we are done by \Cref{obs:combining}. Indeed, we apply it to $G$ with $v_1=z$, $v_2=m(y_{|I|})$ and $v_3=m(x_{|I|})$, while $S_1$ is the set of internal vertices of $Q^*$, $S_2=S$ and $S_3$ are the internal vertices of $Q$ to get all cycle lengths in the interval $\left[(2+\varepsilon) k,\frac{1000}{\varepsilon^2}k\right]$. \hfill \qedsymbol \section{Concluding remarks}\label{sec:concludingrem} In this paper we proved that every Hamiltonian graph on $n \geq 2k^2+o(k^2)$ vertices with independence number $k$ is pancyclic, which is tight up to the $o(k^2)$ error term. Furthermore, our methods allow us to give a short proof of Erd\H{o}s's conjecture that $n = \Omega(k^2)$ vertices are enough for $G$ to be pancyclic. For this, we first note that while getting a bound of $n = \Omega(k^3)$, Keevash and Sudakov \cite{keevash2010pancyclicity} implicitly proved the following result. \begin{lem}[\cite{keevash2010pancyclicity}]\label{lem:ks} There exists a large constant $C$ such that every Hamiltonian graph on $n \geq Ck^2$ vertices with independence number $k$, contains all cycle lengths in $[3,n/C]$. \end{lem} \noindent This reduces Erd\H{o}s's conjecture to the following problem. \begin{itemize} \item[($\ast$)] Does there exist $C' >0$ such that every Hamiltonian graph on $n \geq C'k^2$ vertices with independence number $k$, contains a cycle of length $n-1$? \end{itemize} \noindent Indeed, suppose that the above is true for some large constant $C'$ and let $G$ be a Hamiltonian graph on $n$ vertices with independence number $k$. Then, combining this with the above lemma of Keevash and Sudakov, it follows that if $n \geq CC'k^2$ then $G$ is pancyclic, thus proving Erd\H{o}s's conjecture. Indeed, note that by the lemma above, $G$ contains all cycle lengths up to $n/C$ and one can see that it contains all cycle lengths from $n/C$ to $n$ by iteratively applying the assumption that whenever $n' \geq C'k^2$, there is a cycle of length $n'-1$. The previous results \cite{keevash2010pancyclicity}, \cite{lee2012hamiltonicity} and \cite{dankovics2020low} are all improvements towards question ($\ast$) above. As discussed in the beginning of Section \ref{sec:upperrange}, applying Proposition \ref{lem:jumpwithzigzag} with $c=1$ and $U = \emptyset$ solves this problem in the following stronger form. \begin{thm}\label{thm:outlinen-1} Every Hamiltonian graph on $n > 2k^2+2k$ vertices with independence number $k$, contains a cycle of length $n-1$. \end{thm} Let us note that although \Cref{thm:outlinen-1} shows the existence of a cycle of length $n-1$ already with $n>2k^2+2k$, this is not sufficient to prove that $n>2k^2+\varepsilon k^2$ implies pancyclicity. At this threshold, \Cref{lem:ks} does not apply, so one needs a different argument to find the cycle lengths in the interval $[3,2k^2+2k]$. It turns out that in this setting, the cycle lengths which are hardest to find are those around $2k$, that is, precisely the cycle lengths which are missed by the lower bound construction given in the introduction. Finding them is the most technical part of our proof, given in \Cref{sec:middlerange}. A very interesting open question is to understand the best bound on the number of vertices $n$ in ($\ast$) which guarantees the cycle of length $n-1$. Here the answer might be linear in $n$ as the following question asked in \cite{keevash2010pancyclicity} suggests. \begin{prob} Does there exist a constant $C$ such that every Hamiltonian graph with independence number $k$ and $n \geq Ck$ vertices contain a cycle of length $n-1$? \end{prob}
{ "timestamp": "2022-09-08T02:21:05", "yymm": "2209", "arxiv_id": "2209.03325", "language": "en", "url": "https://arxiv.org/abs/2209.03325", "abstract": "An $n$-vertex graph is Hamiltonian if it contains a cycle that covers all of its vertices, and it is pancyclic if it contains cycles of all lengths from $3$ up to $n$. In 1972, Erdős conjectured that every Hamiltonian graph with independence number at most $k$ and at least $n = \\Omega(k^2)$ vertices is pancyclic. In this paper we prove this old conjecture in a strong form by showing that if such a graph has $n = (2+o(1))k^2$ vertices, it is already pancyclic, and this bound is asymptotically best possible.", "subjects": "Combinatorics (math.CO)", "title": "Pancyclicity of Hamiltonian graphs", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9898303410461385, "lm_q2_score": 0.8104789086703225, "lm_q1q2_score": 0.8022366145798474 }
https://arxiv.org/abs/2212.08285
When is a numerical semigroup a quotient?
A natural operation on numerical semigroups is taking a quotient by a positive integer. If $\mathcal S$ is a quotient of a numerical semigroup with $k$ generators, we call $\mathcal S$ a $k$-quotient. We give a necessary condition for a given numerical semigroup $\mathcal S$ to be a $k$-quotient, and present, for each $k \ge 3$, the first known family of numerical semigroups that cannot be written as a $k$-quotient. We also examine the probability that a randomly selected numerical semigroup with $k$ generators is a $k$-quotient.
\section{Introduction \label{sec:intro We denote $\NN=\{0,1,2,\dots\}$, and we define a \emph{numerical semigroup} to be a set $\nsg\subseteq\NN$ that is closed under addition and contains~0. A numerical semigroup can be defined by a set of generators, \[\langle a_1,\ldots,a_n\rangle = \{a_1x_1+\cdots a_nx_n:\ x_i\in\NN\},\] and if $a_1,\ldots,a_n$ are the minimal set of generators of $\nsg$, we say that $\nsg$ has \emph{embedding dimension} $\mathsf e(\nsg) = n$. For example, \[\langle 3,5\rangle = \{0,3,5,6,8,9,10,\ldots\}\] has embedding dimension 2. If $\nsg$ is a numerical semigroup, then an interesting way to create a new numerical semigroup is by taking the \emph{quotient} \[ \frac{\nsg}{d} = \{ t \in \NN:\ dt \in \nsg\} \] by some positive integer $d$. Note that $\frac{1}{d}\nsg$ is itself a numerical semigroup, one that in particular satisfies $\nsg \subseteq \frac{1}{d}\nsg \subseteq \NN$. For example, \[ \frac{\langle 3,5\rangle}{2}=\{0,3,4,5,\ldots\}=\langle 3,4,5\rangle. \] Quotients of numerical semigroups appear through the literature over the past couple of decades~\cite{symmetriconeelement,symmetricquotient} as well as recently~\cite{harrisquotient,nsquotientgens}; see~\cite[Chapter~5]{numerical} for a thorough overview. \begin{defn}\label{def:quotientrank} We say a numerical semigroup $\nsg$ is a \emph{$k$-quotient} if $\nsg=\langle a_1,\ldots,a_k\rangle/d$ for some positive integers $d, a_1, \ldots, a_k$. The \emph{quotient rank} of $\nsg$ is the smallest $k$ such that $\nsg$ is a $k$-quotient, and we say $\nsg$ has \emph{full quotient rank} if its quotient rank is $\mathsf e(\nsg)$ (since $\nsg=\frac{\nsg}{1}$, its quotient rank is at most $\mathsf e(S)$). \end{defn} Numerical semigroups of quotient rank 2 are precisely the \emph{proportionally modular} numerical semigroups~\cite{openmodularns}, which have been well-studied~\cite{propmodtree,propmodular}. This includes arithmetical numerical semigroups (whose generators have the form $a, a + d, \ldots, a + kd$ with $\gcd(a,d) = 1$), which have a rich history in the numerical semigroup literature~\cite{diophantinefrob,setoflengthsets,nsfreeresarith}. In fact, generalized arithmetical numerical semigroups~\cite{omidalirahmati}, whose generating sets have the form $a, ah + d, \ldots, ah + kd$, can also be shown to have quotient rank 3. For quotient rank $k \ge 3$, much less is known. It is identified as an open problem in~\cite{nsgproblems} that no numerical semigroup had been proven to have quotient rank at least 4. Since then, the only progress in this direction is~\cite{ksquashed}, wherein it is shown there exist infinitely many numerical semigroups with quotient rank at least 4, though no explicit examples are given. With this in mind, we state the main question of the present paper. \begin{mainprob}\label{mainprob:whenaquotient} When is a given numerical semigroup $\nsg$ a $k$-quotient? \end{mainprob} Our main structural results, which are stated in Section~\ref{sec:necessary}, are as follows. \begin{itemize} \item We prove a sufficient condition for full quotient rank (Theorem~\ref{thm:necessary}), which we use to obtain, for each $k$, a numerical semigroup of embedding dimension $k + 1$ that is not a $k$-quotient (Theorem~\ref{thm:noquotient}). When $k \ge 3$, this is the first known example of a numerical semigroup that is not a $k$-quotient. We~also construct, for each $k$, a numerical semigroup that cannot be written as an intersection of $k$-quotients (Theorem~\ref{thm:nointersection}), settling a conjecture posed in~\cite{ksquashed}. \item We prove quotient rank is sub-additive whenever the denominators are coprime. This provides a new method of proving a given numerical semigroup is a quotient:\ partition its generating set, and prove that each subset generates a quotient, e.g., \begin{align*} \langle 11,12,13,17,18,19,20 \rangle &= \langle 11,12,13 \rangle + \langle 17,18,19,20 \rangle = \frac{\langle 11,13\rangle}{2}+\frac{\langle 17,20\rangle}{3} \\ &= \frac{3\langle 11,13\rangle + 2\langle 17,20\rangle}{2 \cdot 3} = \frac{\langle 33,34,39,40\rangle}{6}. \end{align*} We~use this result to prove that any numerical semgiroup with \emph{maximal embedding dimension} (that is, the smallest generator equals the embedding dimension) fails to have full quotient rank (Theorem~\ref{thm:maxembdim}). \end{itemize} Our remaining results are probabilistic in nature. We examine two well-studied models for ``randomly selecting'' a numerical semigroup:\ the ``box'' model, where the number of generators and a bound on the generators are fixed~\cite{expectedfrob,arnoldfrob,burgeinsinaifrob}; as well as a model where the smallest generator and the number of gaps are fixed~\cite{kaplancounting}, whose prior study has yielded connections to enumerative combinatorics~\cite{kunzcoords} and polyhedral geometry~\cite{kunzfaces1,kunz}. We prove that under the first model, asymptotically all semigroups have full quotient rank (Theorem~\ref{thm:numericalbox}), while under the second model, asymptotically no semigroups have full quotient rank (Theorem~\ref{thm:maxembdim}). Our results also represent partial progress on the following question, which has proved difficult. \begin{prob}\label{prob:algorithm} Given a numerical semigroup $\nsg$ and a positive number $k$, is there an algorithm to determine whether $\nsg$ is a $k$-quotient? \end{prob} \begin{remark}\label{rem:relprime} Some texts require that the generators of a numerical semigroup be relatively prime, so that $\NN\setminus \nsg$ is finite. This assumption is harmless, since any numerical semigroup can be written as $m\nsg$, where the generators of $\nsg$ are relatively prime, and it also doesn't affect $k$-quotientability: given a positive integer $d$, one can readily check that \[ \frac{m\nsg}{d} = m' \left(\frac{\nsg}{d'}\right), \] where $m' = m/\gcd(m,d)$ and $d' = d/\gcd(m,d)$. \end{remark} \section{When is $\nsg$ not a $k$-quotient? \label{sec:necessary In this section, we give two structural results. The first (Theorem~\ref{thm:necessary}) is a necessary condition for a given numerical semigroup $\nsg$ to be a $k$-quotient, which forms the backbone of the constructions in Section~\ref{sec:fullquotientrank} and the probabilistic results in Section~\ref{sec:randomsgps}. The second (Theorem~\ref{thm:sums}) is a constructive proof that quotient rank is sub-additive, provided the denominators are relatively prime. In what follows, we write $[p] = \{1,2,\dots,p\}$ for any positive integer $p$, and given a collection of vectors $\{\vec v_i\}$ and a set of indices $I$, we define $\vec v_I=\sum_{i\in I}\vec v_i$. \begin{thm} \label{thm:necessary} Suppose \[\nsg =\frac{\langle b_1,\ldots,b_k\rangle}{d}\] for some $b_i\in \NN$ and positive integer $d$. Given any elements $s_1,\ldots,s_p \in \nsg$ with $p > k$, there exists a nonempty subset $I\subseteq [p]$ such that $s_I/2\in \nsg$. \end{thm} \begin{proof} Let $\vec b=(b_1,\ldots,b_k)$. For $1\le i\le p$, let $\vec c_i=(c_{i1},\ldots,c_{ik})\in\NN^k$ be such that \[s_i = d (c_{i1}b_1+\cdots c_{ik}b_k),\] which exist since $s_i\in\nsg$. For a vector $\vec v\in\ZZ^k$, define $\vec v\bmod 2\in\ZZ_2^k$ to be the coordinate-wise reduction of $\vec v$ modulo 2. For $J\subseteq [p]$, examine $\vec c_J \bmod 2$. There are $2^p$ possible $J$ and $2^k$ possible values for $\vec c_J \bmod 2$, with $p>k$, so there must be two distinct $J_1$ and $J_2$ such that \[\vec c_{J_1}\bmod 2=\vec c_{J_2}\bmod 2.\] Let $I=(J_1\setminus J_2)\cup (J_2\setminus J_1)$ be their symmetric difference, which is nonempty. Then \[\vec c_{I}\bmod 2=\vec c_{J_1}+\vec c_{J_2}-2\vec c_{J_1\cap J_2}\bmod 2=\vec 0,\] so $\vec c_{I}$ has even coordinates. Let $\vec c_I = (2q_1,\ldots,2q_k)$ where $q_i\in\NN$. Then \begin{align*} s_I/2&=\sum_{i\in I}\big(d\cdot (c_{i1}b_1+\cdots c_{ik}b_k)\big)/2 = d\sum_{j=1}^kb_j\sum_{i\in I}c_{ij}/2 =d\sum_{j=1}^k q_jb_j \end{align*} is an element of $\nsg$, as desired. \end{proof} \begin{cor} \label{cor:necessary} Let $\nsg = \langle a_1, \dots, a_n \rangle$ be a numerical semigroup. If $\nsg$ does not have full quotient rank, then there exists $I \subseteq [n]$ such that \[ a_I \in\langle a_j:\ j\notin I \rangle.\] \end{cor} \begin{proof} By applying Theorem~\ref{thm:necessary} to the generating set $\{a_1, \dots, a_n\}$, we obtain that for some $J \subseteq [n]$, $a_J/2 \in \nsg$. So there exist $c_r\in\NN$ such that \[ \sum_{j\in J} a_j=\sum_{r\in R} 2c_r a_r \] where $R=\{r:\ c_r>0\}$. Letting $I=J\setminus R$ and subtracting each $a_j$ with $j\in J\cap R$ from both sides, we have \[ a_I=\sum_{i\in I} a_i = \sum_{r\in J\cap R} (2c_r-1) a_r + \sum_{r\in R\setminus J}2c_r a_r \] is an element of $\langle a_j:\ j\notin I\rangle$, as desired. Note that $I$ is nonempty, as otherwise \[ 0 = a_I = \sum_{r\in J\cap R} (2c_r-1) a_r + \sum_{r\in R\setminus J}2c_r a_r \ge \sum_{r\in J\cap R} a_r = \sum_{r\in J} a_r > 0 \] since $J$ is nonempty, which is a contradiction. \end{proof} \begin{thm} \label{thm:sums} If $\nsg$ and $\nsgtwo$ are numerical semigroups and $\gcd(c,d) = 1$, then \[ \frac{\nsg}{c} + \frac{\nsgtwo}{d} = \frac{d\nsg + c \nsgtwo}{cd}. \] \end{thm} \begin{proof} First suppose that $x \in \frac{1}{c}\nsg + \frac{1}{d}\nsgtwo$. Then $x = s+t$ where $cs \in \nsg$ and $dt \in \nsgtwo$, so \[ cdx = d(cs) + c(dt) \in d \nsg + c \nsgtwo \] which implies $x \in \frac{1}{cd}(d\nsg + c \nsgtwo)$. Note this containment does not require $\gcd(c,d) = 1$. On the other hand, suppose $cdx \in d \nsg + c \nsgtwo$, so \begin{equation} \label{eq:sums} cdx = ds + ct \qquad \text{for some} \qquad s \in \nsg, t \in \nsgtwo. \end{equation} In particular, $ct = d(cx-s)$ is a multiple of $d$. Since $c$ and $d$ are relatively prime, this implies that $t$ is a multiple of $d$, say $t = bd$. Since $t \in \nsgtwo$, we conclude that $b \in \frac{1}{d}\nsgtwo$. Similarly, we can write $s = ac$ for some $a$ and so $a \in \frac{1}{c}\nsg$. Substituting $t=bd$ and $s=ac$ into \eqref{eq:sums}, we obtain \[ cdx = dac + cbd = cd(a+b). \] By cancellation, we obtain $x = a + b$ with $a \in \frac{1}{c}\nsg$ and $b \in \frac{1}{d}\nsgtwo$, as desired. \end{proof} Given the ease of proving Theorem~\ref{thm:sums}, it is surprisingly more difficult when the denominators do have a common factor. In a follow-up to this current paper, we will translate the quotient operation into a geometric setting, which will allow us to generalize Theorem~\ref{thm:sums} to drop the ``coprime denominators'' hypothesis. Intriguingly, the translation can cause a large blow-up in the numbers, e.g., \[ \frac{\langle 11,13\rangle}{2} + \frac{\langle 17,19\rangle}{2} = \frac{\langle 2416656, 2894591, 3441983, 3869571 \rangle}{25357536}. \] Based on experimentation, this blow-up seems necessary. \section{Some families of numerical semigroups with full quotient rank \label{sec:fullquotientrank In this section, we produce two families of numerical semigroups:\ those in the first have embedding dimension $k+1$ but are not $k$-quotients, so in particular have full quotient rank (Theorem~\ref{thm:noquotient}); and those in the second are not even \emph{intersections} of $k$-quotients (Theorem~\ref{thm:nointersection}). \begin{thm} \label{thm:noquotient} Given a positive integer $k$, let $a\ge 2^k$ be an integer. Define $a_i=2a+2^i$ for $i=0,1,\dots,k$. Then the numerical semigroup \[\nsg=\langle a_0,a_1,\ldots,a_k\rangle\] is not a $k$-quotient. \end{thm} \begin{proof} For $1\le j\le 2^k-1$, let $b_j=\omega(j)a+j$, where $\omega(j)$ is the number of 1's in the binary representation of $j$. We first prove that, if $\nsgtwo$ is \emph{any} $k$-quotient the contains $a_0,\ldots,a_k$ (so $\nsgtwo=\nsg$ will be an example), then there exists $j$ ($1\le j\le 2^k-1$) such that $b_j\in \nsgtwo$. Indeed, we apply Theorem~\ref{thm:necessary}. We know that there exists a nonempty $I\subseteq\{0,1,\ldots,k\}$ such that $a_I/2\in \nsgtwo$. If $0\in I$, then $a_I$ is odd and $a_I/2$ is not an integer, so we know $I\subseteq\{1,\ldots,k\}$. Let \[j=\sum_{i\in I}2^{i-1}.\] We have that $1\le j\le 2^k-1$, and \[a_I/2=\sum_{i\in I}\left(2a+2^i\right)/2 = \abs{I}a + \sum_{i\in I}2^{i-1} = \omega(j)a+j=b_j,\] so $b_j\in \nsgtwo$. Now we this apply to $\nsgtwo=\nsg$. Seeking a contradiction, suppose $\nsg$ is a $k$-quotient, and therefore we have some $b_j\in \nsg$, that is, $b_j=\sum_{i=0}^ka_ix_i$ with $x_i\in\NN$. Examining this sum modulo $a$, and noting that $b_j=j\pmod a$ and $a_i=2^i\pmod a$, we see that \[\sum_{i=0}^k x_i\ge \omega(j).\] But a sum of $\omega(j)$ generators of $\nsg$ is too large: \[\omega(j)a+j=b_j\ge \omega(j)\cdot a_0 = \omega(j)(2a+1)\ge \omega(j)a+a\ge \omega(j)a+2^k,\] a contradiction. Therefore $b_j\notin \nsg$, and so $\nsg$ cannot be a $k$-quotient. \end{proof} \begin{thm} \label{thm:nointersection} Given a positive integer $k\ge 2$, let $a\ge k2^k$ be an integer. As before, define $a_i=2a+2^i$ and $b_j=\omega(j)a+j$, where $\omega(j)$ is the number of 1's in the binary representation of $j$. Let $N=(2k+1)a$. Then \[\nsg=\langle a_0,a_1,\ldots,a_k,N-b_1,N-b_2,\ldots,N-b_{2^k-1}\rangle\] cannot be written as an intersection of $k$-quotients. \end{thm} \begin{proof} Suppose, seeking a contradiction, that $\nsg=\bigcap_{\ell=1}^p \nsg_\ell$, where the $\nsg_\ell$ are $k$-quotients. Each $\nsg_\ell$ must contain $a_0,a_1,\ldots,a_k$, and we noted in the proof of Theorem~\ref{thm:noquotient} that this implies that $\nsg_\ell$ must contain $b_j$ for some $j$. But then $\nsg_\ell$ contains both $b_j$ and $N-b_j$, and so additive closure implies that it contains $N$. This means $N \in \bigcap_{\ell=1}^p \nsg_\ell = \nsg$. Let \begin{equation}\label{eq:sumforN} N=\sum_{i=1}^k a_ix_i + \sum_{j=1}^{2^k-1}(N-b_j)y_j, \end{equation} where $x_i,y_j \in \NN$. We break into three cases. \begin{itemize} \item If $\sum_{j}y_j\ge 2$, then~\eqref{eq:sumforN} would be too large, as for some $j_1,j_2$, \begin{align*}(2k+1)a = N &\ge (N-b_{j_1})+(N-b_{j_2})\\ &= 2N-(\omega(j_1)+\omega(j_2))a -(j_1+j_2)\\ &> 2\cdot (2k+1)a -2ka-2\cdot 2^k\\ &= (2k+2)a-2^{k+1}, \end{align*} which is impossible since $a\ge 2^{k+1}$. \item If $\sum_{j}y_j=1$, then~\eqref{eq:sumforN} uses exactly one $N-b_j$. But then $N=(N-b_j)+b_j$ implies that $b_j\in \langle a_0,a_1\ldots,a_k\rangle$, which we saw was impossible in the proof of Theorem~\ref{thm:noquotient} since $a\ge 2^k$. \item If $\sum_{j}y_j=0$, then $N = \sum_{i} a_ix_i$. If $\sum_i x_i \le k$, then \[(2k+1)a=N\le k(2a+2^k),\] which is impossible since $a>k2^k$. On the other hand, if $\sum_i x_i > k$, then \[(2k+1)a=N\ge (k+1)(2a+1)>(2k+1)a,\] which is also impossible. \end{itemize} In each case, we obtain a contradiction. \end{proof} \section{How often do numerical semigroups have full quotient rank? \label{sec:randomsgps In this section, we consider the question ``how likely is a randomly selected numerical semigroup to have full quotient rank?'' We consider two sampling methods. The first is the ``box'' method, wherein a fixed number of generators are selected uniformly and independently from an interval $[1,M]$. Numerical semigroups selected under this model have high probability (i.e., approaching 1 as $M \to \infty$) of having full quotient rank. \begin{thm} \label{thm:numericalbox} Fix a positive integer $n$. If $\nsg = \langle a_1, \dots, a_n \rangle$ where $a_1, \ldots, a_n \in [M]$ are uniformly and independently chosen, then the probability that $\nsg$ has full quotient rank tends to 1 as $M \to \infty$. More precisely, this probability is $1 - O(M^{-\frac{1}{n}})$. \end{thm} \begin{proof} By Corollary~\ref{cor:necessary}, it suffices to bound the probability that there exists $I \subseteq [n]$ such that $a_I \in \langle a_j:\ j\notin I\rangle$. Let $A$ be this event, and let $B$ be the event that $a_i \leq M^{\frac{n-1}{n}}$ for some $i$. We will use that \[ \Pr(A) = \Pr(B)\Pr(A \mid B) + \Pr(B^c)\Pr(A \mid B^c) \leq \Pr(B) + \Pr(A \mid B^c). \] For the first term, the union bound gives us that \[ \Pr(B) \leq n \left( \frac{M^{\frac{n-1}{n}}}{M} \right) = \frac{n}{M^{\frac{1}{n}}}. \] For the second term, fix a nontrivial subset $I \subsetneq [n]$ and $b_i \in \NN$ for $i \notin I$. If $b_i > nM^{\frac{1}{n}}$ for some $i \notin I$, then since every $a_i$ is greater than $M^{\frac{n-1}{n}}$, we have \[ \sum_{j \notin I} b_ja_j \geq b_ia_i > \left( nM^{\frac{1}{n}} \right) M^\frac{n-1}{n} = nM. \] But $a_I$ cannot be this large because it is the sum of at most $n-1$ integers that are each at most $M$. So we need only consider $b_i \le nM^{\frac{1}{n}}$. Letting $i^\ast = \min(I)$ and $m = nM^{\frac{1}{n}}$, \begin{align*} \Pr(A \mid B^c) &\le \sum_{\substack{I \subsetneq [n] \\ I \ne \emptyset}} \sum_{\substack{b_j \le m \\ j \notin I}} \Pr \bigg( \sum_{i \in I}a_i = \sum_{i \notin I}b_ia_i \biggm\vert a_1, \ldots, a_n > M^{\frac{n}{n-1}} \bigg) \\ & = \sum_{\substack{I \subsetneq [n] \\ I \ne \emptyset}} \sum_{\substack{b_j \le m \\ j \notin I}} \Pr \bigg( a_{i^\ast} = \sum_{i \notin I}b_ia_i - \!\!\! \sum_{i \in I \setminus \{i^\ast\}} \!\!\! a_i \biggm\vert a_1, \ldots, a_n > M^{\frac{n}{n-1}} \bigg) \\ & \le \sum_{\substack{I \subsetneq [n] \\ I \ne \emptyset}} \sum_{\substack{b_j \le m \\ j \notin I}} \frac{1}{M-M^{\frac{n-1}{n}}} \le \frac{\left( 2^n - 2 \right) \left( nM^{\frac{1}{n}} \right)^{n-1}} {M-M^{\frac{n-1}{n}}} = \frac{\left( 2^n - 2 \right)n^{n-1}}{M^{\frac{1}{n}}-1}, \end{align*} where the second inequality comes from the fact that for any choice of the $a_i$ with $i \neq i^\ast$, there is at most one choice of $a_i^\ast$ that makes the linear equation hold. Thus, \[ \Pr(A) \leq \frac{\left( 2^n - 2 \right)n^{n-1}}{M^{\frac{1}{n}}-1} + \frac{n}{M^{\frac{1}{n}}-1} = O(M^{-\frac{1}{n}}), \] which completes the proof. \end{proof} \begin{remark}\label{rem:numericalbox} The ``minimally generated'' and ``finite complement'' conditions, which are often imposed on numerical semigroups, do not affect Theorem~\ref{thm:numericalbox}. Indeed, under this ``box'' probability model, the chosen generators $a_1, \dots, a_n$ need not form a minimal generating set. Since the quotient rank is at most the embedding dimension, the (asymptotically rare) event that the rank of $\nsg$ is less than $n$ contains the event that the chosen generating set is not minimal. Additionally, the probability that $a_1,\ldots, a_n$ are relatively prime approaches the positive constant $1/\zeta(n)$ by~\cite{Nym}, where $\zeta(n)$ is the Reimann zeta function $\sum_{i=1}^{\infty}1/i^n$. Therefore, even if one restricts to those $a_1,\ldots, a_n$ that are relatively prime, the conditional probability that the quotient rank of the resulting numerical semigroup is less than $n$ still tends to 0. \end{remark} Under the second model, a numerical semigroup $\nsg$ is selected uniformly at random from among the (finitely many) with fixed smallest generator $m$ and number of gaps~$g$. Such numerical semigroups have high probability (i.e., tending to 1 as $g \to \infty$) of having embedding dimension $m$ (such numerical semigroups are said to have \emph{maximal embedding dimension}). We prove that maximal embedding dimension numerical semigroups never have full quotient rank, illustrating a stark contrast in asymptotic behavior to the first model. We first recall a characterization of quotient rank 2 numerical semigroups, which appears in~\cite{numerical} as a characterization of proportionally modular numerical semigroups in the case $\gcd(\nsg) = 1$. Our statement here is more general, thanks to Remark~\ref{rem:relprime}. \begin{thm}\label{thm:pmcriterion} A numerical semigroup $\nsg$ with $\gcd(\nsg) = D$ has quotient rank 2 if and only if there exists an ordering $b_1, \ldots, b_n$ of its minimal generators such that: \begin{enumerate}[(a)] \item $\gcd(b_i, b_{i+1}) = D$ for $1 \leq i \leq n-1$; and \item $b_{i-1} + b_{i+1}$ is divisible by $b_i$ for $2 \leq i \leq n-1$. \end{enumerate} \end{thm} \begin{lemma}\label{lem:plusminusone} For any $a, b, m \ge 1$, the numerical semigroup $\nsg = \langle m,am-1, bm+1 \rangle$ is a 2-quotient. \end{lemma} \begin{proof} If $\mathsf e(\nsg) \le 2$, then $\nsg$ is clearly a 2-quotient. Otherwise, letting $b_1 = am-1$, $b_2 = m$, and $b_3 = bm+1$, it is clear that $\gcd(b_1, b_2) = \gcd(b_2, b_3) = 1$ and that $b_2 \mid (b_1 + b_3)$. As such, $\nsg$ is a 2-quotient by Theorem \ref{thm:pmcriterion}. \end{proof} \begin{thm}\label{thm:maxembdim} If $m = \min(\nsg \setminus \{0\})$, then $\nsg$ is an $(m-1)$-quotient. In particular, if $\mathsf e(\nsg) = m$, then $\nsg$ does not have full quotient rank. \end{thm} \begin{proof} If $\nsg=\frac{\nsg}{1}$ has embedding dimension less than $m$, then the proof is immediate. If not, then $\nsg$ has $m$ minimal generators, and so they must all have distinct residues modulo $m$. That is, \[ \nsg = \langle m, b_1m+1, \dots, b_{k-1}m + (m-1) \rangle \] for some positive integers $b_1, \dots, b_{m-1}$. Write $\nsg = \nsg_1 + \nsg_2$ where \[ \nsg_1 = \langle m, b_1m + 1, b_{k-1}m + (m-1) \rangle, \: \nsg_2 = \langle b_2m + 2, \dots, b_{m-2}m + (m-2) \rangle. \] Now by Lemma~\ref{lem:plusminusone}, $\nsg_1$ as a 2-quotient, and $\nsg_2 = \frac{\nsg_2}{1}$ is trivially an $(m-3)$-quotient. Since $1$ is coprime to every integer, Theorem~\ref{thm:sums} implies $\nsg$ is an $(m-1)$-quotient. \end{proof} \section*{Acknowledgements Tristram Bogart was supported by internal research grant INV-2020-105-2076 from the Faculty of Sciences of the Universidad de los Andes.
{ "timestamp": "2022-12-19T02:06:32", "yymm": "2212", "arxiv_id": "2212.08285", "language": "en", "url": "https://arxiv.org/abs/2212.08285", "abstract": "A natural operation on numerical semigroups is taking a quotient by a positive integer. If $\\mathcal S$ is a quotient of a numerical semigroup with $k$ generators, we call $\\mathcal S$ a $k$-quotient. We give a necessary condition for a given numerical semigroup $\\mathcal S$ to be a $k$-quotient, and present, for each $k \\ge 3$, the first known family of numerical semigroups that cannot be written as a $k$-quotient. We also examine the probability that a randomly selected numerical semigroup with $k$ generators is a $k$-quotient.", "subjects": "Commutative Algebra (math.AC); Combinatorics (math.CO)", "title": "When is a numerical semigroup a quotient?", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.978384664716301, "lm_q2_score": 0.8198933337131076, "lm_q1q2_score": 0.8021710644080291 }
https://arxiv.org/abs/2206.05396
A systematic approach on some relevant theorems that follows from Kolmogorov's axioms
A selection of the relevant theorems of Probability Theory that comes directly from Kolmogorov's axioms, Set Theory basic results, definitions and rules of inference are listed and proven in a systematic approach, aiming the student who seeks a self-contained account on the matter before moving to more advanced material.
\section{Introduction} \label{introduction} Most of the Probability Theory and Statistics books presents the rules of probability as consequences of Andrei Kolmogorov's axioms \cite{ross2010,rozanov1969,magalhaes2006,degroot1989,shiryaev1996,sinai1992,gnedenko1997}. Although they show proofs of the most relevant relations between probabilities of different kinds of events, either directly or through of exercises, I've found no systematic list of relations and proofs. The lacking of some generalizations is also present, so I've made a selection that includes the most common and relevant theorems (and their consequences) that arise, direct or indirectly, from the axioms. The list is neither complete nor fundamentally rigorous, but provides a secure step for the student to base its researches and prove more theorems even before the introduction of random variables. At the end of this paper, in section \ref{probdiagram}, a diagram relating axioms and main results is presented to increase the broad view of the connections among them. The discussion is intentionally didactic in order to help the student to follow the reasoning. It demands some contact with proof theory and logic beforehand, but nothing more, alongside some Set Theory equations listed briefly in section \ref{sets}. It's perhaps important to emphasize that the axiomatic system proposed by Kolmogorov, somewhat inspired in the frequentist view of statistics \cite{shafer2006}, was not the only one proposed, as described, for instance, by Terenin and Draper \cite{terenin2017}. \section{Sets} \label{sets} For the proofs to follow, some relations between sets are necessary. I'll present them here, without proof, since they are not the subject of this paper. However, they can be found easily in introductions to mathematical proof \cite{wohlgemuth2011}, for instance, and Probability Theory textbooks: \begin{itemize} \item \emph{Empty set}: An empty set, $\emptyset$, have the following properties for any set $A$: $A\cup \emptyset = A$ and $A\cap \emptyset = \emptyset$; \item \emph{Space set}: An space set, $\Omega$, is the union of all possible sets. In other words, the properties $A \subset \Omega$, $A\cup \Omega = \Omega$ and $A\cap \Omega = A$ are valid for any set $A$; \item \emph{Complementary set}: A complementary set of $A$, $\overline{A}$, has the following properties: $A \cup \overline{A}= \Omega$ and $A \cap \overline{A}= \emptyset$; \item \emph{Associative laws for sets}: Given three sets $A$, $B$ and $C$, it can be proven that: $(A\cup B) \cup C = A\cup (B \cup C)$ and $(A\cap B) \cap C = A\cap (B \cap C)$; \item \emph{Distributive laws for sets}: Given three sets $A$, $B$ and $C$, the following equations are valid: $A\cup(B\cap C) =(A\cup B)\cap (A\cup C)$ and $A\cap(B\cup C) =(A\cap B)\cup (A\cap C)$. \end{itemize} \section{Probabilities} \label{probabilities} \begin{defi} $\Omega$ is the set that represents the sample space, the space of all possible events. \label{definition1} \end{defi} \begin{defi} Events related to the sample space are all subsets of $\Omega$. \label{definition2} \end{defi} \begin{defi} Two events $A$ e $B$ are pairwise mutually exclusive (PME) if $A\cap B=\emptyset$, that is, are disjoint sets. \label{definition3} \end{defi} \begin{defi} A class of subsets of $\Omega$, represented by $\mathcal{F}$, is considered a $\sigma$-algebra if it has the following properties \cite{magalhaes2006,sinai1992}: \begin{enumerate} \item $\Omega\in\mathcal{F}$. \item If $A\in\mathcal{F}$, then $\overline{A}\in\mathcal{F}$. \item (\emph{Closure with respect to countable unions}). If a countable collection $\{A_1,A_2,...\}=\{A_i\}_{i=1}^{\infty}$ of sets $A_i$ is such that $A_i\in\mathcal{F}$ for all $i$, then $\displaystyle\bigcup_{i=1}^{\infty}A_i\in\mathcal{F}$. \end{enumerate} \label{definition4} \end{defi} \begin{defi} A partition of the sample space $\Omega$ is defined according the following property: \begin{equation} \displaystyle\bigcup_{i=1}^{n}A_i=\Omega \label{equation1} \end{equation} Being $A_i$ and $A_j$ mutually exclusive (ME) for all combinations of sets. \label{definition5} \end{defi} \begin{defi} Probability $P$ is a function of the subsets of the sample space correspondent to $\mathcal{F}$: $P=P(\mathcal{F})$. Additionally, it should obey the axioms that follow. \label{definition6} \end{defi} \begin{center} \fbox{\parbox{15cm}{ \begin{ax}[Non-negativity] $P(A_i)\geq 0$, for all event $A_i\subset\mathcal{F}$. \label{axiom1} \end{ax} \begin{ax}[Normalization] $P(\Omega)=1$. \label{axiom2} \end{ax} \begin{ax}[Countable additivity] If $A_i$ and $A_j$ are two PME events (disjoint sets) for all $i\neq j$, then $P\displaystyle\left(\bigcup_{i=1}^{\infty}A_i\right)=\displaystyle\sum_{i=1}^{\infty}P(A_i)$ for all $(A_i,A_j)\in\mathcal{F}$. \label{axiom3} \end{ax} } } \end{center} \begin{defi} If $\mathcal{F}$ is a $\sigma$-algebra of set $\Omega$, and $P$ a function of $\mathcal{F}$ with the properties described by axioms \ref{axiom1}, \ref{axiom2} and \ref{axiom3}, then the triple $\{\Omega,\mathcal{F},P\}$ is called \emph{probability space}. \end{defi} \begin{thm} $P(\emptyset)=0$. \label{theorem1} \end{thm} \begin{proof} According to axiom \ref{axiom3}, we can choose the sets $A_i$ for $i\geq 2$ such that $A_i=\emptyset_i=\emptyset$. Consequently: \begin{equation} P\left(\bigcup_{i=1}^{\infty}A_i\right)=P\left[A_1\cup\left(\bigcup_{i=2}^{\infty}A_i\right)\right]=P\left[A_1\cup\left(\bigcup_{i=2}^{\infty}\emptyset_i\right)\right]=P(A_1)+P\left(\bigcup_{i=2}^{\infty}\emptyset_i\right) \label{equation2} \end{equation} Considering that $\displaystyle\bigcup_{i=2}^{\infty}\emptyset_i=\emptyset$, it follows that: \begin{equation} P(A_1\cup\emptyset)=P(A_1)+P(\emptyset) \label{equation3} \end{equation} Given the property of the empty set $A_i\cup\emptyset=A_i$ for all $i$, then $P(A_1\cup\emptyset)=P(A_1)$ which can be substituted in Eq. \ref{equation3}: \begin{equation} P(A_1)=P(A_1)+P(\emptyset) \label{equation4} \end{equation} Finally proving that: \begin{equation} P(\emptyset)=0 \label{equation5} \end{equation} \end{proof} \subsection{Combination of events} \label{combinationsofevents} \begin{thm} If $A_i$ and $A_j$ are PME events, then $P\displaystyle\left(\bigcup_{i=1}^{n}A_i\right)=\displaystyle\sum_{i=1}^{n}P(A_i)$ for $i\neq j$ and $n\geq 1$. \label{theorem2} \end{thm} \begin{proof} From axiom \ref{axiom3}, consider that from $i=1$ to $i=n$ we have the sets $A_1$, $A_2$, etc, $A_n$, and for $i>n$ we have $A_i=\emptyset_i$. Therefore: $$P\left(\bigcup_{i=1}^{\infty}A_i\right)=P\left[\left(\bigcup_{i=1}^{n}A_i\right)\cup\left(\bigcup_{i=n+1}^{\infty}A_i\right)\right]=P\left[\left(\bigcup_{i=1}^{n}A_i\right)\cup\left(\bigcup_{i=n+1}^{\infty}\emptyset_i\right)\right]$$ \begin{equation} P\left[\left(\bigcup_{i=1}^{n}A_i\right)\cup\left(\bigcup_{i=n+1}^{\infty}\emptyset_i\right)\right]=\sum_{i=1}^{n}P(A_i)+\sum_{i=n+1}^{\infty}P(\emptyset_i) \label{equation6} \end{equation} Since $\bigcup_{i=n+1}^{\infty}\emptyset_i=\emptyset$, and $A\cup\emptyset=A$ for all $A$, including $A=\bigcup_{i=1}^{\infty}A_i$, according to theorem \ref{theorem1} $P(\emptyset_i)=P(\emptyset)=0$, thus: \begin{equation} P\left[\left(\bigcup_{i=1}^{n}A_i\right)\cup\left(\bigcup_{i=n+1}^{\infty}\emptyset_i\right)\right]=P\left(\bigcup_{i=1}^{n}A_i\right)=\sum_{i=1}^{n}P(A_i) \label{equation7} \end{equation} \end{proof} \begin{thm}[Normalization condition] Let the sets $A_1$, $A_2$, ..., $A_n$ be a partition in $\Omega$. It implies that: \begin{equation} \displaystyle\sum_{i=1}^nP(A_i)=1 \label{equation8} \end{equation} \end{thm} \begin{proof} Given the definition \ref{definition5}, $A_i\cap A_j=\emptyset$ for all $i\neq j$. Therefore, following the result of the theorem \ref{theorem2}: \begin{equation} \displaystyle\sum_{i=1}^nP(A_i)=P\left(\bigcup_{i=1}^nA_i\right)=P(\Omega)=1 \label{equation9} \end{equation} The second equality in Eq. \ref{equation9} follows from definition \ref{definition5}, and the third is a consequence of the axiom \ref{axiom2}. \end{proof} \begin{lem} Any pair among the sets $A_1$, $A_2$, ..., $A_n$ are mutually exclusive (that is, they are PME) if and only if they are mutually exclusive as a whole (ME for any combination of any number of these sets, including all of them). \label{lemma1} \end{lem} \begin{proof} If $A_i\cap A_j=\emptyset$ for all $i\neq j$ in the presented sequence of sets, then, given that $A_i\cap\emptyset =\emptyset$: $$A_1\cap A_2\cap A_3\cap ... \cap A_n=(A_1\cap A_2)\cap A_3\cap ... \cap A_n=\emptyset\cap A_3\cap ... \cap A_n$$ $$A_1\cap A_2\cap A_3\cap ... \cap A_n=(\emptyset\cap A_3)\cap ... \cap A_n=\emptyset\cap ... \cap A_n$$ $$ \ldots $$ \begin{equation} A_1\cap A_2\cap A_3\cap ... \cap A_n=\emptyset \label{equation10} \end{equation} And since $\bigcap_{i=1}^{n}A_i=\emptyset$, we can deduce that for any $A_i$ and $A_j$ with $i\neq j$: \begin{equation} A_1\cap ... \cap A_i ... \cap A_j \cap ... \cap A_n=(A_i\cap A_j)\cap (A_1\cap ... \cap A_n)=\emptyset \label{equation11} \end{equation} Hence, the property $A\cap \emptyset = \emptyset$ for all $A$ can be applied: $$(A_i\cap A_j)\cap (A_1\cap ... \cap A_n)=\emptyset=\emptyset\cap (A_1\cap ... \cap A_n)$$ \begin{equation} (A_i\cap A_j)=\emptyset \label{equation12} \end{equation} \end{proof} \begin{lem} If $A$ and $B$ are events, then: \begin{equation} P(A\cup B)=P(A)+P(B)-P(A\cap B) \label{equation13} \end{equation} \label{lemma2} \end{lem} \begin{proof} By using the relations $A\cup B=A\cup(\overline{A}\cap B)$ and $B=(\overline{A}\cap B)\cup (A\cap B)$, and since $A$ and $\overline{A}\cap B$ are PME, and $\overline{A}\cap B$ and $A\cap B$ too, we can use the theorem \ref{theorem2} for $n=2$ in order to obtain the relations: \begin{equation} P(A\cup B)=P(A)+P(\overline{A}\cap B) \label{equation14} \end{equation} \begin{equation} P(B)=P(\overline{A}\cap B)+P(A\cap B) \label{equation15} \end{equation} Subtracting Eq. \ref{equation15} from \ref{equation14} leads to $P(A\cup B)=P(A)+P(B)-P(A\cap B)$. \end{proof} \begin{thm}[Rule of addition of probabilities, or inclusion-exclusion principle, or Poincaré's theorem] \begin{equation} P\left(\displaystyle\bigcup_{i=1}^{n}A_i\right)=\sum_{i=1}^{n}P(A_i)-\sum_{i=1}^{n}\sum_{j=i+1}^{n-1}P(A_i\cap A_j)+... \label{equation16} \end{equation} \label{theorem4} \end{thm} \begin{proof} The proof follows from mathematical induction. Eq. \ref{equation16} is refered as proposition $Q(n)$. The case for $n=2$, that is, $Q(2)$, was already proved (lemma \ref{lemma2}), if we change the notation to $A=A_1$ and $B=A_2$. The case $Q(1)$ is trivial, with $P(A)=P(A)$ (or $P(A_1)=P(A_1)$). We can prove the validity of $Q(n+1)$ if the validity of $Q(n)$ is presumed or, equivalently, we might prove $Q(n)$ from $Q(n-1)$. But first we'll prove $Q(3)$ to better understand the structure of $Q(n)$, and of its terms. As already proven: \begin{equation} P(A_1\cup A_2)=P(A_1)+P(A_2)-P(A_1\cap A_2) \label{equation17} \end{equation} In order to demonstrate $Q(3)$, we first use $Q(2)$ (lemma \ref{lemma2}) and the associative properties of sets: \begin{equation} P(A_1\cup A_2\cup A_3)=P[A_1\cup (A_2\cup A_3)]=P(A_1)+P(A_2\cup A_3)-P[A_1\cap (A_2\cup A_3)] \label{equation18} \end{equation} Then using the distributive property, $A_1\cap (A_2\cup A_3)=(A_1\cap A_2)\cup(A_1\cap A_3)$, and two more applications of $Q(2)$: $$P(A_2\cup A_3)=P(A_2)+P(A_3)-P(A_2\cap A_3)$$ $$P[A_1\cap (A_2\cup A_3)]=P(A_1\cap A_2)+P(A_1\cap A_3)-P(A_1\cap A_2\cap A_3)$$ \begin{eqnarray} P(A_1\cup A_2\cup A_3) & = & P(A_1)+P(A_2)+P(A_3) \nonumber \\ & - & P(A_1\cap A_2)-P(A_1\cap A_3)-P(A_2\cap A_3) \nonumber \\ & + & P(A_1\cap A_2\cap A_3) \nonumber \\ \label{equation19} \end{eqnarray} Eq. \ref{equation19} can be written in terms of summations: \begin{equation} P\displaystyle\left(\bigcup_{i=1}^{3}A_i\right)=\sum_{i=1}^{3}P(A_i)-\sum_{i=1}^{3-1}\sum_{j=i+1}^{3}P(A_i\cap A_j)+\sum_{i=1}^{3-2}\sum_{j=i+1}^{3-1}\sum_{k=j+1}^{3}P(A_i\cap A_j\cap A_k) \label{equation20} \end{equation} Which for $n$ events can be generalized to: \begin{eqnarray} P\displaystyle\left(\bigcup_{i=1}^{n}A_i\right) & = & \sum_{i=1}^{n}P(A_i)-\sum_{i=1}^{n-1}\sum_{j=i+1}^{n}P(A_i\cap A_j)+\ldots \nonumber \\ & + & (-1)^{L-1}\sum_{i=1}^{n-(L-1)}\sum_{j=i+1}^{n-(L-2)}...\sum_{l=m+1}^{n-(L-L)}P(A_i\cap A_j\cap ...\cap A_m\cap A_l)+\ldots \nonumber \\ & + & (-1)^{n-1}\sum_{i=1}^{n-(n-1)}\sum_{j=i+1}^{n-(n-2)}...\sum_{e=d+1}^{n-(n-n)}P(A_i\cap A_j\cap ...\cap A_d\cap A_e) \nonumber \\ \label{equation21} \end{eqnarray} A simpler notation for $Q(n)$ can be: \begin{eqnarray} P\displaystyle\left(\bigcup_{i=1}^{n}A_i\right) & = & \sum_{1\leq i\leq n}P(A_i)-\sum_{1\leq i < j\leq n}P(A_i\cap A_j)+\ldots \nonumber \\ & + & (-1)^{L-1}\sum_{1\leq i < j < ...< m < l\leq n}P(A_i\cap A_j\cap ...\cap A_m\cap A_l) +\ldots \nonumber \\ \label{equation22} \end{eqnarray} The equation that corresponds to $Q(n-1)$ can be written as: \begin{eqnarray} P\displaystyle\left(\bigcup_{i=2}^{n}A_i\right) & = & \sum_{2\leq i\leq n}P(A_i)-\sum_{2\leq i < j\leq n}P(A_i\cap A_j)+\ldots \nonumber \\ & + & (-1)^{L-1}\sum_{2\leq i < j < ...< m < l\leq n}P(A_i\cap A_j\cap ...\cap A_m\cap A_l)+\ldots \nonumber \\ \label{equation23} \end{eqnarray} And assuming $Q(n-1)$ we should prove $Q(n)$ in order to complete the proof by induction. From the left side of Eq. \ref{equation16}: \begin{equation} P\left(\bigcup_{i=1}^{n}A_i\right)=P\left[A_{1}\cup\left(\bigcup_{i=2}^{n}A_i\right)\right] \label{equation24} \end{equation} By applying $Q(2)$: \begin{eqnarray} P\left(\bigcup_{i=1}^{n}A_i\right) & = & P(A_{1})+P\left(\bigcup_{i=2}^{n}A_i\right) - P\left[A_{1}\cap\left(\bigcup_{i=2}^{n}A_i\right)\right] \nonumber \\ & = & P(A_{1})+P\left(\bigcup_{i=2}^{n}A_i\right) - P\left[\bigcup_{i=2}^{n}\left(A_1\cap A_i\right)\right] \nonumber \\ \label{equation25} \end{eqnarray} If we apply $Q(n-1)$ in the last two terms on the right of Eq. \ref{equation25}: \begin{eqnarray} P\left(\displaystyle\bigcup_{i=1}^{n}A_i\right) & = & P(A_1)+\displaystyle\sum_{2\leq i\leq n}P(A_i)-\displaystyle\sum_{2\leq i < j\leq n}P(A_i\cap A_j)+\ldots \nonumber \\ & + & (-1)^{L-1}\sum_{2\leq i < j < ...< m < l\leq n}P(A_i\cap A_j\cap ...\cap A_m\cap A_l) +\ldots \nonumber \\ & - & \left\{\displaystyle\sum_{2\leq i \leq n}P(A_1\cap A_i)+\ldots+(-1)^{L-2}\sum_{2\leq i < j < ...< m \leq n}P(A_1\cap A_i\cap ...\cap A_m)+...\right\} \nonumber \\ \label{equation26} \end{eqnarray} Performing the substitutions: \begin{equation} P(A_1)+\displaystyle\sum_{2\leq i\leq n}P(A_i)=\sum_{1\leq i\leq n}P(A_i) \label{equation27} \end{equation} \begin{equation} -\displaystyle\sum_{2\leq i \leq n}P(A_1\cap A_i)-\displaystyle\sum_{2\leq i < j\leq n}P(A_i\cap A_j)=-\displaystyle\sum_{1\leq i < j\leq n}P(A_i\cap A_j) \label{equation28} \end{equation} \begin{center} \ldots \end{center} \begin{eqnarray} \displaystyle -(-1)^{L-2}\sum_{2\leq i < j < ...< m \leq n}P(A_1\cap A_i\cap ...\cap A_m)\nonumber \\ +(-1)^{L-1}\sum_{2\leq i < j < ...< m < l\leq n}P(A_i\cap A_j\cap ...\cap A_m\cap A_l)\nonumber \\ =(-1)^{L-1}\sum_{1\leq i < j < ...< m < l\leq n}P(A_i\cap A_j\cap ...\cap A_m\cap A_l) \nonumber \\ \label{equation29} \end{eqnarray} We reach $Q(n)$, and the theorem is proved by mathematical induction. \end{proof} \begin{lem} If $A$ and $B$ are mutually exclusive events, then: \begin{equation} P(A\cup B)=P(A)+P(B) \label{equation30} \end{equation} \label{lemma3} \end{lem} \begin{proof}[Proof 1 (without theorem \ref{theorem2})] Since the sets are PME, according to definition \ref{definition3} and the theorem \ref{theorem1}: $P(A\cap B)=P(\emptyset)=0$. Considering the lemma \ref{lemma2} the implication is that $P(A\cup B)=P(A)+P(B)-P(A\cap B)=P(A)+P(B)$. \end{proof} \begin{proof}[Proof 2 (by using theorem \ref{theorem2})] By using the theorem \ref{theorem2} for $n=2$, and changing the notation of $A_1=A$ and $A_2=B$, PME, $P(A\cup B)=P(A)+P(B)$. \end{proof} \begin{lem}[Rule of addition of a finite number of ME events] Let the events $A_1$, $A_2$, \ldots, $A_n$ be ME. The following relation is valid: \begin{equation} P\left(\displaystyle\bigcup_{i=1}^{n}A_i\right)=\sum_{i=1}^{n}P(A_i) \label{equation31} \end{equation} \label{lemma4} \end{lem} \begin{proof} Groups of events that are ME as a whole are PME (lemma \ref{lemma4}). Then we can apply the theorem \ref{theorem2} in order to finish the proof. \end{proof} \begin{lem} \begin{equation} P(\overline{A})=1-P(A) \label{equation32} \end{equation} \label{lemma5} \end{lem} \begin{proof}[Proof 1 (based on lemma \ref{lemma3} and two axioms)] Given that $A$ and $\overline{A}$ are PME, then $P(A\cup\overline{A})=P(A)+P(\overline{A})$, according to axiom \ref{axiom3}. Considering the property of complementary events, $A\cup\overline{A}=\Omega$, we can use the axiom \ref{axiom2}: $P(A\cup\overline{A})=P(\Omega)=1=P(A)+P(\overline{A})$, therefore $P(\overline{A})=1-P(A)$. \end{proof} \begin{proof}[Proof 2 (based on \ref{lemma2} and one axiom)] Assume $B=\overline{A}$, then $P(A\cup\overline{A})=P(A)+P(\overline{A})-P(A\cap\overline{A})$. Since $P(A\cup\overline{A})=P(\Omega)$, according to the axiom \ref{axiom2}, we have $P(A\cup\overline{A})=1$, and given the fact that $A$ and $\overline{A}$ are PME (another property of the complementary sets), $P(A\cap\overline{A})=0$ (definition \ref{definition3}). Therefore, $1=P(A)+P(\overline{A})$, leading to Eq. \ref{equation32}. \end{proof} \begin{lem} If $A\subset B$, then $P(A)\leq P(B)$. \label{lemma6} \end{lem} \begin{proof} If $A\subset B$, then $B=A\cup(\overline{A}\cap B)$. Being $A$ and $\overline{A}\cap B$ PME, then lemma \ref{lemma3} applies, and $P(B)=P(A)+P(\overline{A}\cap B)$. Following the axiom \ref{axiom1}, $P(\overline{A}\cap B)\geq 0$, then $P(B)\geq P(A)$ and the proof is complete. \end{proof} \begin{lem} \begin{equation} P(A)\leq 1 \label{equation33} \end{equation} \label{lemma7} \end{lem} \begin{proof} By the definition of the space set, $A\subset\Omega$, and from lemma \ref{lemma6} $P(A)\leq P(\Omega)$. The direct application of the axiom \ref{axiom2}, leads to the result: $P(A)\leq P(\Omega)=1$. \end{proof} \subsection{Dependency among events} \label{dependencyevents} \begin{defi} The conditional probability of the event $A$ given the event $B$, $P(A|B)$, is defined for $P(B)>0$ as: \begin{equation} P(A|B)=\displaystyle\frac{P(A\cap B)}{P(B)} \label{equation34} \end{equation} \label{definition8} \end{defi} \begin{lem} Given $A$ and $B$, and $P(B) > 0$, it's true that $0\leq P(A|B)\leq 1$. \label{lemma8} \end{lem} \begin{proof} According the definition of $P(A|B)$, that depends on $P(A\cap B)$ and $P(B)>0$, once $0\leq P(C)\leq 1$ for any $C$ (axiom \ref{axiom1} and lemma \ref{lemma7}), then both $0\leq P(A|B)\leq 1$ and $0\leq P(A\cap B) \leq 1$ are true. The first condition is the proof of the lemma, and the second can be used to determine the limits of $P(A\cap B)$, leading to $0\leq P(A\cap B)\leq P(B)$. When $P(A\cap B)=P(B)$, $P(A|B)=1$. And if $P(A\cap B)=0$, that is, the events $A$ and $B$ are PME, then $P(A|B)=0$. \end{proof} \begin{prop} If $A$ and $B$ are PME events, then $P(A|B)=0$. \label{proposition1} \end{prop} \begin{proof} If $A$ and $B$ are PME, $A\cap B=\emptyset$ (definition \ref{definition3}), and since $P(\emptyset)=0$ (theorem \ref{theorem1}), then $P(A|B)=0$. Based on the definition of $P(A|B)$ (equation \ref{equation34}), we have $P(A|B)=0$. \end{proof} \begin{prop} If the event $B$ implies the event $A$, that is, $B\subset A$, then $P(A|B)=1$. \label{proposition2} \end{prop} \begin{proof} If $B\subset A$, $P(A\cap B)=P(B)$. By definition, $P(A|B)=P(A\cap B)/P(B)$, hence $P(A|B)=P(B)/P(B)=1$. \end{proof} \begin{prop}[Rule of addition for conditional probabilities] If $A_1$, $A_2$, ..., $A_n$ are ME with union $A=\displaystyle\bigcup_{i=1}^nA_i$, then: \begin{equation} P(A|B)=\displaystyle\sum_{i=1}^nP(A_i|B) \label{equation35} \end{equation} \label{proposition3} \end{prop} \begin{proof} According the definition of $P(A|B)$ (definition \ref{definition8}) and its relation with $P(B)>0$ and $P(A\cap B)$: \begin{equation} P(A|B)=\displaystyle\frac{P(A\cap B)}{P(B)}=\frac{1}{P(B)}P\displaystyle\left(B\cap\bigcup_{i=1}^{n}A_i\right)=\frac{1}{P(B)}P\displaystyle\left[\bigcup_{i=1}^{n}(A_i\cap B)\right] \label{equation36} \end{equation} Since $A_i$ and $A_j$ are PME for $i\neq j$, then $A_i\cap B$ and $A_j\cap B$ also are PME: \begin{equation} A_i\cap A_j=\emptyset \Rightarrow (A_i\cap A_j)\cap B=\emptyset\cap B =\emptyset\Rightarrow (A_i\cap B)\cap (A_j\cap B)=\emptyset \label{equation37} \end{equation} Therefore we can apply lemma \ref{lemma4}: \begin{equation} P(A|B)=\displaystyle\frac{1}{P(B)}P\displaystyle\left[\bigcup_{i=1}^{n}(A_i\cap B)\right]=\displaystyle\frac{1}{P(B)}\sum_{i=1}^{n}P(A_i\cap B)=\sum_{i=1}^{n}\frac{P(A_i\cap B)}{P(B)}=\displaystyle\sum_{i=1}^{n}P(A_i|B) \label{equation38} \end{equation} \end{proof} \begin{defi} The event $A$ is said to be independent of the event $B$, or statistically independent (SI), if and only if $P(A|B)=P(A)$. \label{definition9} \end{defi} \begin{thm}[Rule of the product of probabilities] Let the events $A_1$, $A_2$, ..., $A_n$ with $P\left(\displaystyle\bigcup_{i=1}^{n}A_i\right)\geq 0$ and $P\left(\displaystyle\bigcup_{i=1}^{k}A_i\right)> 0$ for $1\leq k < n$. It can be shown that: \begin{equation} P\left(\displaystyle\bigcap_{i=1}^{n}A_i\right)=P(A_1)\times P(A_2|A_1)\times \ldots \times P(A_n|A_1\cap ...\cap A_{n-1}) \label{equation39} \end{equation} \label{theorem5} \end{thm} \begin{proof} Using the definition of $P(A|B)$ repeatedly for each factor in the product: $$P\displaystyle\left(\bigcap_{i=1}^{n}A_i\right)=\frac{P\displaystyle\left(\bigcap_{i=1}^{n}A_i\right)}{P\displaystyle\left(\bigcap_{i=1}^{n-1}A_i\right)}\times\frac{P\displaystyle\left(\bigcap_{i=1}^{n-1}A_i\right)}{P\displaystyle\left(\bigcap_{i=1}^{n-2}A_i\right)}\times ...\times \frac{P(A_1\cap A_2)}{P(A_1)}P(A_1)$$ \begin{equation} P\displaystyle\left(\bigcap_{i=1}^{n}A_i\right)=P\left[A_n|\displaystyle\left(\bigcap_{i=1}^{n-1}A_i\right)\right]\times P\left[A_{n-1}|\displaystyle\left(\bigcap_{i=1}^{n-2}A_i\right)\right]\times ... \times P(A_2|A_1)P(A_1) \label{equation40} \end{equation} \end{proof} \begin{lem} If $A$ and $B$ are SI, and both $P(A)$ and $P(B)$ are not zero, then $P(A\cap B)=P(A)P(B)$ \label{lemma9} \end{lem} \begin{proof} The definition for $P(A|B)=P(A\cap B)/P(B)$ also applies backwards, $P(B|A)=P(A\cap B)/P(A)$. If $A$ and $B$ are SI, given the definition \ref{definition9} $P(A|B)=P(A)$ and $P(B|A)=P(B)$. Thus, in both cases one can show that $P(A\cap B)=P(A)P(B)$, since $P(A\cap B)=P(A|B)P(B)=P(A)P(B)$ and $P(A\cap B)=P(B|A)P(A)=P(B)P(A)$. \end{proof} \begin{defi} The events $A_1$, $A_2$, ..., $A_n$ are defined as mutually independents (MI) if $P\displaystyle\left(\bigcap_{i=1}^{n}A_i\right)=\prod_{i=1}^{n}P(A_i)$ for all combinations of sets between $1$ e $n$. \label{definition10} \end{defi} \begin{lem} If the events $A_1$, $A_2$, ..., $A_n$ are MI, then any pair $A_i$ e $A_j$, for $i\neq j$, are SI. \label{lemma10} \end{lem} \begin{proof} According the definition \ref{definition10} $A_i$ and $A_j$ must be SI, for the independence is valid for the combination of any number of sets, including pairs. Hence $P(A_i\cap A_j)=P(A_i)P(A_j)$ for any $i\neq j$ if the events $A_1$, $A_2$, ..., $A_n$ are MI. Notice that the opposite is not necessarily true: by assuming $P(A_i\cap A_j)=P(A_i)P(A_j)$ for any pair $(i,j)$ one do not prove $P(A_i\cap A_j\cap A_k)=P(A_i)P(A_j)P(A_k)$ for all $(i,j,k)$, and higher order groups. \end{proof} \begin{lem} If the sets $A$ and $B$ are SI, then they are not PME, and vice versa. \label{lemma11} \end{lem} \begin{proof} If $A$ and $B$ are SI, then $P(A\cap B)=P(A)P(B)$ (lemma \ref{lemma9}). PME events are such that $A\cap B=\emptyset$, and $P(A\cap B)=P(\emptyset)$. According theorem $\ref{theorem1}$, $P(\emptyset)=0$, thus PME events ($P(A\cap B)=0$) cannot be SI ($P(A\cap B)=P(A)P(B)$). Notice that the events $A$ and $B$ are such that $P(A)>0$ and $P(B)>0$, respectively (lemma \ref{lemma9}). \end{proof} \begin{lem} If the set $\{C_1,...,C_n\}$ is a partition of the sample space, $\Omega$, then for any event $A$: \begin{equation} P(A)=\displaystyle\sum_{i=1}^{n}P(A|C_i)P(C_i) \label{equation41} \end{equation} \label{lemma12} \end{lem} \begin{proof} The definition of $P(A|C_i)$ (Eq. \ref{equation34}) implies $P(A|C_i)=P(A\cap C_i)/P(C_i)$. Making the summation from $i=1$ to $n$ on both sides of the equation $P(A|C_i)P(C_i)=P(A\cap C_i)$: \begin{equation} \displaystyle\sum_{i=1}^nP(A|C_i)P(C_i)=\sum_{i=1}^nP(A\cap C_i) \label{equation42} \end{equation} Since $C_i$ and $C_j$ are disjoint for any $i\neq j$, then is also true that $(A\cap C_i)\cap (A\cap C_j)=\emptyset$ (see proof of proposition \ref{proposition3}). Therefore, by using the theorem \ref{theorem2}: \begin{eqnarray} \displaystyle\sum_{i=1}^nP(A|C_i)P(C_i) & = & P\left[\bigcup_{i=1}^n (A\cap C_i)\right]=P\left[A\cap \left(\bigcup_{i=1}^n C_i\right)\right]\nonumber \\ & = & P(A\cap \Omega) = P(A|\Omega)P(\Omega)=P(A) \nonumber \\ \label{equation43} \end{eqnarray} Where we have used the definition of a partition in $\Omega$ (definition \ref{definition5}), the axiom \ref{axiom2}, the definition of conditional probability and the implicit identity $P(A|\Omega)$ for any $A$ ($A\subset \Omega$, so any element of the set $A$ is also an element of the sample space). \end{proof} \begin{thm}[Bayes' theorem] Given the event $A$ is such that $P(A)\geq 0$, and the set $\{C_1, ...,C_n\}$, which defines a partition in $\Omega$, with $P(C_i)\geq 0$ for every $i$, it's possible to prove that: \begin{equation} P(C_i|A)=\displaystyle\frac{P(A|C_i)P(C_i)}{\displaystyle\sum_{i=1}^{n}P(A|C_i)P(C_i)} \label{equation44} \end{equation} \label{theorem6} \end{thm} \begin{proof} Due to the fact that $P(C_i|A)=P(A\cap C_i)/P(A)$ and $P(A|C_i)=P(A\cap C_i)/P(C_i)$ (definition \ref{definition8}), then: \begin{equation} P(C_i|A)=\frac{P(A|C_i)P(C_i)}{P(A)} \label{equation45} \end{equation} We can substitute the result of lemma \ref{lemma12} for $P(A)$ in Eq. \ref{equation45} to prove the theorem. \end{proof} \subsection{Probabillity theorems diagram} \label{probdiagram} In order to represent the main content of this paper, namely, the most relevant equations presented so far, and the relations among them, the diagram in Fig. \ref{diagram} was prepared. It omits some results and assumptions, focusing on the fundamental relations between axioms and results. A line divides the figure in two parts: below are the results proved in section \ref{combinationsofevents}, and above it the results from section \ref{dependencyevents} are listed and interrelated. \begin{figure}[!htbp] \centering \includegraphics[width=\textwidth]{diagram} \caption{Main results from Kolmogorov's axioms (in gray boxes). Those above the dashed line correspond to dependent events, and those below the line are related to combinations of events.} \label{diagram} \end{figure} \section{Conclusion} \label{conclusion} I hope the set of proofs and the choice of theorems/lemmas/propositions and definitions, as much as the order they are presented, help students of mathematics, statistics, engineering, chemistry and physics to se a broad picture of the Kolmogorov's axiomatic system, even if in a simplified and incomplete form. \section{Acknowledgements} \label{acknowledgements} I must thank CNPq and CAPES for the PhD scholarship which allowed me to investigate the theme.
{ "timestamp": "2022-06-14T02:04:32", "yymm": "2206", "arxiv_id": "2206.05396", "language": "en", "url": "https://arxiv.org/abs/2206.05396", "abstract": "A selection of the relevant theorems of Probability Theory that comes directly from Kolmogorov's axioms, Set Theory basic results, definitions and rules of inference are listed and proven in a systematic approach, aiming the student who seeks a self-contained account on the matter before moving to more advanced material.", "subjects": "General Mathematics (math.GM)", "title": "A systematic approach on some relevant theorems that follows from Kolmogorov's axioms", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9867771805808551, "lm_q2_score": 0.8128673133042217, "lm_q1q2_score": 0.8021189156086744 }
https://arxiv.org/abs/math-ph/0511032
A second eigenvalue bound for the Dirichlet Schroedinger operator
Let $\lambda_i(\Omega,V)$ be the $i$th eigenvalue of the Schrödinger operator with Dirichlet boundary conditions on a bounded domain $\Omega \subset \R^n$ and with the positive potential $V$. Following the spirit of the Payne-Pólya-Weinberger conjecture and under some convexity assumptions on the spherically rearranged potential $V_\star$, we prove that $\lambda_2(\Omega,V) \le \lambda_2(S_1,V_\star)$. Here $S_1$ denotes the ball, centered at the origin, that satisfies the condition $\lambda_1(\Omega,V) = \lambda_1(S_1,V_\star)$.Further we prove under the same convexity assumptions on a spherically symmetric potential $V$, that $\lambda_2(B_R, V) / \lambda_1(B_R, V)$ decreases when the radius $R$ of the ball $B_R$ increases.We conclude with several results about the first two eigenvalues of the Laplace operator with respect to a measure of Gaussian or inverted Gaussian density.
\section{Introduction} In an earlier publication \cite{AB2}, Ashbaugh and one of us have proven the Payne-P\'olya-Weinberger (PPW) conjecture, which states that the first two eigenvalues $\lambda_1, \lambda_2$ of the Dirichlet-Laplacian on a bounded domain $\Omega\subset \mathbb R^n$ ($n\ge 2$) obey the bound \begin{equation} \label{EqPPW} \lambda_2 / \lambda_1 \le j_{n/2,1}^2/j_{n/2-1,1}^2. \end{equation} Here $j_{\nu,k}$ stands for the $k$th positive zero of the Bessel function $J_\nu$. Thus the right hand side of (\ref{EqPPW}) is just the ratio of the first two eigenvalues of the Dirichlet-Laplacian on an $n$-dimensional ball of arbitrary radius. This result is optimal in the sense that equality holds in (\ref{EqPPW}) if and only if $\Omega$ is a ball. The proof of the PPW conjecture has been generalized in several ways. In \cite{AB1} a corresponding theorem has been established for the Laplacian operator on a domain $\Omega$ that is contained in a hemisphere of the $n$-dimensional sphere $\mathbb S^n$. More precisely, it has been shown that $\lambda_2(\Omega) \le \lambda_2(S_1)$, where $S_1$ is the $n$-dimensional geodesic ball in $\mathbb S^n$ that has $\lambda_1(\Omega)$ as its first Dirichlet eigenvalue. A further variant of the PPW conjecture has been considered by Haile. In \cite{H} he compares the second eigenvalue $\lambda_2(\Omega,kr^\alpha)$ of the Schr\"o\-din\-ger operator with the potential $V = kr^\alpha$ ($k>0, \alpha\ge 2$) with $\lambda_2(S_1,kr^\alpha)$, where $S_1$ is the ball, centered at the origin, that satisfies the condition $\lambda_1(\Omega,kr^\alpha) = \lambda_1(S_1,kr^\alpha)$. Here and in the following we denote by $\lambda_i(\Omega,V)$ the $i$th eigenvalue of the Schr\"o\-din\-ger operator $-\Delta + V(\vec r)$ with Dirichlet boundary conditions on a bounded domain $\Omega \subset \mathbb{R}^n$. We have to mention a gap in \cite{H}, which occurs in the proof of Lemma 3.2. The author claims (and uses) that all derivatives of the function $Z(\theta)$ (which is equal to $T'(\theta)$ where $T(\theta)=0$) coincide with the derivatives of $T'(\theta)$ in the points where $T(\theta)=0$. This is not proven and there seems to be no reason why it should be true. The same problem occurs in the proof of Lemma 3.3. In the present paper we will prove a theorem that includes Haile's theorem as a special case and thus remedies the situation. One very important difference between the original PPW conjecture and the extended problems in \cite{AB1,H} is that in the later cases the ratio $\lambda_2/\lambda_1$ is not scaling invariant anymore. While $\lambda_2/\lambda_1$ is the same for any ball in $\mathbb R^n$, it is an increasing function of the radius for balls in $\mathbb S^n$ \cite{AB1}. On the other hand, we will see that $\lambda_2(B_R,V)/\lambda_1(B_R,V)$ on the ball $B_R$ is a decreasing function of the radius $R$, if $V$ has certain convexity properties. This rises the question which is the `right size' of the comparison ball in the PPW estimate. We will make some remarks on this problem below. The main objective of the present work is to prove a PPW type result for a Schr\"odinger operator with a positive potential. We will state the corresponding theorem in the following section. In Section \ref{SectionGauss} we will transfer our results to the case of a Laplacian operator with respect to a metric of Gaussian or inverted Gaussian measure, the two cases of which are closely related to the harmonic oscillator. The rest of the article will be devoted to the proofs of our results. \section{Main Results} \label{SectionResults} Let $\Omega \subset \mathbb{R}^n$ (with $n\ge 2$) be some bounded domain and $V : \Omega \rightarrow \mathbb{R}^+$ some positive potential such that the Schr\"odinger operator $-\Delta + V$ (subject to Dirichlet boundary conditions) is self-adjoint in $L^2(\Omega)$. We call $\lambda_i(\Omega,V)$ its $i$th eigenvalue. Further, we denote by $V_\star$ the radially increasing rearrangement of $V$. Then the following PPW type estimate holds: \begin{Theorem} \label{TheoremPPW} Let $S_1 \subset \mathbb{R}^n$ be a ball centered at the origin and of radius $R_1$ and let $\tilde V: S_1 \rightarrow \mathbb{R}^+$ be some radially symmetric positive potential such that $\tilde V(r) \le V_\star(r)$ for all $0 \le r \le R_1$ and $\lambda_1(\Omega,V) = \lambda_1(S_1,\tilde V)$. If $\tilde V(r)$ satisfies the conditions \begin{enumerate} \item[a)] $\tilde V(0) = \tilde V'(0)=0$ and \item[b)] $\tilde V'(r)$ exists and is increasing and convex, \end{enumerate} then \begin{equation}\label{EqTheorem} \lambda_2(\Omega,V) \le \lambda_2(S_1,\tilde V). \end{equation} \end{Theorem} If $V$ is such that $V_\star$ satisfies the convexity conditions stated in the theorem, the best bound is obtained by choosing $\tilde V = V_\star$. In this case the theorem is a typical PPW result and optimal in the sense that equality holds in (\ref{EqTheorem}) if $\Omega$ is a ball and $V = V_\star$. For a general potential $V$ we still get a non-trivial bound on $\lambda_2(\Omega,V)$ though it is not sharp anymore. To show that our Theorem \ref{TheoremPPW} contains Haile's result \cite{H} as a special case, we state the following corollary: \begin{Corollary} Let $\tilde V: \mathbb{R}^n \rightarrow \mathbb{R}^+$ be a radially symmetric positive potential that satisfies the conditions a) and b) of Theorem \ref{TheoremPPW} and let $S_1 \subset \mathbb{R}^n$ be the ball (centered at the origin) such that $\lambda_1(\Omega,\tilde V) = \lambda_1(S_1,\tilde V)$. Then $$\lambda_2(\Omega,\tilde V) \le \lambda_2(S_1,\tilde V).$$ \end{Corollary} The proof of Theorem \ref{TheoremPPW} follows the lines of the proof in \cite{AB2} and will be presented in Section \ref{SectionProof}. Let us make a few remarks on the conditions that $\tilde V$ has to satisfy. Condition a) is not a very serious restriction, because any bounded potential can be shifted such that $V_\star(0) = 0$. Also $V_\star'(0) = 0$ holds if $V$ is somewhat regular where it takes the value zero. Moreover, our method relies heavily on the fact that \begin{equation} \label{EqLambda} \lambda_2(B_R,\tilde V) \ge \left(1+\frac 2n\right) \lambda_1(B_R,\tilde V), \end{equation} which is a byproduct of our proof and holds for any ball $B_R$ and any potential $\tilde V$ that satisfies the conditions of Theorem \ref{TheoremPPW}. The conditions a) and b) will be needed to show the above inequality, which is equivalent to $q''(0) \le 0$ for a function $q$ to be defined in the proof. Numerical studies indicate that b) is somewhat sharp in the sense that, for example, a potential $r^{2-\epsilon}$ (which violates b) only `slightly') does not satisfy (\ref{EqLambda}) for every $R$. In this case the statement of Theorem \ref{TheoremPPW} may still be true, but the typical scheme of the PPW proof will fail. Furthermore, condition a) and b) will allow us to employ the crucial Baumgartner-Grosse-Martin (BGM) inequality \cite{BGM,AB3}: From a) and b) we see that $V(r) + rV'(r)$ is increasing. Consequently $rV(r)$ is convex, which is just the condition needed to apply the BGM inequality. As mentioned above, one has to chose carefully the size of the comparison ball in a PPW estimate if $\lambda_2/\lambda_1$ is a non-constant function of the ball's radius. In the case of the Laplacian on $\mathbb S^n$, one compares the second eigenvalues on $\Omega$ and $S_1$, the ball that has the same first eigenvalue as $\Omega$. By the Rayleigh-Faber-Krahn (RFK) inequality for $\mathbb S^n$ it is clear that $S_1 \subset \Omega^\star$, where $\Omega^\star$ is the spherically symmetric rearrangement of $\Omega$. It has also be shown in \cite{AB1} that $\lambda_2/\lambda_1$ on a geodesic ball in $\mathbb S^n$ is an increasing function of the ball's radius. One can conclude from these two facts that in $\mathbb S^n$ an estimate of the type (\ref{EqTheorem}) is stronger than the inequality \begin{equation} \label{EqLL} \lambda_2(\Omega)/\lambda_1(\Omega) \le \lambda_2(\Omega^\star) / \lambda_1(\Omega^\star). \end{equation} It has also been argued in \cite{AB3} why the situation is different in the hyperbolic space $\mathbb H^n$. Here an estimate of the type (\ref{EqLL}) is not possible, for the following reason: One can show that $\lambda_2/\lambda_1$ on geodesic balls in $\mathbb H^n$ is a decreasing function of the radius. Now suppose, for example, that $\Omega$ is the ball $B_R$ with very long and thin tentacles attached to it. Then the first and the second eigenvalue of the Laplacian on $\Omega$ and $B_R$ are almost the same, while the ratio $\lambda_2/\lambda_1$ on $\Omega^\star$ can be considerably less than on $B_R$ (and thus on $\Omega$). We will prove a PPW inequality of the type $\lambda_2(\Omega) \le \lambda_2(S_1)$ for $\mathbb H^n$ and the monotonicity of $\lambda_2/\lambda_1$ on geodesic balls in a future publication. To shed light on the question which is the right type of PPW inequality for the Schr\"odinger operator on $\Omega$, we state \begin{Theorem} \label{TheoremMonotonicity} Let $V: \mathbb{R}^n \rightarrow \mathbb{R}^+$ be a spherically symmetric potential that satisfies the conditions of Theorem \ref{TheoremPPW}, i.e. \begin{enumerate} \item[a)] $V(0) = V'(0)=0$ and \item[b)] $V'(r)$ exists and is increasing and convex. \end{enumerate} Then the ratio $$\frac{\lambda_2(B_R, V)}{\lambda_1(B_R, V)}$$ is a decreasing function of $R$. \end{Theorem} This theorem shows that one can not replace equation (\ref{EqTheorem}) in our Theorem \ref{TheoremPPW} by an inequality of the type (\ref{EqLL}), following the same reasoning as in the case of the Laplacian on $\mathbb H^n$. Theorem \ref{TheoremMonotonicity} will be proven in Section \ref{SectionProof2}. \section{Connection to the Laplacian operator in Gaussian space} \label{SectionGauss} Recently, there has been some interest in isoperimetric inequalities in $\mathbb{R}^n$ endowed with a measure of Gaussian ($\,\mathrm{d} \mu_- = e^{-r^2/2}\,\mathrm{d}^nr$) or inverted Gaussian ($\,\mathrm{d} \mu_+ = e^{+r^2/2}\,\mathrm{d}^nr$) density. For the Gaussian space it has been known for several years that a classical isoperimetric inequality holds. Yet the ratio of Gaussian perimeter and Gaussian measure is minimized by half-spaces instead of spherical domains \cite{B}. The `inverted Gaussian' case, i.e., $\mathbb{R}^n$ with the measure $\mu_+$, is more similar to the Euclidean case: It has been shown recently that a classical isoperimetric inequality holds and that the minimizers are balls centered at the origin \cite{MPB}. We consider the Dirichlet-Laplacians $-\Delta_\pm$ on $L^2(\Omega,\,\mathrm{d}\mu_\pm)$, where $\Omega\varsubsetneqq\mathbb{R}^n$ is a domain of finite measure $\,\mathrm{d}\mu_\pm(\Omega)$. These two operators are defined by their quadratic forms \begin{equation} \label{EqQuadForm} h_\pm[\Psi] = \int_\Omega |\nabla \Psi(\vec r)|^2 \,\mathrm{d} \mu_\pm, \quad \Psi \in W^{1,2}_0(\Omega,\,\mathrm{d} \mu_\pm). \end{equation} The eigenfunctions $\Psi^\pm_i$ and eigenvalues $\lambda^\pm_i(\Omega)$ in question are determined by the the differential equation \begin{equation} \label{EqEVProblem} -\sum\limits_{k=1}^n \frac{\partial}{\partial r_k} \left(e^{\pm r^2} \frac{\partial \Psi^\pm_i}{\partial r_k}\right) = \lambda^\pm_i(\Omega) e^{\pm r^2} \Psi^\pm_i(\vec r). \end{equation} There is a tight connection between the operators $-\Delta_\pm$ on a domain $\Omega$ and the harmonic oscillator $-\Delta+r^2$ restricted to $\Omega$. Their eigenfunctions and eigenvalues are related by \cite{BCF} \begin{eqnarray} \Psi^\pm_i(\vec r) &=& \Psi_i(\vec r) \cdot e^{\mp r^2/2} \quad \textmd{and} \nonumber\\ \lambda^\pm_i(\Omega) &=& \lambda_i(\Omega, r^2) \pm n, \label{EqCon} \end{eqnarray} denoting by $\Psi_i$ the Dirichlet eigenfunctions of $-\Delta + r^2$ on $\Omega$. There is an equivalent of the RFK inequality in Gaussian space \cite{BCF} stating that $\lambda^-_1(\Omega)$ is minimized for given $\mu_-(\Omega)$ if $\Omega$ is a half-space. The corresponding fact for the `inverted' Gaussian space is that $\lambda^+_1(\Omega)$ is minimized for given $\mu_+(\Omega)$ by the ball centered at the origin. It can be seen by the RFK inequality for Schr\"odinger operators \cite{L} in combination with (\ref{EqCon}). Concerning the second eigenvalue, we will now show what our results from Section \ref{SectionResults} imply for the operators $-\Delta_\pm$. We state \begin{Theorem} \label{TheoremMonotonicity2} For the operator $-\Delta_+$ on a ball $B_R$ of radius $R$ (centered at the origin) the ratio $\lambda_2^+(B_R)/\lambda_1^+(B_R)$ is a strictly decreasing function of $R$. \end{Theorem} In Section \ref{SectionProof3} we will derive Theorem \ref{TheoremMonotonicity2} from Theorem \ref{TheoremMonotonicity} in a purely algebraic way using only the relation (\ref{EqCon}). Repeating the argument for $\mathbb H^n$ from the previous section, we see that by Theorem \ref{TheoremMonotonicity2} the best PPW result we can expect to get is \begin{Theorem} \label{TheoremPPW2} Be $S_1$ the ball (centered at the origin) that satisfies the condition $\lambda^+_1(S_1) = \lambda^+_1(\Omega)$. Then \begin{equation*} \lambda_2^+(\Omega) \le \lambda_2^+(S_1). \end{equation*} \end{Theorem} Theorem \ref{TheoremPPW2} follows immediately from Theorem \ref{TheoremPPW} and (\ref{EqCon}). In the same way we easily get the corresponding version for $-\Delta_-$: \begin{Theorem} \label{TheoremPPW3} Be $S_1$ the ball (centered at the origin) that satisfies the condition $\lambda^-_1(S_1) = \lambda^-_1(\Omega)$. Then \begin{equation*} \lambda_2^-(\Omega) \le \lambda_2^-(S_1). \end{equation*} \end{Theorem} Yet in this case it is not clear anymore whether $S_1$ is the optimal comparison ball: First, in contrast to the `inverted' Gaussian case the ratio $\lambda_2^-(B_R)/\lambda_1^-(B_R)$ is not a decreasing function of $R$ anymore. This can be seen by comparing the values of $\lambda_2^-(B_R)/\lambda_1^-(B_R)$ for $R\rightarrow 0$ and $R\rightarrow \infty$: For small $R$ the ratio is close to the Euclidean value ($\approx 2.539$) while for large $R$ it approaches infinity (by (\ref{EqCon})). Second, the RFK inequality in Gaussian space states that $\lambda_1^-(\Omega)$ is minimized by half-spaces, not circles. This means that for general $\Omega$ we don't know whether $\Omega^\star$ is bigger or smaller than $S_1$. For these differences it remains unclear what is the most natural way to generalize the PPW conjecture to Gaussian space. \section{A monotonicity lemma} \label{SectionLemma} In our proof of Theorem \ref{TheoremPPW} we will need \begin{Lemma}[Monotonicity of $g$ and $B$] \label{LemmaMonotonicityOfgAndB} Let $\tilde V$, $S_1$ and $R_1$ be as in Theorem \ref{TheoremPPW} and call $z_1(r)$ and $z_2(r)$ the radial parts (both chosen positive) of the first two Dirichlet eigenfunctions of $-\Delta + \tilde V$ on $S_1$. Set \begin{eqnarray*} g(r) &=& \frac{z_2(r)}{z_1(r)} \quad \textmd{and}\\ B(r) &=& g'(r)^2 + (n-1)\frac{g(r)^2}{r^2} \end{eqnarray*} for $0 < r < R_1$. Then $g(r)$ is increasing on $(0,R_1)$ and $B(r)$ is decreasing on $(0,R_1)$. \end{Lemma} \begin{proof} \cite{H,AB0} In this section we abbreviate $\lambda_i = \lambda_i(S_1,\tilde V)$. The functions $z_1$ and $z_2$ are solutions of the differential equations \begin{eqnarray} -z''_{1} - \frac{n-1}{r} z'_{1} + \left(\tilde V -\lambda_1\right) z_{1} &=& 0, \label{EquN1}\\ -z''_{2} - \frac{n-1}{r} z'_{2} + \left(\frac{n-1}{r^2} + \tilde V - \lambda_2\right) z_{2} &=& 0 \nonumber \end{eqnarray} with the boundary conditions \begin{equation}\label{EqBoundary} z'_{1}(0) = 0, \quad z_{1}(R_1) = 0, \quad z_{2}(0) = 0, \quad z_{2}(R_1) = 0. \end{equation} This is assured by the BGM inequality \cite{AB0,BGM}, which is applicable because $r\tilde V$ is convex. As in \cite{AB0} we define the function $$q(r) := \frac{rg'(r)}{g(r)}.$$ Proving the lemma is thus reduced to showing that $0 < q(r) < 1$ and $q'(r) < 0$ for $r \in [0,R]$. Using the definition of $g$ and the equations (\ref{EquN1}), one can show that $q(r)$ is a solution of the Riccati differential equation \begin{equation} \label{EqRic} q' = (\lambda_1-\lambda_2) r + \frac{(1-q)(q+n-1)}r - 2q\frac{z_1'}{z_1}. \end{equation} It is straightforward to establish the boundary behavior $$q(0) = 1, \quad q'(0) = 0, \quad q''(0) = \frac 2n \left(\left(1+\frac 2n\right)\lambda_1 - \lambda_2\right)$$ and $$q(R_1) = 0.$$ \begin{Fact} \label{Fact1} For $0 \le r \le R$ we have $q(r) \ge 0$. \end{Fact} \begin{proof} Assume the contrary. Then there exist two points $0 < r_1 < r_2 \le R_1$ such that $q(r_1) = q(r_2) = 0$ but $q'(r_1) \le 0$ and $q'(r_2) \ge 0$. If $r_2 < R_1$ then the Riccati equation (\ref{EqRic}) yields $$0 \ge q'(r_1) = (\lambda_1-\lambda_2) r_1 + \frac{n-1}{r_1} > (\lambda_1-\lambda_2) r_2 + \frac{n-1}{r_2} = q'(r_2) \ge 0,$$ which is a contradiction. If $r_2 = R_1$ then we get a contradiction in a similar way by $$0 \ge q'(r_1) = (\lambda_1-\lambda_2) r_1 + \frac{n-1}{r_1} > (\lambda_1-\lambda_2) R_1+ \frac{n-1}{R_1} = 3q'(R_1) \ge 0.$$ \end{proof} In the following we will analyze the behavior of $q'$ according to (\ref{EqRic}), considering $r$ and $q$ as two independent variables. For the sake of a compact notation we will make use of the following abbreviations: \begin{equation*} \begin{array}{rclrcl} p(r) &=& z_1'(r) /z_1(r) \quad \quad & N_y &=& y^2 - n + 1 \cr % \nu &=& n-2 & M_y &=& N_y^2/(2y) - \nu^2y/2 \cr % E &=& \lambda_2-\lambda_1 & Q_y &=& 2y\lambda_1 + EN_yy^{-1} - 2E \end{array} \end{equation*} We further define the function \begin{equation}\label{EqT} T(r,y) := -2 p(r) y - \frac{\nu y + N_y}{r} - E r. \end{equation} Then we can write (\ref{EqRic}) as \begin{equation*} q'(r) = T(r,q(r)) \end{equation*} The definition of $T(r,y)$ allows us to analyze the Riccati equation for $q'$ considering $r$ and $q(r)$ as independent variables. For $r$ going to zero, $p$ is $\mathcal{O}(r)$ and thus $$T(r,y) = \frac{1}{r}\left((\nu+1+y)(1-y)\right) + \mathcal{O}(r) \quad \textmd{for }y\textmd{ fixed}.$$ Consequently, \begin{equation*} \begin{array}{rcll} \lim_{r\rightarrow 0} T(r,y) &=& +\infty \quad \quad & \textmd{for } 0 \le y < 1 \,\textmd{ fixed,}\cr% \lim_{r\rightarrow 0} T(r,y) &=& 0 & \textmd{for }\, y = 1 \textmd{ and }\cr% \lim_{r\rightarrow 0} T(r,y) &=& -\infty & \textmd{for }\, y > 1 \,\textmd{ fixed.} \end{array} \end{equation*} For $r$ approaching $R_1$, the function $p(r)$ goes to minus infinity, while all other terms in (\ref{EqT}) are bounded. Therefore $$\lim_{r\rightarrow R_1} T(r,y) = +\infty \quad \textmd{ for } y > 0 \textmd{ fixed}.$$ The partial derivative of $T(r,y)$ with respect to $r$ is given by \begin{equation} \label{EqTprime} T' = \frac{\partial}{\partial r} T(r,y) = -2yp' + \frac{\nu y}{r^2} + \frac{N_y}{r^2}- E. \end{equation} In the points $(r,y)$ where $T(r,y) = 0$ we have, by (\ref{EqT}), \begin{equation}\label{EqPAtTZero} p|_{T=0} = -\frac {\nu}{2r}-\frac{N_y}{2yr}-\frac{Er}{2y}. \end{equation} From (\ref{EquN1}) we get the Riccati equation \begin{equation}\label{EqRicp} p'+p^2+\frac{\nu+1}{r}p+\lambda_1-\tilde V=0. \end{equation} Putting (\ref{EqPAtTZero}) into (\ref{EqRicp}) and the result into (\ref{EqTprime}) yields \begin{equation} T'|_{T=0} = \frac{M_y}{r^2} + \frac{E^2r^2}{2y}+Q_y-2y\tilde V. \end{equation} If we define the function $$Z_y(r) := \frac{M_y}{r^2} + \frac{E^2r^2}{2y}+Q_y-2y\tilde V,$$ it is clear that $T'(r,y) = Z_y(r)$ for any $r,y$ with $T(r,y)=0$. The behavior of $Z_y(r)$ at $r=0$ is determined by $M_y$. From the definition of $M_y$ we get \begin{equation}\label{EqMMM} yM_y = \frac 12(y^2-1)\cdot[(y-1)-(n-2)]\cdot[(y+1)+(n-2)]. \end{equation} This implies that \begin{eqnarray*} M_y &>& 0 \quad \textmd{for } 0<y<1,\\ M_1 &=& 0. \end{eqnarray*} and therefore \begin{eqnarray*} \lim_{r\rightarrow 0} Z_y(r) &=& \infty \quad \textmd{for } 0<y<1. \end{eqnarray*} \begin{Fact} \label{Fact2} There is some $r_0>0$ such that $q(r) \le 1$ for $0<r<r_0$ and $q(r_0) < 1$. \end{Fact} \begin{proof} Suppose the contrary, i.e., $q(r)$ first increases away from $r=0$. Then, because $q(0) = 1$ and $q(R) = 0$ and because $q$ is continuous and differentiable, we can find two points $r_1 < r_2$ such that $\hat q := q(r_1) = q(r_2) > 1$ and $q'(r_1) > 0 > q'(r_2)$. Even more, we can chose $r_1$ and $r_2$ such that $\hat q$ is arbitrarily close to one. Writing $\hat q = 1 + \epsilon$ with $\epsilon > 0$, we can calculate from the definition of $Q_y$ that $$Q_{1+\epsilon} = Q_1 + \epsilon n\left(\lambda_2 - \left(1-2/n\right)\lambda_1\right) + {\mathcal O}(\epsilon^2).$$ The term in brackets can be estimated by $$\lambda_2 - (1-2/n)\lambda_1 > \lambda_2 - \lambda_1 >0.$$ We can also assume that $Q_1 \ge 0$, because otherwise $q''(0) = \frac{2}{n^2}Q_1 < 0$ and Fact \ref{Fact2} is immediately true. Thus, choosing $r_1$ and $r_2$ such that $\epsilon$ is sufficiently small, we can make sure that $Q_{\hat q} > 0.$ We further note, that in view of (\ref{EqMMM}) the constant $M_{\hat q}$ can be positive or negative (depending on $n$), but not zero because $1<\hat q<2$. Now consider the function $T(r,\hat q)$. We have $T(r_1,\hat q) > 0 > T(r_2,\hat q)$ and the boundary behavior $T(0,\hat q) = -\infty$ and $T(R_1,\hat q) = +\infty$. Thus $T(r,\hat q)$ changes its sign at least thrice on $[0,R_1]$. Consequently, we can find three points $0 < \hat r_1 < \hat r_2 < \hat r_3 < R_1$ such that \begin{equation}\label{EqZBeh} Z_{\hat q}(\hat r_1) \ge 0, \quad Z_{\hat q}(\hat r_2) \le 0, \quad Z_{\hat q}(\hat r_3) \ge 0. \end{equation} Let us define $$h(r) = \frac{E^2 r^2}{2 \hat q} - 2 \hat q\tilde V(r).$$ Then \begin{equation} \label{EqZg} Z_{\hat q}(r) = \frac{M_{\hat q}}{r^2} + Q_{\hat q} + h(r). \end{equation} By condition b) on $\tilde V$, the function $h'(r)$ is concave. Also $h(0) = h'(0) = 0$. We conclude that if $h'(r_0) < 0$ or $h(r_0) < 0$ for some $r_0> 0$, then $h'(r)$ is negative and decreasing for all $r>r_0$. We will now show that $Z_{\hat q}$ cannot have the properties (\ref{EqZBeh}), a contradiction that proves Fact \ref{Fact2}: Case 1: Assume $M_{\hat q} > 0$. Then from $Z_{\hat q}(\hat r_2) \le 0$ we see that $$-h(\hat r_2) \ge \frac{M_{\hat q}}{\hat r_2^2} + Q_{\hat q} > 0.$$ By what has been said above about $h(r)$, we conclude that $-h(r)$ is a strictly increasing function on $[\hat r_2,\hat r_3]$. Therefore $$-h(\hat r_3) > -h(\hat r_2) \ge \frac{M_{\hat q}}{\hat r_2^2} + Q_{\hat q} > \frac{M_{\hat q}}{\hat r_3^2} + Q_{\hat q},$$ such that $Z_{\hat q}(\hat r_3) < 0$, contradicting (\ref{EqZBeh}). Case 2: Assume $M_{\hat q} < 0$. Then from $Z_{\hat q}(\hat r_1) \ge 0 \ge Z_{\hat q}(\hat r_2)$ follows that $Z_{\hat q}'(\hat r) \le 0$ for some $\hat r\in [\hat r_1,\hat r_2]$. In view of (\ref{EqZg}) we have $h'(\hat r) < 0$. But this means by our above concavity argument that $h'(r)$ is decreasing and thus $h'(r) < 0$ for all $r>\hat r$. Then $Z'_{\hat q}$ is strictly decreasing for $r\ge \hat r$. Together with $ Z_{\hat q}(\hat r_2) \le 0$ and $Z'_{\hat q}(\hat r) \le 0$ this implies that $ Z_{\hat q}(\hat r_3) < 0$, a contradiction to (\ref{EqZBeh}). \end{proof} \begin{Fact} \label{Fact3} For all $0\le r\le R_1$ the inequality $q'(r) \le 0$ holds. \end{Fact} \begin{proof} Assume the contrary. Then there are three points $r_1<r_2<r_3$ in $(0,R_1)$ with $0 < \hat q := q(r_1) = q(r_2) =q(r_3) < 1$ and $q'(r_1) < 0$, $q'(r_2)>0$, $q'(r_3) < 0$. Consider the function $T(r,\hat q)$, which is equal to $q'(r)$ at $r_1,r_2,r_3$. Taking into account its boundary behavior at $r = 0$ and $r=R_1$, it is clear that $T(r,\hat q)$ must have at least the sign changes positive-negative-positive-negative-positive. Thus $T(r,\hat q)$ has at least four zeros $\hat r_1 < \hat r_2 < \hat r_3 <\hat r_4$ with the properties $$Z_{\hat q}(\hat r_1) \le 0, \quad Z_{\hat q}(\hat r_2) \ge 0, \quad Z_{\hat q}(\hat r_3) \le 0, \quad Z_{\hat q}(\hat r_4) \ge 0.$$ We also know that $Z_{\hat q}(0) = +\infty$. To satisfy all these requirements, $Z_{\hat q}$ must either have at least three extremal points where $Z_{\hat q}'$ crosses zero or $Z_{\hat q}$ must vanish on a finite interval. But we have $$Z'_{\hat q}(r) = -\frac{2M_{\hat q}}{r^3} + \frac{E^2r}{\hat q} -2 \hat q\tilde V'(r),$$ which is a strictly concave function (recall $M_{\hat q} > 0$ for $0<\hat q<1$). A strictly concave function can only cross zero twice and not be zero on a finite interval, which is a contradiction that proves Fact \ref{Fact3}. \end{proof} Altogether we have shown that $0 < q(r) < 1$ and $q'(r) \le 0$ for all $r\in[0,R]$, proving Lemma \ref{LemmaMonotonicityOfgAndB}. \end{proof} \section{Proof of Theorem \ref{TheoremPPW}} \label{SectionProof} \begin{proof}[Proof of Theorem \ref{TheoremPPW}] We start from the basic gap inequality \begin{equation} \label{EqGap} \lambda_2(\Omega,V) - \lambda_1(\Omega,V) \le \frac{\int_\Omega |\nabla P|^2 u_1^2 \,\mathrm{d}^nr}{\int_\Omega P^2 u_1^2 \,\mathrm{d}^n r}, \end{equation} where $u_1$ is the first Dirichlet eigenfunction of $-\Delta + V$ on $\Omega$ and $P$ is a suitable test function that satisfies the condition $\int_\Omega Pu_1^2 \,\mathrm{d}^n r = 0$. We set \begin{equation} P_i(\vec r) = g(r) \frac{r_i}{r} \quad \textmd{for } i = 1,2,...,n, \end{equation} where \begin{equation} g(r) = \left\{\begin{array}{ll} \frac{z_2(r)}{z_1(r)} & \textmd{for } r < R_1\cr % \lim\limits_{t \uparrow R_1} g(t) & \textmd{for } r \ge R_1.\end{array} \right. \end{equation} Here $z_1$ and $z_2$ are the radial parts (both chosen positive) of the first two eigenfunctions of $-\Delta + \tilde V$ on $S_1$. More precisely, $z_2(r) r_i r^{-1}$ for $i=1,\dots,n$ is a basis of the space of second eigenfunctions. It follows from the convexity of $r\tilde V$ and the BGM inequality \cite{AB0, BGM} that the second eigenfunctions can be written in that way. According to an argument in \cite{AB2} one can always chose the origin of the coordinate system such that $\int_\Omega P_iu_1^2 \,\mathrm{d}^n r = 0$ is satisfied for all $i$. Putting the functions $P_i$ into (\ref{EqGap}) and summing over all $i$ yields \begin{equation} \label{EqDifLambda} \lambda_2(\Omega,V) - \lambda_1(\Omega,V) \le \frac{\int_\Omega B(r) u_1^2 \,\mathrm{d}^nr}{\int_\Omega g(r)^2 u_1^2 \,\mathrm{d}^nr} \end{equation} with \begin{equation*} B(r) = g'(r)^2 + (n-1)\frac{g(r)^2}{r^2}. \end{equation*} By Lemma \ref{LemmaMonotonicityOfgAndB} we know that $B$ is a decreasing and $g$ an increasing function of $r$. Thus, denoting by $u_1^\star$ the spherically decreasing rearrangement of $u_1$ with respect to the origin, we have \begin{eqnarray} \label{EqChain1} \int_\Omega B(r) u_1^2 \,\mathrm{d}^nr &\le& \int_{\Omega^\star} B^\star(r)\, {u_1^\star}^2 \,\mathrm{d}^nr\\ &\le& \int_{\Omega^\star} B(r) \,{u_1^\star}^2 \,\mathrm{d}^nr \le \int_{S_1} B(r)\, z_1^2 \,\mathrm{d}^nr \nonumber \end{eqnarray} and \begin{eqnarray} \label{EqChain2} \int_\Omega g(r)^2 u_1^2 \,\mathrm{d}^nr &\ge& \int_{\Omega^\star} g_\star(r)^2\, {u_1^\star}^2 \,\mathrm{d}^nr\\ &\ge& \int_{\Omega^\star} g(r)^2 \,{u_1^\star}^2 \,\mathrm{d}^nr \ge \int_{S_1} g(r)^2\, z_1^2 \,\mathrm{d}^nr \nonumber \end{eqnarray} In each of the above chains of inequalities the first step follows from general properties of rearrangements and the second from the monotonicity properties of $g$ and $B$. The third step is justified by a comparison result that we state below and the monotonicity of $g$ and $B$ again. Putting (\ref{EqChain1}) and (\ref{EqChain2}) into (\ref{EqDifLambda}) we get $$\lambda_2(\Omega,V) - \lambda_1(\Omega,V) \le \frac{\int_{S_1} B(r)\, z^2 \,\mathrm{d}^nr}{\int_{S_1} g(r)^2\, z^2 \,\mathrm{d}^nr} = \lambda_2(S_1,\tilde V) - \lambda_1(S_1,\tilde V).$$% Keeping in mind that $\lambda_1(\Omega,V) = \lambda_1(S_1,\tilde V)$, Theorem \ref{TheoremPPW} is proven by this last inequality. \end{proof} \begin{Lemma}[Chiti Comparison result] \label{LemmaChiti} Let $u_1^\star$ be the radially decreasing rearrangement of the first eigenfunction of $-\Delta + V$ on $\Omega$ and $z_1$ the first eigenfunction of $-\Delta + \tilde V$ on $S_1$. Assume both functions to be positive and normalized in $L^2(\Omega^\star)$. Then there exists an $r_0$ such that \begin{eqnarray*} u_1^\star(r) &\le& z_1(r) \quad \textmd{for } r \le r_0 \textmd{ and}\\ u_1^\star(r) &\ge& z_1(r) \quad \textmd{for } r_0 < r \le R_1. \end{eqnarray*} \end{Lemma} \begin{proof} By a version of the RFK inequality for Schr\"odinger operators \cite{L} and by domain monotonicity of the first eigenvalue it is clear that $S_1 \subset \Omega^\star$. This is why we can view $z_1(r)$ as a function in $L^2(\Omega^\star)$, setting $z_1(r) = 0$ for $r > R_1$. Both $u_1^\star$ and $z_1$ are positive and spherically symmetric. Moreover, $u_1^\star(r)$ and $z_1(r)$ are decreasing functions of $r$. For $u_1^\star$ this is clear by definition of the rearrangement. For $z_1$ it follows from a simple comparison argument using $z_1^\star$ as a test function in the Rayleigh quotient for $\lambda_1$. (Here and in the sequel we write short-hand $\lambda_1 = \lambda_1(\Omega,V) = \lambda_1(S_1,\tilde V)$.) We introduce a change of variables via $s = C_n r^n$ and write $u_1^\#(s) \equiv u_1^\star(r)$, $z_1^\#(s) \equiv z_1(r)$ and $\tilde V_\#(s) \equiv \tilde V(r)$. \begin{Fact} \label{FactDE} For the functions $u_1^\#(s)$ and $z_1^\#(s)$ we have \begin{eqnarray} -\frac{\,\mathrm{d} u_1^\#}{\,\mathrm{d} s} &\le& n^{-2}C_n^{-2/n}s^{n/2-2} \int_0^s (\lambda_1 - \tilde V_\#(w))\, u_1^\#(w) \,\mathrm{d} w, \label{EqFact1}\\ -\frac{\,\mathrm{d} z_1^\#}{\,\mathrm{d} s} &=& n^{-2}C_n^{-2/n}s^{n/2-2} \int_0^s (\lambda_1 - \tilde V_\#(w))\, z_1^\#(w) \,\mathrm{d} w. \label{EqFact2} \end{eqnarray} \end{Fact} \begin{proof} We integrate both sides of $-\Delta u_1 + Vu_1 = \lambda_1 u_1$ over the level set $\Omega_t := \{\vec r\in \Omega: u_1(\vec r) > t\}$ and use Gauss' Divergence Theorem to obtain \begin{equation} \label{EqHH} \int_{\partial\Omega_t} |\nabla u_1| H_{n-1}(\,\mathrm{d} r) = \int_{\Omega_t} (\lambda_1-V(\vec r))\, u_1(\vec r) \,\mathrm{d}^n r, \end{equation} where $\partial\Omega_t = \{\vec r\in \Omega: u_1(\vec r) = t\}$. Now we define the distribution function $\mu(t) = |\Omega_t|$. Using the coarea formula, the Cauchy-Schwarz inequality and the classical isoperimetric inequality, Talenti derives (\cite{T}, p.709, eq. (32)) \begin{equation} \label{EqHH2} \int_{\partial\Omega_t} |\nabla u_1| H_{n-1}(\,\mathrm{d} r) \ge - n^2C_n^{2/n} \frac{\mu(t)^{2-2/n}}{\mu'(t)}. \end{equation} The left sides of (\ref{EqHH}) and (\ref{EqHH2}) are the same, thus \begin{eqnarray*} - n^2C_n^{2/n} \frac{\mu(t)^{2-2/n}}{\mu'(t)} &\le& \int_{\Omega_t} (\lambda_1-V(\vec r))\,u_1(\vec r) \,\mathrm{d}^n r\\ &\le& \int_{\Omega_t^\star} (\lambda_1 - V_\star(\vec r))\, u_1^\star(\vec r) \,\mathrm{d}^n r\\ &\le& \int_{\Omega_t^\star} (\lambda_1 - \tilde V(\vec r))\, u_1^\star(\vec r) \,\mathrm{d}^n r\\ &=& \int_0^{(\mu(t)/C_n)^{1/n}} n C_n r^{n-1} (\lambda_1 - \tilde V(r)) u_1^\star(r) \,\mathrm{d} r.\\ \end{eqnarray*} Now we perform the change of variables $r \rightarrow s$ on the right hand side of the above chain of inequalities. We also chose $t$ to be $u_1^\#(s)$. Using the fact that $u_1^\#$ and $\mu$ are essentially inverse functions to one another, this means that $\mu(t) = s$ and $\mu'(t)^{-1} = (u_1^\#)'(s)$. The result is (\ref{EqFact1}). Equation (\ref{EqFact2}) is proven analogously. \end{proof} Fact \ref{FactDE} enables us to prove Lemma \ref{LemmaChiti}. We have $u_1^\#(|S_1|) > z_1^\#(|S_1|) = 0$. Being equally normalized, $u_1^\star$ and $z_1$ must have at least one intersection on $[0,R]$. Thus $u_1^\#$ and $z_1^\#$ have at least one intersection on $[0,|S_1|]$. Now assume that they intersect at least twice. Then there is an interval $[s_1,s_2] \subset [0,|S_1|]$ such that $u_1^\#(s) > z^\#(s)$ for $s \in (s_1,s_2)$, $u_1^\#(s_2) = z_1^\#(s_2)$ and either $u_1^\#(s_1) = z_1^\#(s_1)$ or $s_1=0$. There is also an interval $[s_3,s_4] \subset [s_2,|S_1|]$ with $u_1^\#(s) < z_1^\#(s)$ for $s \in (s_3,s_4)$, $u_1^\#(s_3) = z_1^\#(s_3)$ and $u_1^\#(s_4) = z_1^\#(s_4)$. Be further $\tilde s$ the point where $\tilde V_\#(s) - \lambda_1(S_1,\tilde V)$ crosses zero (Set $\tilde s = |S_1|$ if $\tilde V_\#(s) - \lambda_1$ doesn't cross zero on $[0,|S_1|]$). To keep our notation compact we will write $$I_a^b[u] = \int_a^b (\lambda_1-\tilde V_\#(w))\,u(w)\,\mathrm{d} w.$$ \emph{Case 1:} Assume $\tilde s \ge s_2$. Then $\tilde V_\#(s) - \lambda_1(S_1,\tilde V)$ is negative for $s<s_2$. Set \begin{equation*} v(s) = \left\{ \begin{array}{ll} u_1^\#(s) & \textmd{on } [0,s_1] \textmd{ if } I_0^{s_1}[u_1^\#] > I_0^{s_1}[z_1^\#],\cr% z_1^\#(s) & \textmd{on } [0,s_1] \textmd{ if } I_0^{s_1}[u_1^\#] \le I_0^{s_1}[z_1^\#],\cr % u_1^\#(s) & \textmd{on } [s_1,s_2]\cr % z_1^\#(s) & \textmd{on } [s_2,|S_1|]\cr % \end{array}\right. \end{equation*} Using Fact \ref{FactDE}, one can check that then $v(s)$ fulfills the inequality \begin{equation} \label{Eqv} -\frac{\,\mathrm{d} v}{\,\mathrm{d} s} \le n^{-2}C_n^{-2/n}s^{n/2-2} \int_0^s (\lambda_1 - \tilde V_\#(s)) v(w) \,\mathrm{d} w. \end{equation} \emph{Case 2:} Assume $\tilde s < s_2$. Then $\tilde V_\#(s) - \lambda_1(S_1,\tilde V)$ is positive for $s\ge s_3$. Set \begin{equation*} v(s) = \left\{ \begin{array}{ll} u_1^\#(s) & \textmd{on } [0,s_3] \textmd{ if } I_0^{s_3}[u_1^\#] > I_0^{s_3}[z_1^\#],\cr% z_1^\#(s) & \textmd{on } [0,s_3] \textmd{ if } I_0^{s_3}[u_1^\#] \le I_0^{s_3}[z_1^\#],\cr % u_1^\#(s) & \textmd{on } [s_3,s_4]\cr % z_1^\#(s) & \textmd{on } [s_4,|S_1|]\cr % \end{array}\right. \end{equation*} Again using Fact \ref{FactDE}, one can check that also in this case $v(s)$ fulfills the inequality (\ref{Eqv}). Now define the test function $$\Psi(\vec r) = v(C_nr^n) = v(s).$$ Then we use the Rayleigh characterization of $\lambda_1$, equation (\ref{Eqv}) and integration by parts to calculate \begin{eqnarray*} \lambda_1 {\int_{S_1} \Psi(\vec r)^2 \,\mathrm{d}^n r} &<& \int_{S_1} \left(|\nabla \Psi|^2 + \tilde V(\vec r) \Psi^2\right) \,\mathrm{d}^n r\\ &=& \int_0^{|S_1|} \left( v'(s)^2 n^2 s^{2-2/n} C_n^{2/n} + \tilde V_\#(s) v^2(s)\right) \,\mathrm{d} s\\ &\le& \int_0^{|S_1|} \left(-v'(s) \int_0^s (\lambda_1-\tilde V_\#(w)) v(w) \,\mathrm{d} w + \tilde V_\#(s) v^2(s)\right) \,\mathrm{d} s\\ &=& \int_0^{|S_1|} \left(v(s) (\lambda_1 - \tilde V_\#(s)) v(s) + \tilde V_\#(s) v^2(s) \right) \,\mathrm{d} s\\ &=& \lambda_1 {\int_{S_1} \Psi(\vec r)^2 \,\mathrm{d}^n r}. \end{eqnarray*} This is a contradiction to our original assumption that $u_1^\#(r)$ and $z_1^\#(r)$ have more than one intersection, thus proving Lemma \ref{LemmaChiti}. \end{proof} \section{Proof of Theorem \ref{TheoremMonotonicity}} \label{SectionProof2} \begin{proof}[Proof of Theorem \ref{TheoremMonotonicity}] The first eigenfunction of $-\Delta + V$ on $B_R$ is radially symmetric and will be called $z_1(r)$. Further, a standard separation of variables and the Baumgartner-Grosse-Martin \cite{BGM, AB3} inequality imply that we can write a basis of the space of second eigenfunctions in the form $z_2(r) \cdot r_i \cdot r^{-1}$. The radial parts $z_1$ and $z_2$ of the first and the second eigenfunction, which we assume to be positive, solve the differential equations \begin{eqnarray} -z''_1(r) - \frac{n-1}{r} z'_1(r) + \left(V(r) -\lambda_1\right) z_1(r) &=& 0, \label{EqDEs}\\ -z''_2(r) - \frac{n-1}{r} z'_2(r) + \left(\frac{n-1}{r^2} + V(r)- \lambda_2\right) z_2(r) &=& 0 \nonumber \end{eqnarray} with the boundary conditions \begin{equation}\label{EquM3} z'_1(0) = 0, \quad z_1(R) = 0, \quad z_2(0) = 0, \quad z_2(R) = 0. \end{equation} We define the rescaled functions $\tilde z_{1/2}(r) = z_{1/2}(\beta r)$. Putting $\beta r$ (with $\beta>0$) instead of $r$ into the equations (\ref{EqDEs}) and multiplying by $\beta^2$ yields the rescaled equations \begin{eqnarray*} -\tilde z''_1(r) - \frac{n-1}{r} \tilde z'_1(r) + \left(\beta^2 V(\beta r) -\beta^2 \lambda_1\right) \tilde z_1(r) &=& 0, \nonumber\\ -\tilde z''_2(r) - \frac{n-1}{r} \tilde z'_2(r) + \left(\frac{n-1}{r^2} + \beta^2 V(\beta r)- \beta^2 \lambda_2\right) \tilde z_2(r) &=& 0. \nonumber \end{eqnarray*} We conclude that $\tilde z_1$ and $\tilde z_2$ are the radial parts of the first two eigenfunctions of $-\Delta + \beta^2 V(\beta r)$ on $B_{R/\beta}$ to the eigenvalues $\beta^2 \lambda_1$ and $\beta^2\lambda_2$. Consequently, if we replace $R$ by $R/\beta$ and $V(r)$ by $\beta^2 V(\beta r)$, then the ratio $\lambda_2/\lambda_1$ doesn't change. For the rest of this section we shall write $\lambda_{1/2}(R,V)$ instead of $\lambda_{1/2}(B_R,V)$. We also fix two radii $0 < R_1 < R_2$ and let $\rho(\beta)$ for $\beta > 1$ be the function defined implicitly by \begin{equation} \label{EqLam} \lambda_1(\rho(\beta), V(r)) = \lambda_1(R_2/\beta, \beta^2 V(\beta r)). \end{equation} Then we have $\rho(1) = R_2$. By domain monotonicity of $\lambda_1$ and because $V(r)$ is increasing and positive we see that the right hand side of (\ref{EqLam}) is increasing in $\beta$. Therefore, again by domain monotonicity, $\rho(\beta)$ must be decreasing in $\beta$. One can also check that $\rho(\beta)$ is a continuous function and that $\rho(\beta)$ goes to zero for $\beta \rightarrow \infty$. Thus we can find $\beta_0 > 1$ such that $\rho(\beta_0) = R_1$. Then we can apply Theorem \ref{TheoremPPW}, with $B_{R_2/\beta_0}$ for $\Omega$ and $B_{\rho(\beta_0)}$ for $S_1$, as well as $\beta_0^2 V(\beta_0 r)$ for $V$ and $V(r)$ for $\tilde V$, to get \begin{equation} \label{EqLam2} \lambda_2(R_2/\beta_0, \beta_0^2 V(\beta_0 r) \le \lambda_2(\rho(\beta_0), V(r)) = \lambda_2(R_1, V(r)) . \end{equation} But by what has been said above about the scaling properties of the problem, we have \begin{equation} \label{EqLam3} \frac{\lambda_2(R_2/\beta_0, \beta_0^2 V(\beta_0 r))}{\lambda_1(R_2/\beta_0, \beta_0^2 V(\beta_0 r))} = \frac{\lambda_2(R_2, V(r))}{\lambda_1(R_2, V(r))}. \end{equation} Combining (\ref{EqLam}) for $\beta=\beta_0$, (\ref{EqLam2}) and(\ref{EqLam3}), we get \begin{equation} \label{EqLam4} \frac{\lambda_2(R_1,V(r))}{\lambda_1(R_1, V(r))} \ge \frac{\lambda_2(R_2, V(r))}{\lambda_1(R_2, V(r))}. \end{equation} Because $R_1$ and $R_2$ were chosen arbitrarily, this proves Theorem \ref{TheoremMonotonicity}. \end{proof} \section{Proof of Theorem \ref{TheoremMonotonicity2}} \label{SectionProof3} Before we prove Theorem \ref{TheoremMonotonicity2} we need to state the following technical Lemma: \begin{Lemma}\label{LemmaAlgebraic} Be $a,b,c,d > 0$ with $a\ge b$, $d\ge b$ and $\frac a b < \frac cd$. Then \begin{equation} \frac{a+x}{b+x} < \frac{c+x}{d+x} \end{equation} holds for any $x>0$. \end{Lemma} \begin{proof} Define the function $$f(x) := \frac{c+x}{d+x} - \frac{a+x}{b+x}.$$ then $f(0) > 0$. A straightforward calculation shows that $f$ has exactly one zero at $$x_0 = -\frac{bc-ad}{b+c-a-d}.$$ The numerator $bc-ad$ in the expression for $x_0$ is positive because of the condition $\frac a b < \frac cd$. For the denominator we get $$b+c-a-d > c+b-\frac{bc}{d}-d = \frac{(d-b)(c-d)}{d} \ge 0.$$ This means that $x_0<0$, such that $f(x)>0$ for all $x>0$. \end{proof} \begin{proof}[Proof of Theorem \ref{TheoremMonotonicity2}] Choose some $x > 0$. From Theorem \ref{TheoremMonotonicity} we know that $$\frac{\lambda_2(B_{R+x},r^2)}{\lambda_1(B_{R+x},r^2)} < \frac{\lambda_2(B_{R},r^2)}{\lambda_1(B_{R},r^2)} \quad \textmd{for }x>0.$$ % Moreover, $\lambda_1(B_R,r^2) \ge \lambda_1(B_{R+x},r^2)$ and $\lambda_2(B_{R+x},r^2) > \lambda_1(B_{R+x},r^2)$. Thus we can apply first (\ref{EqCon}), then Lemma \ref{LemmaAlgebraic} and then (\ref{EqCon}) again, to get \begin{equation*} \frac{\lambda^+_2(B_{R+x})}{\lambda^+_1(B_{R+x})} = \frac{\lambda_2(B_{R+x},r^2) + n}{\lambda_1(B_{R+x},r^2) + n} < \frac{\lambda_2(B_{R},r^2) + n}{\lambda_1(B_{R},r^2) + n} = \frac{\lambda^+_2(B_{R})}{\lambda^+_1(B_{R})}.% \end{equation*} \end{proof}
{ "timestamp": "2005-11-09T23:22:23", "yymm": "0511", "arxiv_id": "math-ph/0511032", "language": "en", "url": "https://arxiv.org/abs/math-ph/0511032", "abstract": "Let $\\lambda_i(\\Omega,V)$ be the $i$th eigenvalue of the Schrödinger operator with Dirichlet boundary conditions on a bounded domain $\\Omega \\subset \\R^n$ and with the positive potential $V$. Following the spirit of the Payne-Pólya-Weinberger conjecture and under some convexity assumptions on the spherically rearranged potential $V_\\star$, we prove that $\\lambda_2(\\Omega,V) \\le \\lambda_2(S_1,V_\\star)$. Here $S_1$ denotes the ball, centered at the origin, that satisfies the condition $\\lambda_1(\\Omega,V) = \\lambda_1(S_1,V_\\star)$.Further we prove under the same convexity assumptions on a spherically symmetric potential $V$, that $\\lambda_2(B_R, V) / \\lambda_1(B_R, V)$ decreases when the radius $R$ of the ball $B_R$ increases.We conclude with several results about the first two eigenvalues of the Laplace operator with respect to a measure of Gaussian or inverted Gaussian density.", "subjects": "Mathematical Physics (math-ph)", "title": "A second eigenvalue bound for the Dirichlet Schroedinger operator", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.986777179025415, "lm_q2_score": 0.8128673110375458, "lm_q1q2_score": 0.802118912107604 }
https://arxiv.org/abs/1908.09375
Theoretical Issues in Deep Networks: Approximation, Optimization and Generalization
While deep learning is successful in a number of applications, it is not yet well understood theoretically. A satisfactory theoretical characterization of deep learning however, is beginning to emerge. It covers the following questions: 1) representation power of deep networks 2) optimization of the empirical risk 3) generalization properties of gradient descent techniques --- why the expected error does not suffer, despite the absence of explicit regularization, when the networks are overparametrized? In this review we discuss recent advances in the three areas. In approximation theory both shallow and deep networks have been shown to approximate any continuous functions on a bounded domain at the expense of an exponential number of parameters (exponential in the dimensionality of the function). However, for a subset of compositional functions, deep networks of the convolutional type can have a linear dependence on dimensionality, unlike shallow networks. In optimization we discuss the loss landscape for the exponential loss function and show that stochastic gradient descent will find with high probability the global minima. To address the question of generalization for classification tasks, we use classical uniform convergence results to justify minimizing a surrogate exponential-type loss function under a unit norm constraint on the weight matrix at each layer -- since the interesting variables for classification are the weight directions rather than the weights. Our approach, which is supported by several independent new results, offers a solution to the puzzle about generalization performance of deep overparametrized ReLU networks, uncovering the origin of the underlying hidden complexity control.
\section{Introduction} \dropcap{I}n the last few years, deep learning has been tremendously successful in many important applications of machine learning. However, our theoretical understanding of deep learning, and thus the ability of developing principled improvements, has lagged behind. A satisfactory theoretical characterization of deep learning is emerging. It covers the following areas: 1) {\it approximation} properties of deep networks 2) {\it optimization} of the empirical risk 3) {\it generalization} properties of gradient descent techniques -- why the expected error does not suffer, despite the absence of explicit regularization, when the networks are overparametrized? \subsection{When Can Deep Networks Avoid the Curse of Dimensionality?} We start with the first set of questions, summarizing results in \cite{HierarchicalKernels2015,Hierarchical2015,poggio2015December}, and \cite{Mhaskaretal2016,MhaskarPoggio2016}. The main result is that deep networks have the theoretical guarantee, which shallow networks do not have, that they can avoid the {\it curse of dimensionality} for an important class of problems, corresponding to {\it compositional functions}, that is functions of functions. An especially interesting subset of such compositional functions are {\it hierarchically local compositional functions} where all the constituent functions are local in the sense of bounded small dimensionality. The deep networks that can approximate them without the curse of dimensionality are of the deep convolutional type -- though, importantly, weight sharing is not necessary. Implications of the theorems likely to be relevant in practice are: a) {\it Deep convolutional architectures} have the theoretical guarantee that they can be {\it much better} than one layer architectures such as kernel machines for certain classes of problems; b) the problems for which certain deep networks are guaranteed to avoid the {\it curse of dimensionality} (see for a nice review \cite{Donoho00high-dimensionaldata}) correspond to input-output mappings that are {\it compositional with local constituent functions}; c) the key aspect of convolutional networks that can give them an exponential advantage is {\it not weight sharing} but {\it locality} at each level of the hierarchy. \subsection{Related Work} Several papers in the '80s focused on the approximation power and learning properties of one-hidden layer networks (called shallow networks here). Very little appeared on multilayer networks, (but see \cite{mhaskar1993approx, mhaskar1993neural, chui1994neural, chui1996, Pinkus1999}). By now, several papers \cite{poggio03mathematics,MontufarBengio2014, DBLP:journals/corr/abs-1304-7045} have appeared. \cite{Anselmi2014,anselmi2015theoretical,poggioetal2015, LiaoPoggio2016, Mhaskaretal2016} derive new upper bounds for the approximation by deep networks of certain important classes of functions which avoid the curse of dimensionality. The upper bound for the approximation by shallow networks of general functions was well known to be exponential. It seems natural to assume that, since there is no general way for shallow networks to exploit a compositional prior, lower bounds for the approximation by shallow networks of compositional functions should also be exponential. In fact, examples of specific functions that cannot be represented efficiently by shallow networks have been given, for instance in \cite{Telgarsky2015, SafranShamir2016, Theory_I}. An interesting review of approximation of univariate functions by deep networks has recently appeared \cite{2019arXiv190502199D}. \begin{figure \centering \includegraphics[trim=10 33 30 73, width=0.9\linewidth,clip]{Figures/Example_2_functions.pdf} \caption{The top graphs are associated to {\it functions}; each of the bottom diagrams depicts the ideal {\it network} approximating the function above. In a) a shallow universal network in 8 variables and $N$ units approximates a generic function of $8$ variables $f(x_1, \cdots, x_8)$. Inset b) shows a hierarchical network at the bottom in $n=8$ variables, which approximates well functions of the form $f(x_1, \cdots, x_8) = h_3(h_{21}(h_{11} (x_1, x_2), h_{12}(x_3, x_4)), \allowbreak h_{22}(h_{13}(x_5, x_6), h_{14}(x_7, x_8))) $ as represented by the binary graph above. In the approximating network each of the $n-1$ nodes in the graph of the function corresponds to a set of $Q =\frac{N}{n-1}$ ReLU units computing the ridge function $\sum_{i=1}^Q a_i(\scal{\mathbf{v}_i}{\mathbf{x}}+t_i)_+$, with $\mathbf{v}_i, \mathbf{x} \in \R^2$, $a_i, t_i\in\R$. Each term in the ridge function corresponds to a unit in the node (this is somewhat different from todays deep networks, but equivalent to them \cite{Theory_I}). Similar to the shallow network, a hierarchical network is universal, that is, it can approximate any continuous function; the text proves that it can approximate a compositional functions exponentially better than a shallow network. Redrawn from \cite{MhaskarPoggio2016}. } \label{example_functions} \end{figure} \subsection{Degree of approximation}\label{approxsect} The general paradigm is as follows. We are interested in determining how complex a network ought to be to {\it theoretically guarantee} approximation of an unknown target function $f$ up to a given accuracy $\epsilon>0$. To measure the accuracy, we need a norm $\|\cdot\|$ on some normed linear space $\mathbb{X}$. As we will see the norm used in the results of this paper is the $sup$ norm in keeping with the standard choice in approximation theory. As it turns out, the results of this section require the sup norm in order to be independent from the unknown distribution of the input data. Let $V_N$ be the be set of all networks of a given kind with $N$ units (which we take to be or measure of the complexity of the approximant network). The \textit{degree of approximation} is defined by $\mathsf{dist}(f, V_N)=\inf_{P\in V_N}\|f-P\|.$ For example, if $\mathsf{dist}(f, V_N)=\O(N^{-\gamma})$ for some $\gamma>0$, then a network with complexity $N=\O(\epsilon^{-\frac{1}{\gamma}})$ will be sufficient to guarantee an approximation with accuracy at least $\epsilon$. The only a priori information on the class of target functions $f$, is codified by the statement that $f\in W$ for some subspace $W\subseteq \mathbb{X}$. This subspace is a smoothness and compositional class, characterized by the parameters $m$ and $d$ ($d=2$ in the example of Figure \ref{example_functions} ; it is the size of the kernel in a convolutional network). \subsection{Shallow and deep networks} \label{subprevious} This section characterizes conditions under which deep networks are ``better'' than shallow network in approximating functions. Thus we compare shallow (one-hidden layer) networks with deep networks as shown in Figure \ref{example_functions}. Both types of networks use the same small set of operations -- dot products, linear combinations, a fixed nonlinear function of one variable, possibly convolution and pooling. Each node in the networks corresponds to a node in the graph of the function to be approximated, as shown in the Figure. A unit is a neuron which computes $(\scal{x}{w}+b)_+$, where $w$ is the vector of weights on the vector input $x$. Both $w$ and the real number $b$ are parameters tuned by learning. We assume here that each node in the networks computes the linear combination of $r$ such units $\sum_{i=1}^r c_i (\scal{x}{w_i}+b_i)_+$. Notice that in our main example of a network corresponding to a function with a binary tree graph, the resulting architecture is an idealized version of deep convolutional neural networks described in the literature. In particular, it has only one output at the top unlike most of the deep architectures with many channels and many top-level outputs. Correspondingly, each node computes a single value instead of multiple channels, using the combination of several units. However our results hold also for these more complex networks (see \cite{Theory_I}). \noindent The sequence of results is as follows. \begin{itemize} \item {\it Both shallow (a) and deep (b) networks are universal}, that is they can approximate arbitrarily well any continuous function of $n$ variables on a compact domain. The result for shallow networks is classical. \item We consider a special class of functions of $n$ variables on a compact domain that are {\it hierarchical compositions of local functions}, such as $f(x_1, \cdots, x_8) = h_3(h_{21}(h_{11} (x_1, x_2), h_{12}(x_3, x_4)), \allowbreak h_{22}(h_{13}(x_5, x_6), h_{14}(x_7, x_8))) $ \noindent The structure of the function in Figure \ref{example_functions} b) is represented by a graph of the binary tree type, reflecting dimensionality $d=2$ for the constituent functions $h$. In general, $d$ is arbitrary but fixed and independent of the dimensionality $n$ of the compositional function $f$. \cite{Theory_I} formalizes the more general compositional case using directed acyclic graphs. \item The approximation of functions with a {\it compositional structure} -- can be achieved with the same degree of accuracy by deep and shallow networks but the number of parameters are much smaller for the deep networks than for the shallow network with equivalent approximation accuracy. \end{itemize} We approximate functions with networks in which the activation nonlinearity is a smoothed version of the so called ReLU, originally called {\it ramp} by Breiman and given by $\sigma (x) = x_+ = max(0, x)$ . The architecture of the deep networks reflects the function graph with each node $h_i$ being a ridge function, comprising one or more neurons. Let $I^n=[-1,1]^n$, $\mathbb{X}=C(I^n)$ be the space of all continuous functions on $I^n$, with $\|f\|=\max_{x\in I^n}|f(x)|$. Let $\mathcal{S}_{N,n}$ denote the class of all shallow networks with $N$ units of the form $$ x\mapsto\sum_{k=1}^N a_k\sigma(\scal{{w}_k}{x}+b_k), $$ where ${w}_k\in\R^n$, $b_k, a_k\in\R$. The number of trainable parameters here is $(n+2)N\sim n$. Let $m\ge 1$ be an integer, and $W_m^n$ be the set of all functions of $n$ variables with continuous partial derivatives of orders up to $m < \infty$ such that $\|f\|+\sum_{1\le |\k|_1\le m} \|D^\k f\| \le 1$, where $D^\k$ denotes the partial derivative indicated by the multi-integer $\k\ge 1$, and $|\k|_1$ is the sum of the components of $\k$. For the hierarchical binary tree network, the analogous spaces are defined by considering the compact set $W_m^{n,2}$ to be the class of all compositional functions $f$ of $n$ variables with a binary tree architecture and constituent functions $h$ in $W_m ^2$. We define the corresponding class of deep networks $\mathcal{D}_{N,2}$ to be the set of all deep networks with a binary tree architecture, where each of the constituent nodes is in $\mathcal{S}_{M,2}$, where $N=|V|M$, $V$ being the set of non--leaf vertices of the tree. We note that in the case when $n$ is an integer power of $2$, the total number of parameters involved in a deep network in $\mathcal{D}_{N,2}$ is $4N$. The first theorem is about shallow networks. \begin{theorem} \label{optneurtheo} Let $\sigma :\R\to \R$ be infinitely differentiable, and not a polynomial. For $f\in W_m^n$ the complexity of shallow networks that provide accuracy at least $\epsilon$ is \begin{equation} N= \O(\epsilon^{-n/m})\,\, and\,\, is\,\, the\,\, best\,\, possible. \end{equation} \end{theorem} The estimate of Theorem \ref{optneurtheo} is the best possible if the only a priori information we are allowed to assume is that the target function belongs to $f\in W_m^n$. The exponential dependence on the dimension $n$ of the number $e^{-n/m}$ of parameters needed to obtain an accuracy $\O(\epsilon)$ is known as the {\it curse of dimensionality}. Note that the constants involved in $\O$ in the theorems will depend upon the norms of the derivatives of $f$ as well as $\sigma$. Our second and main theorem is about deep networks with smooth activations (preliminary versions appeared in \cite{poggio2015December,Hierarchical2015,Mhaskaretal2016}). We formulate it in the binary tree case for simplicity but it extends immediately to functions that are compositions of constituent functions of a fixed number of variables $d$ (in convolutional networks $d$ corresponds to the size of the kernel). \begin{theorem} \label{deeptheo} For $f\in W_m^{n,2}$ consider a deep network with the same compositional architecture and with an activation function $\sigma :\R\to \R$ which is infinitely differentiable, and not a polynomial. The complexity of the network to provide approximation with accuracy at least $\epsilon$ is \begin{equation} N =\mathcal{O}((n-1)\epsilon^{-2/m}). \label{deepnetapprox} \end{equation} \end{theorem} The proof is in \cite{Theory_I}. The assumptions on $\sigma$ in the theorems are not satisfied by the ReLU function $x\mapsto x_+$, but they are satisfied by smoothing the function in an arbitrarily small interval around the origin. The result of the theorem can be extended to non-smooth ReLU\cite{Theory_I}. In summary, when the only a priori assumption on the target function is about the number of derivatives, then to {\it guarantee} an accuracy of $\epsilon$, we need a shallow network with $\O(\epsilon^{-n/m})$ trainable parameters. If we assume a hierarchical structure on the target function as in Theorem~\ref{deeptheo}, then the corresponding deep network yields a guaranteed accuracy of $\epsilon$ with $\O(\epsilon^{-2/m})$ trainable parameters. Note that Theorem~\ref{deeptheo} applies to all $f$ with a compositional architecture given by a graph which correspond to, or is a subgraph of, the graph associated with the deep network -- in this case the graph corresponding to $W_m^{n,d}$. \section{ The Optimization Landscape of Deep Nets with Smooth Activation Function} \label{BezoutBoltzman} The main question in optimization of deep networks is to the landscape of the empirical loss in terms of its global minima and local critical points of the gradient. \subsection{Related work} There are many recent papers studying optimization in deep learning. For optimization we mention work based on the idea that noisy gradient descent \cite{DBLP:journals/corr/Jin0NKJ17, DBLP:journals/corr/GeHJY15, pmlr-v49-lee16, s.2018when} can find a global minimum. More recently, several authors studied the dynamics of gradient descent for deep networks with assumptions about the input distribution or on how the labels are generated. They obtain global convergence for some shallow neural networks \cite{Tian:2017:AFP:3305890.3306033, s8409482, Li:2017:CAT:3294771.3294828, DBLP:conf/icml/BrutzkusG17, pmlr-v80-du18b, DBLP:journals/corr/abs-1811-03804}. Some local convergence results have also been proved \cite{Zhong:2017:RGO:3305890.3306109, DBLP:journals/corr/abs-1711-03440, 2018arXiv180607808Z}. The most interesting such approach is \cite{DBLP:journals/corr/abs-1811-03804}, which focuses on minimizing the training loss and proving that randomly initialized gradient descent can achieve zero training loss (see also \cite{NIPS2018_8038, du2018gradient, DBLP:journals/corr/abs-1811-08888}). In summary, there is by now an extensive literature on optimization that formalizes and refines to different special cases and to the discrete domain our results of \cite{theory_II, theory_IIb}. \subsection{Degeneracy of global and local minima under the exponential loss} The {\it first part} of the argument of this section relies on the obvious fact (see \cite{theory_III}), that for RELU networks under the hypothesis of an exponential-type loss function, there are {\it no local minima that separate the data} -- the only critical points of the gradient that separate the data are the global minima. Notice that the global minima are at $\rho = \infty$, when the exponential is zero. As a consequence, the Hessian is identically zero with all eigenvalues being zero. On the other hand any point of the loss at a finite $\rho$ has nonzero Hessian: for instance in the linear case the Hessian is proportional to $\sum_n^N x_n x^T_n$. The local minima which are not global minima must misclassify. How degenerate are they? Simple arguments \cite{theory_III} suggest that the critical points which are not global minima cannot be completely degenerate. We thus have the following \begin{property} Under the exponential loss, global minima are completely degenerate with all eigenvalues of the Hessian ($W$ of them with $W$ being the number of parameters in the network) being zero. The other critical points of the gradient are less degenerate, with at least one -- and typically $N$ -- nonzero eigenvalues. \end{property} For the general case of non-exponential loss and smooth nonlinearities instead of the RELU the following conjecture has been proposed \cite{theory_III}: \begin{conjecture}: For appropriate overparametrization, there are a large number of global zero-error minimizers which are degenerate; the other critical points -- saddles and local minima -- are generically (that is with probability one) degenerate on a set of much lower dimensionality. \end{conjecture} \subsection{SGD and Boltzmann Equation} The second part of our argument (in \cite{theory_IIb}) is that SGD concentrates in probability on the most degenerate minima. The argument is based on the similarity between a Langevin equation and SGD and on the fact that the Boltzmann distribution is exactly the asymptotic ``solution'' of the stochastic differential Langevin equation and also of SGDL, defined as SGD with added white noise (see for instance \cite{raginskyetal17}). The Boltzmann distribution is \begin{equation} p(f) = \frac{1}{Z}e^{-\frac{L}{T}}, \label{Bolzman} \end{equation} \noindent where $Z$ is a normalization constant, $L(f)$ is the loss and $T$ reflects the noise power. The equation implies that SGDL prefers degenerate minima relative to non-degenerate ones of the same depth. In addition, among two minimum basins of equal depth, the one with a larger volume is much more likely in high dimensions as shown by the simulations in \cite{theory_IIb}. Taken together, these two facts suggest that SGD selects degenerate minimizers corresponding to larger isotropic flat regions of the loss. Then SDGL shows concentration -- {\it because of the high dimensionality} -- of its asymptotic distribution Equation \ref{Bolzman}. Together \cite{theory_II} and \cite{theory_III} suggest the following \begin{conjecture}: For appropriate overparametrization of the deep network, SGD selects with high probability the global minimizers of the empirical loss, which are highly degenerate. \end{conjecture} \begin{figure} \centering \includegraphics[width=1.0\linewidth ]{Figures/SGD_vs_SGDL.pdf} \caption{ Stochastic Gradient Descent and Langevin Stochastic Gradient Descent (SGDL) on the $2$D potential function shown above leads to an asymptotic distribution with the histograms shown on the left. As expected from the form of the Boltzmann distribution, both dynamics prefer degenerate minima to non-degenerate minima of the same depth. From \cite{theory_III}. } \label{wedge_rbf_sgdl} \end{figure} \section{Generalization} \label{generalization} Recent results by \cite{2017arXiv171010345S} illuminate the apparent absence of ''overfitting” (see Figure \ref{no-overfitting}) in the special case of linear networks for binary classification. They prove that minimization of loss functions such as the logistic, the cross-entropy and the exponential loss yields asymptotic convergence to the maximum margin solution for linearly separable datasets, independently of the initial conditions and without explicit regularization. Here we discuss the case of nonlinear multilayer DNNs under exponential-type losses, for several variations of the basic gradient descent algorithm. The main results are: \begin{itemize} \item classical uniform convergence bounds for generalization suggest a form of complexity control on the dynamics of the weight {\it directions $V_k$}: minimize a surrogate loss subject to a unit $L_p$ norm constraint; \item gradient descent on the exponential loss with an explicit $L_2$ unit norm constraint is equivalent to a well-known gradient descent algorithms {\it weight normalization} which is closely related to batch normalization; \item unconstrained gradient descent on the exponential loss yields a dynamics with the same critical points as weight normalization: the dynamics implicitly respect a $L_2$ unit constraint on the directions of the weights $V_k$. \end{itemize} We observe that several of these results {\it directly apply to kernel machines} for the exponential loss under the separability/interpolation assumption, because kernel machines are one-homogeneous. \subsection{Related work} A number of papers have studied gradient descent for deep networks \cite{NIPS2017_6836, DBLP:journals/corr/abs-1811-04918, Arora2019FineGrainedAO}. Close to the approach summarized here (details are in \cite{theory_III}) is the paper \cite{Wei2018OnTM}. Its authors study generalization assuming a regularizer because they are -- like us -- interested in normalized margin. Unlike their assumption of an explicit regularization, we show here that commonly used techniques, such as weight and batch normalization, in fact minimize the surrogate loss margin while controlling the complexity of the classifier without the need to add a regularizer or to use weight decay. Surprisingly, we will show that even standard gradient descent on the weights implicitly controls the complexity through an ``implicit'' unit $L_2$ norm constraint. Two very recent papers (\cite{2019arXiv190507325S} and \cite{DBLP:journals/corr/abs-1906-05890}) develop an elegant but complicated margin maximization based approach which lead to some of the same results of this section (and many more). The important question of which conditions are necessary for gradient descent to converge to the maximum of the margin of $\tilde{f}$ are studied by \cite{2019arXiv190507325S} and \cite{DBLP:journals/corr/abs-1906-05890}. Our approach does not need the notion of maximum margin but our theorem \ref{margin-maxTheorem} establishes a connection with it and thus with the results of \cite{2019arXiv190507325S} and \cite{DBLP:journals/corr/abs-1906-05890}. Our main goal here (and in \cite{theory_III}) is to achieve a simple understanding of where the complexity control underlying generalization is hiding in the training of deep networks. \subsection{Deep networks: definitions and properties} We define a deep network with $K$ layers with the usual coordinate-wise scalar activation functions $\sigma(z):\quad \mathbf{R} \to \mathbf{R}$ as the set of functions $f(W;x) = \sigma (W^K \sigma (W^{K-1} \cdots \sigma (W^1 x)))$, where the input is $x \in \mathbf{R}^d$, the weights are given by the matrices $W^k$, one per layer, with matching dimensions. We sometime use the symbol $W$ as a shorthand for the set of $W^k$ matrices $k=1,\cdots,K$. For simplicity we consider here the case of binary classification in which $f$ takes scalar values, implying that the last layer matrix $W^K$ is $W^K \in \mathbf{R}^{1,K_l}$. The labels are $y_n\in\{-1,1\}$. The weights of hidden layer $l$ are collected in a matrix of size $h_l\times h_{l-1}$. There are no biases apart form the input layer where the bias is instantiated by one of the input dimensions being a constant. The activation function in this section is the ReLU activation. For ReLU activations the following important positive one-homogeneity property holds $\sigma(z)=\frac{\partial \sigma(z)}{\partial z} z$. A consequence of one-homogeneity is a structural lemma (Lemma 2.1 of \cite{DBLP:journals/corr/abs-1711-01530}) $\sum_{i,j} W^{i,j}_k \left(\frac{\partial f(x)}{\partial W^{i,j}_k}\right)= f(x)$ where $W_k$ is here the vectorized representation of the weight matrices $W_k$ for each of the different layers (each matrix is a vector). For the network, homogeneity implies $f(W;x)=\prod_{k=1}^K \rho_k f(V_1,\cdots,V_K; x_n)$, where $W_k=\rho_k V_k$ with the matrix norm $||V_k||_p=1$. Another property of the Rademacher complexity of ReLU networks that follows from homogeneity is $\mathbb{R}_N(\mathbb{F}) = \rho \mathbb{R}_N(\tilde{\mathbb{F}})$ where $\rho=\rho_1 \prod_{k=1}^K \rho_k$, $\mathbb{F}$ is the class of neural networks described above. We define $f= \rho \tilde{f}$; $\tilde{\mathbb{F}}$ is the associated class of normalized neural networks (we call $f(V;x)=\tilde{f}(x)$ with the understanding that $f(x)=f(W;x)$). Note that $\frac{\partial f}{\partial \rho_k} = \frac{\rho}{\rho_k}\tilde{f} \label{rho}$ and that the definitions of $\rho_k$, $V_k$ and $\tilde{f}$ all depend on the choice of the norm used in normalization. In the case of training data that can be separated by the networks $f(x_n) y_n>0 \quad \forall n=1,\cdots,N$. We will sometime write $f(x_n)$ as a shorthand for $y_n f(x_n)$. \subsection{Uniform convergence bounds: minimizing a surrogate loss under norm constraint} \label{Early stopping} Classical {\it generalization bounds for regression} \cite{Bousquet2003} suggest that minimizing the empirical loss of a loss function such as the cross-entropy subject to constrained {\it complexity of the minimizer} is a way to to attain generalization, that is an expected loss which is close to the empirical loss: \begin{proposition} The following generalization bounds apply to $\forall f \in \mathbb{F}$ with probability at least $(1-\delta)$: \begin{equation} L(f) \leq \hat{L}(f) + c_1\mathbb{R}_N(\mathbb{F}) + c_2 \sqrt \frac{\ln(\frac{1}{\delta})}{2N} \label{bound} \end{equation} \end{proposition} \vskip0.1in \noindent where $L(f) = \mathbf E [\ell(f(x), y)]$ is the expected loss, $\hat{L}(f)$ is the empirical loss, $\mathbb{R}_N(\mathbb{F})$ is the empirical Rademacher average of the class of functions $\mathbb{F}$, measuring its complexity; $c_1, c_2$ are constants that depend on properties of the Lipschitz constant of the loss function, and on the architecture of the network. Thus minimizing under a constraint on the Rademacher complexity a surrogate function such as the cross-entropy (which becomes the logistic loss in the binary classification case) will minimize an upper bound on the expected classification error because such surrogate functions are upper bounds on the $0-1$ function. We can choose a class of functions $\mathbf{\tilde{F}}$ with normalized weights and write $f(x)=\rho \tilde{f}(x)$ and $\mathbb{R}_N(\mathbb{F})=\rho \mathbb{R}_N(\mathbb{\tilde{F}})$. One can choose any fixed $\rho$ as a (Ivanov) regularization-type tradeoff. In summary, the problem of generalization may be approached by minimizing the exponential loss -- more in general an exponential-type loss, such the logistic and the cross-entropy -- under a unit norm constraint on the weight matrices, since we are interested in the directions of the weights: \begin{equation} \lim_{\rho \to \infty} \arg\min_{||V_k||=1, \ \forall k} L(\rho \tilde{f}) \label{UnitNormMin} \end{equation} \noindent where we write $f(W) = \rho \tilde{f}(V)$ using the homogeneity of the network. As it will become clear later, gradient descent techniques on the exponential loss automatically increase $\rho$ to infinity. We will typically consider the sequence of minimizations over $V_k$ for a sequence of increasing $\rho$. The key quantity for us is $\tilde{f}$ and the associated weights $V_k$; $\rho$ is in a certain sense an auxiliary variable, a constraint that is progressively relaxed. In the following we explore the implications for deep networks of this classical approach to generalization. \subsubsection{Remark: minimization of an exponential-type loss implies margin maximization } Though not critical for our approach to the question of generalization in deep networks it is interesting that constrained minimization of the exponential loss implies margin maximization. This property relates our approach to the results of several recent papers \cite{2017arXiv171010345S, 2019arXiv190507325S,DBLP:journals/corr/abs-1906-05890}. Notice that our theorem \ref{margin-maxTheorem} as in \cite{DBLP:conf/nips/RossetZH03} is a {\it sufficient condition for margin maximization}. Necessity is not true for general loss functions. To state the margin property more formally, we adapt to our setting a different result due to \cite{DBLP:conf/nips/RossetZH03} (they consider for a linear network a vanishing $\lambda$ regularization term whereas we have for nonlinear networks a set of unit norm constraints). First we recall the definition of the empirical loss $L(f)=\sum_{n=1}^N \ell(y_n f(x_n))$ with an exponential loss function $\ell(yf)= e^{-yf}$. We define $\eta(f)$ a the {\it margin} of $f$, that is $\eta(f)=\min_n f(x_n)$. Then our margin maximization theorem (proved in \cite{theory_III}) takes the form \begin{theorem} Consider the set of $V_k, k=1,\cdots, K$ corresponding to \begin{equation} \min_{{||V_k||}=1} L(f(\rho_k, V_k)) \label{V(rho)} \end{equation} \noindent where the norm $||V_k||$ is a chosen $L_p$ norm and $L(f)(\rho_k, V_K) = L(\tilde{f}(\rho)) = \sum_ n \ell(y_n \rho f(V; x_n))$ is the empirical exponential loss. For each layer consider a sequence of increasing $\rho_k$. Then the associated sequence of $V_k$ defined by Equation \ref{V(rho)}, converges for $\rho \to \infty$ to the maximum margin of $\tilde{f}$, that is to $\max_{||V_k|| \leq 1} \eta(\tilde{f})$ . \label{margin-maxTheorem} \end{theorem} \subsection{Minimization under unit norm constraint: weight normalization} \label{OurWeightNormalization} The approach is then to minimize the loss function $ L(f(w))=\sum_{n=1}^N e^{- f(W;x_n) y_n }= \sum_{n=1}^N e^{- \rho f(V_k;x_n) y_n }$, with $\rho= \prod \rho_k$, subject to $||V_k||^p_p =1 \ \forall k$, that is under a unit norm constraint for the weight matrix at each layer (if $p=2$ then $\sum_{i,j} (V_k)_{i,j}^2= 1$ is the Frobenius norm), since $V_k$ are the directions of the weights which are the relevant quantity for classification. The minimization is understood as a sequence of minimizations for a sequence of increasing $\rho_k$. Clearly these constraints imply the constraint on the norm of the product of weight matrices for any $p$ norm (because any induced operator norm is a sub-multiplicative matrix norm). The standard choice for a loss function is an exponential-type loss such the cross-entropy, which for binary classification becomes the logistic function. We study here the exponential because it is simpler and retains all the basic properties. There are several gradient descent techniques that given the unconstrained optimization problem transform it into a {\it constrained} gradient descent problem. To provide the background let us formulate the standard unconstrained gradient descent problem for the exponential loss as it is used in practical training of deep networks: \begin{equation} \dot{W}^{i,j}_k = -\frac{\partial L}{\partial W^{i,j}_k}= \sum_{n=1}^N y_n \frac{\partial{f(x_n; w)}} {\partial W^{i,j}_k} e^{- y_n f(x_n;W)} \label{standardynamicsW} \end{equation} \noindent where $W_k$ is the weight matrix of layer $k$. Notice that, since the structural property implies that at a critical point we have $\sum_{n=1}^N y_n f(x_n; w) e^{- y_n f(x_n;W)} = 0$, the only critical points of this dynamics that separate the data (i.e. $y_n f(x_n; w)>0 \ \forall n$) are global minima at infinity. Of course for separable data, while the loss decreases asymptotically to zero, the norm of the weights $\rho_k$ increases to infinity, as we will see later. Equations \ref{standardynamicsW} define a dynamical system in terms of the gradient of the exponential loss $L$. The set of gradient-based algorithms enforcing a unit-norm constraints \cite{845952} comprises several techniques that are equivalent for small values of the step size. They are all good approximations of the true gradient method. One of them is the {\it Lagrange multiplier method}; another is the {\it tangent gradient method} based on the following theorem: \begin{theorem} \cite{845952} Let $||u||_p$ denote a vector norm that is differentiable with respect to the elements of $u$ and let $g(t)$ be any vector function with finite $L_2$ norm. Then, calling $\nu(t)=\frac{\partial ||u||_p}{\partial u}_{u=u(t)}$, the equation \begin{equation} \dot{u}=h_g(t)=Sg(t)= (I-\frac{\nu \nu^T}{||\nu||_2^2}) g(t) \label{dot_u} \end{equation} \noindent with $||u(0)|| =1$, describes the flow of a vector $u$ that satisfies $||u(t)||_p=1$ for all $t \ge 0$. \label{Theorem1} \end{theorem} In particular, a form for $g$ is $g(t)= \mu(t) \nabla_u L$, the gradient update in a gradient descent algorithm. We call $Sg(t)$ the tangent gradient transformation of $g$. In the case of $p=2$ we replace $\nu$ in Equation \ref{dot_u} with $u$ because $\nu(t)=\frac{\partial ||u||_2}{\partial u}=u$. This gives $S= I-\frac{u u^T}{||u||_2^2}$ and $\dot{u}=Sg(t).$ Consider now the empirical loss $L$ written in terms of $V_k$ and $\rho_k$ instead of $W_k$, using the change of variables defined by $W_k=\rho_k V_k$ but without imposing a unit norm constraint on $V_k$. The flows in $\rho_k,V_k$ can be computed as $\dot{\rho_k}=\frac{\partial W_k}{\partial \rho_k} \frac{\partial L}{\partial W_k}= V_k^T \frac{\partial L}{\partial W_k}$ and $\dot{V_k}=\frac{\partial W_k}{\partial V_k} \frac{\partial L}{\partial W_k} = \rho_k \frac{\partial L}{\partial W_k}$, with $\frac{\partial L}{\partial W_k}$ given by Equations \ref{standardynamicsW}. We now enforce the unit norm constraint on $V_k$ by using the tangent gradient transform on the $V_k$ flow. This yields \begin{equation} \dot{\rho_k}= V_k^T \frac{\partial L}{\partial W_k} \quad \dot{V_k}= S_k \rho_k \frac{\partial L}{\partial W_k}. \label{v-flow-withunitnorm} \end{equation} Notice that the dynamics above follows from the classical approach of controlling the Rademacher complexity of $\tilde{f}$ during optimization (suggested by bounds such as Equation \ref{bound}. The approach and the resulting dynamics for the directions of the weights may seem different from the standard unconstrained approach in training deep networks. It turns out, however, that the dynamics described by Equations \ref{v-flow-withunitnorm} is the same dynamics of {\it Weight Normalization}. The technique of {\it Weight normalization} \cite{SalDied16} was originally proposed as a small improvement on standard gradient descent ``to reduce covariate shifts''. It was defined for each layer in terms of $w=g \frac{v}{||v||}$, as \begin{equation} \dot{g}=\frac{v}{||v||} \frac{\partial L}{\partial w} \dot{v}=\frac{g}{||v||} S \frac{\partial L}{\partial w} \label{W-normalization} \end{equation} \noindent with $S=I- \frac{v v^T}{||v||^2}$. It is easy to see that Equations \ref{v-flow-withunitnorm} are the same as the weight normalization Equations \ref{W-normalization}, if $||v||_2=1$. We now observe, multiplying Equation \ref{v-flow-withunitnorm} by $v^T$, that $v^T \dot{v}=0$ because $v^T S=0$, implying that $||v||^2$ is constant in time with a constant that can be taken to be $1$. Thus the two dynamics are the same. \subsection{Generalization with hidden complexity control} \label{nounitnorm} Empirically it appears that GD and SGD converge to solutions that can generalize even without batch or weight normalization. Convergence may be difficult for quite deep networks and generalization may not be as good as with batch normalization but it still occurs. How is this possible? We study the dynamical system $\dot{W_k}^{i,j}$ under the reparametrization $W^{i,j}_k = \rho_k V^{i,j}_k$ with $||V_k||_2=1$. We consider for each weight matrix $W_k$ the corresponding ``vectorized'' representation in terms of vectors $W_k^{i,j} = W_k$. We use the following definitions and properties (for a vector $w$): \begin{itemize} \item Define $\frac{w}{||w||_2}=\tilde{w}$; thus $w=||w||_2\tilde{w}$ with $||\tilde{w}||_2=1$. Also define $S={I-\tilde{w}\tilde{w}^T}=I- \frac {w w^T}{||w||_2^2}$. \item The following relations are easy to check: \begin{enumerate} \item $\frac{\partial ||w||_2}{\partial w}=\tilde{w}$ \item $\frac{\partial \tilde{w}}{\partial w}=\frac{S}{||w||_2}$. \item $Sw=S \tilde{w}=0$ \item $S^2=S$ \label{Relations} \end{enumerate} \end{itemize} The gradient descent dynamic system used in training deep networks for the exponential loss is given by Equation \ref{standardynamicsW}. Following the chain rule {\it for the time derivatives}, the dynamics for $W_k$ is exactly (see \cite{theory_III}) equivalent to the following dynamics for $||W_k||=\rho_k$ and $V_k$: \begin{equation} \dot{\rho_k}= \frac{\partial ||W_k||}{\partial W_k} \frac{\partial W_k}{\partial t}= V^T_k \dot{W_k} \label{rhodot} \end{equation} \noindent and \begin{equation} \dot{V_k}= \frac{\partial V_k}{\partial W_k} \frac{\partial W_k}{\partial t}= \frac {S_k}{\rho_k} \dot{W_k} \label{vdot} \end{equation} \noindent where $S_k= I- V_kV_k^T$. We used property 1 in \ref{Relations} for Equation \ref{rhodot} and property 2 for Equation \ref{vdot}. The key point here is that the dynamics of $\dot{V_k}$ includes a unit $L_2$ norm constraint: using the tangent gradient transform will not change the equation because $S^2=S$. As separate remarks , notice that if for $t>t_0$, $f$ separates all the data, $\frac{d}{dt}{\rho_k} >0$, that is $\rho$ diverges to $\infty$ with $\lim_{t \to \infty}\dot{\rho}=0$. In the 1-layer network case the dynamics yields $\rho \approx \log t$ asymptotically. For deeper networks, this is different. \cite{theory_III} shows (for one support vector) that the product of weights at each layer diverges faster than logarithmically, but each individual layer diverges slower than in the 1-layer case. The norm of the each layer grows at the same rate $\dot{\rho_k^2}$, independent of $k$. The $V_k$ dynamics has stationary or critical points given by \begin{equation} \sum \alpha_n(\rho(t) \left(\frac{\partial{\tilde{f}(x_n)}} {\partial V_k^{i,j}}-V_k^{i,j} \tilde{f}(x_n) \right), \label{wdot4} \end{equation} \noindent where $\alpha_n= e^{-y_n \rho(t) \tilde{f}(x_n)}$. We examine later the linear one-layer case $\tilde{f}(x)=v^T x$ in which case the stationary points of the gradient are given by $\sum \alpha_n(\rho(t) (x_n - v v^T x_n)$ and of course coincide with the solutions obtained with Lagrange multipliers. In the general case the critical points correspond for $\rho \to \infty$ to degenerate zero ``asymptotic minima'' of the loss. To understand whether there exists an implicit complexity control in standard gradient descent of the weight directions, we check whether there exists an $L_p$ norm for which unconstrained normalization is equivalent to constrained normalization. From Theorem \ref{Theorem1} we expect the constrained case to be given by the action of the following projector onto the tangent space: \begin{equation} S_{p} = I-\frac{\nu \nu^T}{||\nu||_2^2} \quad\textnormal{with}\quad \nu_i=\frac{\partial ||w||_p}{\partial w_i} = \textnormal{sign}(w_i)\circ\left(\frac{|w_i|}{||w||_p}\right)^{p-1}. \end{equation} The constrained Gradient Descent is then \begin{equation} \dot{\rho_k}= V^T_k \dot{W_k} \quad \dot{V_k} = \rho_k S_p \dot{W_k}. \label{ConstrainedGradP} \end{equation} On the other hand, reparametrization of the unconstrained dynamics in the $p$-norm gives (following Equations \ref{rhodot} and \ref{vdot}) \begin{equation} \begin{split} \dot{\rho_k}&= \frac{\partial ||W_k||_p}{\partial W_k} \frac{\partial W_k}{\partial t}= \textnormal{sign}(W_k)\circ\left(\frac{|W_k|}{||W_k||_p}\right)^{p-1} \cdot \dot{W_k} \\ \dot{V_k}&= \frac{\partial V_k}{\partial W_k} \frac{\partial W_k}{\partial t}= \frac {I - \textnormal{sign}(W_k) \circ\left(\frac{|W_k|}{||W_k||_p}\right)^{p-1}W_k^T}{||W_k||_p^{p-1}} \dot{W_k}. \end{split} \end{equation} These two dynamical systems are clearly different for generic $p$ reflecting the presence or absence of a regularization-like constraint on the dynamics of $V_k$. As we have seen however, for $p=2$ the 1-layer dynamical system obtained by minimizing $L$ in $\rho_k$ and $V_k$ with $W_k=\rho_k V_k$ under the constraint $||V_k||_2=1$, is the weight normalization dynamics \begin{equation} \dot{\rho_k}=V_k^T \dot{W_k} \quad \dot{V_k}= S \rho_k \dot{W_k} , \label{WN} \end{equation} \noindent which is quite similar to the standard gradient equations \begin{equation} \dot{\rho_k}= V_k^T \dot{W_k} \quad \dot{v} =\frac{S}{\rho_k} \dot{W_k}. \label{StandardGrad} \end{equation} \begin{figure*}[t!]\centering \includegraphics[width=1.0\textwidth]{Figures/fig2.pdf} \caption{\it The top left graph shows testing vs training cross-entropy loss for networks each trained on the same data sets (CIFAR10) but with a different initializations, yielding zero classification error on training set but different testing errors. The top right graph shows the same data, that is testing vs training loss for the same networks, now normalized by dividing each weight by the Frobenius norm of its layer. Notice that all points have zero classification error at training. The red point on the top right refers to a network trained on the same CIFAR-10 data set but with randomized labels. It shows zero classification error at training and test error at chance level. The top line is a square-loss regression of slope $1$ with positive intercept. The bottom line is the diagonal at which training and test loss are equal. The networks are 3-layer convolutional networks. The left can be considered as a visualization of Equation \ref{bound} when the Rademacher complexity is not controlled. The right hand side is a visualization of the same relation for normalized networks that is $L(\tilde{f}) \leq \hat{L}(\tilde{f}) + c_1\mathbb{R}_N(\mathbb{\tilde{F}}) + c_2 \sqrt \frac{\ln(\frac{1}{\delta})}{2N}$. Under our conditions for $N$ and for the architecture of the network the terms $c_1\mathbb{R}_N(\mathbb{\tilde{F}}) + c_2 \sqrt \frac{\ln(\frac{1}{\delta})}{2N}$ represent a small offset. From \cite{DBLP:journals/corr/abs-1807-09659}. } \label{main} \end{figure*} The two dynamical systems differ only by a $\rho_k^2$ factor in the $\dot{V_k}$ equations. However, the critical points of the gradient for the $V_k$ flow, that is the point for which $\dot{V_k}=0$, are the same in both cases since for any $t>0$ $\rho_k(t)>0$ and thus $\dot{V_k}=0$ is equivalent to $S\dot{W_k}=0$. Hence, gradient descent with unit $L_p$-norm constraint is equivalent to the standard, unconstrained gradient descent but only when $p = 2$. Thus \begin{fact} The standard dynamical system used in deep learning, defined by $\dot{W_k}=-\frac{\partial L}{\partial W_k}$, implicitly respectss a unit $L_2$ norm constraint on $V_k$ with $\rho_k V_k =W_k$. Thus, under an exponential loss, if the dynamics converges, the $V_k$ represent the minimizer under the $L_2$ unit norm constraint. \label{w(T)} \end{fact} Thus standard GD implicitly enforces the $L_2$ norm constraint on $V_k=\frac{W_k}{||W_k||_2}$, consistently with Srebro's results on implicit bias of GD. Other minimization techniques such as coordinate descent may be biased towards different norm constraints. \subsection{Linear networks and rates of convergence} The linear ($f(x)=\rho v^T x$) networks case \cite{2017arXiv171010345S} is an interesting example of our analysis in terms of $\rho$ and $v$ dynamics. We start with unconstrained gradient descent, that is with the dynamical system \begin{equation} \dot{\rho}= \frac{1}{\rho} \sum_{n=1}^N e^{- \rho v^Tx_n} v^Tx_n \quad \dot{v}=\frac{1}{\rho}\sum_{n=1}^N e^{- \rho v^T x_n} (x_n- v v^T x_n). \label{Ttt} \end{equation} If gradient descent in $v$ converges to $\dot{v}=0$ at finite time, $v$ satisfies $ v v^T x = x$, where $x= \sum_{j=1}^C \alpha_j x_j$ with positive coefficients $\alpha_j$ and $x_j$ are the $C$ support vectors (see \cite{theory_III}). A solution $v^T = ||x|| x^\dagger$ then {\it exists} ($x^\dagger$, the pseudoinverse of $x$, since $x$ is a vector, is given by $x^\dagger= \frac{x^T}{||x||^2}$). On the other hand, the operator $T$ in $v(t+1)=T v(t)$ associated with equation \ref{Ttt} is non-expanding, because $||v||=1,\ \forall t$. Thus there is a fixed point $v \propto x$ which is {\it independent of initial conditions} \cite{Ferreira1996} and unique (in the linear case) The rates of convergence of the solutions $\rho(t)$ and $v(t)$, derived in different way in \cite{2017arXiv171010345S}, may be read out from the equations for $\rho$ and $v$. It is easy to check that a general solution for $\rho$ is of the form $\rho \propto C \log t$. A similar estimate for the exponential term gives $e^{- \rho v^T x_n} \propto \frac{1}{t}$. Assume for simplicity a single support vector $x$. We claim that a solution for the error $\epsilon= v-x$, since $v$ converges to $x$, behaves as $\frac{1}{\log t}$. In fact we write $v =x+\epsilon$ and plug it in the equation for $v$ in \ref{T}. We obtain (assuming normalized input $||x||=1$) \begin{equation} \dot{\epsilon}=\frac{1}{\rho} e^{- \rho v^T x} (x- (x+\epsilon) (x+\epsilon)^T x) \approx \frac{1}{\rho} e^{- \rho v^T x} ( x- x - x \epsilon^T - \epsilon x^T), \label{T} \end{equation} \noindent which has the form $\dot{\epsilon}=-\frac{1}{t \log t} (2 x \epsilon^T)$. Assuming $\epsilon$ of the form $\epsilon \propto \frac{1}{\log t}$ we obtain $-\frac{1}{t \log^2 t }= -B \frac{1}{t \log^2 t}$. Thus the error indeed converges as $\epsilon \propto \frac{1}{\log t}$. A similar analysis for the weight normalization equations \ref{WN} considers the same dynamical system with a change in the equation for $v$, which becomes \begin{equation} \dot{v} \propto e^{-\rho} \rho (I- v v^T) x. \label{T-WN} \end{equation} This equation differs by a factor $\rho^2$ from equation \ref{T}. As a consequence equation \ref{T-WN} is of the form $\dot{\epsilon}=-\frac{\log t}{t} \epsilon$, with a general solution of the form $\epsilon \propto t^{-\frac{1}{2}\log t}$. In summary, {\it GD with weight normalization converges faster to the same equilibrium than standard gradient descent: the rate for $\epsilon= v- x$ is $t^{-\frac{1}{2} log(t)}$ vs $\frac{1}{\log t}$.} Our goal was to find $\lim_{\rho \to \infty} \arg \min_{||V_k||=1, \ \forall k} L(\rho \tilde{f}) $. We have seen that various forms of gradient descent enforce different paths in increasing $\rho$ that empirically have different effects on convergence rate. It will be an interesting theoretical and practical challenge to find the optimal way, in terms of generalization and convergence rate, to grow $\rho\rightarrow \infty$. Our analysis of simplified batch normalization \cite{theory_III} suggests that several of the same considerations that we used for weight normalization should apply (in the linear one layer case BN is identical to WN). However, BN differs from WN in the multilayer case in several ways, in addition to weight normalization: it has for instance separate normalization for each unit, that is for each row of the weight matrix at each layer. \begin{figure \centering \includegraphics[width=0.9\linewidth]{Figures/no-overfitting.pdf} \caption{Empirical and expected error in CIFAR 10 as a function of number of neurons in a 5-layer convolutional network. The expected classification error does not increase when increasing the number of parameters beyond the size of the training set in the range we tested.} \label{no-overfitting} \end{figure} \section{Discussion} A main difference between shallow and deep networks is in terms of {\it approximation} power or, in equivalent words, of the ability to learn good representations from data based on the compositional structure of certain tasks. Unlike shallow networks, deep local networks -- in particular convolutional networks -- can avoid the curse of dimensionality in approximating the class of hierarchically local compositional functions. This means that for such class of functions deep local networks represent an appropriate hypothesis class that allows good approximation with a minimum number of parameters. It is not clear, of course, why many problems encountered in practice should match the class of compositional functions. Though we and others have argued that the explanation may be in either the physics or the neuroscience of the brain, these arguments are not rigorous. Our conjecture at present is that compositionality is imposed by the wiring of our cortex and, critically, is reflected in language. Thus compositionality of some of the most common visual tasks may simply reflect the way our brain works. {\it Optimization} turns out to be surprisingly easy to perform for overparametrized deep networks because SGD will converge with high probability to global minima that are typically much more degenerate for the exponential loss than other local critical points. More surprisingly, gradient descent yields {\it generalization} in classification performance, despite overparametrization and even in the absence of explicit norm control or regularization, because standard gradient descent in the weights enforces an implicit unit ($L_2$) norm constraint on the {\it directions of the weights} in the case of exponential-type losses. In summary, it is tempting to conclude that the practical success of deep learning has its roots in the almost magic synergy of unexpected and elegant theoretical properties of several aspects of the technique: the deep convolutional network architecture itself, its overparametrization, the use of stochastic gradient descent, the exponential loss, the homogeneity of the RELU units and of the resulting networks. Of course many problems remain open on the way to develop a full theory and, especially, in translating it to new architectures. More detailed results are needed in approximation theory, especially for densely connected networks. Our framework for optimization is missing at present a full classification of local minima and their dependence on overparametrization for general loss functions. The analysis of generalization should include an analysis of convergence of the weights for multilayer networks (see \cite{2019arXiv190507325S} and \cite{DBLP:journals/corr/abs-1906-05890}). A full theory would also require an analysis of the trade-off between approximation and estimation error, relaxing the separability assumption. \showmatmethods{} \acknow{We are grateful to Sasha Rakhlin and Nate Srebro for useful suggestions about the structural lemma and about separating critical points. Part of the funding is from the Center for Brains, Minds and Machines (CBMM), funded by NSF STC award CCF-1231216, and part by C-BRIC, one of six centers in JUMP, a Semiconductor Research Corporation (SRC) program sponsored by DARPA.} \showacknow{}
{ "timestamp": "2019-08-27T02:16:15", "yymm": "1908", "arxiv_id": "1908.09375", "language": "en", "url": "https://arxiv.org/abs/1908.09375", "abstract": "While deep learning is successful in a number of applications, it is not yet well understood theoretically. A satisfactory theoretical characterization of deep learning however, is beginning to emerge. It covers the following questions: 1) representation power of deep networks 2) optimization of the empirical risk 3) generalization properties of gradient descent techniques --- why the expected error does not suffer, despite the absence of explicit regularization, when the networks are overparametrized? In this review we discuss recent advances in the three areas. In approximation theory both shallow and deep networks have been shown to approximate any continuous functions on a bounded domain at the expense of an exponential number of parameters (exponential in the dimensionality of the function). However, for a subset of compositional functions, deep networks of the convolutional type can have a linear dependence on dimensionality, unlike shallow networks. In optimization we discuss the loss landscape for the exponential loss function and show that stochastic gradient descent will find with high probability the global minima. To address the question of generalization for classification tasks, we use classical uniform convergence results to justify minimizing a surrogate exponential-type loss function under a unit norm constraint on the weight matrix at each layer -- since the interesting variables for classification are the weight directions rather than the weights. Our approach, which is supported by several independent new results, offers a solution to the puzzle about generalization performance of deep overparametrized ReLU networks, uncovering the origin of the underlying hidden complexity control.", "subjects": "Machine Learning (cs.LG); Machine Learning (stat.ML)", "title": "Theoretical Issues in Deep Networks: Approximation, Optimization and Generalization", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.975576912786245, "lm_q2_score": 0.8221891283434877, "lm_q1q2_score": 0.8021087315557535 }
https://arxiv.org/abs/1801.01692
Algebraic stories from one and from the other pockets
We present a number of questions in commutative algebra posed on the problem solving seminar in algebra at Stockholm University during the period Fall 2014 - Spring 2017.
\section{The Waring problem for complex-valued forms} The following famous result on binary forms was proven by J.~J.~Sylvester in 1851. (Below we use the terms ``forms" and ``homogeneous polynomials" as synonyms.) \begin{Theorem}[Sylvester's Theorem \cite{Syl2}]\label{th:Sylv} \rm{(i)} A general binary form $f$ of odd degree $k=2s-1$ with complex coefficients can be written as $$ f(x, y) =\sum_{j=1}^s(\al_jx+\be_jy)^k.$$ \noindent \rm{(ii)} A general binary form $f$ of even degree $k=2s$ with complex coefficients can be written as $$ f(x,y)=\la x^k +\sum_{j=1}^s(\al_jx+\be_jy)^k.$$ \end{Theorem} Sylvester's result was the starting point of the study of the so-called \emph{Waring problem for polynomials} which we discuss below. \medskip Let $S = \bC[x_1,\ldots,x_n]$ be the polynomial ring in $n$ variables with complex coefficients. With respect to the standard grading, we have $S = \bigoplus_{d\geq 0} S_d$, where $S_d$ denotes the vector space of all forms of degree $d$. \begin{defi+}\label{def1} Let $f$ be a form of degree $k$ in $S$. A presentation of $f$ as a sum of $k$-th powers of linear forms, i.e., $f=l_1^k+\ldots+l_s^k$, where $l_1,\ldots,l_s \in S_1$, is called a \emph{Waring decomposition} of $f$. The minimal length of such a decomposition is called the \emph {Waring rank} of $f$, and we denote it as $\rk(f)$. By $\rk^\circ(k,n)$ we denote the Waring rank of a {\em general} complex-valued form of degree $k$ in $n$ variables. \end{defi+} \begin{Rmk} Besides being a natural question from the point of view of algebraic geometry, the Waring problem for polynomials is partly motivated by its celebrated prototype, i.e., the Waring problem for natural numbers. The latter was posed in 1770 by a British number theorist E. Waring who claimed that, for any positive integer $k$, there exists a minimal number $g(k)$ such that every natural number can be written as sum of at most $g(k)$ $k$-th powers of positive integers. The famous Lagrange's four-squares Theorem (1770) claims that $g(2) = 4$ while the existence of $g(k)$, for any integer $k\ge 2$, is due to D. Hilbert (1900). Exact values of $g(k)$ are currently known only in a few cases. \end{Rmk} \subsection{Generic $k$-rank} In terms of Definition~\ref{def1}, Sylvester's Theorem claims that the Waring rank of a general binary complex-valued form of degree $k$ equals $\left\lfloor \frac{k}{2} \right\rfloor$. More generally, the important result of J. Alexander and A. Hirschowitz \cite{al-hi} completely describes the Waring rank $\rk^\circ(k,n)$ of general forms of any degree and in any number of variables. \begin{Theorem}[Alexander-Hirschowitz Theorem, 1995]\label{th:AH} For all pairs of positive integers $(k,n)$, the generic Waring rank $\rk^\circ(k,n)$ is given by \begin{equation}\label{rgen} \rk^\circ(k,n)=\left\lceil \frac{\binom{n+k-1}{n-1}}{n} \right\rceil, \end{equation} except for the following cases: \begin{enumerate} \item $k = 2$, where $\rk^{\circ}(2,n) = n$; \item $k=4,\; n=3,4,5$; and $k=3, n=5$, where, $r_{gen}(k,n)$ equals the r.h.s of \eqref{rgen} plus $1$. \end{enumerate} \end{Theorem} Going further, R.F. and B.S. jointly with G.~Ottaviani considered the following natural version of the Waring problem for complex-valued forms, see \cite{FOS}. \begin{defi+} Let $k,d$ be positive integers. Given a form $f$ of degree $kd$, a {\em $k$-Waring decomposition} is a presentation of $f$ as a sum of $k$-th powers of forms of degree $d$, i.e., $f = g_1^k + \cdots + g_s^k$, with $g_i \in S_d$. The minimal length of such an expression is called the {\em $k$-rank} of $f$ and is denoted by $\rk_k(f)$. We denote by $\rk_k^\circ(kd,n)$ the $k$-rank of a general complex-valued form of degree $kd$ in $n$ variables. \end{defi+} In this notation, the case $d = 1$ corresponds to the classical Waring rank, i.e., if $k = \deg(f)$, then $\rk(f) = \rk_{k}(f)$ and $\rk^\circ(k,n)=\rk_k^\circ(k,n)$. Since the case $k = 1$ is trivial, we assume below that $k \geq 2$. \begin{problem} Given a triple of positive integers $(k,d,n)$, calculate $\rk_k^\circ(kd,n).$ \end{problem} The main result of \cite{FOS} states that, for any triple $(k,d,n)$ as above, \begin{equation}\label{ineq:main} \rk^\circ_k(kd,n)\le k^{n-1}. \end{equation} At the same time, by a simple parameter count, one has an obvious lower bound for $\rk^\circ(k,n)$ given by \begin{equation}\label{ineq:lower bound} \rk^\circ_k(kd,n)\geq \left\lceil \frac{\binom{n+kd-1}{n-1}} {\binom{n+d-1}{n-1}} \right\rceil. \end{equation} \medskip A remarkable fact about the upper bound given by \eqref{ineq:main} is that it is independent of $d$. Therefore, since the right-hand side of \eqref{ineq:lower bound} equals $k^n$ when $d \gg 0$, we get that for large values of $d$, the bound in \eqref{ineq:main} is actually sharp. As a consequence of this remark, for any fixed $n\ge 1$ and $k\ge 2$, there exists a positive integer $d_{k,n}$ such that $\rk^\circ_k(kd,n)= k^n,$ for all $d\ge d_{k,n}$. In the case of binary forms, it has been proven that \eqref{ineq:lower bound} is actually an equality \cite{Re1, LuOnReSh}. Exact values of $d_{k,n}$, and the behaviour of $\rk_k(kd,n)$ for $d \leq d_{k,n}$, have also been computed in a few other cases, see \cite[Section 3.3]{On}. These results agree with the following illuminating conjecture suggested by G.~Ottaviani in 2014. \begin{conjecture}\label{conj:main} The $k$-rank of a general form of degree $kd$ in $n$ variables is given by \begin{equation}\label{eq:main} \rk_k^\circ(kd,n)=\begin{cases} \min \left\{s\ge 1 ~|~ s\binom{n+d-1}{n-1}-\binom {s}{2}\ge \binom{n+2d-1}{n-1}\right\}, & \text{ for } k=2;\\ \min \left\{s\ge 1 ~|~ s\binom{n+d-1}{n-1}\ge \binom{n+kd-1}{n-1}\right\}, & \text{ for } k\ge 3. \end{cases} \end{equation} \end{conjecture} Observe that, for $k\ge 3$, Conjecture~\ref{conj:main} claims that the na\"ive bound \eqref{ineq:lower bound} obtained by a parameter count is actually sharp, while, for $k = 2$, due to an additional group action there are many {\it defective} cases where the inequality is strict. \begin{Rmk} Problems about additive decompositions including the above Waring problems can be usually rephrased geometrically in terms of {\em secant varieties}. In the case of $k$-Waring decompositions, we need to consider the {\em variety of powers} $V_{n,kd}^{(k)}$, i.e., the variety of $k$-th powers of forms of degree $d$ inside the (projective) space of forms of degree $kd$. The {\em $s$-th secant variety} $\sigma_s(V_{n,kd}^{(k)})$ is the Zariski closure of the union of all linear spaces spanned by $s$-tuples of points lying on $V_{n,kd}^{(k)}$. In other words, it is the closure of the set of forms whose $k$-rank is at most $s$. Since the variety of powers is non-degenerate, i.e., it is not contained in any proper linear subspace of the space of forms of degree $kd$, the sequence of secant varieties stratifies the latter space and coincides with it for all sufficiently large $s$. Hence, the $k$-rank of a general form is the smallest value of $s$ for which the $s$-th secant variety of $V_{n,kd}^{(k)}$ coincides with the space of all forms of degree $kd$. Hence, Conjecture \ref{conj:main} can be rephrased as a conjecture about the dimensions of the secant varieties of $V_{n,kd}^{(k)}$. (We refer to \cite[Section 1.3.3]{On} for more details.) \end{Rmk} \subsection{Maximal $k$-rank} A harder problem which is largely open even in the classical case of Waring decompositions, deals with the computation of the $k$-rank of an {\em arbitrary} complex-valued form of degree divisible by $k$. \begin{defi+} Given a triple $(k,d,n)$, denote by $\rk^{\max}_k(kd,n)$ the minimal number of terms such that {\em every} form of degree $kd$ in $n+1$ variables can be represented as the sum of at most $\rk^{\max}_k(kd,n)$ $k$-th powers of forms of degree $d$. The number $\rk^{\max}_k(kd,n)$ is called the {\it maximal $k$-rank}. (Similarly to the above, we omit the subscript when considering the classical Waring rank, i.e., for $d = 1$.) \end{defi+} In \cite[Theorem 5.4]{Re1}, B. Reznick proved that the maximal Waring rank of binary forms of degree $k$ equals $k$. Moreover, the maximal value $k$ is attained exactly on the binary forms representable as $\ell_1 \ell_2^{k-1},$ where $\ell_1$ and $\ell_2$ are any two non-proportional linear binary forms. (Apparently these claims have been known much earlier, but have never been carefully written down with a complete proof.) \begin{problem} Given a triple of positive integers $(k,d,n)$, calculate $\rk_k^{\max}(kd,n).$ \end{problem} At the moment, we have an explicit conjecture about the maximal $k$-rank only in the case of binary forms. \begin{conjecture}\label{conj:maxrank} For any positive integers $k,d$, the maximal $k$-rank $\rk^{\max}_k(kd,2)$ of binary forms equals $k$. Additionally, in the above notation, binary forms representable by $\ell_1 \ell_2^{kd-1}$, where $\ell_1$ and $\ell_2$ are non-proportional linear forms, have the latter maximal $k$-rank. \end{conjecture} Conjecture~\ref{conj:maxrank} is obvious for $k=2$ since, for any binary form $f$ of degree $2d$, we can write \begin{equation} \label{eq:k2} f = g_1 g_2 = \left(\frac{1}{2} (g_1+g_2)\right)^2 + \left(\frac{i}{2} (g_1-g_2)\right)^2 \text{ with } g_1,g_2 \in S_d. \end{equation} The first non-trivial case is the one of binary sextics, i.e., $k=3, d=2$, which has been settled in \cite{LuOnReSh} where it has also been shown that the $4$-rank of $x_1x_2^7$ is equal to $4$. \begin{Rmk} The best known general result about maximal ranks is due to G. Bleckherman and Z. Teitler, see \cite{BT} where they prove that the maximal rank is always at most twice as big as the generic rank. (This fact is true both for the classical ($d=1$) and for the higher ($d\geq 2$) Waring ranks.) In the classical case of Waring ranks, this bound is (almost) sharp for binary forms, but in many other cases it is rather crude. At present, better bounds are known only in few special cases of low degrees \cite{BD13, Je14}. To the best of our knowledge, the exact values of the maximal Waring rank are only known for binary forms (classical, see \cite{Re1}), quadrics (classical), ternary cubics (see \cite{Seg42, LT10}), ternary quartics \cite{Kl99}, ternary quintics \cite{DeP15} and quaternary cubics \cite{Seg42}. \end{Rmk} \subsection{The $k$-rank of monomials} Let $m = x_1^{a_1} \cdots x_n^{a_n}$ be a monomial with $0 < a_1 \leq a_2 \leq \cdots \leq a_n.$ It has been shown in \cite{ca-ch-ge} that the classical Waring rank of $m$ is equal to $\frac{1}{(a_1+1)}\prod_{i=1,\ldots,n} (a_i+1)$. \medskip Later E. Carlini and A.O. settled the case of the $2$-rank, see \cite{ca-on}. Namely, if $m$ is a monomial of degree $2d$, then we can write $m=m_1 m_2$, where $m_1$ and $m_2$ are monomials of degree $d$. From identity (\ref{eq:k2}), it follows that the $2$-rank of $m$ is at most two. On the other hand, $m$ has rank one exactly when we can choose $m_1 = m_2$, i.e., when the power of each variable in $m$ is even. While the cases $k=1$ and $k=2$ are solved, for $k\ge 3$, the question about the $k$-rank of monomials of degree $kd$, is still open. At present, we are only aware of two general results in this direction. Namely, \cite{ca-on} contains the bound $\rk_k(m) \leq 2^{k-1}$, and recently, S.L., A.O., B.S., together with B. Reznick, have shown that $\rk_k(m) \leq k$ when $d \geq n(k-2)$, see \cite{LuOnReSh}. Thus, for fixed $k$ and $n$, all but a finite number of monomials of degree divisible by $k$ have $k$-rank less than $k$. \begin{problem} \label{prob:monrk} Given $k \geq 3$ and a monomial $m$ of degree $kd$, determine the monomial $k$-rank $\rk_k(m)$. \end{problem} In the case of binary forms, a bit more is currently known which motivates the following question. \begin{problem}\label{prob:C} Given $k \geq 3$ and a monomial $x^a y^b$ of degree $a+b=kd$, it is known that $\rk_k(x^{a}y^{b}) \leq \max(s,t)+1$, where $s$ and $t$ are the remainders of the division of $a$ and $b$ by $k$, see \cite{ca-on}. Is it true that the latter inequality is, in fact, an equality? \end{problem} \subsection{Degree of the Waring map} Here again, we concentrate on the case of binary forms (i.e., $n=2$). As we mentioned above, in this case, it is proven that $$\rk_k^\circ(kd,2)= \left \lceil \frac{\dim S_{kd}}{\dim S_{d}}\right \rceil= \left \lceil\frac{kd+1}{d+1}\right \rceil.$$ \begin{defi+} We say that a pair $(k,d)$ is {\it perfect} if $\frac{kd+1}{d+1}$ is an integer. \end{defi+} All perfect pairs are easy to describe. \begin{Lemma}\label{lm:perfect} The set of all pairs $(k,d)$ for which $\frac{kd+1}{d+1}\in\bN$ splits into the disjoint sequences $E_j := \{(jd+j+1,d) ~|~ d = 1,2,\ldots \}$. In each $E_j$, the corresponding quotient equals $jd+1$. \end{Lemma} Given a perfect pair $(k,d)$, set $s:=\frac{kd+1}{d+1}$. Consider the map $$W_{k,d}: S_d \times \ldots \times S_d \to S_{kd}, ~~(g_1,\ldots,g_s) \mapsto g_1^k + \ldots + g_s^k.$$ Let $\widetilde{W}_{k,d}$ be the same map, but defined up to a permutation of the $g_i$'s. We call it the {\it Waring map}. By \cite[Theorem 2.3]{LuOnReSh}, $\widetilde{W}_{k,d}$ is a generically finite map of complex linear spaces of the same dimension. By definition, its {\it degree} is the cardinality of the inverse image of a generic form in $S_{kd}$. \begin{problem}\label{pr:deg} Calculate the degree of $\widetilde{W}_{k,d}$ for perfect pairs $(k,d)$. \end{problem} For the classical Waring decomposition ($d = 1$), we have a perfect pair if and only if $k$ is odd. From Sylvester's Theorem, we know that in this case the degree of the Waring map is $1$, i.e., the general binary form of odd degree has a {\it unique} Waring decomposition, up to a permutation of its summands. \begin{Rmk} For the case of the classical Waring decomposition, the latter problem has also been considered in the case of more variables. In modern terminology, the cases where the general form of a given degree has a unique decomposition up to a permutation of the summands are called {\it identifiable}. Besides the case of binary forms of odd degree, some other identifiable cases are classically known. These are the quaternary cubics (Sylvester's Pentahedral Theorem \cite{Syl2}) and the ternary quintics \cite{Hilb, Pal03, Ri04, MM13}. Recently, F.~Galuppi and M.~Mella proved that these are the only possible identifiable cases, \cite{GM16}. \end{Rmk} \begin{Rmk} Problems dealing with additive decompositions of homogeneous polynomials similar to those we consider in this section, have a very long story going back to J.~J.~Sylvester and the Italian school of algebraic geometry of the late 19-th century. In the last decades, these problems received renewed attention due to their potential applications. Namely, homogeneous polynomials can be naturally identified with {\it symmetric tensors} and in several applied branches of science where such tensors are used, for example, to encode multidimensional data sets, {\it additive decompositions of tensors} play a crucial role as an efficient way to code those. We refer to \cite{Lan} for an extensive exposition of these connections. \end{Rmk} \section{Ideals of generic forms} Let $I$ be a homogeneous ideal in $S$, i.e., an ideal generated by homogeneous polynomials. The ideal $I$ and the quotient algebra $R = S/I$ inherit the {\it grading} of the polynomial ring. \begin{defi+} Given a homogeneous ideal $I \subset S$, we call the function $$ \HF_R(i) := \dim_{\bC} R_i = \dim_{\bC} S_i - \dim_\bC I_i $$ the {\it Hilbert function} of $R.$ The power series $$ \HS_R(i) := \sum_{i\in\bN} \HF_R(i) t^i \in \bC[[t]] $$ is called the {\em Hilbert series} of $R$. \end{defi+} Let $I$ be a homogeneous ideal generated by forms $f_1, \dots, f_r$ of degrees $d_1, \dots, d_r$, respectively. It was shown in \cite{fr-lo} that, for fixed parameters $(n,d_1,\dots, d_r)$, there exists only a finite number of possible Hilbert series for $S/I$, and that there is a Zariski open subset in the space of coefficients of the $f_i$'s on which the Hilbert series of $S/I$ is one and the same and, in the appropriate sense, it is minimal among all possible Hilbert series, see below. We call algebras with this Hilbert series {\it generic}. There is a longstanding conjecture about this minimal Hilbert series formulated by the first author, see \cite{fr}. \begin{conjecture}[Fr\"oberg's Conjecture, 1985]\label{conj:fr} Let $f_1, \dots, f_r$ be generic forms of degrees $d_1, \dots, d_r$, respectively. Then the Hilbert series of the quotient algebra $R = S/(f_1,\ldots,f_r)$ is given by \begin{equation}\label{eq:RALF} \HF_R(t)=\left[\frac{\prod_{i=1}^r(1-t^{d_i})}{(1-t)^n}\right]_+. \end{equation} Here $[\sum_{i\ge0}a_iz^i]_+:=\sum_{i\ge0}b_iz^i$, with $b_i=a_i$ if $a_j\ge0$ for all $j\le i$ and $b_i=0$. In other words, $[\sum_{i\ge0}a_iz^i]_+$ is the truncation of a power series at its first non-positive coefficient. \end{conjecture} Conjecture~\ref{conj:fr} has been proven in the following cases: for $r\le n$ (easy exercise, since in this case $I$ is a complete intersection); for $n\le2$, \cite{fr}; for $n=3$, \cite{an3}, for $r=n+1$, which follows from \cite{st}. Additionally, in \cite{ho-la} it has been proven that \eqref{eq:RALF} is correct in the first nontrivial degree $\min_{i=1}^r(d_i+1)$. There are also other special results in the case $d_1=\dots =d_r$, see \cite{ni, au, fr-ho, mi-mi, ne}. We should also mention that \cite{fr-lu} contains a survey of the existing results on the generic series for various algebras and also it studies the (opposite) problem of finding the maximal Hilbert series for fixed parameters $(n,d_1,\ldots,d_r)$. \smallskip It is known that the actual Hilbert series of the quotient ring of any ideal with the same numerical parameters is lexicographically larger than or equal to the conjectured one. This fact implies that if for a given discrete data $(n, d_1, \ldots, d_r)$, one finds just a single example of an algebra with the Hilbert series as in \eqref{eq:RALF}, then Conjecture~\ref{conj:fr} is settled in this case. \medskip Although algebras with the minimal Hilbert series constitute a Zariski open set, they are hard to find constructively. We are only aware of two explicit constructions giving the minimal series in the special case $r=n+1$, namely R.~Stanley's choice $x_1^{d_1}, \ldots,x_n^{d_n}, (x_1+\cdots + x_n)^{d_{n+1}}$, and C. Gottlieb's choice $x_1^{d_1}, \ldots,x_n^{d_n}, h_{d_{n+1}}$, where $h_{d}$ denotes the complete homogeneous symmetric polynomial of degree $d$, (private communication). To the best of our knowledge, already in the next case $r=n+2$ there is no concrete guess about how to construct a similar example. There is however a substantial computer-based evidence pointing towards the possibility of replacing generic forms of degree $d$ by a product of generic forms of much smaller degrees. We present some problems and conjectures related to such pseudo-concrete constructions below. \subsection{Hilbert series of generic power ideals.} Differently from the situation occurring in R.~Stanley's result, if we consider ideals generated by more than $n+1$ powers of generic linear forms, there are known examples of $(n,d_1,...,d_r)$ for which algebras generated by powers of generic linear forms fail to have the Hilbert series as in \eqref{eq:RALF}. \medskip Recall that ideals generated by powers of linear forms are usually called {\it power ideals}. Due to their appearance in several areas of algebraic geometry, commutative algebra and combinatorics, they have been studied more thoroughly. In the next section, we will discuss their relation with the so-called {\it fat points}. (For a more extensive survey of power ideals, we refer to a nice paper by F. Ardila and A. Postnikov \cite{AP10}.) \medskip Studying Hilbert functions of generic power ideals, A.~Iarrobino formulated the following conjecture, usually referred to as the Fr\"oberg-Iarrobino Conjecture, see \cite{ia, Ch}. \begin{conjecture}[Fr\"oberg-Iarrobino Conjecture]\label{conj:fr-ia} Given generic linear forms $\ell_1, \ldots, \ell_r$ and a positive integer $d$, let $I$ be the power ideal generated by $\ell_1^d,\ldots,\ell_r^d$. Then the Hilbert function of $R = S/I$ is as in \eqref{eq:RALF}, except for the cases $(n,r) = (3,7), (3,8), (4,9), (5,14)$ and possibly for $r = n+2$ and $r = n+3$. \end{conjecture} This conjecture is still largely open. In \cite{fr-ho} R.F. and J. Hollman checked it for low degrees and low number of variables using the first version of the software package {\it Macaulay2}. In the last decades, some progress has been made in reformulation of Conjecture~\ref{conj:fr-ia} in terms of the ideals of {\it fat points} and {\it linear systems}. We will return to this topic in the next section. \subsection{Hilbert series of other classes of ideals} Computer experiments suggest that in order to always generically get the Hilbert function as in \eqref{eq:RALF} we need to replace power ideals by slightly less special ideals. \medskip For example, given a partition $\mu = (\mu_1,\ldots,\mu_k) \vdash d$, we call by a {\it $\mu$-power ideal} an ideal generated by forms of the type $({\bf l}_1^\mu,\ldots,{\bf l}_r^\mu)$, where ${\bf l}_i^\mu=l_{i,1}^{\mu_1}\cdots l_{i,k}^{\mu_k}$ and $l_{i,j}$'s are distinct linear forms. \begin{problem}\label{problem: F} For $\mu \neq (d)$, does a generic $\mu$-power ideal have the same Hilbert function as in \eqref{eq:RALF}? \end{problem} Performed computer experiments suggest a positive answer to the latter problem. L. Nicklasson has also conjectured that ideals generated by powers of generic forms of degree $\geq 2$ have the Hilbert series as in \eqref{eq:RALF}. \begin{conjecture}[\cite{ni}] \label{conj:nic} For generic forms $g_1,\ldots,g_r$ of degree $d>1$, the ideal $(g_1^k,\ldots,g_r^k)$ has the same Hilbert series as the one generated by $r$ generic forms of degree $dk$. \end{conjecture} It was observed in \cite[Theorem A.3]{LuOnReSh} that Conjecture \ref{conj:nic} implies Conjecture \ref{conj:main}, connecting the two first sections of the present paper. It was also shown that Conjecture \ref{conj:nic} holds in the case of binary form by specializing the $g_i$'€™s to be $d$-th powers of linear forms and applying the fact that generic power ideals in two variables have the generic Hilbert series \cite{GeSh}. The same idea gives a positive answer to Problem \ref{problem: F} in the case of binary forms, by specializing $l_{i,1} = \ldots = l_{i,k}$, for $i = 1,\ldots,r$. \subsection{Lefschetz properties of graded algebras} We say that a graded algebra $A$ has the {\it weak Lefschetz property} (WLP) (respectively, the {\it strong Lefschetz property} (SLP)) if the multiplication map $\times l:A_i\rightarrow A_{i+1}$ (respectively, $\times l^k:A_i\rightarrow A_{i+k}$) has the maximal rank, i.e., it is either injective or surjective, for a generic linear form $l$ and all $i$ (resp., for all $i$ and $k$). (For more references and open problems about the Lefschetz properties, see \cite{MiNa}.) \begin{problem} It has been conjectured that each complete intersection $R=S/(f_1,\ldots,f_n)$ satisfies the WLP and also the SLP, see \cite{HMNW}. Does the same hold for $R=S/(f_1,\ldots,f_r)$, with $f_1,\ldots,f_r$ being generic forms, and $r > n$? \end{problem} It follows from \cite{St} that monomial complete intersections satisfy the SLP. In \cite{BFL} the following situation has been studied. For the ring $T_{n,d,k}=S/(x_1^d,\ldots,x_n^d)^k$, it is shown that for $k\ge d^{n-2}$, $n\ge3$, $(n,d)\ne(3,2)$, $T_{n,d,k}$ fails the WLP. For $n=3$, there is an explicit conjecture when the WLP holds. Additionally, there is some information about $n>3$. \begin{problem} When are the WLP and the SLP true for $T_{n,d,k}$? \end{problem} We now introduce the concept of the \emph{$\mu$-Lefschetz properties}. Let $\mu=(\mu_1,\ldots,\mu_k)$ be a partition of $d$, i.e., $\sum_{i=1}^k\mu_i=d$. We say that an algebra has the {\it $\mu$-Lefschetz property} if $\times{\bf l}^\mu:A_i\rightarrow A_{i+d}$ has maximal rank for all $i$, where ${\bf l}^\mu=l_1^{\mu_1}\cdots l_k^{\mu_k}$, and $l_i$'s are generic linear forms. \begin{problem} For $R= S/(f_1,\ldots,f_r)$, where $f_1,\ldots,f_r$ are generic forms, does $R$ satisfy the $\mu$-Lefschetz property for all partitions $\mu$? \end{problem} \section{Symbolic powers} For a prime ideal $\wp$ in a Noetherian ring $R$, define its {\it $m$-th symbolic power} $\wp^{(m)}$ as $$\wp^{(m)}=\wp^mR_\wp\cap R.$$ It is the $\wp$-primary component of $\wp^m$. For a general ideal $I$ in $R$, its {\it $m$-th symbolic power} is defined as $I^{(m)}=\cap_{\wp\in{\rm Ass}(I)}(I^mR_\wp\cap R)$. \subsection{Hilbert functions of fat points.} Let $I_X$ be the ideal in $\bC[x_1,\ldots,x_n]$ defining a scheme of reduced points $X = P_1 + \ldots + P_s$ in $\mathbb P^{n-1}$, say $I_X = \wp_1 \cap \ldots \cap \wp_s$ where $\wp_i$ is the prime ideal defining the point $P_i$. Then, the $m$-th symbolic power $I^{(m)}$ is the ideal $I_X^{(m)} = \wp_1^m \cap \ldots \cap \wp_s^m$ which defines the scheme of {\it fat points} $X = mP_1 + \ldots + mP_s$. Ideals of $0$-dimensional schemes are classical objects of study since the beginning of the last century. Their Hilbert functions are of particular interest. Study of these ideals and calculation of their Hilbert functions can be often related to the so-called {\it polynomial interpolation problem}. Indeed, the homogeneous part of degree $d$ of the ideal $I_X$ is the space of hypersurfaces of degree $d$ in $\mathbb P^{n-1}$ passing through the $P_i$'s up to order $m-1$, i.e., the space of polynomials of degree $d$ whose partial differentials up to order $m-1$ vanish at every $P_i$. It is well-known that the Hilbert function of $0$-dimensional schemes is strictly increasing until it reaches the {\it multiplicity} of the scheme, see \cite[Theorem 1.69]{IK06}. Hence, since the degree of a $m$-fat point in $\mathbb P^{n-1}$ is ${n-1+m-1 \choose n-1}$, the expected Hilbert function is $$\HF_{S/I_X}(d) = \min\left\{{n-1+d \choose n-1}, s{n-1+m-1 \choose n-1}\right\}.$$ In the case of simple generic points, i.e., for $m=1$, it is known that the actual Hilbert function is as expected. In the case of double points ($m=2$), counterexamples were known since the end of the 19-th century. In 1995, after a series of important papers, J. Alexander and A. Hirschowitz proved that the classically known examples were the only counterexamples. For higher multiplicity, very little is known at present. In the case of projective plane, a series of equivalent conjectures have been given by B. Segre \cite{Seg61}, B. Harbourne \cite{Har86}, A. Gimigliano \cite{Gim87} and A. Hirschowitz \cite{Hir89}. These are known as the {\it SHGH-Conjecture}, see \cite{Har00} for a survey of this topic. \smallskip {\it Apolarity Theory} is a very useful tool in studying of the ideals of fat points and it is connecting all the algebraic stories we have told above. In particular, the following lemma is crucial. (We refer to \cite{IK06} and \cite{Ger96} for an extensive description of this issue.) \begin{Lemma}[Apolarity Lemma] Let $X = P_1+\ldots+P_s$ be a scheme of reduced points in $\mathbb P^{n-1}$ and let $L_{1},\ldots, L_{s}$ be linear forms in $\bC[x_0,\ldots,x_n]$ such that, for any $i$, the coordinates of $P_i$ are the coefficients of $L_i$. Then, for every $m\geq d$, $$ \HF_{S/I^{(m)}_X}(d) = \dim_{\bC}[(L_1^{m-d+1},\ldots,L_s^{m-d+1})]_d. $$ \end{Lemma} Using this statement we obtain that calculation of the Hilbert function of a scheme of fat points is equivalent to the calculation of the Hilbert function of the corresponding power ideal. In particular, Fr\"oberg-Iarrobino conjecture (Conjecture \ref{conj:fr-ia}) can be rephrased as a conjecture about the Hilbert function of ideals of generic fat points. \medskip Recently R.F. raised the question about what happens in case of the ideals of generic fat points in a multi-graded space. A point in multi-projective space $P \in \mathbb P^{n_1-1}\times\ldots\times \mathbb P^{n_t-1}$ is defined by a prime ideal $\wp$ in the multi-graded polynomial ring $S = \bC[x_{1,1},\ldots,x_{1,n_1};\ldots;x_{t,1},\ldots,x_{t,n_t}] = \bigoplus_{I \subset \bN^t} S_I$, where $S_I$ is the vector space of multi-graded polynomials of multi-degree $I = (i_1,\ldots,i_t) \in \bN^t$. A scheme of fat points $X = mP_1+\ldots+mP_s$ is the scheme associated with the multi-graded ideal $\wp_1^m\cap\ldots\cap\wp_s^m$. \begin{problem} Given a scheme of generic fat points $X \subset \mathbb P^{n_1-1}\times\ldots\times \mathbb P^{n_t-1}$, what is the multi-graded Hilbert function $\HF_{S/I_X}(I)$, for $I \in \bN^t$? \end{problem} This question was first considered by M. V. Catalisano, A. V. Geramita and A. Gimigliano who solved it in the case of double points, i.e., for $m=2$ in $\mathbb P^1 \times \ldots \times \mathbb P^1$. Recently, A.O. jointly with E. Carlini and M. V. Catalisano resolved the case of triple points ($m=3$) in $\mathbb P^1 \times \mathbb P^1$ and computed the Hilbert function for an arbitrary multiplicity except for a finite region in the space of multi-indices, see \cite{CCO17}. \subsection{Symbolic powers vs. ordinary powers.} As we mentioned above, if $I$ is the ideal defining a set $X$ of points, the $m$-th symbolic power of $I$ is the ideal of polynomials vanishing up to order $m-1$ at all points in $X$ or, in other words, the space of hypersurfaces which are singular at all points in $X$ up to order $m-1$. For this reason, symbolic powers are interesting from a geometrical point of view, but they are more difficult to study compared to the usual powers which carry less geometrical information. Hence, it is important to find relations between them. Observe that the inclusion $I^m \subset I^{(m)}$ is trivial. \smallskip {\it Containment problems} between the ordinary and the symbolic powers of ideals of points have been studied in substantial details. One particularly interesting question is to understand for which pairs of positive integers $(m,r)$, $I^{(m)} \subset I^r$. A very important result in this direction is the fact that, for any ideal $I$ of reduced points in $\mathbb P^n$ and any $r > 1$, we have $I^{(nr)}\subset I^r$. This statement was proven in \cite{ELS01} by L. Ein, R. Lazersfeld and K. Smith for characteristic $0$ and by M. Hochster and C. Huneke in positive characteristic, see \cite{HH02}. At present, the important question is whether the bound in the latter statement is sharp. In \cite{DSTG13}, M. Dumicki, T. Szemberg and H. Tutaj-Gasi\'nska provided the first example of a configuration of points such that $I^{(3)} \not\subset I^2$. (We refer to \cite{SS17} for a complete account on this topic.) \medskip In the recent paper \cite{GGSVT16}, F. Galetto, A. V. Geramita, Y. S. Shin and A. Van Tuyl defined the {\it $m$-th symbolic defect of an ideal} as the number of minimal generators of the quotient ideal $I^{(m)} / I^m$. If $I$ defines a set of general points in projective space, it was already known that $I^{(m)} = I^m$ if and only if $I$ is a complete intersection. Additionally, in \cite[Theorem 6.3]{GGSVT16} the authors characterize all cases of $s$ points in $\mathbb P^2$ having the $2$-nd symbolic defect equal to $1$. These cases are exactly $s = 3,5,7,8$. \begin{problem} For the ideal $I$ of $s$ general points in $\mathbb P^{n-1}$, what is the difference between the Hilbert series of the $m$-th symbolic power and the $m$-th ordinary power? \end{problem} \section{Miscellanea} \subsection{Hilbert series of numerical semigroup rings} Let $\mathcal{S}=\langle s_1,\ldots,s_k\rangle$ be a numerical semigroup, i.e. $\mathcal{S}$ consists of all linear combinations with non-negative integer coefficients of the positive integers $s_i$, and let $k[x^{s_1},\ldots,x^{s_k}]=k[\mathcal{S}]$ be the semigroup ring. The Hilbert series of $k[\mathcal{S}]$ is of the form $p(t)/q(t)$, where $p,q$ are polynomials with integer coefficients. A semigroup is called \emph {cyclotomic} if the polynomial $p(t)$ has all its roots in the unit circle (which in fact, implies that they lie on the unit circle). (Detailed information about numerical semigroups can be found in e.g., \cite{Ci}.) \begin{conjecture} $\mathcal{S}$ is cyclotomic if and only if $k[\mathcal{S}]$ is a complete intersection. \end{conjecture} \subsection{Non-negative forms} The next circle of problems is related to the celebrated article \cite{Hi} of D.~Hilbert and to a number of results formulated in \cite{CLR}. Denote by $P_{n,m}$ the set of all non-negative real forms, i.e., real homogeneous polynomials of (an even) degree $m$ in $n$ variables which never attain negative values; denote by $\Sigma_{n.m}\subseteq P_{n,m}$ the subset of non-negative forms which can be represented as sums of squares of real forms of degree $\frac{n}{2}$. (In \cite{Hi} D.~Hilbert proved that $\Delta_{n,m}=P_{n,m}\setminus\Sigma_{n,m}$ is non-empty unless the pair $(n,m)$ is of the form $(n,2)$, $(2,m)$ or $(4,3)$.) Finally, if $\Z(p)$ stands for the real zero locus of a real form $p$, denote by $B_{n,m}$ (resp. $B'_{n,m}$) the supremum of $|\Z(p)|$ over $p\in P_{n,m}$ such that $|\Z(p)|<\infty$ (resp. over $p\in\Sigma_{n,m}$ such that $|\Z(p)|<\infty$). In other words, $B(n,m)$ is the supremum of the number of zeros of non-degenerate forms under the assumption that all these roots are isolated (and similarly for $B'_{n,m}$). Obviously, $B'_{n,m}<B_{n,m}$. \medskip The following basic question was posed in \cite{CLR}. \begin{problem} Are $B_{n,m}$ and $B'_{n,m}$ finite for any pair $(n,m)$ with even $n$? \end{problem} In \cite{CLR} it was shown that the answer to this problem is positive for $m=2,3$ and for the pair $(4,4)$. Relatively recently, in \cite{Stu} the following upper bound for $B_{n,m}$ was established $$B_{n,m}\le 2\frac{(m-1)^{n+1}-1}{m-2}.$$ However this bound can not be sharp, as shown in \cite{Ko}. In case of $B'_{n,m}$, the following guess seems quite plausible and is proven for $m=3$. \begin{conjecture} For any given pair $(n,m)$ with even $n$, $B'_{n,m}=\left(\frac{n}{2}\right)^{m-1}$. \end{conjecture} For $B_{n,m}$, no similar guess is known, but some intriguing information is available in the case $m=3$, see \cite{CLR}. The following problem is related to the classical Petrovski-Oleinik upper bound on the number of real ovals of real plane algebraic curves. \begin{problem} Determine $\lim_{n\to\infty}\frac{B_{n,3}}{n^2}$. \end{problem} The latter limit exists and lies in the interval $\left[\frac{5}{18},\frac{1}{2}\right]$, see \cite{CLR}. \subsection{Polynomial generation} Let $p$ be a prime number and let $\mathbb{F}_p$ denote the field with $p$ elements. Consider the two maps $$\phi: \mathbb{F}_p[x_1,\ldots,x_n] \to \mathbb{F}_p[x_1,\ldots,x_n], f \mapsto \sum_{a \in Z(f)} x^a,$$ $$\psi: \mathbb{F}_p[x_1,\ldots,x_n] \to \mathbb{F}_p[x_1,\ldots,x_n], f \mapsto \sum_{a \in \mathbb{F}_p^n} f(a) x^a.$$ Here $x^a := x_1^{a_1} \cdots x_n^{a_n}$, where each $a_i$ is regarded as an integer, and $Z(f)$ is the zero locus of $f$ in $F_p^n$, i.e., $Z(f) := \{a \in \mathbb{F}_p^n \,|\, f(a) = 0\}$. When $p = 2$, then $\phi$ is a bijective map on the vector space of polynomials of degree at most one in each variable, and $\phi^4(f) = f$, see \cite{lu}. The map $\psi$, suggested by M. Boij, is a linear bijective map on the vector space of polynomials with degree at most $p-1$ in each variable, and when $p = 2$, these two maps are closely related in the sense that $\phi(f) = \psi(f) + \sum_{a \in \mathbb{F}_2^n} x^a.$ Consider now the case $n = 1$ and $p > 2$. The map $\phi$ is no longer a bijection, but the sequence $\phi(f),\phi^2(f),\ldots$ will eventually become periodic. It is an easy exercise to show that $0 \mapsto 1 + x + \cdots + x^{p-1} \mapsto x \mapsto 1 \mapsto 0$. When $p \leq 17$, this is the only period, i.e., $\phi(f)^{d(f)} = 0$ for some $d(f)$. For $p=71$, we have found a period of length two; $1 +x^{63} \mapsto x^{23} + x^{26} + x^{34} + x^{39} + x^{41} + x^{51} + x^{70} \mapsto 1 + x^{63}.$ One can show that the length of the period is always an even number, but it is not clear which even numbers that can occur as lengths of periods. \begin{problem} For $n=1$ and given $p$, what are the (lengths of the) possible periods of $\phi$? \end{problem} Let us now turn to the map $\psi$ and the case $n=1$. For $p =3$, $\psi^8(f) = f$ for all polynomials $f$ in $\mathbb{F}_3[x]$ of degree at most two. For $p = 5$, the least $i$ such that $\psi^{i} = {\rm Id}$ on the space of polynomials of degree at most four, is equal to $124$. For $p = 7$, the corresponding number is $1368$. \begin{problem} For $n = 1$ and given $p$, find the minimal positive integer $i$ such that $\psi^i$ is the identity map on the space of polynomials of degree at most $p-1$. \end{problem} \subsection{Exterior algebras} Let $f$ be a generic form of even degree in the exterior algebra $E$ over $\mathbb{C}$ with $n$ generators. Moreno and Snellman showed that the Hilbert series of $E/(f)$ is equal to the expected series $[(1+t)^d(1-t^d)]_+$, see \cite{M-S}. When the degree of $f$ is odd, we have $(f)\subseteq{\rm Ann}{f}$. This annihilator ideal shows an unexpected behaviour. The most striking case is when $(n,d) = (9,3)$. It turns out that $\dim_{\mathbb{C}}({\rm Ann}{f})_3 = 4$, see \cite{lu-ni}. At the same time a naive guess is that $({\rm Ann}{f})_3$ is spanned by $f$ only. Additionally, computer experiments suggest that $(f)$ and ${\rm Ann}{f}$ agree in low degrees. \begin{problem} Let $f$ be a form of odd degree $d$ in $E$. Is it true that $({\rm Ann}(f))_i = (f)_i$, for $i < (n-d)/2$? \end{problem} \medskip We finish our list of problems with the following conjecture stated in \cite{cr-lu-ne}, which connects the question about the Hilbert series of generic forms in the exterior algebra with the Hilbert series of power ideals in the commutative setting. \begin{conjecture} Let $f$ and $g$ be generic quadratic forms in $E$ and let $\ell_1$ and $\ell_2$ be two generic linear forms in $S$. Then the Hilbert series of $E/(f,g)$ is equal to the Hilbert series of $S/(x_1^2,\ldots,x_n^2,\ell_1^2,\ell_2^2)$ and is given by $1 + a(n,1) t + a(n,2) t^2 + \cdots + a(n,s)t^s + \cdots$, where $a(n,s)$ is the number of lattice paths inside the rectangle $(n+2-2s)\times (n+2)$ starting from the bottom-left corner and ending at the top-right corner by using only moves of two types: either $(x,y)\rightarrow (x+1,y+1) \textrm{\ or\ } (x-1,y+1)$. \end{conjecture} \medskip\noindent {\bf Acknowledgements.} The authors want to thank all the participants of the problem-solving seminar at Stockholm University for their contributions and patience. The fourth author is sincerely grateful to Dr.~Kh.~Khozhasov for pointing out references \cite{Stu, Ko}.
{ "timestamp": "2018-01-08T02:15:16", "yymm": "1801", "arxiv_id": "1801.01692", "language": "en", "url": "https://arxiv.org/abs/1801.01692", "abstract": "We present a number of questions in commutative algebra posed on the problem solving seminar in algebra at Stockholm University during the period Fall 2014 - Spring 2017.", "subjects": "Commutative Algebra (math.AC)", "title": "Algebraic stories from one and from the other pockets", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9838471670723236, "lm_q2_score": 0.8152324915965392, "lm_q1q2_score": 0.802064177362567 }
https://arxiv.org/abs/1804.02465
Reconstructing Point Sets from Distance Distributions
We address the problem of reconstructing a set of points on a line or a loop from their unassigned noisy pairwise distances. When the points lie on a line, the problem is known as the turnpike; when they are on a loop, it is known as the beltway. We approximate the problem by discretizing the domain and representing the $N$ points via an $N$-hot encoding, which is a density supported on the discretized domain. We show how the distance distribution is then simply a collection of quadratic functionals of this density and propose to recover the point locations so that the estimated distance distribution matches the measured distance distribution. This can be cast as a constrained nonconvex optimization problem which we solve using projected gradient descent with a suitable spectral initializer. We derive conditions under which the proposed distance distribution matching approach locally converges to a global optimizer at a linear rate. Compared to the conventional backtracking approach, our method jointly reconstructs all the point locations and is robust to noise in the measurements. We substantiate these claims with state-of-the-art performance across a number of numerical experiments. Our method is the first practical approach to solve the large-scale noisy beltway problem where the points lie on a loop.
\subsection{Related Work} In the noiseless case, Lemke and Werman \cite{Lemke:1988} addressed the turnpike problem via polynomial factorization. Namely, the polynomial $Q_\mathcal{D}(a) = N+\sum_{k=1}^K(a^{d_k}+a^{-d_k})$ is invariant to permutations of pairwise distances. If one can factorize it as $Q_\mathcal{D}(a)=R(a)R(a^{-1})$ where $R(a)=\sum_{n=1}^N a^{u_n}$, then the point locations can be read off from the exponents. When the distances are all integers, the factorization runs in a time that is polynomial in the degree of $Q(a)$ \cite{Lenstra1982}, which is the largest pairwise distance. However, this approach quickly becomes impractical, and is brittle in the presence of noise. The more practical backtracking algorithm by Skiena et al. \cite{Skiena:1990} produces a solution for typical instances in time $\mathcal{O}(n^2\log n)$. It progressively finds the assignment for the remaining largest unassigned distance in $\mathcal{D}$, and adopts the branch-and-bound search strategy to recover the point locations in a depth-first manner. However, there exist examples with exponential runtime \cite{Skiena:1990,ZhangExp:1994}. Abbas and Bahig \cite{Abbas2016} later demonstrated that some of the worst-case scenarios could be avoided by performing a breadth-first search instead. An alternative to clever combinatorial search is to formulate the problem as a binary integer program \cite{Miller:1960,IBARAKI197639,Papadimitriou:1982}, and then relax it to obtain a convex semidefinite program \cite{Dakic:2000}. One drawback of this scheme is that it is computationally infeasible for large-scale problems. In this paper we propose to relax the integer program to a constrained nonconvex optimization problem that can be solved efficiently using projected gradient descent with a spectral initializer. To address the noisy case where the turnpike problem becomes NP-hard \cite{Cieliebak:2004}, Skiena and Sundaram proposed a modification of the backtracking algorithm where an interval is associated with each recovered point to account for the uncertainty \cite{Skiena:1994}. As a consequence, the number of backtracking paths could grow exponentially large. Pruning can be performed on the paths when the relative errors in the distances are small; however, it requires careful adaptive tuning and could lead to no solution sometimes. Our approach naturally incorporates noise into the problem formulation, thus exhibiting better performance compared to the current state-of-the-art backtracking approach. We mention that the turnpike problem is also related to the problem of string reconstruction from substring compositions which arises in protein mass spectrometry \cite{Acharya2015StringRF,Bulteau:2014,LEE2013}. The advances presented here for the turnpike problem might inspire similar approaches to solve its string variant. The beltway problem is more difficult than the turnpike problem \cite{Skiena:1990, Lemke2003}. Due to the loop structure, it can no longer be formulated as a polynomial factorization problem. It is also impossible for the backtracking approach to rely on the remaining largest unassigned distance to find the point locations progressively \cite{Lemke2003}. For small problems, Fomin \cite{Fomin:2016:1,Fomin:2016:2} proposed to avoid an exhaustive search in the noiseless case by further removing the redundant distances from $\mathcal{H}$ sequentially, and later extended it to handle noisy measurements\cite{Fomin:2019:3}. To the best of our knowledge, our work in this paper offers an alternative by providing the first practical approach to solve the large-scale noisy beltway problem. \subsection{Uniqueness} One complication with the turnpike problem is that the solution is not necessarily unique (up to a relabeling of the points and up to a congruence). Fortunately, the solution to the uDGP in any dimension is known to be \textit{generically} unique, in the sense made precise in the form of the reconstructability tests for the point configurations by Boutin and Kemper in \cite[Theorem 2.6 and Proposition 2.11]{BOUTIN2004709}. For example, if the points are sampled iid from an absolutely continuous probability distribution, then almost surely the distance distribution specifies their geometry uniquely (up to a relabeling and a congruence). Boutin and Kemper worked with complete distance measurements. Gortler et al. \cite{GUGR2018} later relaxed the completeness assumption and only required the underlying graph to be generically globally rigid \cite{CGGR2010}. Under this sufficient condition, they proved that the reconstruction of a generic point configuration is unique. Importantly, beyond uniqueness, Boutin and Kemper \cite{BOUTIN2004709} showed that when the multiset $\mathcal{D}$ in the turnpike problem contains only distinct distances, there is a suitably defined neighborhood around each uniquely reconstructable point configuration such that all configurations within the neighborhood are also uniquely reconstructable, and the forward and backward mappings between the different distance multisets are continuous. To the best of our knowledge, there has not been much work on the uniqueness of beltway reconstructions. In the remainder of this paper, we will assume that the measured distances correspond to a uniquely reconstructable configuration. \subsection{Our Approach} The combinatorial turnpike problem can be formulated as an assignment problem \cite{Assignment:1957,Burkard:2009} or a general integer program (when the domain is discrete) \cite{Miller:1960,IBARAKI197639,Papadimitriou:1982}. Most of the prior approaches described in the above sections try to first find the correct assignments of the distances to pairs of points $\alpha(k)$, and then recover the point locations $u_n$. On the other hand, the approach by Daki\'c \cite{Dakic:2000} adopts the integer programming formulation where the point locations are represented by a binary vector in the noiseless case, and directly recovered via semidefinte programming (SDP). Assignments are then a byproduct of the process. We proceed along the line of integer programming to solve the turnpike problem and the beltway problem in section \ref{sec:main_turnpike} and \ref{sec:main_beltway} respectively. Instead of relaxing the integer program to a convex SDP, we relax it to a constrained minimization of a nonconvex objective, which leads to the proposed approach that is more efficient and suitable for large-scale problems. Importantly, measurement noise is naturally incorporated into our formulation by smoothing the target distance distribution. A key ingredient in the proposed approach is a suitably constructed initializer inspired by the spectral initialization strategy \cite{Netrapalli2015:RPAM,WF:2015}. We analyze the convergence of the projected gradient method to a global optimum. In order to have a fast method, we also propose a computationally efficient projection onto the relaxed constraint set. Both the turnpike and beltway problems can be formulated in similar ways. Starting with the easier turnpike problem, we shall present the proposed approach in detail in section \ref{sec:main_turnpike}, and then demonstrate how it can be adapted to solve the beltway problem as well in section \ref{sec:main_beltway}. Numerical experiments in section \ref{sec:exp} show that our method achieves state-of-the-art performances for the turnpike recovery, and is the first practical approach to solve the large-scale noisy beltway problem. We conclude this paper with a discussion of our results in Section \ref{sec:con}. The proofs of the formal results can be found in the Appendix. \subsection{Distance Distribution Matching} \label{subsec:ddm} \begin{figure*}[tbp] \centering \subfigure[]{ \label{fig:prob_approx_cont} \includegraphics[width=2.5in]{figures/prob_gaussian_convolve}} \subfigure[]{ \label{fig:prob_approx_disc} \includegraphics[width=2.5in]{figures/prob_gaussian_convolve_histogram}} \caption{(a) The approximated distribution $p(d)$ based on the distance multiset $\mathcal{D}$; (b) The discretized distance distribution $p(y)$ from $p(d)$.} \label{fig:prob_dist_discretization} \end{figure*} Depending on how the distance $d_k$ is measured in various applications, a variety of noise models for $w_k$ may be appropriate \cite{Oppenheim:2009,Vaseghi:2006}. Here we model $w_k$ as iid zero-mean Gaussian noise with variance $\xi^2$ \[w_k\sim\mathcal{N}(0,\xi^2)=\frac{1}{\sqrt{2\pi\xi^2}}\exp\left(-\frac{1}{2\xi^2}w_k^2\right)\,.\] Ideally, we would like to find a set of point locations so that the \emph{estimated} distance distribution matches the \emph{oracle} distance distribution $g(d)$ \begin{align} \label{eq:oracle_dist} g(d)=\frac{1}{K}\cdot\sum_{k=1}^{K}\mathcal{N}\left(d\ \left|\ s_k,\xi^2\right.\right) :=\frac{1}{K}\sum_{k=1}^K\frac{1}{2\pi\xi^2}\exp\left(-\frac{(d-s_k)^2}{2\xi^2}\right)\,. \end{align} Since $g(d)$ is in general unknown, we approximate it here using the following distribution $p(d)$ based on the measured distances in $\mathcal{D}$. \begin{align} \label{eq:approx_dist} p(d)=\frac{1}{K}\cdot\sum_{k=1}^{K}\mathcal{N}\left(d\ \left|\ d_k, \sigma^2\right.\right)\,, \end{align} where $\sigma^2$ should be tuned according to an a priori estimate of the level of noise in the data and the grid resolution. As shown in Fig. \ref{fig:prob_dist_discretization}, the distribution $p(d)$ is further discretized to $p(y)$ in order to perform distribution matching with respect to the quantized distance $y$. \begin{align} \label{eq:obs_dist} p(y)=\int_{(y-0.5)\Delta l}^{(y+0.5)\Delta l}p(d)\ \textnormal{d}d\,. \end{align} Similar to \eqref{eq:quad_form_p}, the estimated distribution $q(y)$ can also be expressed in terms of the solution ${\bm{z}}$ \begin{align} \label{eq:quad_form_q} q(y)=\frac{1}{K}\cdot{\bm{z}}^T{\bm{A}}_y{\bm{z}}\,, \end{align} where $z_m$ is the estimated (unnormalized) probability that a point is located at $l_m$. We can solve for it by minimizing the mean-squared error between $q(y)$ and $p(y)$ subject to the constraints in \eqref{eq:relaxed_constraint} and \eqref{eq:l1_constraint}: \begin{align} \label{eq:constrained_nonconvex} \begin{split} \min_{{\bm{z}}}\quad &f({\bm{z}})=\frac{1}{M}\sum_{y=0}^{M-1}\big(q(y)-p(y)\big)^2 \end{split}\\ \label{eq:relaxed_constraint} \textnormal{subject to}\quad &0\leq z_m\leq 1,\ \forall\ m\in\set{1,\cdots,M}\\ \label{eq:l1_constraint} &\|{\bm{z}}\|_1 = N\,. \end{align} \subsection{Extracting Point Locations from the Estimated Distribution} In general, the recovered vector ${\bm{z}}$ will not be supported on exactly $N$ indices. In the following we discuss how to extract the $N$ point location estimates when this is the case. \paragraph{Noiseless case.} If we assume that there is no measurement noise in $d_k$ and no quantization error in the quantized distance $y_k=\left\lfloor\frac{d_k}{\Delta l}\right\rceil$, the vector ${\bm{x}}$ is then binary: ${\bm{x}}\in\set{0,1}^M$. Suppose ${\bm{z}}^\dagger$ is one of the global optimizers of \eqref{eq:constrained_nonconvex} that is different from ${\bm{x}}$ and $f({\bm{z}}^\dagger)=0$. We then have from $q(y=0)=p(y=0)$ that \[ \|{\bm{z}}^\dagger\|_2^2={{\bm{z}}^\dagger}^T{\bm{A}}_0{\bm{z}}^\dagger = {\bm{x}}^T{\bm{A}}_0{\bm{x}}=\|{\bm{x}}\|_2^2=N\,. \] From \eqref{eq:l1_constraint}, we can get \begin{align} \label{eq:noiseless_global_opt} \|{\bm{z}}^\dagger\|_2^2=\|{\bm{z}}^\dagger\|_1=N\,. \end{align} If the $m$-th entry $z^\dagger_m\in(0,1)$, then $\|{\bm{z}}^\dagger\|_2^2<\|{\bm{z}}^\dagger\|_1$ which is in contradiction with \eqref{eq:noiseless_global_opt}. Hence $z^\dagger_m\notin(0,1)$, and the global optimizer is integer-valued, ${\bm{z}}^\dagger\in\set{0,1}^M$. The points are estimated at the segments that correspond to the $1$-entries in ${\bm{z}}^\dagger$. If the solution ${\bm{z}}$ is not a global optimizer, then ${\bm{z}}\in[0,1]^M$. The point locations can be extracted in the same way as in the noisy case which we describe next. \begin{figure*}[tbp] \centering \subfigure{ \label{fig:z_vec} \includegraphics[width=\textwidth]{figures/z_vec.png}}\\ \subfigure{ \label{fig:h_clustering} \includegraphics[width=\textwidth]{figures/h_clustering.png}} \caption{Illustration of agglomerative clustering for $N=5$. The agglomerative clustering produces $8$ clusters, only the centroids of the $5$ clusters with the highest weights are taken as the point locations.} \label{fig:hierarchical_clustering} \end{figure*} \paragraph{Noisy case.} In the noisy case we have ${\bm{x}}\in[0,1]^M$. The $m$-th entry $z_m$ of ${\bm{z}}$ is the \emph{estimated} probability that a point is located at the $m$-th segment $l_m$. Extracting $N$ point locations from ${\bm{z}}$ can be posed as a clustering problem. As illustrated in Fig. \ref{fig:hierarchical_clustering}, each $l_m$ is viewed as a cluster with the weight $z_m$. We can cluster the $M$ segments using the agglomerative clustering approach \cite{Rokach2005} summarized in Algorithm \ref{alg:agg}. The centroids of the $N$ clusters with the largest weights are taken as the estimated point locations. \begin{algorithm}[tbp] \caption{Extracting the point locations via agglomerative clustering } \label{alg:agg} \begin{algorithmic}[1] \REQUIRE The solution ${\bm{z}}$, the smallest distance between two different points $d_{\min}$. \STATE Treat each segment $l_m$ with a nonzero weight $\omega_m=z_m$ as one cluster $C_m=\set{l_m}$ \STATE Compute the centroid $c_m$ of every cluster $C_m\in\mathcal{C}=\set{C_1,C_2,\cdots}$ \WHILE{$|\mathcal{C}|>N$} \STATE Merge the two closest clusters\footnotemark $\set{C_i,C_j}$ with weights $\set{w_i<1,w_j<1}$ and centroids $\|c_i-c_j\|<d_{\min}$ into one cluster $C_i$ \STATE Update the weight $w_i$ and the centroid $c_i$ of the new cluster $C_i$ \IF{the clusters cannot be merged further} \STATE \textbf{break} \ENDIF \ENDWHILE \STATE {\bfseries Return} the set of centroids $\set{c_1,c_2,\cdots}$ \end{algorithmic} \end{algorithm} \footnotetext{Randomly pick a pair of clusters in case of a draw.} \subsection{Projected Gradient Descent} Let $\mathcal{S}$ denote the convex set defined by the constraints \eqref{eq:relaxed_constraint},\eqref{eq:l1_constraint}. Given a proper initialization ${\bm{z}}_0$, we propose to solve \eqref{eq:constrained_nonconvex} via the projected gradient descent method: \begin{align} \label{eq:pgd_update} {\bm{z}}_{t+1} = \mathscr{P}_\mathcal{S}\big({\bm{z}}_t-\eta\cdot\nabla f({\bm{z}}_t)\big)\,, \end{align} where $\eta>0$ is the step size, $\mathscr{P}_\mathcal{S}(\cdot)$ is the projection of the gradient descent update onto $\mathcal{S}$, and $\nabla f({\bm{z}}_t)$ is the gradient \begin{align} \nabla f({\bm{z}}_t) = \frac{2}{MK^2}\sum_{y=0}^{M-1}\left({\bm{z}}_t^T{\bm{A}}_y{\bm{z}}_t-{\bm{x}}^T{\bm{A}}_y{\bm{x}}\right)\cdot\left({\bm{A}}_y+{\bm{A}}_y^T\right){\bm{z}}_t\,, \end{align} where both $q(y)$ and $p(y)$ are replaced with their quadratic forms in \eqref{eq:quad_form_p} and \eqref{eq:quad_form_q}. An adaptive strategy can be used to determine some suitable step size $\eta>0$ to minimize the objective function. The proposed approach is finally summarized by Algorithm \ref{alg:pdp_pgd}. \begin{algorithm}[tbp] \caption{The noisy turnpike problem via projected gradient descent} \label{alg:pdp_pgd} \begin{algorithmic}[1] \REQUIRE the distance multiset $\mathcal{D}$, the number of points $N$, quantization step size $\Delta l$, gradient descent step size $\eta$, adaptive rate $\beta\in(0,1)$, convergence threshold $\epsilon$, the maximum number of iterations $T$. \STATE Compute the discrete approximated distribution $q(y)$ from $\mathcal{D}$ \STATE Compute the spectral initializer ${\bm{z}}_0$ \FOR{$t=\{0,1,\cdots,T\}$} \WHILE{true} \STATE Compute the projected gradient descent update ${\bm{z}}_{t+1}=\mathscr{P}_\mathcal{S}\big({\bm{z}}_t-\eta\cdot\nabla f({\bm{z}}_t)\big)$ \IF {$f({\bm{z}}_{t+1})\leq f({\bm{z}}_t)$} \STATE Increase the step size $\eta=\frac{1}{\beta}\cdot\eta$ \STATE \textbf{break} \ELSE \STATE Decrease the step size $\eta=\beta\cdot\eta$ \ENDIF \ENDWHILE \IF {$\frac{\|{\bm{z}}_{t+1}-{\bm{z}}_t\|_2}{\|{\bm{z}}_t\|_2}<\epsilon$} \STATE Convergence is reached, set ${\bm{z}}={\bm{z}}_{t+1}$ \STATE \textbf{break} \ENDIF \ENDFOR \STATE {\bfseries Return} ${\bm{z}}$ \end{algorithmic} \end{algorithm} \subsubsection{Spectral Initialization} \label{subsec:spec_init} \begin{figure*}[tbp] \centering \includegraphics[width=\textwidth]{figures/spec_init_1000} \caption{An example of the obtained spectral initializer ${\bm{z}}_0$. The entries corresponding to the neighbourhood of the true point locations (illustrated by vertical lines) in general have larger values, indicating higher confidence in those locations.} \label{fig:spec_init} \end{figure*} A suitable initialization is needed to solve the constrained noncovex problem in \eqref{eq:constrained_nonconvex} via projected gradient descent. Here we can borrow ideas from another problem with quadratic measurements, the phase retrieval problem \cite{Gerchberg:72,Fienup:82}. In phase retrieval, the task is to compute a complex signal ${\bm{v}}\in\mathbb{C}^M$ from its quadratic measurements of the form $\mu_i = |\langle{\bm{v}}, {\bm{a}}_i\rangle|^2$ for $1 \leq i \leq I$. Since $\mu_i = {\bm{v}}^* {\bm{a}}_i {\bm{a}}_i^* {\bm{v}}$, spectral initialization for phase retrieval is based on a weighted sum of the rank-1 measurement matrices ${\bm{a}}_i {\bm{a}}_i^*$. Namely, using matrix concentration results, Netrapalli et al. \cite{Netrapalli2015:RPAM} showed that the leading eigenvector of $\sum_{i=1}^I \mu_i {\bm{a}}_i {\bm{a}}_i^*$ is close to the the true ${\bm{v}}$. Similar arguments can be used for quadratic systems of full-rank random matrices \cite{QuadFR:2019}. In our formulation of the turnpike problem \eqref{eq:quad_form_p}, the rank-1 matrices ${\bm{a}}_i {\bm{a}}_i^*$ are replaced by ${\bm{A}}_y$ which are not necessarily PSD nor rank-1; they are also deterministic. Notwithstanding, we can use the spectral initialization strategy. As we shall see from the numerical experiments in Section \ref{sec:exp}, this strategy works well empirically, although a rigorous proof remains an open question. Let $\beta_y = \frac{p(y)\cdot K}{\|{\bm{A}}_y\|_F}$ and ${\bm{H}}_y=\frac{{\bm{A}}_y}{\|{\bm{A}}_y\|_F}$. We can rewrite \eqref{eq:quad_form_p} as \begin{align} \beta_y={\bm{x}}^T{\bm{H}}_y{\bm{x}}=\langle{\bm{H}}_y,\ {\bm{x}}\vx^T\rangle =: \langle {\bm{H}}_y,\ {\bm{X}}\rangle\,. \end{align} The set $\left\{{\bm{H}}_y,\ 0\leq y\leq M-1\right\}$ can be viewed as an orthonormal basis for the matrix subspace $\mathrm{span} \, \set{{\bm{H}}_1, \ldots {\bm{H}}_{M-1}}$, \begin{align} \langle {\bm{H}}_i,\ {\bm{H}}_j \rangle=\left\{\begin{array}{l} 1\\ 0 \end{array} \quad \begin{array}{l} \textnormal{if }i=j\\ \textnormal{if }i\neq j\, \end{array}. \right. \end{align} With this interpretation, $\beta_y$ becomes the expansion coefficient of ${\bm{X}}$ in the direction of ${\bm{H}}_y$. The least squares estimate of ${\bm{X}}$ is then \begin{align} \widehat{{\bm{X}}}=\sum_{y=0}^{M-1}\beta_y\cdot{\bm{H}}_y\,, \end{align} which is nothing but the orthogonal projection of ${\bm{X}}$ on the subspace spanned by the ${\bm{H}}_y$. Finally, we find the spectral initializer ${\bm{z}}_0$ so that ${\bm{z}}_0 {\bm{z}}_0^T$ is close to $\widehat{{\bm{X}}}$ in Frobenius norm subject to the constraint that $\|{\bm{z}}_0\|^2_2=N$. Let the spectral initializer ${\bm{z}}_0=\sqrt{N}{\bm{e}}_{\max}$, where $\|{\bm{e}}_{\max}\|_2=1$. We have \begin{align} \begin{split} {\bm{e}}_{\max}&=\mathrm{argmin}_{{\bm{e}}}\ \|\widehat{{\bm{X}}}-N{\bm{e}}\ve^T\|_F^2\\ &=\mathrm{argmin}_{{\bm{e}}}\ \|\widehat{{\bm{X}}}\|_F^2+N^2\|{\bm{e}}\ve^T\|_F^2-2N\langle\widehat{{\bm{X}}},{\bm{e}}\ve^T\rangle\\ &=\mathrm{argmax}_{{\bm{e}}}\ {\bm{e}}^T\widehat{{\bm{X}}}{\bm{e}}\,. \end{split} \end{align} \begin{itemize} \item When $\widehat{{\bm{X}}}$ is symmetric, ${\bm{e}}_{\max}$ is given by the leading singular vector of $\widehat{{\bm{X}}}$ that corresponds to the largest singular value. \item When $\widehat{{\bm{X}}}$ is not symmetric, we use the method of Lagrange multipliers and find the stationary points of the Lagrangian $\mathcal{L}({\bm{e}},\lambda)={\bm{e}}^T\widehat{{\bm{X}}}{\bm{e}}-\lambda({\bm{e}}^T{\bm{e}}-1)$. Setting the gradients to $0$, we have \begin{align} \left(\widehat{{\bm{X}}}+\widehat{{\bm{X}}}^T\right){\bm{e}}&=2\lambda{\bm{e}}\\ {\bm{e}}^T{\bm{e}}&=1 \end{align} The stationary points are given by the eigenvectors of $\widehat{{\bm{X}}}+\widehat{{\bm{X}}}^T$, with ${\bm{e}}_{\max}$ being the one that corresponds to the largest eigenvalue. In practice we find it via the power iteration. \end{itemize} \remove{ \begin{algorithm}[tbp] \caption{Spectral initialization} \label{alg:spec_init_max} \begin{algorithmic}[1] \REQUIRE the least square estimate $\widehat{{\bm{X}}}$. \STATE Initialize ${\bm{e}}_0$ with all $1$s. \FOR{$t=\{0,1,\cdots,T\}$} \STATE Compute ${\bm{e}}_{t+1}=\frac{\left(\widehat{{\bm{X}}}+\widehat{{\bm{X}}}^T\right){\bm{e}}_t}{\left\|\left(\widehat{{\bm{X}}}+\widehat{{\bm{X}}}^T\right){\bm{e}}_t\right\|_2}$ \IF {Convergence is reached} \STATE Set ${\bm{z}}_0=\sqrt{{\bm{e}}_{t+1}^T\widehat{{\bm{X}}}{\bm{e}}_{t+1}}\cdot{\bm{e}}_{t+1}$ \STATE \textbf{break} \ENDIF \ENDFOR \STATE {\bfseries Return} ${\bm{z}}_0$ \end{algorithmic} \end{algorithm} } \subsubsection{Efficient Projection onto the $l_1$-ball with Box Constraints} \begin{figure*}[tbp] \centering \includegraphics[height=2.5in]{figures/projection} \caption{ The gradient descent update ${\bm{z}}={\bm{z}}_t-\eta \nabla f({\bm{z}}_t)$ is projected back to the convex set $\mathcal{S}$.} \label{fig:projection} \end{figure*} As shown in Fig. \ref{fig:projection}, the gradient descent update ${\bm{z}}={\bm{z}}_t-\eta\cdot\nabla f({\bm{z}}_t)$ is projected back onto the convex set $\mathcal{S}$, which is the $l_1$-ball with box constrains defined by \eqref{eq:relaxed_constraint} and \eqref{eq:l1_constraint}. The projection is the solution to the following convex problem \begin{align} \label{eq:projection_box_constrains} \begin{split} \min_{\bm{s}}\quad&\frac{1}{2}\|{\bm{s}}-{\bm{z}}\|_2^2\\ \textnormal{subject to}\quad&0\leq s_m\leq 1,\ \forall\ m\in[M]\\ &\|{\bm{s}}\|_1=N\,. \end{split} \end{align} Duchi et al. \cite{projection06,Duchi:2008} proposed an efficient algorithm to compute the projection onto the $l_1$-ball when $s_m$ is only lower-bounded by $0$. Gupta et al. \cite{l1_box10, l1_box12} later extended that approach to handle projections with box constraints, when $s_m$ is both lower-bounded and upper-bounded. However, their approach is based on a sequential search for an optimal threshold $\kappa$, which is inefficient and cannot be parallelized for large-scale problems. Building on the work of \cite{projection06}, we address these issues by deriving a closed-form expression for the optimal $\kappa$ in \eqref{eq:kappa_compute} in terms of the entry index $r$ of a sorted ${\bm{s}}$. Specifically, the Lagrangian of \eqref{eq:projection_box_constrains} is: \begin{align} \mathcal{L}=\frac{1}{2}\|{\bm{s}}-{\bm{z}}\|_2^2+\kappa\left(\|{\bm{s}}\|_1-N\right)-{\boldsymbol\zeta}^T\cdot{\bm{s}}+{\boldsymbol\xi}^T\cdot({\bm{s}}-\boldsymbol 1)\,, \end{align} where $\kappa\in\mathbb{R}$ is a real Lagrange multiplier, ${\boldsymbol\zeta}\in\mathbb{R}_+^M$, ${\boldsymbol\xi}\in\mathbb{R}_+^M$ are the nonnegative Lagrange multipliers. Taking the subgradient of $\mathcal{L}$ w.r.t. ${\bm{s}}$, and setting it to $0$, we have \begin{align} \frac{\partial\mathcal{L}}{\partial s_m}=s_m-z_m+\kappa-\zeta_m+\xi_m=0\,. \end{align} Since $\mathcal{S}$ is a closed convex set, the projection solution ${\bm{s}}$ exists and is unique. We need to consider the following two cases. \begin{enumerate}[label={\arabic*)}] \item If the solution ${\bm{s}}$ contains only $[0,1]$-entries, there are $N$ entries in ${\bm{s}}$ that equal $1$, and their indices correspond to the top $N$ entries of ${\bm{z}}$. \item If at least one entry of ${\bm{s}}$ is between $0$ and $1$, the complementary slackness KKT condition indicates that when $0< s_m < 1$, the Lagrange multipliers $\zeta_m=\xi_m=0$. We then have: \begin{align} \label{eq:project_one} s_m=z_m-\kappa\quad\quad\textnormal{if } 0< s_m < 1\,. \end{align} The above \eqref{eq:project_one} gives us an efficient way to compute $s_m$ if it happens to be between $0$ and $1$: simply subtract the threshold $\kappa$ from $z_m$. In order to find the optimal solution ${\bm{s}}$, we need to compute $\kappa$ and identify the three types of entries of ${\bm{s}}$: those that equal $0$, those that equal $1$, and those that are between $0$ and $1$. We will make use of the following lemma from \cite{projection06} about the entries of ${\bm{s}}$ that equal $0$: \begin{lemma} \label{lemma:lower_bound} \cite {projection06} Let ${\bm{s}}$ be the optimal solution to the minimization problem in (\ref{eq:projection_box_constrains}). Let $i$ and $j$ be two indices such that $z_i>z_j$. If $s_i=0$ then $s_j$ must be $0$ as well. \end{lemma} Similarly, we can prove the following lemma about the entries of ${\bm{s}}$ that equal $1$ (the proof can be found in Appendix \ref{proof:lemma:upper_bound}). \begin{lemma} \label{lemma:upper_bound} Let ${\bm{s}}$ be the optimal solution to the minimization problem in (\ref{eq:projection_box_constrains}). Let $i$ and $j$ be two indices such that $z_i>z_j$. If $s_j=1$ then $s_i$ must be $1$ as well. \end{lemma} Since reordering of the entries of ${\bm{z}}$ does not change the value of (\ref{eq:projection_box_constrains}), and adding some constant to ${\bm{z}}$ does not change the solution of (\ref{eq:projection_box_constrains}), without loss of generality we can assume that the entries of ${\bm{z}}$ are all positive in a non-increasing order: $z_1\geq z_2\geq\cdots\geq z_M\geq N$. Lemma \ref{lemma:lower_bound} and \ref{lemma:upper_bound} imply that for the optimal solution ${\bm{s}}$: \begin{itemize} \item The entries of ${\bm{s}}$ are in a non-increasing order. \item The first $\rho$ entries of ${\bm{s}}$ satisfy $0<s_m\leq 1$; the rest of the entries are $0$s. \end{itemize} Since $\exists\ s_m\in(0,1)$, we have $\rho>N$ and that at most $N-1$ entries of ${\bm{s}}$ could equal $1$. Suppose the first $r-1$ entries of ${\bm{s}}$ are all $1$s, the following must hold for $1\leq r\leq N<\rho$ \begin{align} \label{eq:cst_r} &0<z_r-\kappa< 1\\ \label{eq:cst_rm1} &1\leq z_{r-1}-\kappa,\quad\textnormal{if }2\leq r\leq N<\rho\,. \end{align} \begin{algorithm}[tbp] \caption{Efficient projection onto the $l_1$-ball with box constraints} \label{alg:ep_box} \begin{algorithmic}[1] \STATE Shift ${\bm{z}}$ s.t. $z_m\geq N$, $\forall\ m\in\set{1,\cdots,M}$; and sort ${\bm{z}}$ in a non-increasing order. \FOR{$r=1:N$} \STATE Construct ${\bm{v}}$ out of ${\bm{z}}$ by removing the first $r-1$ entries, and compute $\rho_v$ according to \cite{Duchi:2008}:\\ $\rho_v=\max\left\{l\in[N-r+1]\,:\, v_l-\textstyle\frac{1}{l}\left(\textstyle\sum_{m=1}^lv_m-(N-r+1)\right)>0\right\} $ \STATE Compute $\kappa_v=\frac{1}{\rho_v}\left(\sum_{m=1}^{\rho_v}v_m-(N-r+1)\right)$ \STATE Check if $(\rho_v, \kappa_v)$ satisfy \eqref{eq:cst_r} by examining $\widehat{s}_r=z_r-\kappa_v$ \IF{$0<\widehat{s}_r< 1$} \IF{$r=1$} \STATE Set $\kappa=\kappa_v$, $\rho=\rho_v+r-1$ and \textbf{break} \ELSE \STATE Check if $(\rho_v, \kappa_v)$ satisfy \eqref{eq:cst_rm1} by examining $\widehat{s}_{r-1}=z_{r-1}-\kappa_v$ \IF{$\widehat{s}_{r-1}\geq 1$} \STATE Set $\kappa=\kappa_v$, $\rho=\rho_v+r-1$ and \textbf{break} \ENDIF \ENDIF \ELSE \STATE \textbf{continue} \ENDIF \ENDFOR \IF{$(r,\rho,\kappa)$ can be found} \STATE Compute ${\bm{s}}=\max\{{\bm{z}}-\kappa, 0\}$ followed by ${\bm{s}}=\min\{{\bm{s}},1\}$ \ELSE \STATE Compute ${\bm{s}}$ by setting the top $N$ entries of ${\bm{z}}$ to $1$ and the rest entries to $0$ \ENDIF \STATE {\bfseries Return} ${\bm{s}}$ \end{algorithmic} \end{algorithm} We can write the sum of ${\bm{s}}$ as follows: \begin{align} \sum_{m=1}^Ms_m=\sum_{m=1}^\rho s_m=(r-1)+\sum_{m=r}^\rho (z_m-\kappa)=N\,. \end{align} We then have: \begin{align} \label{eq:kappa_compute} \kappa=\frac{1}{\rho-r+1}\left(\sum_{m=r}^\rho z_m-(N-r+1)\right)\,. \end{align} Finally, we can write the minimizing ${\bm{s}}$ as \begin{align} \label{eq:min_s} {\bm{s}}=\left\{ \begin{array}{l} 1, \\ z_m-\kappa, \\ 0, \end{array} \quad \begin{array}{l} \textnormal{if }m\leq r-1\\ \textnormal{if }r\leq m\leq \rho\\ \textnormal{if }\rho+1\leq m\leq M\,. \end{array} \right. \end{align} If $r$ is known, we can find the value of $\rho$ efficiently using the approach in \cite{projection06,Duchi:2008}, and thus identify the three types of entries in ${\bm{s}}$. The threshold $\kappa$ and the solution ${\bm{s}}$ can be then computed using \eqref{eq:kappa_compute} and \eqref{eq:min_s} respectively. According to the following lemma (proved in Appendix \ref{proof:lemma:unique_r}), we can find $r$ by checking the integers in the set $\set{1,\ldots,N}$ one by one or in parallel until the computed $(\rho, \kappa)$ satisfy the two constraints \eqref{eq:cst_r} and \eqref{eq:cst_rm1}. \begin{lemma} \label{LEMMA:UNIQUE_R} If the solution ${\bm{s}}$ has at least one entry $s_m\in(0,1)$, there is one and only one $r\in\set{1,\ldots,N}$ that produces the $(\rho,\kappa)$ satisfying \eqref{eq:cst_r} and \eqref{eq:cst_rm1}. \end{lemma} \end{enumerate} In practice we do not know beforehand what the solution ${\bm{s}}$ is like. Given the uniqueness of the solution, we could start by trying to look for the right $(r,\rho,\kappa)$-values. If they can be found, ${\bm{s}}$ can then be computed using \eqref{eq:min_s}. If they can not be found, it means that ${\bm{s}}$ contains only $[0,1]$-entries and can be obtained straightforwardly. The proposed approach to perform efficient projection on the $l_1$-ball with box constraints is summarized in Algorithm \ref{alg:ep_box}. \subsection{Convergence Analysis} \label{subsec:cvg} In this section we study the convergence behavior of the proposed approach in the neighbourhood $\mathcal{E}(\tau)$ around a global optimizer ${\bm{x}}$. \[\mathcal{E}(\tau)=\set{{\bm{z}}\ |\ \|{\bm{z}}-{\bm{x}}\|_2<\tau\,,\ {\bm{z}}\in\mathcal{S}}\,,\] \paragraph{Noiseless recovery.} In this case we have ${\bm{x}}\in\set{0,1}^M$. There exists some $\tau>0$ that depends on ${\bm{x}}$, such that if the $t$-th iterate ${\bm{z}}_t\in\mathcal{E}(\tau)$, the projected gradient descent update in \eqref{eq:pgd_update} is guaranteed to converge linearly to ${\bm{x}}$. \begin{theorem} \label{THM:CVG_SED} In the noiseless case, let ${\bm{h}}={\bm{z}}_t-{\bm{x}}$ and ${\bm{B}}_y={\bm{A}}_y+{\bm{A}}_y^T$. If ${\bm{z}}_t$ satisfies \[ \|{\bm{h}}\|_2=\|{\bm{z}}_t-{\bm{x}}\|_2<\tau=\left(2-\frac{1}{q}\right)\cdot\sqrt{\frac{\lambda_{\bm{E}}}{4}}\,, \] where $q\in\big(\frac{1}{2},1\big)$ is some fixed constant and $\lambda_{\bm{E}}>0$ depends on the matrix ${\bm{E}}=\sum_{y=0}^{M-1}{\bm{B}}_y{\bm{x}}\vx^T{\bm{B}}_y^T$, the projected gradient descent update in \eqref{eq:pgd_update} converges linearly to ${\bm{x}}$, \begin{equation} \label{eq:converge} \|{\bm{z}}_{t+1}-{\bm{x}}\|_2<\mu^{\frac{1}{2}}\|{\bm{z}}_t-{\bm{x}}\|_2\,, \end{equation} where $\mu\in(0,1)$ and the step size $\eta>0$ both depend on $\left\{q,\ \tau,\ {\bm{z}}_t\right\}$. \end{theorem} As proved in Appendix \ref{proof:thm:cvg_sed}, the size of the convergence neighbourhood varies for different signals ${\bm{x}}$. According to Lemma \ref{LEMMA:E_MIN}, $\lambda_{\bm{E}}$ can be computed via the following convex program: \begin{align} \begin{split} \lambda_{\bm{E}} &= \min_{\widehat{{\bm{h}}}\in\mathcal{G}}\ \widehat{{\bm{h}}}^T{\bm{E}}\widehat{{\bm{h}}}=\min_{\widehat{{\bm{h}}}\in\mathcal{G}}\ \textstyle\sum_{y=0}^{M-1}\left(\widehat{{\bm{h}}}^T{\bm{B}}_y{\bm{x}}\right)^2\,, \end{split} \end{align} where $\widehat{{\bm{h}}}=\frac{{\bm{z}}-{\bm{x}}}{\|{\bm{z}}-{\bm{x}}\|_1}$, ${\bm{z}}\in\mathcal{S}$, ${\bm{z}}\neq{\bm{x}}$, and $\mathcal{G}$ is the convex set defined in Lemma \ref{LEMMA:E_MIN}. Note that $\lambda_{\bm{E}}>0$ in the noiseless case. To see this, let us assume that $\lambda_{\bm{E}}=0$. We then have \begin{align} \label{eq:noiseless_lambda_E_zero} ({\bm{z}}-{\bm{x}})^T{\bm{B}}_y{\bm{x}}=0,\ \forall\ y\in\set{0,\cdots,M-1}. \end{align} Using ${\bm{B}}_0=2{\bm{I}}$, where ${\bm{I}}$ is the identity matrix, we can get ${\bm{z}}^T{\bm{x}}={\bm{x}}^T{\bm{x}}=N$. Since ${\bm{z}}\in\mathcal{S}$ and ${\bm{x}}$ is a binary vector containing exactly $N$ $1$-entries, the vector ${\bm{z}}$ must equal ${\bm{x}}$ to ensure ${\bm{z}}^T{\bm{x}}=N$. This is in contradiction with the assumption that ${\bm{z}}\neq{\bm{x}}$\footnote{If ${\bm{z}}={\bm{x}}$, then we already have a global optimal solution.}. Hence $\lambda_{\bm{E}}\neq 0$. Since ${\bm{E}}$ is a positive semidefinite matrix, we can get that $\lambda_{\bm{E}}>0$ and $\tau>0$. \paragraph{Noisy recovery.} In this case we have ${\bm{x}}\in[0,1]^M$. From the proof of Theorem \ref{THM:CVG_SED}, we know that $\lambda_{\bm{E}}>0$ is a sufficient condition for the convergence neighbourhood to exist. We next show how the signal ${\bm{x}}$ plays a role in ensuring $\lambda_{\bm{E}}>0$. First, let us assume $\lambda_{\bm{E}}=0$. From \eqref{eq:noiseless_lambda_E_zero}, we have \begin{align} \label{eq:noisy_lambda_E_zero} {\bm{x}}^T{\bm{B}}_y{\bm{z}} = {\bm{x}}^T{\bm{B}}_y{\bm{x}},\quad\forall\ y\in\set{0,\cdots,M-1}. \end{align} Let ${\bm{s}}_y^T={\bm{x}}^T{\bm{B}}_y$ and ${\bm{S}}^T=\left[{\bm{s}}_0\ {\bm{s}}_1\ \cdots\ {\bm{s}}_{M-1}\right]$, we can rewrite \eqref{eq:noisy_lambda_E_zero} as follows: \begin{align} \label{eq:noisy_R_z_x} {\bm{S}}({\bm{z}}-{\bm{x}})={\bm{0}}\,. \end{align} Let $\mathcal{V}$ and $\mathcal{T}$ denote the following two sets \begin{align} \mathcal{V} &= \textnormal{Null}({\bm{S}})\ \backslash\ \set{{\bm{0}}}\\ \mathcal{T} &= \left\{{\bm{h}}\ |\ {\bm{h}}={\bm{z}}-{\bm{x}}\right\}\,. \end{align} If $\mathcal{V}\cap\mathcal{T}=\varnothing$, we can get that ${\bm{z}}={\bm{x}}$ according to \eqref{eq:noisy_R_z_x}. This is in contradiction with the assumption that ${\bm{z}}\neq{\bm{x}}$, hence $\lambda_{\bm{E}}\neq0$. Since ${\bm{E}}$ is positive semidefinite, we further have $\lambda_{\bm{E}}>0$ and $\tau>0$. We can see that $\mathcal{V}\cap\mathcal{T}=\varnothing$ is thus a sufficient condition for the convergence neighbourhood to exist. \subsection{Noiseless Partial Digestion on Real Genome Data} \begin{figure*}[t] \centering \includegraphics[height=2.5in]{figures/partial_digestion} \caption{Partial digestion of a DNA with the restriction enzyme when $N=5$.} \label{fig:partial_digestion} \end{figure*} As illustrated in Fig. \ref{fig:partial_digestion}, the DNA strands are partially digested at $N$ restriction sites by the restriction enzymes, producing all possible $\tbinom{N}{2}$ fragments. Here we perform experiments on \emph{E. Coli} K12 MG1655 genome data from the GenBank$^{\tiny{\text{\textregistered}}}$ assembly \cite{E:Coli:K12:MG1655}, which is a nucleotide sequence of length$=4,641,652$. Four letters \texttt{A}, \texttt{C}, \texttt{G}, \texttt{T} are used to represent the four nucleotide bases of the DNA strand \cite{NucleotidesSymbol:1985}. The list of restriction enzymes used in the experiments and the number of restriction sites (including the two ends \texttt{5'} and \texttt{3'} as dummy restriction sites) are shown in Table \ref{tab:restriction_enzyme}. Note that the recognition sequence could also be in a reverse order depending on which way the nucleotide sequence is read. Since there are four nucleotide bases that cannot be further digested, the unlabeled pairwise distances are all integers in this case, and the DNA sequence has a total of $M=4,641,653$ equally spaced possible locations for the restriction sites. Note that the matrix ${\bm{A}}_y$ has a simple structure and thus needs not be stored during computation. Using our proposed approach, we can reconstruct all the locations of the sites in Table \ref{tab:restriction_enzyme} successfully. \begin{table}[tbp] \caption{The list of restriction enzymes used in the partial digest experiments.} \label{tab:restriction_enzyme} \centering \begin{tabular}{llllll} \toprule Enzyme &Recognition sequence &$N$ &Enzyme &Recognition sequence &$N$ \\ \midrule \multirow{2}{*}{SmaI} &\texttt{5'---CCC \textcolor{magenta}{|} GGG---3'} & \multirow{2}{*}{$495$} &\multirow{2}{*}{BamHI} &\texttt{5'---G \textcolor{magenta}{|} GATCC---3'} & \multirow{2}{*}{$512$}\\ &\texttt{3'---GGG \textcolor{magenta}{|} CCC---5'} & & &\texttt{3'---CCTAG \textcolor{magenta}{|} G---5'} &\\ \bottomrule \end{tabular} \vspace{-3mm} \end{table} \subsection{Turnpike Recovery on Simulated Data} \label{sec:exp:turnpike} In the turnpike recovery experiments where the points are located on a line, we compare the proposed approach and the state-of-the-art backtracking approach by \cite{Skiena:1994} through simulated noisy recovery experiments. We first uniformly sample $N=10$ points from the interval $[0,1]$ with the minimum pairwise distance between two different points set to $d_{\min}=1e^{-2}$ and the maximum pairwise distance set to $d_{\max}=1$. The length $L$ of the line $\boldsymbol l$ thus equals $d_{\max}$. The quantization step is set to $\Delta l=1e^{-3}$ to balance the trade-off between reducing the quantization error and computational complexity, creating $M=\frac{L}{\Delta l}=1e^3$ possible locations for the $10$ points. The distance measurement $d_k$ is corrupted with white Gaussian noise, $w\sim\mathcal{N}(0,\xi^2)$. We control the noise level by varying the standard deviation of the noise: $\xi\in\{0,1e^{-3}, 3e^{-3}, 5e^{-3}, 7e^{-3}, 9e^{-3}\}$. The results obtained when $\xi=0$ correspond to the case where there is only quantization error and no measurement noise. For the proposed approach, the unlabeled pairwise distance measurements are collected and extended to form the multiset $\mathcal{D}$. As discussed in section \ref{subsec:ddm}, the parameter $\sigma$ in the approximated distribution $p(d)$ is unknown, and can be tuned in practice to obtain best performance. In the experiments, $\sigma$ is tuned in the interval $(0, d_{\min}=1e^{-2})$, producing multiple solutions corresponding to each $\sigma$. We shall choose the solution whose distance distribution is closest to the observed distance distribution in terms of the earth mover's distance \cite{EMD:2001}. The exact recovered point locations $\{\widehat{u}_1,\widehat{u}_2,\cdots,\widehat{u}_N\}$ are obtained using the aforementioned agglomerative clustering method in section \ref{subsec:ddm}. For each noise level specified by $\xi$, $100$ random simulations are performed and the number of correctly recovered points is recorded for each random run. For the backtracking approach, the search path for every distance $d_k$ is performed in an interval $[d_k-\Delta d,\ d_k+\Delta d]$. In order to make a fair comparison, we need to ensure that both approaches are evaluating the distance $d_k$ within roughly the same range. Here we choose $\Delta d=5\sigma_{\max}=5e^{-2}$, where $\sigma_{\max}$ is the largest $\sigma$ tuned by the proposed approach. We should note that the best results are obtained by choosing $\Delta d=1$, i.e. the maximum pairwise distance. However, this essentially becomes performing an exhaustive search over all possible paths, and is simply impractical when the number of points $N$ and the number of possible locations $M$ are large. Since there are only $10$ points to be recovered in this case, we also compute the solution obtained via the exhaustive search as a comparison, which corresponds to the best solution one can hope to achieve given noisy measurements. \begin{figure*}[tbp] \centering \subfigure{ \label{fig:n10_compare} \includegraphics[height=2.2in]{figures/compare_methods_10.pdf}} \subfigure{ \label{fig:n100_compare} \includegraphics[height=2.2in]{figures/compare_methods_100.pdf}} \caption{The distribution and the mean of the number of correctly recovered points across $100$ random runs in the \emph{``turnpike''} recovery experiments using the proposed approach ({\bfseries P}), the backtracking approach ({\bfseries B}) and the exhaustive search ({\bfseries E}). In each random run, $N$ points are uniformly sampled from the interval $[0,1]$. When $N=10$, the smallest distance between two different points is set to $d_{\min}=1e^{-2}$. When $N=100$, we set $d_{\min}=1e^{-4}$. The distances are further corrupted with white Gaussian noise $w\sim\mathcal{N}(0,\xi^2)$, where we control $\xi<d_{\min}$.} \label{fig:tp_compare} \end{figure*} \begin{figure*}[tbp] \centering \subfigure{ \label{fig:n10_compare_bw} \includegraphics[height=2.2in]{figures/compare_methods_10_bw.pdf}} \subfigure{ \label{fig:n100_compare_bw} \includegraphics[height=2.2in]{figures/compare_methods_100_bw.pdf}} \caption{The distribution and the mean of the number of correctly recovered points across $100$ random runs in the \emph{``beltway''} recovery experiments using the proposed approach ({\bfseries P}). In each random run, $N$ points are uniformly sampled from a loop of length $L=d_{\min}+d_{\max}$, where the largest pairwise distance $d_{\max}$ is set to $1$, the smallest distance $d_{\min}$ between two different points is set to $1e^{-2}$ when $N=10$ and $1e^{-4}$ when $N=100$. The distances are further corrupted with white Gaussian noise $w\sim\mathcal{N}(0,\xi^2)$, where we control $\xi<d_{\min}$.} \label{fig:bw_compare} \end{figure*} The recovered point locations can be matched to the true point locations efficiently using the Hungarian algorithm \cite{Hungarian:1955}. If the distance between a recovered location $\hat{u}_n$ and the true location $u_n$ is less than half the smallest pairwise distance $d_{\min}$, the recovery of the $n$-th point is considered to be a success. The recovery results when $N=10$ are shown in Fig. \ref{fig:tp_compare}: the distribution and the mean of the number of correctly recovered points across $100$ random runs are shown for each approach. We can see that the proposed approach is significantly more robust to noise compared to the conventional backtracking approach, and offers a competitive alternative to the search approach. In order to test how the two approaches are holding up against large-scale problems, we then uniformly sample $N=100$ points from the interval $[0,1]$ as before, with the minimum pairwise distance set to $d_{\min}=1e^{-4}$ and the maximum pairwise distance set to $d_{\max}=1$. The distance measurement $d_k$ is also corrupted with white Gaussian noise $w\sim\mathcal{N}(0,\xi^2)$, where $\xi\in\{0,1e^{-5}, 3e^{-5}, 5e^{-5}, 7e^{-5}, 9e^{-5}\}$. The quantization step is set to $\Delta l=1e^{-5}$, creating $M=\frac{L}{\Delta l}=1e^5$ possible locations for the $100$ points. For the proposed approach, the standard deviation $\sigma$ in the noise model is tuned in the interval $\sigma\in(0,d_{\min}=1e^{-4})$. For the backtracking approach, the tolerance threshold $\tau_d$ is chosen to be $\tau_d=5\sigma_{\max}=5e^{-4}$. Since $N$ and $M$ are much larger in this case, we are not able to perform an exhaustive search for comparison here. The recovery results when $N=100$ are shown in Fig. \ref{fig:tp_compare}. We can see that the proposed approach is more robust and has greater advantage over the backtracking approach for large-scale problems. When the noise level is high, the backtracking approach is not able to produce solutions, whereas the proposed approach can still produce partially correct solutions. \subsection{Beltway Recovery on Simulated Data} \label{sec:exp:beltway} We next use the proposed approach to perform the beltway recovery experiments where the points lie on a loop. To the best of our knowledge, our approach is the first practical approach that can solve the large-scale beltway problem efficiently. Note that the exhaustive search is impractical even when $N$ is small but $M$ is large \cite{Lemke2003}. Hence we only present the recovery results obtained using the proposed approach here. We uniformly sample $N$ points from a loop of length $L=d_{\min}+d_{\max}$, where $d_{\min}$ is the minimum distance between two different points and $d_{\max}$ is the maximum pairwise distance. When $N=10$, we set $d_{\min}=1e^{-2}$ and $d_{\max}=1$. The distance $d_k$ is also corrupted with a white Gaussian noise: $w_k\sim\mathcal{N}(0,\xi^2)$, where $\xi\in\set{0,1e^{-3}, 3e^{-3}, 5e^{-3},7e^{-3},9e^{-3}}$. The quantization step is set to $\Delta l=1e^{-3}$, creating $M=\frac{L}{\Delta l}=1.01e^3$ possible locations for the $10$-points case. When $N=100$, we set $d_{\min}=1e^{-4}$ and $d_{\max}=1$. The standard deviation of the white Gaussian noise is chosen from $\xi\in\set{0,1e^{-5},3e^{-5},5e^{-5},7e^{-5},9e^{-5}}$ as before, and the quantization step is set to $\Delta l=1e^{-5}$, creating $M=\frac{L}{\Delta l}=1.0001e^5$ possible locations for the $100$-points case. The recovery results are shown in Fig. \ref{fig:bw_compare}. We can see that the proposed approach is able to reconstruct all the point locations correctly when there is only quantization error and no measurement noise, i.e. $\xi=0$. When measurement noise is added to the distance, the proposed approach could still perform a partial reconstruction. \subsection{Comparison of Initialization Schemes} \begin{figure*}[tbp] \centering \subfigure{ \label{fig:init_tp_n100_compare} \includegraphics[width=.48\textwidth]{figures/compare_methods_init_100.pdf}} \subfigure{ \label{fig:init_bw_n100_compare} \includegraphics[width=.48\textwidth]{figures/compare_methods_init_100_bw.pdf}} \caption{The distribution and the mean of the number of correctly recovered points across $100$ random runs in the turnpike and beltway recoveries comparing the \emph{``three initialization schemes''}: the spectral initialization ({\bfseries S}), the random initialization ({\bfseries R}), and the uniform initialization ({\bfseries U}). In each random run, $N=100$ points are uniformly sampled from the interval $[0,1]$, with the smallest distance between two different points set to $d_{\min}=1e^{-4}$. The distances are further corrupted with white Gaussian noise $w\sim\mathcal{N}(0,\xi^2)$, where we control $\xi<d_{\min}$.} \label{fig:init_compare} \end{figure*} A spectral initialization scheme is adopted in the proposed approach to solve the nonconvex turnpike and beltway recoveries. It is meant to provide a good initializer that highlights the possible point locations. Here we put it to test and compare it with the other two initialization schemes, i.e. the ``random'' initialization and the ``uniform'' initialization. In the random initialization scheme, the entries of the initializer ${\bm{z}}_0$ are generated independently according to the white Gaussian distribution $\mathcal{N}(0,0.01)$. In the uniform initialization scheme, the entries of ${\bm{z}}_0$ are set to all ones. We should note that the initializers from all three schemes are projected to the convex set $\mathcal{S}$ defined by \eqref{eq:relaxed_constraint} and \eqref{eq:l1_constraint} before they can be used with the projected gradient descent. Following the same settings when $N=100$ as in sections \ref{sec:exp:turnpike} and \ref{sec:exp:beltway}, we perform simulated noisy turnpike and beltway recoveries using the three initialization schemes. The recovery results are shown in Fig. \ref{fig:init_compare}. For the turnpike recovery, the spectral initialization is more robust than the other two schemes. For the beltway recovery, the spectral initialization and the random initialization perform almost equally well, and they both perform better than the uniform initialization. \section{Proofs for Projected Gradient Descent} \remove{ \subsection{Proof of Lemma ~\ref{LEMMA:VALID_STEPSIZE}} \label{proof:lemma:valid_stepsize} Since ${\bm{z}}_t\in\mathcal{S}$, we can represent any point ${\bm{v}}$ in the convex set $\mathcal{S}$ by \begin{align} {\bm{v}}&={\bm{z}}_t+{\bm{r}}\,, \end{align} where $\sum_{m=1}^M r_m=0$, $-{z_t}_m\leq r_m\leq 1-{z_t}_m$, $\forall\ m\in\set{1,\cdots,M}$ and ${z_t}_m$ is the $m$-th entry of ${\bm{z}}_t$. \begin{enumerate}[label={\arabic*)}] \item To simplify notation, we use $f^\prime$ to denote the gradient $\nabla f({\bm{z}}_t)$. The projection of ${\bm{z}}_t-\eta\cdot f^\prime$ onto $\mathcal{S}$ can be obtained via the following two equivalent problems: \begin{align} \begin{split} \arg\min_{{\bm{v}}\in\mathcal{S}}\quad\|{\bm{v}}-({\bm{z}}_t-\eta\cdot f^\prime)\|_2^2 \quad \Longleftrightarrow \quad \arg\min_{{\bm{r}}}\quad\|{\bm{r}}+\eta\cdot f^\prime\|_2^2 \end{split} \end{align} The gradient $f^\prime$ can be decomposed into the sum of two vectors $f^\prime_1$ and $f^\prime_2$ as follows. Let $\boldsymbol 1$ denote a vector of length $M$ with all $1$s. We have \begin{align} f^\prime_1&=\frac{{\boldsymbol 1}^Tf^\prime}{M}\cdot\boldsymbol 1\\ f^\prime_2&=f^\prime-f^\prime_1\,. \end{align} Since ${\bm{r}}^Tf^\prime_1=0$, we get: \begin{align} \begin{split} \widehat{{\bm{r}}}=&\arg\min_{{\bm{r}}}\quad\|{\bm{r}}+\eta\cdot(f^\prime_1+f^\prime_2)\|_2^2\\ =&\arg\min_{{\bm{r}}}\quad{\bm{r}}^T{\bm{r}}+2\eta\cdot{\bm{r}}^Tf^\prime_2\,. \end{split} \end{align} Since we can always make ${\bm{r}}^T{\bm{r}}+2\eta\cdot{\bm{r}}^Tf^\prime_2=0$ by setting ${\bm{r}}=\boldsymbol 0$, we then have the following inequality for the minimizing $\widehat{{\bm{r}}}$: \begin{align} \label{eq:convergence_ieq1} \widehat{{\bm{r}}}^T\widehat{{\bm{r}}}+2\eta\cdot\widehat{{\bm{r}}}^Tf^\prime_2\leq 0\,. \end{align} Since $\widehat{{\bm{r}}}^T\widehat{{\bm{r}}}\geq 0$ and $\eta>0$, we must have: \begin{align} \label{eq:ineq_1} \widehat{{\bm{r}}}^Tf^\prime_2\leq 0\,. \end{align} \item Let ${\bm{B}}_y={\bm{A}}_y+{\bm{A}}_y^T$. We can compute $\nabla^2 f({\bm{z}})$ as follows \begin{align} \nabla^2 f({\bm{z}}) = \frac{2}{MK^2}\sum_{y=0}^{M-1}\left[{\bm{B}}_y{\bm{z}}\vz^T{\bm{B}}_y^T+\left({\bm{z}}^T{\bm{A}}_y{\bm{z}}-{\bm{x}}^T{\bm{A}}_y{\bm{x}}\right)\cdot{\bm{B}}_y\right]\,. \end{align} Since $\nabla^2 f({\bm{z}}_t)$ is bounded, $\nabla f({\bm{z}}_t)$ is locally Lipschitz continuous in a neighbourhood $\rho({\bm{z}}_t)$ around ${\bm{z}}_t$. In other words, $\forall {\bm{v}}\in\rho({\bm{z}}_t)$, the objective function $f({\bm{v}})$ has Lipschitz continuous gradient. The projection of ${\bm{z}}_t$ onto $\mathcal{S}$ can be computed as ${\bm{s}}={\bm{z}}_t+\widehat{{\bm{r}}}$. We can then obtain a quadratic upper bound on $f({\bm{s}})$. Suppose the Lipschitz constant is $\mathcal{L}$. We have \begin{align} \label{eq:ub_fvh} \begin{split} f({\bm{s}})&=f\left({\bm{z}}_t+\widehat{{\bm{r}}}\right)\\ &\leq f\left({\bm{z}}_t\right)+\widehat{{\bm{r}}}^Tf^\prime+\frac{\mathcal{L}}{2}\widehat{{\bm{r}}}^T\widehat{{\bm{r}}}\\ &=f\left({\bm{z}}_t\right)+\widehat{{\bm{r}}}^T(f^\prime_1+f^\prime_2)+\frac{\mathcal{L}}{2}\widehat{{\bm{r}}}^T\widehat{{\bm{r}}}\\ &=f\left({\bm{z}}_t\right)+\widehat{{\bm{r}}}^Tf^\prime_2+\frac{\mathcal{L}}{2}\widehat{{\bm{r}}}^T\widehat{{\bm{r}}}\,. \end{split} \end{align} If the step size $\eta$ is chosen such that $0<\eta\leq\frac{1}{\mathcal{L}}$, using \eqref{eq:ineq_1}, we have: \begin{align} \label{eq:ll_2_5} \frac{2}{\mathcal{L}}\cdot\widehat{{\bm{r}}}^Tf^\prime_2 \leq 2\eta\cdot\widehat{{\bm{r}}}^Tf^\prime_2 \,. \end{align} Plug \eqref{eq:ll_2_5} into (\ref{eq:convergence_ieq1}). We can get: \begin{align} \label{eq:ub_fvh_p2} \widehat{{\bm{r}}}^Tf^\prime_2+\frac{\mathcal{L}}{2}\widehat{{\bm{r}}}^T\widehat{{\bm{r}}}\leq 0\,. \end{align} Combining \eqref{eq:ub_fvh} and \eqref{eq:ub_fvh_p2}, we have $f({\bm{s}})\leq f({\bm{z}}_t)$. \end{enumerate} } \subsection{Proof of Lemma \ref{lemma:upper_bound}} \label{proof:lemma:upper_bound} Suppose that $s_i<1$. We construct a vector $\widetilde{{\bm{s}}}\in\mathbb{R}^M$ out of ${\bm{s}}$ by swapping the positions of $s_i$ and $s_j$ in ${\bm{s}}$, i.e. $\widetilde{s}_i=s_j$ and $\widetilde{s}_j=s_i$. $\widetilde{{\bm{s}}}$ also satisfies the constraints in (\ref{eq:projection_box_constrains}). Since $z_i>z_j$ and $s_j=1$, we then have: \begin{align} \begin{split} \|{\bm{s}}-{\bm{z}}\|_2^2-\|\tilde{{\bm{s}}}-{\bm{z}}\|_2^2&=(s_i-z_i)^2+(s_j-z_j)^2-(s_j-z_i)^2-(s_i-z_j)^2\\ &=2(1-s_i)(z_i-z_j)\\ &> 0\,. \end{split} \end{align} This is in contradiction with the fact that ${\bm{s}}$ is the minimizer of (\ref{eq:projection_box_constrains}). Hence $s_i$ must be $1$. \subsection{Proof of Lemma \ref{LEMMA:UNIQUE_R}} \label{proof:lemma:unique_r} Let $\mathcal{S}$ denote the convex set defined by the constraints $0\leq s_m\leq 1$, $\forall\ 1\leq m\leq M$ and $\|{\bm{s}}\|_1=N$. Note that the entries of ${\bm{z}}$ and ${\bm{x}}$ are in a non-increasing order. We will proceed in the following two steps: \begin{enumerate}[label={\arabic*)}] \item Since $\mathcal{S}$ is non-empty, the projection onto it exists, i.e. there is one $r\in\set{1,\cdots,N}$ that produces the $(\rho,\kappa)$ that satisfy $1\leq z_{r-1}-\kappa$ if $2\leq r\leq N<\rho$ and $0<z_r-\kappa<1$. In fact, since $\mathcal{S}$ is a closed convex set, the projection is also unique. \item Without loss of generality, suppose that there are two different $r_1<r_2\in\set{1,\cdots,N}$ that produce the two pairs $(\rho_1,\kappa_1|r_1)$ and $(\rho_2,\kappa_2|r_2)$ that satisfy the constraints (\ref{eq:cst_r}) and (\ref{eq:cst_rm1}). We have: \begin{gather*} r_1<r_2 \,\Rightarrow\, r_1\leq r_2-1 \,\Rightarrow\, z_{r_2-1}-\kappa_1\leq z_{r_1}-\kappa_1< 1 \,\Rightarrow\, z_{r_2-1}-1< \kappa_1\\ 1< z_{r_2-1}-\kappa_2 \,\Rightarrow\, \kappa_2< z_{r_2-1}-1\,. \end{gather*} Hence $\kappa_2<\kappa_1$. We further have: \begin{gather*} z_{\rho_1}-\kappa_1>0 \,\Rightarrow\, z_{\rho_1}>\kappa_1\\ z_{\rho_2+1}-\kappa_2\leq 0 \,\Rightarrow\, z_{\rho_2+1}\leq\kappa_2\\ z_{\rho_2+1}\leq\kappa_2<\kappa_1< z_{\rho_1} \,\Rightarrow\, z_{\rho_2+1} < z_{\rho_1}\,. \end{gather*} Hence $\rho_2+1>\rho_1\,\Rightarrow\,\rho_2\geq\rho_1$. \begin{enumerate} \item If $r_2\leq\rho_1$, we can find the upper bound for the sum of the first $\rho_1$ entries of ${\bm{s}}_1$: \begin{align} \label{eq:r_rho_neq_1} \begin{split} r_1-1+\sum_{m=r_1}^{\rho_1}(z_m-\kappa_1)=\,&r_1-1+\sum_{m=r_1}^{r_2-1}(z_m-\kappa_1)+\sum_{m=r_2}^{\rho_1}(z_m-\kappa_1)\\ <\,&r_1-1+\sum_{m=r_1}^{r_2-1}1+\sum_{m=r_2}^{\rho_1}(z_m-\kappa_1)\\ <\,&r_2-1+\sum_{m=r_2}^{\rho_1}(z_m-\kappa_2)\\ \leq\,&r_2-1+\sum_{m=r_2}^{\rho_2}(z_m-\kappa_2)\,. \end{split} \end{align} \item If $r_2>\rho_1$, we can compute: \begin{align} \label{eq:r_rho_neq_2} \begin{split} r_1-1+\sum_{m=r_1}^{\rho_1}(z_m-\kappa_1)\,\leq\, &r_1-1+\sum_{m=r_1}^{\rho_1}1\\ =\,&\rho_1\\ \leq\,&r_2-1\\ <\,&r_2-1+\sum_{m=r_2}^{\rho_2}(z_m-\kappa_2)\,. \end{split} \end{align} \end{enumerate} Let ${\bm{s}}_1,{\bm{s}}_2$ denote the solutions of \eqref{eq:projection_box_constrains} produced by $r_1,r_2$ respectively. Both (\ref{eq:r_rho_neq_1}) and (\ref{eq:r_rho_neq_2}) show that $\|{\bm{s}}_1\|_1<\|{\bm{s}}_2\|_1$. This is in contradiction with the assumption that they have the same $l_1$ norm, i.e. $\|{\bm{s}}_1\|_1=\|{\bm{s}}_2\|_1=N$. Hence $r_1=r_2$, there is only one $r\in\set{1,\cdots,N}$ that produces the $(\rho,\kappa)$ that satisfy the constraints (\ref{eq:cst_r}) and (\ref{eq:cst_rm1}). \end{enumerate} \section{Proofs for Convergence Analysis} \subsection{Lemma ~\ref{LEMMA:E_MIN}} \label{proof:lemma:e_min} \begin{lemma} \label{LEMMA:E_MIN} Let ${\bm{B}}_y={\bm{A}}_y+{\bm{A}}_y^T$, ${\bm{E}}=\sum_{y=0}^{M-1}{\bm{B}}_y{\bm{x}}{{\bm{x}}}^T{\bm{B}}_y^T$, $\mathcal{S}$ be the convex set defined by the constraints \eqref{eq:relaxed_constraint},\eqref{eq:l1_constraint}. The following problem is convex $\forall\ {\bm{z}}\in\mathcal{S}$, ${\bm{z}}\neq{\bm{x}}$: \begin{align} \label{eq:lambda_min_val} \begin{split} \lambda({\bm{E}}) &= \min_{{\bm{z}}\in\mathcal{S}, {\bm{z}}\neq{\bm{x}}}\,\frac{1}{\|{\bm{z}}-{\bm{x}}\|_1^2}({\bm{z}}-{\bm{x}})^T{\bm{E}}({\bm{z}}-{\bm{x}})\\ &=\min_{\widehat{{\bm{h}}}\in\mathcal{G}}\,\widehat{{\bm{h}}}^T{\bm{E}}\widehat{{\bm{h}}}\,, \end{split} \end{align} where $\lambda({\bm{E}})>0$ and $\widehat{{\bm{h}}}=\frac{{\bm{z}}-{\bm{x}}}{\|{\bm{z}}-{\bm{x}}\|_1}$, $\mathcal{G}$ is a convex set defined by the following constraints: \begin{align} \label{eq:hh_con_2} &\sum_{i=1}^M\widehat{h}_i=0\\ \label{eq:hh_con_3} &\widehat{h}_i\in[0,\,0.5]\quad\textnormal{if $x_i=0$}\\ \label{eq:hh_con_4} &\widehat{h}_i\in[-0.5,\,0]\quad\textnormal{if $x_i=1$}\\ \label{eq:hh_con_1} &\|\widehat{{\bm{h}}}\|_1={\bm{r}}^T\widehat{{\bm{h}}}=1\,, \end{align} where ${\bm{r}}\in\{-1,1\}^M$ depends on ${\bm{x}}$ and is defined as follows: \begin{align} r_i=\left\{ \begin{array}{l} 1\\ -1 \end{array} \quad \begin{array}{l} \textnormal{if $x_i=0$}\\ \textnormal{if $x_i=1$}\,. \end{array} \right. \end{align} \end{lemma} \begin{proof} Since ${\bm{E}}=\sum_{y=0}^{M-1}{\bm{B}}_y{\bm{x}}{{\bm{x}}}^\textrm{T}{\bm{B}}_y^\textrm{T}$, we can see that $\widehat{{\bm{h}}}^T{\bm{E}}\widehat{{\bm{h}}}=\sum_y\left(\widehat{{\bm{h}}}^T{\bm{B}}_y{\bm{x}}\right)^2\geq 0$, $\forall\ \widehat{{\bm{h}}}\in\mathbb{R}^M$. Hence ${\bm{E}}$ is positive-semidefinite. We define the following set $\mathcal{H}$: \begin{align} \label{eq:H_set_def} \mathcal{H}=\left\{\widehat{{\bm{h}}}\,\left|\,\widehat{{\bm{h}}}=\frac{1}{\|{\bm{z}}-{\bm{x}}\|_1}({\bm{z}}-{\bm{x}}),\quad\forall {\bm{z}}\in\mathcal{S}\,,{\bm{z}}\neq{\bm{x}}\right.\right\}\,. \end{align} where $\mathcal{S}$ is the convex set defined by \eqref{eq:relaxed_constraint} and \eqref{eq:l1_constraint}. We then have \begin{align} \begin{split} \lambda({\bm{E}}) &= \min_{{\bm{z}}\in\mathcal{S}, {\bm{z}}\neq{\bm{x}}}\,\frac{1}{\|{\bm{z}}-{\bm{x}}\|_1^2}({\bm{z}}-{\bm{x}})^T{\bm{E}}({\bm{z}}-{\bm{x}})\\ & = \min_{\widehat{h}\in\mathcal{H}}\,\widehat{{\bm{h}}}^T{\bm{E}}\widehat{{\bm{h}}}\,. \end{split} \end{align} \begin{enumerate}[label={\arabic*)}] \item We first prove that $\mathcal{H}$ is a convex set. Let $\widehat{{\bm{h}}}^{(1)},\widehat{{\bm{h}}}^{(2)}\in\mathcal{H}$. We have $\textnormal{sign}(\widehat{h}_i^{(1)})=\textnormal{sign}(\widehat{h}_i^{(2)})$ if $\widehat{h}_i^{(1)}\neq0,\ \widehat{h}_i^{(2)}\neq0$. Let $\widehat{{\bm{h}}}^{(3)}=(1-\rho)\widehat{{\bm{h}}}^{(1)}+\rho\widehat{{\bm{h}}}^{(2)}$, where $\rho\in(0,1)$. We have: \begin{align} \label{eq:l1_norm_z3} \begin{split} \|\widehat{{\bm{h}}}^{(3)}\|_1 & = \left\|(1-\rho)\widehat{{\bm{h}}}^{(1)}+\rho\widehat{{\bm{h}}}^{(2)}\right\|_1\\ &=\sum_{i=1}^M\left|(1-\rho)\widehat{h}_i^{(1)}+\rho \widehat{h}_i^{(2)}\right|\\ &=\sum_{i=1}^M\left|(1-\rho)\widehat{h}_i^{(1)}\right|+\left|\rho \widehat{h}_i^{(2)}\right|\\ &=(1-\rho)\left\|\widehat{{\bm{h}}}^{(1)}\right\|_1+\rho\left\|\widehat{{\bm{h}}}^{(2)}\right\|_1=1\,. \end{split} \end{align} Let $\nu_1=\frac{1-\rho}{\left\|{\bm{z}}^{(1)}-{\bm{x}}\right\|_1}$, $\nu_2=\frac{\rho}{\left\|{\bm{z}}^{(2)}-{\bm{x}}\right\|_1}$. We have \begin{align} \begin{split} \widehat{{\bm{h}}}^{(3)}&=(1-\rho)\widehat{{\bm{h}}}^{(1)}+\rho \widehat{{\bm{h}}}^{(2)}\\ &=\frac{1-\rho}{\left\|{\bm{z}}^{(1)}-{\bm{x}}\right\|_1}\left({\bm{z}}^{(1)}-{\bm{x}}\right)+\frac{\rho}{\left\|{\bm{z}}^{(2)}-{\bm{x}}\right\|_1}\left({\bm{z}}^{(2)}-{\bm{x}}\right)\\ &=\left(\nu_1+\nu_2\right)\left(\frac{\nu_1}{\nu_1+\nu_2}{\bm{z}}^{(1)}+\frac{\nu_2}{\nu_1+\nu_2}{\bm{z}}^{(2)}-{\bm{x}}\right)\\ &=(\nu_1+\nu_2)({\bm{z}}^{(3)}-{\bm{x}})\,. \end{split} \end{align} Using \eqref{eq:l1_norm_z3}, we can see that $\nu_1+\nu_2=\frac{1}{\|{\bm{z}}^{(3)}-{\bm{x}}\|_1}$. Since ${\bm{z}}^{(1)},{\bm{z}}^{(2)}\in\mathcal{S}$, we have ${\bm{z}}^{(3)}\in\mathcal{S}$. We have shown that $\widehat{{\bm{h}}}^{(3)}$ can be written in the same form given in \eqref{eq:H_set_def} and thus belongs to $\mathcal{H}$. \begin{align} \widehat{{\bm{h}}}^{(3)}=\frac{1}{\|{\bm{z}}^{(3)}-{\bm{x}}\|_1}({\bm{z}}^{(3)}-{\bm{x}})\,. \end{align} Hence $\widehat{{\bm{h}}}^{(3)}\in\mathcal{H}$, and $\mathcal{H}\subset\mathbb{R}^M$ is a convex set. Minimizing $\widehat{{\bm{h}}}^\textrm{T}{\bm{E}}\widehat{{\bm{h}}}$ with respect to $\widehat{{\bm{h}}}\in\mathcal{H}$ is a convex problem. \item We next prove that $\lambda({\bm{E}})$ in \eqref{eq:lambda_min_val} is strictly positive. If $\widehat{{\bm{h}}}^T{\bm{E}}\widehat{{\bm{h}}}=\sum_y\left(\widehat{{\bm{h}}}^T{\bm{B}}_y{\bm{x}}\right)^2=0$, we have $({\bm{z}}-{\bm{x}})^T{\bm{B}}_y{\bm{x}}=0$, $\forall\ y\in\set{0,\cdots,M-1}$. When $y=0$, ${\bm{B}}_0=2{\bm{I}}$, where ${\bm{I}}$ is the identity matrix, we get ${\bm{z}}^T{\bm{x}}={{\bm{x}}}^\textrm{T}{\bm{x}}=N$. Since ${\bm{z}}\in\mathcal{S}$ and ${\bm{x}}\in\set{0,1}^M$ in the noiseless case, we have ${\bm{z}}={\bm{x}}$. This is in contradiction with the assumption ${\bm{z}}\neq{\bm{x}}$, hence $\widehat{{\bm{h}}}^T{\bm{E}}\widehat{{\bm{h}}}>0$, $\forall\ \widehat{{\bm{h}}}\in\mathcal{H}$. \item We finally prove that $\mathcal{H}$ and a new set $\mathcal{G}$ defined by the following constraints are the same: \begin{align} \label{eq:hh_con_2} &\sum_{i=1}^M\widehat{h}_i=0\\ \label{eq:hh_con_3} &\widehat{h}_i\in[0,\,0.5]\quad\textnormal{if $x_i=0$}\\ \label{eq:hh_con_4} &\widehat{h}_i\in[-0.5,\,0]\quad\textnormal{if $x_i=1$}\\ \label{eq:hh_con_1} &\|\widehat{{\bm{h}}}\|_1={\bm{r}}^T\widehat{{\bm{h}}}=1\,, \end{align} where ${\bm{r}}\in\{-1,1\}^M$ is defined as follows: \begin{align} r_i=\left\{ \begin{array}{l} 1\\ -1 \end{array} \quad \begin{array}{l} \textnormal{if $x_i=0$}\\ \textnormal{if $x_i=1$}\,. \end{array} \right. \end{align} \begin{itemize} \item It is easy to verify that if $\widehat{{\bm{h}}}\in\mathcal{H}$, \eqref{eq:hh_con_2} and \eqref{eq:hh_con_1} hold. Since $z_i\in[0,1]$ and $x_i\in\{0,1\}$, if $x_i=0$, $\widehat{h}_i\geq 0$; if $x_i=1$, $\widehat{h}_i\leq 0$. On the other hand, if $|\widehat{h}_i|>0.5$, from \eqref{eq:hh_con_2} we have $\sum_{j\neq i}|\widehat{h}_j|\geq|\sum_{j\neq i}\widehat{h}_j|=|-\widehat{h}_i|>0.5$. This means that $\|\widehat{{\bm{h}}}\|_1=|\widehat{h}_i|+\sum_{j\neq i}|\widehat{h}_j|>1$, which contradicts \eqref{eq:hh_con_1}. Hence $|\widehat{h}_i|\leq 0.5$, \eqref{eq:hh_con_3} and \eqref{eq:hh_con_4} hold. This proves that $\widehat{{\bm{h}}}\in\mathcal{G}$. \item If $\widehat{{\bm{h}}}\in\mathcal{G}$, we can construct such a $\widehat{{\bm{z}}}={\bm{x}}+\widehat{{\bm{h}}}$. It is easy to verify that $\widehat{{\bm{z}}}\in\mathcal{S}$ and $\|\widehat{{\bm{z}}}-{\bm{x}}\|_1=\|\widehat{{\bm{h}}}\|_1=1$. Hence $\widehat{{\bm{h}}} = \frac{1}{\|\widehat{{\bm{z}}}-{\bm{x}}\|_1}(\widehat{{\bm{z}}}-{\bm{x}})\in\mathcal{H}$. \end{itemize} \end{enumerate} Computing $\lambda({\bm{E}})=\min_{\widehat{{\bm{h}}}\in\mathcal{G}}\widehat{{\bm{h}}}^T{\bm{E}}\widehat{{\bm{h}}}\,>0$ is thus a convex problem, and can be efficiently solved via quadratic programming. \end{proof} \subsection{Proof of Theorem ~\ref{THM:CVG_SED}} \label{proof:thm:cvg_sed} Let ${\bm{B}}_y={\bm{A}}_y+{\bm{A}}_y^T$. The objective function $f({\bm{z}})$ in \eqref{eq:constrained_nonconvex} can be written as \begin{align} f({\bm{z}}) = \frac{1}{4MK^2}\sum_{y=0}^{M-1}\left({\bm{z}}^T{\bm{B}}_y{\bm{z}}-{\bm{x}}^T{\bm{B}}_y{\bm{x}}\right)^2\,. \end{align} The gradient $\nabla f({\bm{z}})$ is \begin{align} \begin{split} \nabla f({\bm{z}}) &= \frac{1}{MK^2}\sum_{y=0}^{M-1}{\bm{B}}_y{\bm{z}}\cdot\left({\bm{z}}^T{\bm{B}}_y{\bm{z}}-{\bm{x}}^T{\bm{B}}_y{\bm{x}}\right)\\ &= \frac{1}{MK^2}\sum_{y=0}^{M-1}{\bm{B}}_y{\bm{z}}\cdot({\bm{z}}-{\bm{x}})^T{\bm{B}}_y({\bm{z}}+{\bm{x}})\,. \end{split} \end{align} In order to prove the conditions under which the projected gradient descent updates converge to a global optimum, we first establish the conditions on the gradient descent updates. \paragraph{Step 1:} When the distance between the current solution ${\bm{z}}_t$ and a global optimum ${\bm{x}}$ is less than some $\tau>0$, i.e. $\|{\bm{z}}_t-{\bm{x}}\|_2<\tau$, we would like to show that the gradient descent update ${\bm{z}}_t-\eta\nabla f({\bm{z}}_t)$ converges linearly to ${\bm{x}}$. In other words, we need to prove the following \eqref{eq:diff} is less than $0$ for some $\mu\in(0,1)$ and $\eta>0$: \begin{align} \label{eq:diff} \begin{split} \left\|{\bm{z}}_t-\eta\nabla f({\bm{z}}_t)-{\bm{x}} \right\|^2_2-\mu\|{\bm{z}}_t-{\bm{x}}\|^2_2 =\eta^2\|\nabla f({\bm{z}}_t)\|_2^2-2\eta\langle{\bm{z}}_t-{\bm{x}},\, \nabla f({\bm{z}}_t)\rangle+(1-\mu)\|{\bm{z}}_t-{\bm{x}}\|_2^2\,. \end{split} \end{align} We then have: \begin{align} \label{eq:first_term_lb} \begin{split} \|\nabla f({\bm{z}}_t) \|_2^2& = \frac{1}{K^4}\left\|\frac{1}{M}\sum_y{\bm{B}}_y{\bm{z}}_t\cdot({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y({\bm{z}}_t+{\bm{x}})\right\|_2^2\\ &\leq \frac{1}{MK^4}\sum_y\left\|{\bm{B}}_y{\bm{z}}_t\cdot({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y({\bm{z}}_t+{\bm{x}})\right\|_2^2\\ &= \frac{1}{MK^4}\sum_y\left\|{\bm{B}}_y{\bm{z}}_t\right\|_2^2\cdot\left(({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y({\bm{z}}_t+{\bm{x}})\right)^2\\ &\leq \frac{1}{MK^4}\sum_y\sigma_{\max}^2\left({\bm{B}}_y\right)\|{\bm{z}}_t\|_2^2\cdot\left(({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y({\bm{z}}_t+{\bm{x}})\right)^2\\ &\leq \frac{4}{MK^4}\|{\bm{z}}_t\|_2^2\sum_y\left(({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y({\bm{z}}_t+{\bm{x}})\right)^2\\ &= \frac{16}{K^2}\|{\bm{z}}_t\|_2^2\cdot f({\bm{z}}_t)\,, \end{split} \end{align} where $\sigma_{\max}^2\left({\bm{B}}_y\right)\leq 4$, $\forall\ y=\{0,1,\cdots,M-1\}$ according to the Schur's bound \cite{Schur1911}. Using the Cauchy-Schwarz inequality, we also have that \begin{align} \label{eq:second_term} \begin{split} \langle{\bm{z}}_t-{\bm{x}},\, \nabla f({\bm{z}}_t)\rangle&=\frac{1}{MK^2}\sum_y({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y{\bm{z}}_t\cdot({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y({\bm{z}}_t+{\bm{x}})\\ &=4f({\bm{z}}_t)-\frac{1}{MK^2}\sum_y({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y({\bm{z}}_t+{\bm{x}})\cdot({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y{\bm{x}}\\ &\geq 4f({\bm{z}}_t)-\sqrt{4f({\bm{z}}_t)}\sqrt{\frac{1}{MK^2}\sum_y\left(({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y{\bm{x}}\right)^2}\,. \end{split} \end{align} We proceed by further lower-bounding the above \eqref{eq:second_term}. Let ${\bm{h}}={\bm{z}}_t-{\bm{x}}$. For some $\frac{1}{2}<q<1$, we have \begin{align} \label{eq:second_term_lb1} \begin{split} &q^2\sum_y\left(({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y({\bm{z}}_t+{\bm{x}})\right)^2-\sum_y\left(({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y{\bm{x}}\right)^2\\ =\ &q^2\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y({\bm{h}}+2{\bm{x}})\right)^2-\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{x}}\right)^2\\ =\ &q^2\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{h}}\right)^2+4q^2\sum_y{\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{h}}\cdot{\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{x}}+(4q^2-1)\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{x}}\right)^2\\ \geq \ &q^2\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{h}}\right)^2-4q^2\sqrt{\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{h}}\right)^2}\sqrt{\sum_y\left({\bm{h}}^T{\bm{B}}_y{\bm{x}}\right)^2}+(4q^2-1)\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{x}}\right)^2\\ =\ &\left(q\sqrt{\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{h}}\right)^2}-2q\sqrt{\sum_y\left({\bm{h}}^T{\bm{B}}_y{\bm{x}}\right)^2}\right)^2-\left(\sqrt{\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{x}}\right)^2}\right)^2\\ =\ &\left(q\sqrt{\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{h}}\right)^2}-(2q-1)\sqrt{\sum_y\left({\bm{h}}^T{\bm{B}}_y{\bm{x}}\right)^2}\right)\left(q\sqrt{\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{h}}\right)^2}-(2q+1)\sqrt{\sum_y\left({\bm{h}}^T{\bm{B}}_y{\bm{x}}\right)^2}\right)\,. \end{split} \end{align} To make \eqref{eq:second_term_lb1} greater than $0$, either of the following two inequalities should hold: \begin{align} \label{eq:second_term_lb1_one} \sqrt{\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{h}}\right)^2}&>(2+\frac{1}{q})\sqrt{\sum_y\left({\bm{h}}^T{\bm{B}}_y{\bm{x}}\right)^2}\\ \label{eq:second_term_lb1_two} \sqrt{\sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{h}}\right)^2}&<(2-\frac{1}{q})\sqrt{\sum_y\left({\bm{h}}^T{\bm{B}}_y{\bm{x}}\right)^2}\,. \end{align} We can obtain an upper bound on $\|{\bm{h}}\|_2$ to make \eqref{eq:second_term_lb1_two} hold. Specifically, the left-hand side of \eqref{eq:second_term_lb1_two} can be upper bounded via: \begin{align} \label{eq:ub_hDh} \begin{split} \sum_y\left({\bm{h}}^\textrm{T}{\bm{B}}_y{\bm{h}}\right)^2&=\|{\bm{h}}\|_2^4\cdot\sum_y\left({\bm{u}}^\textrm{T}{\bm{B}}_y{\bm{u}}\right)^2\\ &=\|{\bm{h}}\|_2^4\cdot\sum_y\|{\bm{B}}_y\|_{op}^2\cdot\left(\frac{|{\bm{u}}^T{\bm{B}}_y{\bm{u}}|}{\|{\bm{B}}_y\|_{op}}\right)^2\\ &\leq \|{\bm{h}}\|_2^4\cdot\sum_y\|{\bm{B}}_y\|_{op}^2\cdot\left(\frac{|{\bm{u}}^T{\bm{B}}_y{\bm{u}}|}{\|{\bm{B}}_y\|_{op}}\right)\\ &=\|{\bm{h}}\|_2^4\cdot\sum_y\|{\bm{B}}_y\|_{op}\cdot|{\bm{u}}^T{\bm{B}}_y{\bm{u}}|\\ &\leq \|{\bm{h}}\|_2^4\cdot\sum_y\|{\bm{B}}_y\|_{op}\cdot |{\bm{u}}|^T{\bm{B}}_y|{\bm{u}}|\\ &=\|{\bm{h}}\|_2^4\cdot\sum_y\sigma_{\max}({\bm{B}}_y)\cdot |{\bm{u}}|^T{\bm{B}}_y|{\bm{u}}|\\ &\leq 2\|{\bm{h}}\|_2^4\cdot\sum_y|{\bm{u}}|^T{\bm{B}}_y|{\bm{u}}|\\ &=2\|{\bm{h}}\|_2^2\cdot|{\bm{h}}|^T\sum_y{\bm{B}}_y|{\bm{h}}|\\ &=2\|{\bm{h}}\|_2^2\cdot|{\bm{h}}|^T({\bm{1}}_\textnormal{mat}+{\bm{I}})|{\bm{h}}|\\ &=2\|{\bm{h}}\|_2^2\cdot(\|{\bm{h}}\|_1^2+\|{\bm{h}}\|_2^2)\\ &\leq 4\|{\bm{h}}\|_2^2\cdot\|{\bm{h}}\|_1^2\,, \end{split} \end{align} where ${\bm{u}}=\frac{1}{\|{\bm{h}}\|_2}{\bm{h}}$, ${\bm{1}}_\textnormal{mat}$ is a matrix of all $1$s and ${\bm{I}}$ is the identity matrix. The first inequality in \eqref{eq:ub_hDh} is obtained by $|{\bm{u}}^T{\bm{B}}_y{\bm{u}}|=|\langle{\bm{u}},{\bm{B}}_y{\bm{u}}\rangle|\leq\|{\bm{u}}\|_2\|{\bm{B}}_y{\bm{u}}\|_2\leq\|{\bm{B}}_y\|_{op}$ and hence $\frac{|{\bm{u}}^T{\bm{B}}_y{\bm{u}}|}{\|{\bm{B}}_y\|_{op}}\leq 1$; the second inequality is obtained by $|{\bm{u}}^\textrm{T}{\bm{B}}_y{\bm{u}}|=|\sum_{ij}A_y(i,j)u_iu_j|\leq\sum_{ij}A_y(i,j)|u_i||u_j|=|{\bm{u}}|^\textrm{T}{\bm{B}}_y|{\bm{u}}|$. If we choose the operator norm $\|\cdot\|_{op}$ to be the Euclidean norm, then $\|{\bm{B}}_y\|_{op}=\sigma_{\max}({\bm{B}}_y)\leq 2$; the last inequality is obtained via $\|{\bm{h}}\|_2\leq\|{\bm{h}}\|_1$. The right-hand side of \eqref{eq:second_term_lb1_two} can be low-bounded as: \begin{align} \label{eq:lb_hdx} \begin{split} \sum_y\left({\bm{h}}^T{\bm{B}}_y{\bm{x}}\right)^2&=\|{\bm{h}}\|_1^2\cdot{\bm{v}}^\textrm{T}\left(\sum_y{\bm{B}}_y{\bm{x}}{{\bm{x}}}^\textrm{T}{\bm{B}}_y^\textrm{T}\right){\bm{v}}=\|{\bm{h}}\|_1^2\cdot{\bm{v}}^\textrm{T}{\bm{E}}{\bm{v}}\\ &\geq\|{\bm{h}}\|_1^2\cdot\lambda_{\bm{E}}\,, \end{split} \end{align} where ${\bm{v}}=\frac{1}{\|{\bm{h}}\|_1}{\bm{h}}$, ${\bm{E}}=\sum_{y=0}^{M-1}{\bm{B}}_y{\bm{x}}{{\bm{x}}}^T{\bm{B}}_y^T$ and $\lambda_{\bm{E}}>0$ can be computed using Lemma \ref{LEMMA:E_MIN}. Combining \eqref{eq:second_term_lb1_two}, \eqref{eq:ub_hDh} and \eqref{eq:lb_hdx}, we can see that as long as the following \eqref{eq:h_bd} holds, \eqref{eq:second_term_lb1_two} will also hold. \begin{align} \label{eq:h_bd} \|{\bm{h}}\|_2<\tau=\left(2-\frac{1}{q}\right)\cdot\sqrt{\frac{\lambda_{\bm{E}}}{4}}\,. \end{align} The above \eqref{eq:h_bd} guarantees that \eqref{eq:second_term_lb1} is always greater than $0$. We then have \begin{align} \label{eq:lb_gradient_align} -\sqrt{\frac{1}{MK^2}\sum_y\left(({\bm{z}}_t-{\bm{x}})^\textrm{T}{\bm{B}}_y{\bm{x}}\right)^2}>-q\sqrt{4f({\bm{z}}_t)}\,. \end{align} Plug the above \eqref{eq:lb_gradient_align} into \eqref{eq:second_term}. We have: \begin{align} \label{eq:second_term_lb} \langle{\bm{z}}_t-{\bm{x}},\, \nabla f({\bm{z}}_t)\rangle > 4(1-q)f({\bm{z}}_t)\,. \end{align} Plug \eqref{eq:first_term_lb}, \eqref{eq:h_bd} and \eqref{eq:second_term_lb} into \eqref{eq:diff}. We have: \begin{align} \label{eq:diff2} \begin{split} &\left\|{\bm{z}}_t-\eta\nabla f({\bm{z}}_t)-{\bm{x}} \right\|^2_2-\mu\|{\bm{z}}_t-{\bm{x}}\|^2_2\\ <&\frac{16}{K^2}\|{\bm{z}}_t\|_2^2f({\bm{z}}_t)\cdot\eta^2-8(1-q)f({\bm{z}}_t)\cdot\eta+(1-\mu)\tau^2\\ =&\frac{16}{K^2}\|{\bm{z}}_t\|_2^2f({\bm{z}}_t) \cdot \left(\left(\eta-\frac{(1-q)K^2}{4\|{\bm{z}}_t\|_2^2}\right)^2-\frac{(1-q)^2K^4}{16\|{\bm{z}}_t\|_2^4}+\frac{(1-\mu)K^2\tau^2}{16\|{\bm{z}}_t\|_2^2f({\bm{z}}_t)}\right)\,. \end{split} \end{align} In order to make \eqref{eq:diff2} strictly less than $0$, the following should hold: \begin{align} \label{eq:condition1} \left(\eta-\frac{(1-q)K^2}{4\|{\bm{z}}_t\|_2^2}\right)^2 < \frac{(1-q)^2K^4}{16\|{\bm{z}}_t\|_2^4}-\frac{(1-\mu)K^2\tau^2}{16\|{\bm{z}}_t\|_2^2f({\bm{z}}_t)}\,. \end{align} The right hand side of \eqref{eq:condition1} should be strictly greater than $0$ so that a valid $\eta$ can be obtained. This requires \begin{align} \label{eq:mu_lb} \mu>1-\frac{(1-q)^2K^2}{\|{\bm{z}}_t\|_2^2}\frac{f({\bm{z}}_t)}{\tau^2}\,. \end{align} We also need $\mu\in(0,1)$ to ensure convergence towards the global optimum ${\bm{x}}$. Hence \begin{align} \max\left(0,\ 1-\frac{(1-q)^2K^2}{\|{\bm{z}}_t\|_2^2}\frac{f({\bm{z}}_t)}{\tau^2}\right)<\mu<1\,. \end{align} The step size $\eta$ is then: \begin{align} \frac{(1-q)K^2}{4\|{\bm{z}}_t\|_2^2}-\nu<\eta<\frac{(1-q)K^2}{4\|{\bm{z}}_t\|_2^2}+\nu\,, \end{align} where $\nu=\sqrt{\frac{(1-q)^2K^4}{16\|{\bm{z}}_t\|_2^4}-\frac{(1-\mu)K^2\tau^2}{16\|{\bm{z}}_t\|_2^2f({\bm{z}}_t)}}$. \paragraph{Step 2:} We use ${\bm{z}}={\bm{z}}_t-\eta\nabla f({\bm{z}}_t)$ to denote the gradient descent update, and ${\bm{z}}_{t+1}=\mathscr{P}_\mathcal{S}({\bm{z}})\in\mathcal{S}$ is the projected gradient descent update. In step 1 we established conditions under which \eqref{eq:diff}$<0$, i.e. \begin{align} \|{\bm{z}}-{\bm{x}}\|_2^2<\nu\|{\bm{z}}_t-{\bm{x}}\|_2^2\,. \end{align} Let ${\bm{s}}$ be a linear combination of ${\bm{z}}_{t+1}$ and a global optimizer ${\bm{x}}$ such that \begin{align} \label{eq:linear_comb} {\bm{x}}-{\bm{z}}_{t+1} = a({\bm{z}}_{t+1}-{\bm{s}})\,, \end{align} where $a\in\mathbb{R}$, $a\neq 0$ is some constant. We can always find an ${\bm{s}}$ such that the following holds, \begin{align} \label{eq:perpendicular} ({\bm{s}}-{\bm{z}})^T({\bm{s}}-{\bm{z}}_{t+1})=0\,. \end{align} \begin{enumerate}[label={\arabic*.}] \item If ${\bm{z}}={\bm{z}}_{t+1}$, then ${\bm{z}}\in\mathcal{S}$ and \eqref{eq:converge} holds. \item If ${\bm{x}}={\bm{z}}_{t+1}$, \eqref{eq:converge} naturally holds. \item Otherwise, we can choose $a=\frac{\|{\bm{x}}-{\bm{z}}_{t+1}\|_2^2}{\left({\bm{z}}_{t+1}-{\bm{z}}\right)^T\left({\bm{x}}-{\bm{z}}_{t+1}\right)}$. From \eqref{eq:perpendicular}, we can get: \begin{align} \label{eq:sum_four} \|{\bm{s}}\|_2^2-{\bm{z}}^T{\bm{s}}-{\bm{s}}^T{\bm{z}}_{t+1} = -{\bm{z}}^T{\bm{z}}_{t+1}\,. \end{align} We also have \begin{align} \label{eq:sum_one} \|{\bm{z}}-{\bm{z}}_{t+1}\|_2^2 &= \|{\bm{z}}\|_2^2 + \|{\bm{z}}_{t+1}\|_2^2-2{\bm{z}}^T{\bm{z}}_{t+1}\\ \label{eq:sum_two} \|{\bm{z}}-{\bm{s}}\|_2^2&=\|{\bm{z}}\|_2^2+\|{\bm{s}}\|_2^2-2{\bm{z}}^T{\bm{s}}\\ \label{eq:sum_three} \|{\bm{s}}-{\bm{z}}_{t+1}\|_2^2 &= \|{\bm{s}}\|_2^2 + \|{\bm{z}}_{t+1}\|_2^2 - 2{\bm{s}}^T{\bm{z}}_{t+1}\,. \end{align} Combining \eqref{eq:sum_four}-\eqref{eq:sum_three}, we get that \begin{align} \label{eq:sum_five} \|{\bm{z}}-{\bm{z}}_{t+1}\|_2^2=\|{\bm{z}}-{\bm{s}}\|_2^2+\|{\bm{s}}-{\bm{z}}_{t+1}\|_2^2\,. \end{align} Using \eqref{eq:linear_comb}, we have \begin{align} \label{eq:sz_xs} {\bm{s}}-{\bm{z}}_{t+1} = \frac{1}{a+1}({\bm{s}}-{\bm{x}})\,. \end{align} Plug \eqref{eq:sz_xs} into \eqref{eq:perpendicular}. We have \begin{align} ({\bm{s}}-{\bm{z}})^T({\bm{s}}-{\bm{x}})=0\,. \end{align} Similarly, we can get that \begin{align} \label{eq:sum_six} \|{\bm{z}}-{\bm{x}}\|_2^2=\|{\bm{z}}-{\bm{s}}\|_2^2+\|{\bm{s}}-{\bm{x}}\|_2^2\,. \end{align} \begin{enumerate}[label*={\arabic*)}] \item If ${\bm{s}}\in\mathcal{S}$, since ${\bm{z}}_{t+1}$ is the projection of ${\bm{z}}$ in $\mathcal{S}$, we have $\|{\bm{z}}-{\bm{z}}_{t+1}\|_2^2\leq\|{\bm{z}}-{\bm{s}}\|_2^2$. Using \eqref{eq:sum_five}, we have: \begin{align} \|{\bm{s}}-{\bm{z}}_{t+1}\|_2^2=0\,. \end{align} Hence ${\bm{s}}$ and ${\bm{z}}_{t+1}$ is the same point. From \eqref{eq:sum_six}, we can get: \begin{align} \|{\bm{z}}-{\bm{x}}\|_2^2=\|{\bm{z}}-{\bm{z}}_{t+1}\|_2^2+\|{\bm{z}}_{t+1}-{\bm{x}}\|_2^2\,. \end{align} Since ${\bm{z}}\notin\mathcal{S}$, we have $\|{\bm{z}}-{\bm{z}}_{t+1}\|_2^2>0$. Hence $\|{\bm{z}}-{\bm{x}}\|_2^2>\|{\bm{z}}_{t+1}-{\bm{x}}\|_2^2$. \item If ${\bm{s}}\notin\mathcal{S}$, we have: \begin{align} \label{eq:sum_seven} \begin{split} \|{\bm{s}}-{\bm{x}}\|_2^2&=\|{\bm{s}}-{\bm{z}}_{t+1}+{\bm{z}}_{t+1}-{\bm{x}}\|_2^2\\ &=\|{\bm{s}}-{\bm{z}}_{t+1}\|_2^2+\|{\bm{z}}_{t+1}-{\bm{x}}\|_2^2+2\left({\bm{s}}-{\bm{z}}_{t+1}\right)^\textrm{T}\left({\bm{z}}_{t+1}-{\bm{x}}\right)\,. \end{split} \end{align} \begin{itemize} \item If $a\in[0, \infty)$, from \eqref{eq:linear_comb}, we have $\left({\bm{s}}-{\bm{z}}_{t+1}\right)^\textrm{T}\left({\bm{z}}_{t+1}-{\bm{x}}\right)\geq 0$. From \eqref{eq:sum_seven}, we have $\|{\bm{s}}-{\bm{x}}\|_2^2\geq\|{\bm{z}}_{t+1}-{\bm{x}}\|_2^2$. Using \eqref{eq:sum_six}, we have $\|{\bm{z}}-{\bm{x}}\|_2^2\geq\|{\bm{z}}_{t+1}-{\bm{x}}\|_2^2$. \item If $a\in(-1,0)$, from \eqref{eq:linear_comb}, we have ${\bm{z}}_{t+1}-{\bm{x}}=\frac{a}{-1-a}({\bm{x}}-{\bm{s}})$. Since $\frac{a}{-1-a}>0$, $({\bm{z}}_{t+1}-{\bm{x}})^T({\bm{x}}-{\bm{s}})>0$. We then have: \begin{align} \begin{split} \|{\bm{z}}_{t+1}-{\bm{s}}\|_2^2&=\|{\bm{z}}_{t+1}-{\bm{x}}+{\bm{x}}-{\bm{s}}\|_2^2\\ &=\|{\bm{z}}_{t+1}-{\bm{x}}\|_2^2+\|{\bm{x}}-{\bm{s}}\|_2^2+2({\bm{z}}_{t+1}-{\bm{x}})^T({\bm{x}}-{\bm{s}})\\ &>\|{\bm{x}}-{\bm{s}}\|_2^2\,. \end{split} \end{align} Using \eqref{eq:sum_five} and \eqref{eq:sum_six}, we have: \begin{align} \|{\bm{z}}-{\bm{z}}_{t+1}\|_2^2>\|{\bm{z}}-{\bm{x}}\|_2^2\,. \end{align} This is in contradiction with the assumption that ${\bm{z}}_{t+1}$ is the projection of ${\bm{z}}$ in $\mathcal{S}$ so that ${\bm{z}}_{t+1}$ is closest point in $\mathcal{S}$ to ${\bm{z}}$ in terms of $l_2$ norm: $\|{\bm{z}}-{\bm{z}}_{t+1}\|_2^2\leq\|{\bm{z}}-{\bm{x}}\|_2^2$, hence $a\notin(-1,0)$. \item If $a\in(-\infty, -1]$, from \eqref{eq:linear_comb}, we have ${\bm{s}}=-\frac{1}{a}{\bm{x}}+(1+\frac{1}{a}){\bm{z}}_{t+1}$. Since $-\frac{1}{a}\in(0,1]$, ${\bm{s}}\in\mathcal{S}$. This is in contradiction with the assumption ${\bm{s}}\notin\mathcal{S}$, hence $a\notin(-\infty,1]$ \end{itemize} \end{enumerate} In summary, we have \begin{align} \|{\bm{z}}_{t+1}-{\bm{x}}\|_2^2\leq\|{\bm{z}}-{\bm{x}}\|_2^2<\mu\|{\bm{z}}_t-{\bm{x}}\|_2^2\,. \end{align} \end{enumerate} \vspace{0.2cm} \section{Introduction} \label{sec:intro} \input{1_intro.tex} \section{The Noisy Turnpike Problem} \label{sec:main_turnpike} \input{2_propose_turnpike.tex} \section{The Noisy Beltway Problem} \label{sec:main_beltway} \input{3_propose_beltway.tex} \section{Experimental Results} \label{sec:exp} \input{4_experiments.tex} \section{Conclusion} \label{sec:con} \input{5_conclusion.tex} \bibliographystyle{IEEEbib}
{ "timestamp": "2020-01-07T02:24:55", "yymm": "1804", "arxiv_id": "1804.02465", "language": "en", "url": "https://arxiv.org/abs/1804.02465", "abstract": "We address the problem of reconstructing a set of points on a line or a loop from their unassigned noisy pairwise distances. When the points lie on a line, the problem is known as the turnpike; when they are on a loop, it is known as the beltway. We approximate the problem by discretizing the domain and representing the $N$ points via an $N$-hot encoding, which is a density supported on the discretized domain. We show how the distance distribution is then simply a collection of quadratic functionals of this density and propose to recover the point locations so that the estimated distance distribution matches the measured distance distribution. This can be cast as a constrained nonconvex optimization problem which we solve using projected gradient descent with a suitable spectral initializer. We derive conditions under which the proposed distance distribution matching approach locally converges to a global optimizer at a linear rate. Compared to the conventional backtracking approach, our method jointly reconstructs all the point locations and is robust to noise in the measurements. We substantiate these claims with state-of-the-art performance across a number of numerical experiments. Our method is the first practical approach to solve the large-scale noisy beltway problem where the points lie on a loop.", "subjects": "Data Structures and Algorithms (cs.DS); Information Theory (cs.IT); Machine Learning (cs.LG)", "title": "Reconstructing Point Sets from Distance Distributions", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9838471618625456, "lm_q2_score": 0.8152324915965392, "lm_q1q2_score": 0.8020641731153867 }
https://arxiv.org/abs/0809.3482
Elliptic curves with maximal Galois action on their torsion points
Given an elliptic curve E over a number field k, the Galois action on the torsion points of E induces a Galois representation, \rho_E : Gal(\bar{k}/k) \to GL_2(\hat{Z}). For a fixed number field k, we describe the image of \rho_E for a "random" elliptic curve E over k. In particular, if k\neq Q is linearly disjoint from the cyclotomic extension of Q, then \rho_E will be surjective for "most" elliptic curves over k.
\section{Introduction} Let $E$ be an elliptic curve defined over a number field $k$. For each positive integer $m$, let $E[m]$ be the group of $m$-torsion points in $E(\kbar)$. The group $E[m]$ is isomorphic to $(\ZZ/m\ZZ)^2$ and is acted on by the absolute Galois group $G_k:=\operatorname{Gal}(\kbar/k)$. This action may be expressed in terms of a Galois representation, \[ \rho_{E,m} \colon G} %\newcommand{\calG}{\mathcal G_k \to \Aut(E[m]) \cong \GL_2(\ZZ/m\ZZ). \] Taking the inverse limit over all $m$ (ordered by divisibility), we have a Galois representation \[ \rho_E \colon G} %\newcommand{\calG}{\mathcal G_k \to \Aut(E(\kbar)_{\operatorname{tors}}) \cong \GL_2(\widehat\ZZ). \] In this paper, we will describe how large the image of $\rho_E$ is for a ``random'' elliptic curve over $k$. \\ A renowned theorem of Serre (\cite{Serre-Inv72}*{Th\'eor\`eme~3}) says that if $E$ does not have complex multiplication, then $\rho_E(G_k)$ has finite index in $\GL_2(\widehat\ZZ)$. In the same paper, Serre shows that $\rho_E(G} %\newcommand{\calG}{\mathcal G_\mathbb Q)\neq \GL_2(\widehat\ZZ)$ for every elliptic curve $E$ over $\mathbb Q$. The first example of a surjective $\rho_E$ was given in the doctoral thesis of A.~Greicius \cite{GThesis}. Let $\alpha\in \Qbar$ be a root of the polynomial $x^3+x+1$, and let $E/\mathbb Q(\alpha)$ be the elliptic curve given by the Weierstrass equation $y^2+2xy+\alpha y = x^3 - x^2$. In \cite{GThesis}*{\S3.3}, it is shown that $\rho_E(G} %\newcommand{\calG}{\mathcal G_{\mathbb Q(\alpha)})=\GL_2(\widehat\ZZ)$. \subsection{Statement of results} Denote the ring of integers of $k$ by $\mathcal O_k$. For $(a,b) \in \mathcal O_k^2$, define $\Delta_{a,b} = -16(4a^3+27b^2)$. If $\Delta_{a,b}\neq 0$, then let $E(a,b)$ be the elliptic curve over $k$ defined by the Weierstrass equation \[ Y^2=X^3+aX+b. \] Fix a norm $\norm{\cdot}$ on $\mathbb R\otimes_\ZZ \mathcal O_k^2 \cong \mathbb R^{2[k:\mathbb Q]}$. For each real number $x>0$, define the set \[ B_k(x) = \{ (a,b) \in \mathcal O_k^2 : \Delta_{a,b} \neq 0,\, \norm{(a,b)}\leq x \}. \] Thus to each pair $(a,b)\in B_k(x)$, we can associate an elliptic curve $E(a,b)$ over $k$. Note that \begin{equation} \label{E:box asymptotics} |B_k(x)|\sim \kappa x^{2[k:\mathbb Q]} \end{equation} as $x\to \infty$, where $\kappa>0$ is a constant which depends on $k$ and $\norm{\cdot}$. The following theorem addresses a question of Greicius on the surjectivity of the $\rho_E$ (\cite{GThesis}*{\S3.4 Problem~3}). Let $\mathbb Q^{\operatorname{cyc}}\subseteq \kbar$ be the cyclotomic extension of $\mathbb Q$. \begin{thm} \label{T:main theorem 1} Suppose that $k\cap \mathbb Q^{{\operatorname{cyc}}} =\mathbb Q$ and $k\neq \mathbb Q$. Then \[ \lim_{x\to \infty} \frac{ |\{ (a,b) \in B_k(x) : \rho_{E(a,b)}(G} %\newcommand{\calG}{\mathcal G_k) = \GL_2(\widehat\ZZ) \}| }{ |B_k(x)| } = 1. \] \end{thm} Intuitively, the theorem says that for a randomly chosen pair $(a,b)\in \mathcal O_k^2$, the corresponding elliptic curve $E(a,b)$ satisfies $\rho_{E(a,b)}(G} %\newcommand{\calG}{\mathcal G_k)=\GL_2(\widehat\ZZ).$ \\ Let $\chi_k\colon G} %\newcommand{\calG}{\mathcal G_k \to \widehat\ZZ^\times$ be the cyclotomic character of $k$. For each elliptic curve $E$ over $k$, we have $\det \circ \rho_E=\chi_k$. In particular, the assumption $k\cap \mathbb Q^{{\operatorname{cyc}}} =\mathbb Q$ (equivalently $\chi_k(G} %\newcommand{\calG}{\mathcal G_k)=\widehat\ZZ^\times$) is necessary for Theorem~\ref{T:main theorem 1}. For a number field $k$, define the group \[ H_k := \{ A \in \GL_2(\widehat\ZZ) : \det(A) \in \chi_k(G} %\newcommand{\calG}{\mathcal G_k)\}. \] Given an elliptic curve $E$ over $k$, we certainly have $\rho_E(G} %\newcommand{\calG}{\mathcal G_k)\subseteq H_k$. Our main theorem, which generalizes Theorem~\ref{T:main theorem 1}, shows that this is the only general constraint for $k\neq \mathbb Q$. \begin{thm} \label{T:main theorem 2} Let $k\neq \mathbb Q$ be a number field. Then \[ \frac{ |\{ (a,b) \in B_k(x) : \rho_{E(a,b)}(G} %\newcommand{\calG}{\mathcal G_k) \neq H_k \}| }{ |B_k(x)| } \ll_{k,\norm{\cdot}} \frac{\log x}{\sqrt{x}}. \] \end{thm} \begin{remark} Theorem~\ref{T:main theorem 2} shows that the proportion of $(a,b)$ in $B_k(x)$ with $\rho_{E(a,b)}(G} %\newcommand{\calG}{\mathcal G_k)=H_k$, as a function of $x$, quickly approaches $1$. The implicit constant in the theorem is effective and depends only on $k$ and the fixed norm. \end{remark} Before continuing, let us introduce some more notation. Let $E$ be an elliptic curve over a field $k$. For each positive integer $m$, denote the fixed field in $\kbar$ of $\ker(\rho_{E,m})$ by $k(E[m])$. \subsection{The rationals} \label{SS:rationals} For completeness, we now discuss the case $k=\mathbb Q$ which was excluded from Theorem~\ref{T:main theorem 2}. Let $E$ be an elliptic curve over $\mathbb Q$, and let $\Delta$ be the discriminant of some Weierstrass model of $E$ over $\mathbb Q$. There exists an integer $n\geq 1$ such that $\mathbb Q(\sqrt{\Delta}) \subseteq \mathbb Q(\mu_n)$, where $\mu_n$ is the set of $n$-th roots of unity (the assumption $k=\mathbb Q$ is important here!). Using that the field $\mathbb Q(\sqrt{\Delta})$ lies in both $\mathbb Q(E[2])$ and $\mathbb Q(\mu_n)\subseteq \mathbb Q(E[n])$, Serre deduced that the index $[\GL_2(\ZZ/2n\ZZ): \rho_{E,2n}(G} %\newcommand{\calG}{\mathcal G_\mathbb Q)]$ is \emph{even} (for details, see \cite{Serre-Inv72}*{pp.~310-311}) and in particular $\rho_E(G} %\newcommand{\calG}{\mathcal G_\mathbb Q)\neq \GL_2(\widehat\ZZ)$. Following Lang and Trotter, we make the following definition. \begin{definition} An elliptic curve $E$ over $\mathbb Q$ is a \emph{Serre curve} if $[\GL_2(\widehat\ZZ) : \rho_{E}(G} %\newcommand{\calG}{\mathcal G_\mathbb Q) ]=2$. \end{definition} A Serre curve is thus an elliptic curve $E$ over $\mathbb Q$ for which $\rho_E(G} %\newcommand{\calG}{\mathcal G_\mathbb Q)$ is as large as possible. For a Serre curve $E/\mathbb Q$, the group $\rho_{E}(G} %\newcommand{\calG}{\mathcal G_\mathbb Q)$ can be described explicitly in terms of the field $\mathbb Q(\sqrt{\Delta})$. N.~Jones \cite{Jones} (building on work of Duke \cite{Duke}) has shown that ``most'' elliptic curves over $\mathbb Q$ are Serre curves. The analogue of Theorem~\ref{T:main theorem 2} is then the following. \begin{thm}[Jones] \label{T:Jones-Serre curves} There is a constant $\beta>0$ such that \[ \frac{ |\{ (a,b) \in B_\mathbb Q(x) : E(a,b) \text{ is not a Serre curve} \}| }{ |B_\mathbb Q(x)| } \ll_{\norm{\cdot}} \frac{(\log x)^\beta}{\sqrt{x}}. \] \end{thm} \begin{remark} Theorem~\ref{T:Jones-Serre curves} is a special case of \cite{Jones}*{Theorem~4} and will be proven in \S\ref{SS:Serre curves}. Unlike Jones' version, the constants in our proof will be effective. \end{remark} \subsection{Overview of proof} Suppose that $E$ is an elliptic curve over a number field $k\neq \mathbb Q$. There is an exact sequence \[ 1\to \operatorname{SL}_2(\widehat\ZZ) \to \GL_2(\widehat\ZZ) \stackrel{\det}{\to} \widehat\ZZ^\times \to 1, \] and the representation $\det \circ \rho_E\colon G} %\newcommand{\calG}{\mathcal G_k \to \widehat\ZZ^\times$ is the cyclotomic character $\chi_k$ of $k$. Therefore, \[ \rho_E(G} %\newcommand{\calG}{\mathcal G_k) \cap \operatorname{SL}_2(\widehat\ZZ) = \rho_E(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}}). \] Thus the equality $\rho_E(G} %\newcommand{\calG}{\mathcal G_k)=H_k$ is equivalent to $\rho_E(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}}) = \operatorname{SL}_2(\widehat\ZZ)$. A group theoretic argument will show that this in turn is equivalent to having $\rho_{E,m}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\ZZ/m\ZZ)$ whenever $m$ is equal to $4$, $9$, or a prime $\geq5$. For a prime $m=\ell\geq 5$, the condition $\rho_{E,\ell}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\ZZ/\ell\ZZ)$ is equivalent to $\rho_{E,\ell}(G} %\newcommand{\calG}{\mathcal G_{k})\supseteq \operatorname{SL}_2(\ZZ/\ell\ZZ).$ By considering the Frobenius endomorphism for the reduction of $E$ modulo several primes $\mathfrak p\subseteq \mathcal O_k$, we can determine which conjugacy classes of $\GL_2(\ZZ/\ell\ZZ)$ meet $\rho_{E,\ell}(G} %\newcommand{\calG}{\mathcal G_k)$. Combining this modulo $\mathfrak p$ information together, we will use the large sieve to give an asymptotic upper bound for the growth of \[ |\{(a,b) \in B_k(x) : \rho_{E(a,b),\ell}(G} %\newcommand{\calG}{\mathcal G_k)\not\supseteq \operatorname{SL}_2(\ZZ/\ell\ZZ) \}| \] as a function of $x$; see \S\ref{S:exceptional primes}. To understand the distribution of reductions modulo $\mathfrak p$, we will use a recent result of Jones; see \S\ref{S:over finite fields}. Of significant importance is a theorem of Masser and W{\"u}stholz, which is needed to bound the number of $\ell$'s that must be considered. The conditions at $m=4$ or $9$ are more involved. In particular, for $m=4$ we will need to impose the condition that $\sqrt{\Delta}$ is not in the cyclotomic extension of $k$ (this avoids the obstruction of \S\ref{SS:rationals} that always occurs for $k=\mathbb Q$). In \S\ref{S:discriminants}, we again use the large sieve to bound the number of $(a,b)\in B_k(x)$ for which $\sqrt{\Delta_{a,b}}$ (and $\sqrt[3]{\Delta_{a,b}}$ if $\mu_3\subseteq k$) lie in the cyclotomic extension of $k$. Our main theorems will then be deduced in \S\ref{S:proof}. \subsection*{Acknowledgments} Many thanks to Bjorn Poonen for his careful reading of this paper and helpful comments. Thanks also to Aaron Greicius and Nathan Jones. \subsection*{Notation and conventions} \label{S:notation} For each field $k$, let $\kbar$ be an algebraic closure of $k$ and let $G} %\newcommand{\calG}{\mathcal G_k := \operatorname{Gal}(\kbar/k)$ be the absolute Galois group of $k$. For each integer $n\geq 1$, let $\mu_n$ be the group of $n$-th roots of unity in $\kbar$. Let $k^{\operatorname{cyc}}$ (resp.~$k^{\operatorname{ab}}$) be the cyclotomic (resp.~maximal abelian) extension of $k$ in $\kbar$. For a number field $k$, denote its ring of integers by $\mathcal O_k$. Let $\Sigma_k$ be the set of non-zero prime ideals of $\mathcal O_k$. For each $\mathfrak p\in \Sigma_k$, we have a residue field $\mathbb F_\mathfrak p=\mathcal O_k/\mathfrak p$; let $N(\mathfrak p)$ be its cardinality. Let $\Sigma_k(x)$ be the set of primes $\mathfrak p$ in $\Sigma_k$ with $N(\mathfrak p)\leq x$. Let $\ord_\mathfrak p\colon k^\times \to \ZZ$ be the surjective discrete valuation corresponding to $\mathfrak p$. Denote the absolute discriminant of $k$ by $d_k$. Fix a group $G$. Let $G'$ be the derived subgroup of $G$, i.e., the minimal normal subgroup of $G$ for which $G/G'$ is abelian. Equivalently, $G'$ is the group generated by the set $\{xyx^{-1}y^{-1}:x,y\in G\}$. The {abelianization} of $G$ is $G^{\operatorname{ab}}:=G/G'$. Profinite groups will always be considered with their profinite topologies. For $a,b$ in a field $k$, define $\Delta_{a,b}= -16(4a^3+27b^2)$. If $\Delta_{a,b}\neq 0$, then let $E(a,b)$ be the elliptic curve over $k$ defined by the Weierstrass equation $Y^2=X^3+aX+b$. Suppose that $f$ and $g$ are real valued functions of a real variable $x$. By $f\ll g$ (or $g \gg f$), we mean that there are positive constants $C_1$ and $C_2$ such that for all $x\geq C_1$, $|f(x)| \leq C_2 |g(x)|$. We use $O(f)$ to represent an unspecified function $g$ with $g \ll f$. The dependencies of the implied constants will always be indicated by subscripts. Also, all implicit constants occurring in this paper are effective. Finally, the symbols $\ell$ and $p$ will always denote rational primes. \section{Criterion for maximal Galois action} \label{S:criterion} \begin{prop} \label{P:Criterion} Let $E$ be an elliptic curve over a number field $k$, and let $\Delta$ be the discriminant of a Weierstrass model of $E$ over $k$. Suppose that the following conditions hold: \begin{alphenum} \item \label{I:a} $\rho_{E,\ell}(G} %\newcommand{\calG}{\mathcal G_k)\supseteq \operatorname{SL}_2(\ZZ/\ell\ZZ)$ for every prime $\ell\geq 5$, \item \label{I:b} $\rho_{E,4}(G} %\newcommand{\calG}{\mathcal G_k)\supseteq \operatorname{SL}_2(\ZZ/4\ZZ)$ and $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_k)\supseteq \operatorname{SL}_2(\ZZ/9\ZZ)$, \item \label{I:c} $\sqrt{\Delta} \not\in k^{\operatorname{cyc}}$, \item \label{I:d} $\mu_3 \not\subseteq k$ or $\sqrt[3]{\Delta}\not\in k^{\operatorname{cyc}}$. \end{alphenum} Then $\rho_E(G} %\newcommand{\calG}{\mathcal G_k)=H_k$. \end{prop} \begin{remark} \begin{romanenum} \item The image of $\Delta$ in $k^\times/(k^\times)^{12}$ depends only on the isomorphism class of $E/k$. Thus for a positive integer $r$ dividing $12$, the $r$-th root of $\Delta$, up to a factor in $\mu_r\cdot k^\times$, is independent of all choices. In particular, conditions (\ref{I:c}) and (\ref{I:d}) are well-defined. \item The Kronecker-Weber theorem says that $\mathbb Q^{\operatorname{cyc}}=\mathbb Q^{\operatorname{ab}}$, so condition (\ref{I:c}) \emph{never} holds for $k=\mathbb Q$. \end{romanenum} \end{remark} Since $\det\circ \rho_E \colon G} %\newcommand{\calG}{\mathcal G_k \to \widehat\ZZ^\times$ is the cyclotomic character of $k$, we find that $\rho_E(G} %\newcommand{\calG}{\mathcal G_{k})=H_k$ if and only if $\rho_E(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\widehat\ZZ)$. Applying Lemma~\ref{L:SL(Zhat) Goursat} to $\rho_E(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})$, we have $\rho_E(G} %\newcommand{\calG}{\mathcal G_{k})=H_k$ if and only if $\rho_{E,m}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\ZZ/m\ZZ)$ holds whenever $m$ is $4$, $9$, or a prime $\geq 5$. Proposition~\ref{P:Criterion} is then an immediate consequence of the following lemma. \begin{lemma} \label{L:cyclotomic exceptional} Let $E$ be an elliptic curve over a number field $k$ with discriminant $\Delta\in k^\times/(k^\times)^{12}$. \begin{romanenum} \item Let $\ell\geq 5$ be a prime. If $\rho_{E,\ell}(G} %\newcommand{\calG}{\mathcal G_k) \supseteq \operatorname{SL}_2(\ZZ/\ell\ZZ)$, then $\rho_{E,\ell}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}}) = \operatorname{SL}_2(\ZZ/\ell\ZZ)$. \item If $\rho_{E,4}(G} %\newcommand{\calG}{\mathcal G_k) \supseteq \operatorname{SL}_2(\ZZ/4\ZZ)$ and $\sqrt{\Delta} \not\in k^{\operatorname{cyc}}$, then $\rho_{E,4}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\ZZ/4\ZZ)$. \item If $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_k) \supseteq \operatorname{SL}_2(\ZZ/9\ZZ)$ and $\sqrt[3]{\Delta} \not\in k^{\operatorname{cyc}}$, then $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\ZZ/9\ZZ)$. \item If $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_k) \supseteq \operatorname{SL}_2(\ZZ/9\ZZ)$ and $\mu_3 \not\subseteq k$, then $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\ZZ/9\ZZ)$. \end{romanenum} \end{lemma} \begin{proof} Let $m$ be a positive integer such that $\rho_{E,m}(G} %\newcommand{\calG}{\mathcal G_k)\supseteq \operatorname{SL}_2(\ZZ/m\ZZ)$. Since $k^{\operatorname{cyc}}$ is an abelian extension of $k$, we have inclusions \begin{equation} \label{E:cyclotomic inclusions} \operatorname{SL}_2(\ZZ/m\ZZ)' \subseteq \rho_{E,m}(G} %\newcommand{\calG}{\mathcal G_{k})' \subseteq \rho_{E,m}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}}) \subseteq \operatorname{SL}_2(\ZZ/m\ZZ). \end{equation} \noindent (i) Suppose that $m=\ell\geq 5$ is prime. By Lemma~\ref{L:derived subgroup finite} we have $\operatorname{SL}_2(\ZZ/\ell\ZZ)'=\operatorname{SL}_2(\ZZ/\ell\ZZ)$, so from (\ref{E:cyclotomic inclusions}) we deduce that $\rho_{E,\ell}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}}) = \operatorname{SL}_2(\ZZ/\ell\ZZ)$. \noindent (ii) Our assumption $\rho_{E,4}(G} %\newcommand{\calG}{\mathcal G_k) \supseteq \operatorname{SL}_2(\ZZ/4\ZZ)$ implies that $\rho_{E,4}(G} %\newcommand{\calG}{\mathcal G_{k(\mu_4)}) = \operatorname{SL}_2(\ZZ/4\ZZ)$. Thus to prove $\rho_{E,4}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}}) = \operatorname{SL}_2(\ZZ/4\ZZ)$, it suffices to show that $k(E[4])\cap k^{\operatorname{cyc}} = k(\mu_4)$. In \cite{Lang-Trotter}*{Part III \S11}, it is shown that $\sqrt[4]{\Delta}$ is an element of $k(E[4])$. Using $\sqrt{\Delta} \not\in k(\mu_4)$, one finds that $k(\mu_4, \sqrt[4]{\Delta})\subseteq k(E[4])$ is an abelian extension of $k(\mu_4)$ of degree $4$. By Lemma~\ref{L:derived subgroup finite} the group $\operatorname{SL}_2(\ZZ/4\ZZ)^{\operatorname{ab}}$ is cyclic of order $4$, so $k(E[4]) \cap k(\mu_4)^{\operatorname{ab}} = k(\mu_4, \sqrt[4]{\Delta})$. Therefore \begin{equation} \label{E:4 cyclotomic} k(E[4]) \cap k^{\operatorname{cyc}} = (k(E[4])\cap k(\mu_4)^{\operatorname{ab}}) \cap k^{\operatorname{cyc}} = k(\mu_4, \sqrt[4]{\Delta}) \cap k^{\operatorname{cyc}} = k(\mu_4), \end{equation} where the last equality uses $\sqrt{\Delta} \not\in k^{\operatorname{cyc}}$. \noindent (iii) The assumption $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_k)\supseteq \operatorname{SL}_2(\ZZ/9\ZZ)$ implies that $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_{k(\mu_9)})= \operatorname{SL}_2(\ZZ/9\ZZ)$. By Lemma~\ref{L:derived subgroup finite}, the group $\operatorname{SL}_2(\ZZ/9\ZZ)^{\operatorname{ab}}$ has order $3$. Note that $\sqrt[3]{\Delta}$ is an element of $k(E[3])$ (see \cite{Adelmann}*{Proposition 5.4.3} for example). Arguing as in part (ii), we find that $k(E[9]) \cap k(\mu_9)^{\operatorname{ab}} = k(\mu_9, \sqrt[3]{\Delta}).$ Since $\sqrt[3]{\Delta} \not\in k^{\operatorname{cyc}}$, we deduce that $k(E[9]) \cap k^{\operatorname{cyc}}=k(\mu_9)$, and hence $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\ZZ/9\ZZ)$. \noindent (iv) The assumptions imply that $\rho_{E,3}(G} %\newcommand{\calG}{\mathcal G_k)=\GL_2(\ZZ/3\ZZ)$. One readily checks that $\GL_2(\ZZ/3\ZZ)'=\operatorname{SL}_2(\ZZ/3\ZZ)$, and thus $\rho_{E,3}(G} %\newcommand{\calG}{\mathcal G_k)'=\operatorname{SL}_2(\ZZ/3\ZZ)$. Using (\ref{E:cyclotomic inclusions}), with $m=3$, gives $\rho_{E,3}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\ZZ/3\ZZ)$. By Lemma~\ref{L:derived subgroup finite}, the group $\operatorname{SL}_2(\ZZ/9\ZZ)^{\operatorname{ab}}$ has order $3$. So from (\ref{E:cyclotomic inclusions}), with $m=9$, we find that $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})$ is either $\operatorname{SL}_2(\ZZ/9\ZZ)'$ or $\operatorname{SL}_2(\ZZ/9\ZZ)$. If $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\ZZ/9\ZZ)'$, then Lemma~\ref{L:derived subgroup finite} implies that $\rho_{E,3}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})\neq \operatorname{SL}_2(\ZZ/3\ZZ)$. Therefore, $\rho_{E,9}(G} %\newcommand{\calG}{\mathcal G_{k^{\operatorname{cyc}}})=\operatorname{SL}_2(\ZZ/9\ZZ)$. \end{proof} We now state a criterion that applies to $k=\mathbb Q$. \begin{lemma}[Jones] \label{L:Serre curve criterion} Let $E$ be an elliptic curve over $\mathbb Q$ which satisfies the following properties: \begin{alphenum} \item $\rho_{E,\ell}(G} %\newcommand{\calG}{\mathcal G_\mathbb Q)\supseteq \operatorname{SL}_2(\ZZ/\ell\ZZ)$ for every prime $\ell\geq 5$, \item $\rho_{E,72}(G} %\newcommand{\calG}{\mathcal G_\mathbb Q)\supseteq \operatorname{SL}_2(\ZZ/72\ZZ)$. \end{alphenum} Then $E$ is a Serre curve. \end{lemma} \begin{proof} For each $m\geq 1$, we have $\rho_{E,m}(G} %\newcommand{\calG}{\mathcal G_\mathbb Q)=\GL_2(\ZZ/m\ZZ)$ if and only if $\rho_{E,m}(G} %\newcommand{\calG}{\mathcal G_\mathbb Q)\supseteq \operatorname{SL}_2(\ZZ/m\ZZ)$. The lemma is now a special case of \cite{Jones}*{Lemma~5}. \end{proof} \section{Elliptic curve over finite fields} \label{S:over finite fields} Fix a positive integer $m$ and a prime $p\nmid m$. Let $E$ be an elliptic curve over the field $\mathbb F_p$. As before, one has a Galois representation $\rho_{E,m}\colon \operatorname{Gal}({\FFbar}_p/\mathbb F_p) \to \GL_2(\ZZ/m\ZZ)$, which arises from the Galois action on the $m$-torsion of $E$. Let $\operatorname{Frob}_p\in \operatorname{Gal}(\FFbar_p/\mathbb F_p)$ be the $p$-th power Frobenius automorphism. For a subset $C$ of $\GL_2(\ZZ/m\ZZ)$ stable under conjugation, define the set \[ \Omega_C(p) := \big\{(r,s) \in \mathbb F_p^2 : \Delta_{r,s} \neq 0, \, \rho_{E(r,s),m}(\operatorname{Frob}_p) \in C \big\}. \] The following theorem gives a good estimate on the cardinality of this set. \begin{thm}[Jones] \label{T:Jones} Fix a positive integer $m$ and a conjugacy class $C$ of $\GL_2(\ZZ/m\ZZ)$. Let $d$ be the element of $(\ZZ/m\ZZ)^\times$ such that $\det(C)=\{d\}$. Then for all primes $p$ with $p\equiv d \bmod{m}$, \[ \frac{|\Omega_C(p)|}{p^2} = \frac{|C| }{|\operatorname{SL}_2(\ZZ/m\ZZ)|} + O\Bigg(\frac{m^5}{p^{1/2}}\Bigg). \] \end{thm} \begin{proof} This follows from Theorem~8 (and Theorem~7) of \cite{Jones}. A key ingredient is a generalization of results of Hurwitz, see \cite{Jones-Trace}. \end{proof} \section{The large sieve} Let $K$ be a number field, $\Lambda$ a free $\mathcal O_K$-module of rank $n$, and $\norm{\cdot}$ a norm on $\Lambda_\mathbb R = \mathbb R\otimes_\ZZ \Lambda$. Fix a subset $Y$ of $\Lambda$. Let $x\geq1$ and $Q>0$ be real numbers. For every prime ideal $\mathfrak p\in\Sigma_K$, let $\omega_\mathfrak p$ be a real number in the interval $[0,1)$. Assume the following conditions hold: \begin{enumerate} \item The set $Y$ is contained in a ball of radius $x$; i.e.,~there is an $a_0\in \Lambda_\mathbb R$ such that $\norm{a-a_0}\leq x$ for all $a\in Y$. \item For every $\mathfrak p\in \Sigma_K(Q)$, the image $Y_\mathfrak p$ of $Y$ in $\Lambda/\mathfrak p\Lambda$ by reduction modulo $\mathfrak p$ satisfies \[ |Y_\mathfrak p| \leq (1-\omega_\mathfrak p)|\Lambda/\mathfrak p\Lambda|. \] \end{enumerate} \begin{thm}[Large sieve, \cite{SerreMordellWeil}*{\S12.1}] \label{T:large sieve} With assumptions as above, we have \[ |Y| \ll_{K,\Lambda, \norm{\cdot}} \frac{ x^{[K:\mathbb Q]n} + Q^{2n}}{L(Q)} \] where \[ L(Q):=\sum_{ \substack{\aA\subseteq \mathcal O_K \text{squarefree} \\ N(\aA)\leq Q}} \prod_{\mathfrak p|\aA} \frac{\omega_\mathfrak p}{1-\omega_\mathfrak p}. \] \end{thm} \begin{remark} We will apply the large sieve with $\Lambda=\mathcal O_k^2$ and $\norm{\cdot}$ our fixed norm on $\mathbb R\otimes_\ZZ \mathcal O_k^2$. In \S\ref{S:exceptional primes} and \S\ref{S:discriminants}, we will take $K$ to be $k$ and $\mathbb Q$, respectively. \end{remark} \section{Most elliptic curves have large $\ell$-adic Galois images} \label{S:exceptional primes} Throughout this section, fix a number field $k$. \begin{definition} For each positive integer $m$, define the set \[ B_{k,m}(x):=\{ (a,b) \in B_k(x): \rho_{E(a,b),m}(G} %\newcommand{\calG}{\mathcal G_k) \not\supseteq \operatorname{SL}_2(\ZZ/m\ZZ) \}. \] \end{definition} The main goal of this section is to prove the following bound. \begin{prop} \label{P:l-adic surjectivity} There is an absolute constant $\beta\geq 1$ such that \[ \frac{|B_{k,4}(x) \cup B_{k,9}(x) \cup \bigcup_{\ell \geq 5} B_{k,\ell}(x)|}{|B_k(x)|} \ll_{k,\norm{\cdot}} \frac{(\log x)^{\beta}}{x^{[k:\mathbb Q]/2}}. \] \end{prop} \begin{remark} For an elliptic curve $E$ over $k$, we have Galois representations ${\rho}_{E,\ell^\infty} \colon G} %\newcommand{\calG}{\mathcal G_k \to \GL_2(\ZZ_\ell)$ coming from the action on the $\ell$-power torsion. Proposition~\ref{P:l-adic surjectivity} (with Lemma~\ref{L:reduce to finite groups}) shows that for a ``random'' elliptic curve $E$ over $k$, we have ${\rho}_{E,\ell^\infty}(G} %\newcommand{\calG}{\mathcal G_k)\supseteq \operatorname{SL}_2(\ZZ_\ell)$ for all primes $\ell$. Since $\det \circ {\rho}_{E,\ell^\infty} \colon G} %\newcommand{\calG}{\mathcal G_k \to \ZZ_\ell^\times$ is the $\ell$-adic cyclotomic character of $k$, we find that ${\rho}_{E,\ell^\infty}(G} %\newcommand{\calG}{\mathcal G_k)$ is as ``large as possible'' for all $\ell$. \end{remark} \begin{remark} Our proof of Proposition~\ref{P:l-adic surjectivity} is clearly based on Duke's paper \cite{Duke}, which proves the $k=\mathbb Q$ case (with Jones \cite{Jones} handling $4$ and $9$). Unlike Duke's result, the implicit constants in Proposition~\ref{P:l-adic surjectivity} are effective. The source of non-effective constants in \cite{Duke} is the use of the Siegel-Walfisz theorem. We avoid this by applying the pigeonhole principle in the proof of Lemma~\ref{L:exceptional prime bound 1} and then sieving only by conjugacy classes with a fixed determinant. \end{remark} \subsection{Sieving elliptic curves by Frobenius conjugacy classes} For a positive integer $m$ and a conjugacy class $C$ of $\GL_2(\ZZ/m\ZZ)$, define the set \[ Y_C(x) := \{ (a,b) \in B_k(x): \rho_{E(a,b),m}(G} %\newcommand{\calG}{\mathcal G_k) \cap C = \emptyset \big\}. \] For $d\in(\ZZ/m\ZZ)^\times$, let $\Sigma_{k}^1(Q;d,m)$ be the set of $\mathfrak p\in\Sigma_k(Q)$ with degree $1$ (i.e., $N(\mathfrak p)$ prime) and $N(\mathfrak p)\equiv d \bmod{m}$. \begin{prop} \label{P:nasty sieving} Let $m$ be a positive integer and $C$ a conjugacy class of $\GL_2(\ZZ/m\ZZ)$. Let $d$ be the unique element of $(\ZZ/m\ZZ)^\times$ such that $\det(C)=\{d\}$, and assume that $d\in \chi_k(G} %\newcommand{\calG}{\mathcal G_k) \bmod{m} \subseteq (\ZZ/m\ZZ)^\times$. Then \[ \frac{ |Y_C(x)|}{|B_k(x)|} \ll_{k,\norm{\cdot}} \Bigg( \frac{|C|}{|\operatorname{SL}_2(\ZZ/m\ZZ)|} |\Sigma_{k}^1(x^{[k:\mathbb Q]/2};d,m)| + O_k(m^5 x^{[k:\mathbb Q]/4}) \Bigg)^{-1}. \] \end{prop} \begin{proof} Let $\Lambda$ be the $\mathcal O_k$-module $\mathcal O_k^2$. We have already chosen a norm $\norm{\cdot}$ on $\Lambda_\mathbb R:=\mathbb R \otimes_\ZZ \mathcal O_k^2$, and the set $B_k(x)\subseteq \Lambda_\mathbb R$ lies in a ball of radius $x$. Let $Q:=x^{[k:\mathbb Q]/2}$. For each $\mathfrak p\in\Sigma_{k}^1(Q;d,m)$, define \[ \Omega_\mathfrak p = \{ (r,s) \in \mathbb F_\mathfrak p^2 : \Delta_{r,s} \neq 0, \, \rho_{E(r,s),m}(\operatorname{Frob}_{N(\mathfrak p)}) \in C \} \] and $\omega_\mathfrak p = |\Omega_\mathfrak p|/N(\mathfrak p)^2$. Let $Y_{\mathfrak p}$ be the image of $Y_C(x)$ in $\mathbb F_\mathfrak p^2$ via reduction modulo $\mathfrak p$. Suppose that $(a,b)\in B_k(x)$ satisfies $(a,b) \bmod{\mathfrak p} \in \Omega_\mathfrak p$; then $E(a,b)$ has good reduction at $\mathfrak p$ and $\rho_{E(a,b),m}(\operatorname{Frob}_\mathfrak p) \subseteq C$. So $\rho_{E(a,b),m}(G} %\newcommand{\calG}{\mathcal G_k) \cap C \neq \emptyset$, and thus $(a,b) \notin Y_C(x)$. This shows that \begin{equation*} Y_{\mathfrak p} \subseteq \mathbb F_\mathfrak p^2 - \Omega_\mathfrak p, \end{equation*} and hence $|Y_{\mathfrak p}| \leq (1-\omega_\mathfrak p)|\Lambda/\mathfrak p\Lambda|.$ For $\mathfrak p\notin\Sigma_{k}^1(Q;d,m)$, define $\omega_\mathfrak p=0$. By the large sieve (Theorem~\ref{T:large sieve}), \begin{equation} \label{E: result of large sieve} |Y_C(x)| \ll_{k,\norm{\cdot}} \frac{x^{2[k:\mathbb Q]} }{L(Q)}, \end{equation} where \[ L(Q):=\sum_{\substack{\aA\subseteq \mathcal O_k \text{ squarefree} \\ N(\aA)\leq Q}} \prod_{\mathfrak p|\aA} \frac{\omega_\mathfrak p}{1-\omega_\mathfrak p} \geq \sum_{\mathfrak p \in \Sigma_{k}^1(Q;d,m)} \omega_\mathfrak p. \] For $\mathfrak p \in \Sigma_{k}^1(Q;d,m)$, Theorem~\ref{T:Jones} gives \[ \omega_\mathfrak p = \frac{|C|}{|\operatorname{SL}_2(\ZZ/m\ZZ)|} + O(m^5/N(\mathfrak p)^{1/2}). \] Therefore \begin{align*} L(Q) & \geq \sum_{\mathfrak p \in \Sigma_{k}^1(Q;d,m)} \bigg( \frac{|C|}{|\operatorname{SL}_2(\ZZ/m\ZZ)|} + O(m^5/N(\mathfrak p)^{1/2}) \bigg)\\ &=\frac{|C|}{|\operatorname{SL}_2(\ZZ/m\ZZ)|} |\Sigma_{k}^1(Q;d,m)| + O_k(m^5 Q^{1/2}). \end{align*} The assumption $d\in \chi_k(G} %\newcommand{\calG}{\mathcal G_k) \bmod{m}$ is needed to guarantees that $L(Q) \gg_{k,m} 1$. The proposition follows by combining our lower bound of $L(Q)$ and (\ref{E:box asymptotics}) with (\ref{E: result of large sieve}). \end{proof} \subsection{Galois image modulo an integer} The following proposition shows that for a ``random'' elliptic curve $E$ over $k$, $\rho_{E,m}$ has large image. \begin{prop} \label{P:exceptional prime bound m} For a positive integer $m$, \[ \frac{|B_{k,m}(x)|}{|B_{k}(x)|} \ll_{k,\norm{\cdot},m} \frac{\log x}{x^{[k:\mathbb Q]/2}}. \] \end{prop} \begin{proof} Let $C_1,\dots,C_n$ be the conjugacy classes of $\GL_2(\ZZ/m\ZZ)$ with determinant $1$. By Lemma~\ref{L:SL condition for m}, we have $B_{k,m}(x) \subseteq \bigcup_{i=1}^n Y_{C_i}(x)$. Proposition~\ref{P:nasty sieving} gives \[ \frac{|B_{k,m}(x)|}{|B_k(x)|} \leq \sum_{i=1}^n \frac{|Y_{C_i}(x)|}{|B_k(x)|} \ll_{k,\norm{\cdot}} \sum_{i=1}^n \Big( \frac{|C_i|}{|\operatorname{SL}_2(\ZZ/m\ZZ)|} |\Sigma_{k}^1(x^{[k:\mathbb Q]/2};1,m)| + O_k(m^5 x^{[k:\mathbb Q]/4})\Big)^{-1}. \] Using $ |\Sigma_{k}^1(x^{[k:\mathbb Q]/2};1,m)|\gg_{k,m} x^{[k:\mathbb Q]/2}/\log x$, we deduce that \[ \frac{|B_{k,m}(x)|}{|B_k(x)|} \ll_{k,\norm{\cdot},m} \sum_{i=1}^n \big( x^{[k:\mathbb Q]/2}/\log x\big)^{-1} \ll_m \frac{\log x}{x^{[k:\mathbb Q]/2}}. \qedhere \] \end{proof} \subsection{Galois image modulo primes} Let $h$ be the absolute logarithmic height on $\mathbb P^1(\Qbar)$. For an elliptic curve $E$, let $j(E)$ be its $j$-invariant. The following theorem bounds the number of $\ell$'s that we need to consider (in particular, it gives an effective version of a result of Serre \cite{Serre-Inv72}). \begin{thm}[Masser-W\"ustholz \cite{MasserWustholz}] \label{T:MasserWustholz} Let $E$ be an elliptic curve defined over a number field $k$, and assume that $E$ does not have complex multiplication. There are positive absolute constants $c$ and $\gamma$ such that if $\ell > c \big( \max\big\{[k:\mathbb Q], h(j(E))\big\} \big)^\gamma$, then $\rho_{E,\ell}(G} %\newcommand{\calG}{\mathcal G_k) \supseteq \operatorname{SL}_2(\ZZ/\ell\ZZ).$ \end{thm} \begin{lemma} \label{L:j bound} If $(a,b) \in B_k(x)$, then $h(j(E(a,b)))\ll_{k,\norm{\cdot}} \log x$. \end{lemma} \begin{proof} Let $\Sigma_k^\infty$ be the set of archimedean places of $k$. For each $v\in \Sigma_k^\infty$, let $|\!\cdot\!|_v$ be an absolute value on the completion $k_v$ of $k$ at $v$. On $\prod_{v\in\Sigma_k^\infty} k_v^2$, we have a norm $\norm{ (a_v,b_v)_v }_1 = \sup_{v\in \Sigma_k^\infty} |a_v|_v + \sup_{v\in \Sigma_k^\infty} |b_v|_v.$ Using the natural isomorphism $\mathbb R\otimes_\ZZ \mathcal O_k^2 \cong \prod_{v\in\Sigma_k^\infty} k_v^2$, we may view $\norm{\cdot}_1$ as a norm on $\mathbb R\otimes_\ZZ \mathcal O_k^2$. Recall that $j(E(a,b)) = - 1728(4a)^3/ \Delta_{a,b}$. Since $a$ and $b$ are integral, we have \begin{align*} h(j(E(a,b))) & = h([- 1728(4a)^3 : \Delta_{a,b}] ) \ll_k {\sum}_{v\in \Sigma_k^\infty } \log( \max\{ 1728\cdot 4^3|a|^3_v, |\Delta_{a,b}|_v \} ) \end{align*} and thus $h(j(E(a,b))) \ll_k \log \norm{(a,b)}_1$. The norms $\norm{\cdot}$ and $\norm{\cdot}_1$ are equivalent, so \[ h(j(E(a,b))) \ll_k \log \norm{(a,b)}_1 \ll_{k,\norm{\cdot}} \log \norm{(a,b)} \leq \log x. \qedhere \] \end{proof} \begin{lemma} \label{L:MasserWustholz} There is a constant $c_k>0$ (depending only on $k$) and an absolute constant $\gamma>0$ such that \[ \big\{ (a,b) \in B_k(x): \rho_{E(a,b),\ell}(G} %\newcommand{\calG}{\mathcal G_k) \not\supseteq \operatorname{SL}_2(\ZZ/\ell\ZZ) \text{ for some prime $\ell\geq 5$} \big\} = \bigcup_{5\leq \ell \leq c_k (\log x)^\gamma} B_{k,\ell}(x). \] \end{lemma} \begin{proof} For an elliptic curve $E/k$ with complex multiplication, we have $\rho_{E,\ell}(G} %\newcommand{\calG}{\mathcal G_k) \not\supseteq \operatorname{SL}_2(\ZZ/\ell\ZZ)$ for every prime $\ell\geq 5$. The lemma follows by combining Theorem~\ref{T:MasserWustholz} and Lemma~\ref{L:j bound}. \end{proof} \begin{lemma} \label{L:exceptional prime bound 1} Assume that $5\leq \ell \leq c_k (\log x)^\gamma$, where $c_k$ and $\gamma$ are the constants from Lemma~\ref{L:MasserWustholz}. Then \[ \frac{|B_{k,\ell}(x)|}{|B_k(x)|} \ll_{k,\norm{\cdot}} \frac{ (\log x)^{7\gamma+1} }{x^{[k:\mathbb Q]/2}}. \] \end{lemma} \begin{proof} We may assume that $\ell$ satisfies $k\cap\mathbb Q(\mu_\ell)=\mathbb Q$ (this excludes only a finite number of $\ell$, which can be handled with Proposition~\ref{P:exceptional prime bound m}). Define the set $\Sigma^1_k(x) := \{ \mathfrak p \in \Sigma_k(x) : N(\mathfrak p) \text{ is prime}\}$. By the pigeonhole principle, there is an element $d\in (\ZZ/\ell\ZZ)^\times = \chi_k(G} %\newcommand{\calG}{\mathcal G_k) \bmod{\ell}$ such that \[ |\Sigma_{k}^1(x^{[k:\mathbb Q]/2};d,\ell)| \geq \frac{1}{\ell-1}|\Sigma^1_k(x^{[k:\mathbb Q]/2})| +O_k(1) \gg_k \frac{1}{\ell-1} x^{[k:\mathbb Q]/2}/\log x. \] Let $C_1,\dots, C_n$ be the conjugacy classes of $\GL_2(\ZZ/\ell\ZZ)$ with $\det(C_i)=d$. Combining Lemma~\ref{L:SL condition for ell} and Proposition~\ref{P:nasty sieving}, we have \begin{align*} \frac{|B_{k,\ell}(x)|}{|B_k(x)|} \leq \sum_{i=1}^n \frac{|Y_{C_i}(x)|}{|B_k(x)|} &\ll_{k,\norm{\cdot}} \sum_{i=1}^n \left(\frac{|C_i|}{|\operatorname{SL}_2(\ZZ/\ell\ZZ)|} |\Sigma_{k}^1(x^{[k:\mathbb Q]/2};d,\ell)| + O_k(\ell^5x^{[k:\mathbb Q]/4}) \right)^{-1} \\ &\ll_{k,\norm{\cdot}} \sum_{i=1}^n \left(\frac{|C_i|}{|\GL_2(\ZZ/\ell\ZZ)|} x^{[k:\mathbb Q]/2}/\log x + O_k(\ell^5x^{[k:\mathbb Q]/4})\right)^{-1}. \end{align*} The bounds $n\leq \ell^3$, $1\leq |C_i|$, and $\ell\leq c_k(\log x)^\gamma$ imply \[ \frac{|B_{k,\ell}(x)|}{|B_k(x)|} \ll_{k,\norm{\cdot}} n \ell^4 \frac{\log x}{x^{[k:\mathbb Q]/2}} \ll_k \frac{(\log x)^{7\gamma+1}}{x^{[k:\mathbb Q]/2}}. \qedhere \] \end{proof} \subsection{Proof of Proposition~\ref{P:l-adic surjectivity}} Using Lemmas~\ref{L:MasserWustholz} and \ref{L:exceptional prime bound 1}, we obtain the following bounds: \[ \frac{|\bigcup_{\ell \geq 5} B_{k,\ell}(x) |}{|B_k(x)|} \leq \sum_{5\leq \ell \leq c_k (\log x)^\gamma} \frac{|B_{k,\ell}(x)|}{|B_k(x)|} \ll _{k,\norm{\cdot}} \sum_{5\leq \ell \leq c_k (\log x)^\gamma} \frac{(\log x)^{7\gamma+1}}{ x^{[k:\mathbb Q]/2}} \ll_k \frac{(\log x)^{8\gamma + 1}}{ x^{[k:\mathbb Q]/2}}. \] By Proposition~\ref{P:exceptional prime bound m} (with $m=4$ and $9$), we have \[ \frac{|B_{k,4}(x) \cup B_{k,9}(x)|}{|B_k(x)|} \ll_{k,\norm{\cdot}} \frac{\log x}{x^{[k:\mathbb Q]/2}}. \] The proposition follows immediately with $\beta=8\gamma+1$. \section{Discriminants} \label{S:discriminants} \begin{prop} \label{P:Serre obstruction bound} Fix a number field $k\neq \mathbb Q$, an integer $r\geq 2$, and assume that $k$ contains $\mu_r$. Then \[ \frac{|\{ (a,b) \in B_k(x) : \sqrt[r]{\Delta_{a,b}} \in k^{\operatorname{cyc}} \}|}{|B_k(x)|} \ll_{k,\norm{\cdot},r} \frac{\log x}{\sqrt{x}}. \] \end{prop} \begin{remark} From Proposition~\ref{P:Serre obstruction bound}, we find that conditions (\ref{I:c}) and (\ref{I:d}) of Proposition~\ref{P:Criterion} hold for ``most'' elliptic curves over a fixed number field $k\neq \mathbb Q$. \end{remark} For the rest of this section, we shall fix $k$ and $r$ as in Proposition~\ref{P:Serre obstruction bound}. Let $d=[k:\mathbb Q]$. Let $S$ be the finite set of rational primes which satisfies the following conditions with minimal value $\prod_{p\in S} p$; \begin{itemize} \item $S$ contains the primes dividing $6r$, \item $S$ contains the primes that are ramified in $k$, \item $\mathcal O_S$ is a principal ideal domain, where $\mathcal O_S$ is the ring of $S'$-integers of $k$ and $S'=\{\mathfrak p \in \Sigma_k : \mathfrak p | p,\, \text{ for some } p\in S\}$. \end{itemize} Note that the above choice of $S$ depends only on $k$ and $r$. \begin{lemma} \label{L:ramification cyclotomic} Fix a prime $p \notin S$, an element $\Delta\in k^\times$, and let $\Pp_1\dots,\Pp_n$ be the prime ideals of $\mathcal O_{k(\sqrt[r]{\Delta})}$ lying over $p$. If $\sqrt[r]{\Delta} \in k^{\operatorname{cyc}}$, then \[ e(\Pp_1/p) = \dots = e(\Pp_n/p), \] where $e(\Pp_i/p)$ is the ramification index of $\Pp_i$ over $p$. \end{lemma} \begin{proof} Since $\sqrt[r]{\Delta} \in k^{\operatorname{cyc}}= \mathbb Q^{\operatorname{cyc}}\cdot k$, one can show that there is a field $L \subseteq \mathbb Q^{\operatorname{cyc}}$ such that $k(\sqrt[r]{\Delta}) = L \cdot k$. Since $p$ is unramified in $k$, we find that $e(\Pp_i/p) = e(\Pp_i \cap \mathcal O_L/p)$. The value $e(\Pp_i \cap \mathcal O_L/p)$ is independent of $i$, since $L$ is a Galois extension of $\mathbb Q$. \end{proof} \begin{lemma} \label{L:beta description} Let $\mathcal B \subseteq \mathcal O_S^\times$ be a set of representatives for the cosets of $\mathcal O_S^\times/(\mathcal O_S^\times)^r$. Then for any $\Delta \in \mathcal O_k$ with $\sqrt[r]{\Delta} \in k^{\operatorname{cyc}}$, there are $m \in \ZZ$, $\alpha \in \mathcal O_S$, and $\beta \in \mathcal B$ such that $\Delta= m \alpha^r \beta$. \end{lemma} \begin{proof} Fix $\Delta \in \mathcal O_k$ with $\sqrt[r]{\Delta} \in k^{\operatorname{cyc}}$. We first show that $\Delta$ can be written in the form $m\alpha^r \beta$, for some $m\in \ZZ$, $\alpha \in \mathcal O_S$, and $\beta \in \mathcal O_S^\times$. We may assume that $\Delta$ is non-zero. Since $\mathcal O_S$ is a principal ideal domain, there is an element $\alpha\in \mathcal O_S$ such that $0\leq \ord_\mathfrak p(\Delta/\alpha^r) < r$ for all $\mathfrak p \not\in S'$. Take any prime $p\notin S$ and let $\mathfrak p_1,\dots, \mathfrak p_g$ be the prime ideals of $\mathcal O_S$ lying over $p$. Suppose that some $\mathfrak p_i$ divides $\Delta/\alpha^r$ in $\mathcal O_S$. Since $0< \ord_\mathfrak p(\Delta/\alpha^r) < r$, we deduce that the extension $k(\sqrt[r]{\Delta})/k$ is ramified at $\mathfrak p_i$. By Lemma~\ref{L:ramification cyclotomic}, we find that $k(\sqrt[r]{\Delta})/k$ is ramified at all the primes $\mathfrak p_1,\dots,\mathfrak p_g$, and hence $p\mathcal O_S = \mathfrak p_1\dots \mathfrak p_g$ divides $\Delta/\alpha^r$ in $\mathcal O_S$. Dividing by $p$ and repeating the above process, we find that there is an integer $m\geq 1$ such that $\beta:=\Delta/(m\alpha^r)$ is an element of $\mathcal O_S^\times$. We may assume that $\beta$ is in $\mathcal B$ after multiplying $\alpha$ by an appropriate element of $\mathcal O_S^\times$. \end{proof} For each $\beta\in \mathcal O_S^\times$, define the sets \[ W_\beta := \{ (a,b) \in \mathcal O_k^2: \Delta_{a,b} = m \alpha^r \beta, \text{ for some } m \in \ZZ,\, \alpha \in \mathcal O_S \} \] and $W_\beta(x) := W_\beta \cap B_k(x)$. For a set $\mathcal B$ as in Lemma~\ref{L:beta description}, we have \[ \{ (a,b) \in B_k(x) : \sqrt[r]{\Delta_{a,b}} \in k^{\operatorname{cyc}} \} \subseteq \bigcup_{\beta \in \mathcal B} W_\beta(x). \] The set $\mathcal B$ is finite (since the abelian group $\mathcal O_S^\times$ is finitely generated), so \begin{equation} \label{E: W bound for discriminant} |\{ (a,b) \in B_k(x) : \sqrt[r]{\Delta_{a,b}} \in k^{\operatorname{cyc}} \}| \ll_{k,r} \max_{\beta \in \mathcal O_S^\times} |W_\beta(x)|. \end{equation} Thus to prove Proposition~\ref{P:Serre obstruction bound}, it suffices to find bounds for the functions $|W_\beta(x)|$. \begin{lemma} \label{L:Weil conjecture bound} Let $p\nmid 6$ be a prime with $p\equiv 1 \bmod{r}$. For any $\gamma\in \mathbb F_p^\times$, \[ |\{(a,b) \in \mathbb F_p^2 : \Delta_{a,b} = \gamma c^r, \, \text{for some $c\in\mathbb F_p$}\}| = \frac{1}{r} p^2 + O_{r}(p^{3/2}). \] \end{lemma} \begin{proof} Fix $\gamma\in \mathbb F_p^\times$. The equation $\Delta_{a,b} = \gamma c^r$ defines a geometrically irreducible variety $X$ in $\AA^3_{\mathbb F_p}=\Spec(\mathbb F_p[a,b,c])$. Using the Weil conjectures, we find that \begin{equation} \label{E:Weil conjecture} |X(\mathbb F_p)| = p^2 + O_r(p^{3/2}) \end{equation} (that the implicit constant in (\ref{E:Weil conjecture}) depends only on $r$ can be deduced from \cite{Bombieri}). For fixed $(a,b)\in \mathbb F_p^2$, if $ \Delta_{a,b} = \gamma c^r$ has a solution $c\in \mathbb F_p^\times$, then it has exactly $r$ such solutions (this uses the assumption $p\equiv 1\bmod{r}$). Most solutions have $c\neq 0$, since $|\{(a,b)\in\mathbb F_p^2: \Delta_{a,b}=0\}| \ll p$. The lemma is now immediate. \end{proof} \begin{lemma} \label{L:W mod p bound} Take any $\beta \in \mathcal O_S^\times$. Let $p \not\in S$ be a prime that splits completely in $k$, and let $W_{\beta,p}$ be the image of $W_\beta$ in $\mathcal O_k^2/p\mathcal O_k^2$. Then \[ |W_{\beta,p}| \leq \Big( \frac{1}{r^{d-1}} + O_{r,d}(p^{-1/2}) \Big) |\mathcal O_k^2/p\mathcal O_k^2|. \] \end{lemma} \begin{proof} Let $\mathfrak p_1,\dots, \mathfrak p_d \in \Sigma_k$ be the prime ideals lying over $p$. By the Chinese remainder theorem, we have a natural identification $\mathcal O_k/p\mathcal O_k = \prod_{i=1}^d \mathbb F_{\mathfrak p_i}$. Then \[ W_{\beta,p} \subseteq \bigcup_{ m \in R } \prod_{i=1}^d \big\{(a,b) \in \mathbb F_{\mathfrak p_i}^2 : \Delta_{a,b}= m \alpha^r \cdot (\beta \bmod{\mathfrak p_i}),\, \text{for some } \alpha\in \mathbb F_{\mathfrak p_i} \big\}, \] where the union is over a set of coset representatives $R\subseteq \mathbb F_p^\times$ of $\mathbb F_p^\times/(\mathbb F_p^\times)^r$. We have $p\equiv 1 \bmod{r}$, since $p$ splits completely in $k$ and by assumption $\mu_r\subseteq k$. By Lemma~\ref{L:Weil conjecture bound}, \[ |W_{\beta,p}| \leq |R| \big(p^2/r + O_{r}(p^{3/2})\big)^d = p^{2d}/{r^{d-1}} + O_{r,d}(p^{2d-1/2}). \qedhere \] \end{proof} \begin{lemma} \label{L:W bound} For $\beta\in \mathcal O_S^\times$, ${|W_\beta(x)|} \ll_{k,\norm{\cdot},r} {|B_k(x)|} (\log x)/{\sqrt{x}}.$ \end{lemma} \begin{proof} Let $I$ be the set of primes $p\not\in S$ that split completely in $k$. By Lemma~\ref{L:W mod p bound}, for each prime $p \in I$, we have $|W_\beta(x) \bmod{p\mathcal O_k^2}| \leq (1-\omega_p) |\mathcal O_k^2/p\mathcal O_k^2|$, where $\omega_p = 1-1/r^{d-1} + O_{r,d}(p^{-1/2})$. For $p\notin I$, set $\omega_p=0$. We may now apply the large sieve. By Theorem~\ref{T:large sieve} (with $K=\mathbb Q$, $\Lambda=\mathcal O_k^2$, $Q=\sqrt{x}$, and our chosen norm $\norm{\cdot}$ on $\mathbb R\otimes_\ZZ \Lambda$), we have $|W_\beta(x)| \ll_{k,\norm{\cdot}} { x^{2d}}/{L(\sqrt{x})},$ where \[ L(\sqrt{x}) := \sum_{ \substack{J \subseteq I \text{ finite} \\ \prod_{p\in J} p \leq \sqrt{x}} } \prod_{p\in J} \frac{\omega_p}{1-\omega_p}. \] Using $r\geq 2$ and $d\geq 2$ (since $k\neq \mathbb Q$), we have the bound \[ L(\sqrt{x}) \geq \sum_{ \substack{J \subseteq I \text{ finite} \\ \prod_{p\in J} p \leq \sqrt{x} } } \prod_{p\in J} \Big( 1 + O_{r,d}(p^{-1/2}) \Big) \geq \sum_{ \substack{p\in I \\ p \leq \sqrt{x} } } \Big( 1 + O_{r,d}(p^{-1/2}) \Big). \] The set $I$ has positive density in the primes, so $L(\sqrt{x}) \gg_{r,k} \sqrt{x}/\log x$. The lemma follows by using this bound for $L(\sqrt{x})$ and (\ref{E:box asymptotics}) with our upper bound for $|W_\beta(x)|$. \end{proof} \begin{proof}[Proof of Proposition~\ref{P:Serre obstruction bound}] Apply Lemma~\ref{L:W bound} to the bound (\ref{E: W bound for discriminant}). \end{proof} \section{Elliptic curves with maximal Galois action} \label{S:proof} \subsection{Proof of Theorem~\ref{T:main theorem 2}} Define the sets \begin{align*} Y_1(x ) &= B_{k,4}(x) \cup B_{k,9}(x) \cup {\bigcup}_{\ell\geq 5}B_{k,\ell}(x),\\ Y_2(x) &= \{ (a,b) \in B_k(x) : \sqrt{\Delta_{a,b}} \in k^{\operatorname{cyc}} \}, \\ Y_3(x) &= \{ (a,b) \in B_k(x) : \mu_3 \subseteq k \text{ and } \sqrt[3]{\Delta_{a,b}} \in k^{\operatorname{cyc}}\}. \end{align*} By Proposition~\ref{P:Criterion}, we have $\{(a,b) \in B_k(x): \rho_{E(a,b)}(G} %\newcommand{\calG}{\mathcal G_k) \neq H_k \} \subseteq Y_1(x) \cup Y_2(x) \cup Y_3(x)$, and thus \[ |\{(a,b) \in B_k(x): \rho_{E(a,b)}(G} %\newcommand{\calG}{\mathcal G_k) \neq H_k \}| \leq |Y_1(x)| + |Y_2(x)| + |Y_3(x)|. \] By Proposition \ref{P:l-adic surjectivity}, we have ${|Y_1(x)|}/{|B_k(x)|} \ll_{k,\norm{\cdot}} {(\log x)^{\beta}}/{x^{[k:\mathbb Q]/2}}$, where $\beta\geq 1$ is an absolute constant. By Proposition~\ref{P:Serre obstruction bound}, we have \[ \frac{|Y_2(x)|}{|B_k(x)|} \ll_{k,\norm{\cdot}} \frac{\log x}{\sqrt{x}} \text{\quad and \quad} \frac{|Y_3(x)|}{|B_k(x)|} \ll_{k,\norm{\cdot}} \frac{\log x}{\sqrt{x}}. \] Combining everything together gives: \[ \frac{|\{(a,b) \in B_k(x): \rho_{E(a,b)}(G} %\newcommand{\calG}{\mathcal G_k) \neq H_k \}|}{|B_k(x)|} \ll_{k,\norm{\cdot}} \max\Big\{ \frac{(\log x)^{\beta}}{x^{[k:\mathbb Q]/2}}, \frac{\log x}{\sqrt x} \Big\} \ll \frac{\log x}{\sqrt{x}}, \] where the last bound uses $k\neq \mathbb Q$. \subsection{Proof of Theorem~\ref{T:Jones-Serre curves}} \label{SS:Serre curves} The theorem is easily deduced by combining the criterion of Lemma~\ref{L:Serre curve criterion} with Proposition~\ref{P:l-adic surjectivity} and Proposition~\ref{P:exceptional prime bound m} (with $m=72$).
{ "timestamp": "2008-09-20T05:21:53", "yymm": "0809", "arxiv_id": "0809.3482", "language": "en", "url": "https://arxiv.org/abs/0809.3482", "abstract": "Given an elliptic curve E over a number field k, the Galois action on the torsion points of E induces a Galois representation, \\rho_E : Gal(\\bar{k}/k) \\to GL_2(\\hat{Z}). For a fixed number field k, we describe the image of \\rho_E for a \"random\" elliptic curve E over k. In particular, if k\\neq Q is linearly disjoint from the cyclotomic extension of Q, then \\rho_E will be surjective for \"most\" elliptic curves over k.", "subjects": "Number Theory (math.NT)", "title": "Elliptic curves with maximal Galois action on their torsion points", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9895109075027804, "lm_q2_score": 0.8104789063814616, "lm_q1q2_score": 0.801977718165381 }
https://arxiv.org/abs/1802.01621
Un-unzippable Convex Caps
An unzipping of a polyhedron P is a cut-path through its vertices that unfolds P to a non-overlapping shape in the plane. It is an open problem to decide if every convex P has an unzipping. Here we show that there are nearly flat convex caps that have no unzipping. A convex cap is a "top" portion of a convex polyhedron; it has a boundary, i.e., it is not closed by a base.
\section{Introduction} \seclab{Introduction} We define an \emph{unzipping} of a polyhedron ${\mathcal P}$ in ${\mathbb{R}}^3$ to be a non-overlapping, single-piece unfolding of the surface to the plane that results from cutting a continuous path ${\gamma}$ through all the vertices of ${\mathcal P}$. The cut-path ${\gamma}$ need not follow the edges of ${\mathcal P}$, nor even be polygonal, but it must include every vertex of ${\mathcal P}$, passing through $n-2$ vertices and beginning and ending at the other two vertices, where $n$ is the total number of vertices. If ${\gamma}$ does follow edges of ${\mathcal P}$, we call it an \emph{edge-unzipping}. Edge-unzippings are special cases of edge-unfoldings, where the cuts follow a tree of edges that span the $n$ vertices. The interest in edge-unfoldings stems largely from what has become known as D\"urer's problem~\cite{do-gfalop-07}~\cite{o-dp-13}: Does every convex polyhedron have an edge-unfolding? The emphasis here is on a non-overlapping result, what is often called a \emph{net} for the polyhedron. This question was first formally raised by Shephard in~\cite{s-cpcn-75}. In that paper, he already investigated the special case where the cut edges form a Hamiltonian path of the $1$-skeleton of ${\mathcal P}$: Hamiltonian unfoldings. These are exactly what I'm calling edge-unzippings. Shephard noted that the rhombic dodecahedron does not have an edge-unzipping because its $1$-skeleton has no Hamiltonian path. The attractive ``zipping'' terminology stems from the paper~\cite{lddss-zupc-10}, which defined \emph{zipper unfoldings} to be what I'm shortening to unzippings. They showed that all the Platonic and the Archimedean solids have edge-unzippings. And they posed a fascinating question: \begin{center} \fbox{\textbf{Open Problem}: Does every convex polyhedron have an unzipping?} \end{center} \subsection{Nonconvex Polyhedra} \seclab{NonconvexPolyhedra} First we note that not every nonconvex polyhedron has an unzipping. This has been a ``folk theorem'' for years, but has apparently not been explicitly stated in the literature.\footnote{ The closest is~\cite{bdekms-upcf-03}, which notes that ``the neighborhood of a negative-curvature vertex ... requires two or more cuts to avoid self-overlap.''} In any case, it is not difficult to see. Consider the polyhedron illustrated in Fig.~\figref{heartsfan}. The central vertex $v$ has more than $4 \pi$ incident surface angle. In fact, it has well more than $8 \pi$ incident angle, but we only need $> 4 \pi$. An unzipping cut-path ${\gamma}$ cannot terminate at $v$, because the neighbhood of $v$ in the unfolding has more than $2\pi$ incident angle, and so would overlap in the planar development. Nor can ${\gamma}$ pass through $v$, because partitioning the $> 4 \pi$ angle would leave more than $2\pi$ to one side or the other, again forcing overlap in the neighborhood of at least one of the two planar images of $v$. Therefore, no polyhedron with a vertex with more than $> 4 \pi$ incident angle has an unzipping. Indeed, as Stefan Langerman observed,\footnote{ Personal communication, Aug. 2017} similar reasoning shows that for any degree ${\delta}$ there is a polyhedron that cannot be unfolded without overlap by a cut tree of maximum degree ${\delta}$. The polyhedron in Fig.~\figref{heartsfan} requires degree $> 4$ at $v$ to partition the more than $8 \pi$ angle into $< 2\pi$ pieces. \begin{figure}[htbp] \centering \includegraphics[width=0.4\linewidth]{Figures/heartsfan} \caption{A polyhedron that cannot be unzipped. Based on Fig.~24.14, p.370 in~\protect\cite{do-gfalop-07}.} \figlab{heartsfan} \end{figure} \subsection{Open Problem: Conjecture} This negative result for nonconvex polyhedra increases the interest in the open problem for convex polyhedra. In~\cite{o2015spiral} I conjectured the answer is {\sc no}, but it seems far from clear how to settle the problem. For that reason, here we turn to a very special case. \subsection{Convex Caps} \seclab{ConvexCaps} The special case is unzipping ``convex caps.'' I quote the definition from~\cite{o-eunfcc-17}: \begin{quotation} \noindent ``Let ${\mathcal P}$ be a convex polyhedron, and let $\phi(f)$ be the angle the normal to face $f$ makes with the $z$-axis. Let $H$ be a halfspace whose bounding plane is orthogonal to the $z$-axis, and includes points vertically above that plane. Define a \emph{convex cap} ${\mathcal C}$ of angle ${\Phi}$ to be $C={\mathcal P} \cap H$ for some ${\mathcal P}$ and $H$, such that ${\phi}(f) \le {\Phi}$ for all $f$ in ${\mathcal C}$. [...] Note that ${\mathcal C}$ is not a closed polyhedron; it has no ``bottom,'' but rather a boundary ${\partial \mathcal C}$.'' \end{quotation} \noindent The result of this note is: \begin{theorem} For any ${\Phi} > 0$, there is a convex cap ${\mathcal C}$ that has no unzipping. \thmlab{ZipCex} \end{theorem} \noindent Because this holds for any ${\Phi} > 0$, there are arbitrarily flat convex caps that cannot be unzipped. (${\Phi}$ will not otherwise play a role in the proof.) \section{Proof of Theorem~1} \seclab{proof} The convex caps used to prove the theorem are all variations on the cap shown in Fig.~\figref{ZipCex3D}. The base ${\partial \mathcal C} = (b_1,b_2,b_3)$ forms a unit side-length equilateral triangle in the $xy$-plane. The three vertices $a_1,a_2,a_3$ are also the corners of an equilateral triangle, lifted a small amount $z_a$ above the base. In projection to the the $xy$-plane, the ``apron'' of quadrilaterals between $\triangle b_1 b_2 b_3$ and $\triangle a_1 a_2 a_3$ has width ${\varepsilon} > 0$. The vertex $c$, at height $z_c > z_a$, sits over the centroids of the equilateral triangles. The shape of the cap is controlled by three parameters: ${\varepsilon}, z_a, z_b$. Keeping ${\varepsilon}$ fixed and varying $z_a$ and $z_c$ permits controlling the curvatures ${\omega}_a$ at $a_i$ and ${\omega}_c$ at $c$. In Fig.~\figref{ZipCex3D}, ${\varepsilon}=0.1$ and $z_a,z_c = 0.02, 0.1$ leads to ${\omega}_a=1.9^\circ$ and ${\omega}_c=5.6^\circ$. \begin{figure}[htbp] \centering \includegraphics[width=0.75\linewidth]{Figures/ZipCex3D} \caption{A convex cap ${\mathcal C}$ that has no unzipping.} \figlab{ZipCex3D} \end{figure} A typical attempt at an unzipping (of a variant of Fig.~\figref{ZipCex3D}) is shown in Fig.~\figref{Layout_caaa}. In general we will only display what are labeled $L$ and $R$ in this figure, rather than the full unfolding. \begin{figure}[htbp] \centering \includegraphics[width=0.6\linewidth]{Figures/Layout_caaa} \caption{An overlapping unfolding of a convex cap (a variant of Fig.~\protect\figref{ZipCex3D}) from cut-path ${\gamma} =(c, a_2, a_3, a_1, b_1)$. Compare Fig.~\protect\figref{Devel_caaa_linear_5_10} ahead.} \figlab{Layout_caaa} \end{figure} From now on we will illustrate cut-paths and unzippings in the plane, starting from Fig.~\figref{EqTris} (and not always repeating all the labels). \begin{figure}[htbp] \centering \includegraphics[width=0.5\linewidth]{Figures/EqTris} \caption{Projection of Fig.~\protect\figref{ZipCex3D} to $xy$-plane.} \figlab{EqTris} \end{figure} \subsection{Constraints on the cut-path} \seclab{Constraints} Any point $p$ in the relative interior of ${\gamma}$ (i.e., not an endpoint) develops in the plane to two points $p'$ and $p''$, with right and left incident surface angles ${\rho}={\rho}(p)$ and ${\lambda}={\lambda}(p)$. If $p$ is not at a vertex of ${\mathcal C}$, then ${\lambda}+{\rho}=2\pi$. If $p$ is at a vertex of curvature ${\omega}={\omega}(p)$, then ${\lambda}+{\rho}+{\omega}=2\pi$. We will show the development ${\mathcal C}$ as cut by ${\gamma}$ by drawing two directed paths $R$ and $L$, each determined by the ${\rho}$ and ${\lambda}$ angles, which deviate by ${\omega}(v)$ at each vertex $v \in {\gamma}$. The surface of ${\mathcal C}$ is right of $R$ and left of $L$ (see Fig.~\figref{Layout_caaa}), but not explicitly depicted in subsequent figures. The constraints on ${\gamma}$ to be an unzipping are: \begin{enumerate} \setlength{\itemsep}{0pt} \item ${\gamma}$ must be a path, by definition of unzipping. \item ${\gamma}$ must start at one of the vertices $\{c,a_1,a_2,a_3\}$ and terminate on ${\partial \mathcal C}$. \item ${\gamma}$ does not have to include any of the vertices $\{b_1,b_2,b_3\}$, it just needs to exit ${\mathcal C}$ at some point of ${\partial \mathcal C}$. \item ${\gamma}$ can only touch ${\partial \mathcal C}$ at one point, for if it touches at two or more points, the unfolding would be disconnected into more than one piece. \item Between vertices, ${\gamma}$ can follow any path on ${\mathcal C}$, as long as ${\gamma}$ does not self-cross, which would again result in more than one piece. \item And of course, the developments of $R$ and $L$ must not cross in the plane, for $R / L$ crossings imply overlap.\footnote{ The reverse is not always true: It could be that $R$ and $L$ do not cross, but other portions of the surface away from the cut ${\gamma}$ are forced to overlap by, for example, large curvature openings.} \end{enumerate} We think of ${\gamma}$ as directed from its root start vertex to ${\partial \mathcal C}$; the path opens from the root to its boundary exit. The main constraint we exploit is item~4: ${\gamma}$ can only touch ${\partial \mathcal C}$ at one point. We will see that only by leaving ${\mathcal C}$ and returning could the unzipping avoid overlap. Due to the symmetry of ${\mathcal C}$---in particular, the equivalence of $\{a_1,a_2,a_3\}$---there are only four combinatorially distinct possible cut-paths ${\gamma}$, where we use $b$ to represent any point on ${\partial \mathcal C}$: \begin{enumerate} \setlength{\itemsep}{0pt} \item ${\gamma} = (c, a_1, a_2, a_3, b) = caaab$. \item ${\gamma} = (a_1, c, a_2, a_3, b) = acaab$. \item ${\gamma} = (a_1, a_2, c, a_3, b) = aacab$. \item ${\gamma} = (a_1, a_2, a_3, c, b) = aaacb$. \end{enumerate} We abbreviate the path structure with strings $acaab$ and so on, with the obvious meaning. It turns out that the location of $b$, the point at which ${\gamma}$ exits ${\mathcal C}$, plays little role in the proof. We will display the structure of ${\gamma}$ and the developments of $R$ and $L$ as in Fig.~\figref{Path_acaa_linear}. \begin{figure}[htbp] \centering \includegraphics[width=0.75\linewidth]{Figures/Path_acaa_linear} \caption{% Left: path ${\gamma} = (a_1, c, a_2, a_3, b_3) = acaab$. Right: $R$ and $L$ developed. $\{{\omega}_a,{\omega}_c\}=\{5^\circ,10^\circ\}$.} \figlab{Path_acaa_linear} \end{figure} Here ${\gamma}$ is shown following straight segments between vertices, and the developments overlap substantially. But as per item~5 above, ${\gamma}$ can follow potentially any (non-self-intersecting) curve between vertices. However, the developed images of the vertices are independent of the shape of the path between vertices, a condition we exploit in the proof. So once the combinatorial structure of the cut ${\gamma}$ is fixed, the developed locations of the vertex images are determined. We will continue to use $\{{\omega}_a,{\omega}_c\}=\{5^\circ,10^\circ\}$ for illustration, although any smaller curvatures also work in the proofs. \subsection{Radial Monotonicity: Intuition} \seclab{RM} Before beginning the proof details, we provide the intuition behind it. That intuition depends on the notion of a ``radially monotone'' curve, a concept used in~\cite{o-ucprm-16} and~\cite{o-eunfcc-17}. A directed polygonal chain $P$ in the plane with vertices $u_1, u_2, \ldots, u_k$ is \emph{radially monotone with respect to $u_1$} if the distance from $u_1$ to every point $p \in P$ increases monotonically as $p$ moves out along the chain. $P$ is \emph{radially monotone} if it is radially monotone with respect to each vertex $u_i$: concentric circles centered on each $u_i$ are crossed just once by the chain beyond $u_i$. If both the $R$ and $L$ developments are radially monotone, then $L$ and $R$ do not intersect except at their common ``root'' vertex, a fact proved in the cited papers.\footnote{ There are some curvature bound assumptions to this claim that are not relevant here.} This suggests that ${\gamma}$ should be chosen so that $R$ and $L$ are radially monotone. However, if $R$ or $L$ or both are not radially monotone, they do not necessarily overlap: radial monotonicity is sufficient for non-overlap but not necessary. Nevertheless, striving for radial monotonicity makes sense. The sharp turns necessary to span the vertices of ${\mathcal C}$ (visible in Fig.~\figref{Path_acaa_linear}) should be avoided, for they violate radial monotonicity. (Any angle $\angle u_{i-1}, u_i, u_{i+1}$ smaller than $90^\circ$ implies non-monotonicity at $u_i$ with respect to $u_{i-1}$.) Avoiding these sharp turns forces ${\gamma}$ to exit ${\mathcal C}$ before spanning the vertices. Although radial monotonicity is not used in the proofs to follow, it is the intuition behind the proofs. \subsection{Lemmas~1,2,3,4} \seclab{Lemmas} Of the four possible types of ${\gamma}$, $acaab$ is the ``closest'' to being unzippable, so we start with this type. \begin{lemma} For sufficiently small ${\omega}_a$, ${\omega}_c$, and ${\varepsilon}$, any cut-path ${\gamma}$ of type $acaab$ must leave and reenter ${\mathcal C}$ to avoid overlap. Therefore, ${\mathcal C}$ cannot be unzipped with this type of cut-path. \lemlab{acaab} \end{lemma} \begin{proof} We have already seen in Fig.~\figref{Path_acaa_linear} that straight connections between the vertices leads to overlap. Fig.~\figref{RMlogic4frames}(a) repeats the set-up of that figure, with added notation. Let $R_1,R_2,R_3$ be the portions of the right development $R$ between vertices, and similarly for $L_i$. We now imagine that $R_i$ and $L_i$ are arbitrary cuts between their vertex endpoints. We concentrate on $R_3$ and $L_3$. From the fact that the images of the vertices, and in particular, $a_2$, are in their correct developed planar locations, we can derive constraints on the shape of the $R_3$ and $L_3$ paths. The shape of $R_i$ determines $L_i$ and vice versa, because for all non-vertex points of ${\gamma}$, ${\rho}+{\lambda} = 2\pi$. Thus $R_i$ and $L_i$ are congruent as curves, but rigidly rotated differently by the curvatures along ${\gamma}$. There are only two topological possibilities for $R_3$ and $L_3$ to avoid crossing earlier portions of $R$ and $L$, illustrated in Fig.~\figref{RMlogicTopology}. In~(a) of the figure, $R_3$ passes right of $a''_2$ on its way counterclockwise to $a'_3$, and in~(b), $L_3$ passes right of $a'_2$ on its way clockwise to $a''_3$. The situations are analogous in the neighborhood of $a_2$, and we concentrate only on the former more direct route. Knowing that $R_3$ passes to the right of $a''_2$ determines the vector displacement of the tightest possible prefix $(a'_2,a''_2)$ of $R_3$, but not the shape of that prefix. This vector displacement forms an effective angle $\angle c',a'_2,a''_2$ of much larger than the near-$30^\circ$ necessary to stay on the narrow ${\varepsilon}$-apron. Fig.~\figref{RMlogic4frames}(a) and~(b) show that this angle is nearly $70^\circ$, well beyond $30^\circ$. (And the angle is larger if $R_3$ passes further to the right of $a''_2$.) This $70^\circ$ turn implies an effective surface angle ${\rho} = 290^\circ$ to the right of ${\gamma}$ on ${\mathcal C}$ at $a_2$, ``effective'' because the exact shape of ${\gamma}$ is unknown. The exact angle and length of vector displacement depend on $\{{\omega}_a,{\omega}_c\}$, but for any given curvatures, we can choose an ${\varepsilon}$ small enough so that the prefix steps ${\gamma}$ exterior to ${\mathcal C}$. Thus ${\gamma}$ must leave ${\mathcal C}$ to avoid overlap before it completes its tour of the vertices. Although this proves the lemma, we continue the analysis below to reveal a deeper structure. \end{proof} \begin{figure}[htbp] \centering \includegraphics[width=0.75\linewidth]{Figures/RMlogic4frames} \caption{Analysis of ${\gamma}$ of type $acaab$. (a)~Opening at $a_1$ and $c$ causes $R_3 / L_3$ overlap. (b)~$R_3$ bends around $a''_2$. (c)~$L_3$ complements $R_3$, which again intersects $L_3$. (d)~$L_3$ complements $R_3$. $R_3$ is following the arc centered on $x$.} \figlab{RMlogic4frames} \end{figure} \begin{figure}[htbp] \centering \includegraphics[width=0.85\linewidth]{Figures/RMlogicTopology} \caption{Possible paths from $a_2$ to $a_3$. (a)~$R_3$ skirts $a''_2$. (b)~$L_3$ skirts $a'_2$.} \figlab{RMlogicTopology} \end{figure} Knowing this constraint just derived on the prefix of $R_3$, we know $L_3$ must complement $R_3$ on that prefix, which leads to Fig.~\figref{RMlogic4frames}(c). Again there is an overlap intersection further along $R_3$. Altering $R_3$ to again bend around $L_3$ leads to Fig.~\figref{RMlogic4frames}(d). Continuing this process of incrementally determining constraints on $R_3$ that are mirrored in $L_3$, leads to the conclusion that $R_3$ must follow (or be outside of) the arc of a circle centered at $x$, where $x$ is the combined center of rotation of the rotations at $a_1$ and at $c$. This is why we can be sure the angle at $a_2$ is well beyond $30^\circ$ independent of $\{{\omega}_a,{\omega}_c\}$: it is determined by (an approximation to) the tangent to this circle. Finally, $R_3$ and $L_3$ are opened further by the curvature ${\omega}_a$ at $a_2$, which does not alter the previous analysis. Here we pause to discuss the ``combined center of rotation'' just used. Any pair of rotations about two distinct points is equivalent to a single rotation about a combined center. In our situation, the two rotations are ${\omega}_a$ about $a_1$ and ${\omega}_c$ about $c$. For small rotations, they are equivalent to a rotation by ${\omega}_a+{\omega}_c$ about the weighted center $$ x= \frac{ {\omega}_a a_1 + {\omega}_c c }{ {\omega}_a+{\omega}_c} \;. $$ This point is indicated in Figs.~\figref{RMlogic4frames}(a) and~(d). This result on combining rotations is proved in both~\cite{o2017addendum} and~\cite{barvinok2017pseudo} (and likely elsewhere). The above analysis suggests that ${\mathcal C}$ can be unzipped if the apron were large enough to include the circle arc that ${\gamma}$ must follow from $a_2$ to $a_3$. And indeed Fig.~\figref{Devel_acaa_arc_5_10} shows that this is true. An interesting consequence of this unzipping is that, even with a small apron, if we close the convex cap ${\mathcal C}$ by adding an equilateral triangle base to form a closed convex polyhedron ${\mathcal P}$, then ${\mathcal P}$ does have an unzipping. Follow the path shown in Fig.~\figref{Devel_acaa_arc_5_10}, and complete it by extending ${\gamma}$ to cut $(b_3, b_1, b_2)$, leaving $b_2 b_3$ uncut. Then the arc illustrated would lie on the unfolding of the base $\triangle b_1 b_2 b_3$. \begin{figure}[htbp] \centering \includegraphics[width=0.75\linewidth]{Figures/Devel_acaa_arc_5_10} \caption{An $acaab$ unzipping of ${\mathcal C}$ extending outside ${\partial \mathcal C}$. Right: The $R$ and $L$ developments do not cross.} \figlab{Devel_acaa_arc_5_10} \end{figure} We now turn to the other three types of cut-paths ${\gamma}$. The proof for type $caaab$ is similar to Lemma~\lemref{acaab}, and so will only be sketched. \begin{lemma} For sufficiently small ${\omega}_a$, ${\omega}_c$, and ${\varepsilon}$, any cut-path ${\gamma}$ of type $caaab$ must leave and reenter ${\mathcal C}$ to avoid overlap. Therefore, ${\mathcal C}$ cannot be unzipped with this type of cut-path. \lemlab{caaab} \end{lemma} \begin{proof} The cut-path with straight segments overlaps at two spots in development, as shown in Fig.~\figref{Devel_caaa_linear_5_10} (cf.~Fig.~\figref{Layout_caaa}). Using the same reasoning as in Lemma~\lemref{acaab}, except that the rotations are centered on $c$ (rather than on both $a_1$ and $c$), leads to the conclusion that $R_2$ and $R_3$ must both deviate from the $30^\circ$ turn at $a_2$ and the $60^\circ$ turn at $a_3$ needed to stay on an arbitrarily thin apron. In fact, $R_2$ and $R_3$ must follow circle arcs centered on $c$. Doing so would in fact allow ${\mathcal C}$ to be unzipped if apron were large enough, as shown in Fig.~\figref{Devel_caaa_arcs_5_10}. But for an arbitrarily thin ${\varepsilon}$-apron, ${\gamma}$ must exit ${\mathcal C}$ before visiting all vertices, and so cannot be unzipped with this type of cut-path. \end{proof} \begin{figure}[htbp] \centering \includegraphics[width=0.75\linewidth]{Figures/Devel_caaa_linear_5_10} \caption{The cut-path type $caaab$ leads to overlap with straight segments.} \figlab{Devel_caaa_linear_5_10} \end{figure} \begin{figure}[htbp] \centering \includegraphics[width=0.75\linewidth]{Figures/Devel_caaa_arcs_5_10} \caption{An $caaab$ unzipping of ${\mathcal C}$ extending outside ${\partial \mathcal C}$. Right: The $R$ and $L$ developments do not cross.} \figlab{Devel_caaa_arcs_5_10} \end{figure} The third type of cut-path, $aacab$ (Fig.~\figref{Devel_aaca_linear_5_10}), is different in that not even following arcs outside of ${\mathcal C}$ would suffice to unzip it without overlap. \begin{figure}[htbp] \centering \includegraphics[width=0.75\linewidth]{Figures/Devel_aaca_linear_5_10} \caption{The cut-path type $aacab$ leads to $L / R$ overlap.} \figlab{Devel_aaca_linear_5_10} \end{figure} \begin{lemma} For sufficiently small ${\omega}_a$, ${\omega}_c$, and ${\varepsilon}$, any cut-path ${\gamma}$ of type $aacab$ cannot visit all vertices without overlap in the development. Therefore, ${\mathcal C}$ cannot be unzipped with this type of cut-path. \lemlab{aacab} \end{lemma} \begin{proof} We analyze the constraints on $R_2$ in Fig.~\figref{RMlogic_aaca}. The rotation at $a_1$ determines the prefix of $R_2$ following the same reasoning as in Lemma~\lemref{acaab}, and again already in Fig.~\figref{RMlogic_aaca}(b) we have an angle at $a_2$ much larger than the $30^\circ$ turn required to reach $c$. This already establishes the cut-path cannot be an unzipping. But in fact, it is clear that $R_2$ must follow the circle arc shown in Fig.~\figref{RMlogic_aaca}(d), centered on $a_1$. Following this arc makes it impossible for $R_2$ to reach $c$: the path is forced toward $a_3$ instead. \end{proof} \begin{figure}[htbp] \centering \includegraphics[width=0.75\linewidth]{Figures/RMlogic_aaca} \caption{Analysis of ${\gamma}$ of type $aacab$. (a)~Opening at $a_1$ causes $R_2 / L_2$ overlap. (b)~$R_2$ bends around $a''_2$. (c)~$L_2$ complements $R_2$, which again intersects $L_2$. (d)~$L_2$ complements $R_2$. $R_2$ is following the arc centered on $a_1$.} \figlab{RMlogic_aaca} \end{figure} \noindent The last combinatorial cut-path type, $aaacb$, mixes themes in the others: first, ${\gamma}$ must go outside ${\mathcal C}$, which already establishes there is no unzipping, and second, even if the apron were large enough, ${\gamma}$ cannot reach $c$. We rely just on the first impediment. \begin{lemma} For sufficiently small ${\omega}_a$, ${\omega}_c$, and ${\varepsilon}$, any cut-path ${\gamma}$ of type $aaacb$ cannot visit all vertices without leaving and re-entering ${\mathcal C}$. Therefore, ${\mathcal C}$ cannot be unzipped with this type of cut-path. \lemlab{aaacb} \end{lemma} \begin{proof} Fig.~\figref{Devel_aaac_linear_5_10} shows there is overlap when ${\gamma}$ is composed of straight segments. By now familiar reasoning, the portion of ${\gamma}$ from $a_2$ to $a_3$ must follow a circular arc centered on $a_1$. This is illustrated in Fig.~\figref{Devel_aaac_arc_5_10}, and already steps outside and ${\varepsilon}$-thin apron in the neighborhood of $a_2$, where it makes an angle of approximately $90^\circ$ rather than the necessary $60^\circ$. This establishes the claim of the lemma. \end{proof} \begin{figure}[htbp] \centering \includegraphics[width=0.75\linewidth]{Figures/Devel_aaac_linear_5_10} \caption{The cut-path type $aaacb$ leads to $L / R$ overlap with straight segments.} \figlab{Devel_aaac_linear_5_10} \end{figure} \begin{figure}[htbp] \centering \includegraphics[width=0.75\linewidth]{Figures/Devel_aaac_arc_5_10} \caption{The cut-path type $aaacb$ extends outside ${\mathcal C}$ on the $(a_2,a_3)$ arc (and still $R$ crosses $L$ near $a_3)$.} \figlab{Devel_aaac_arc_5_10} \end{figure} \medskip \noindent We restate Theorem~1 in more detail: \setcounter{theorem}{0} \begin{theorem} Convex caps ${\mathcal C}$ with sufficiently small $\{{\omega}_a,{\omega}_c, {\varepsilon}\}$, as depicted in Fig.~\figref{ZipCex3D}, have no unzipping: they are un-unzippable. Thus there are arbitrarily flat convex caps that cannot be unzipped. \end{theorem} \begin{proof} We argued that only four combinatorial types of cut-paths ${\gamma}$ are possible on ${\mathcal C}$. Lemmas~\lemref{acaab}, \lemref{caaab}, \lemref{aacab}, \lemref{aaacb} established that for sufficiently small curvatures $\{{\omega}_a,{\omega}_c\}$ and a sufficiently thin ${\varepsilon}$-apron, each of these cut-path types fails to unzip ${\mathcal C}$. Because the arguments are independent of the exact values of $\{{\omega}_a,{\omega}_c\}$, only requiring a sufficiently small ${\varepsilon}$ to match, the claim holds for arbitrarily flat convex caps. \end{proof} \section{Discussion} \seclab{Discussion} It is tempting to hope that the negative result of Theorem~\thmref{ZipCex} can somehow be used to address the open problem for convex polyhedra. However, as mentioned earlier (Sec.~\secref{Lemmas}), closing the convex cap in Fig.~\figref{ZipCex3D} by adding a base creates a polyhedron that can in fact be unzipped. Perhaps this is not surprising, as the proofs rely crucially on the fact that ${\mathcal C}$ has a boundary ${\partial \mathcal C}$. I have also explored using several un-unzippable convex caps to tile a closed convex polyhedron, but so far to no avail. The open problem from~\cite{lddss-zupc-10} quoted in Sec.~\secref{Introduction} remains open. \newpage \bibliographystyle{alpha}
{ "timestamp": "2018-02-07T02:01:48", "yymm": "1802", "arxiv_id": "1802.01621", "language": "en", "url": "https://arxiv.org/abs/1802.01621", "abstract": "An unzipping of a polyhedron P is a cut-path through its vertices that unfolds P to a non-overlapping shape in the plane. It is an open problem to decide if every convex P has an unzipping. Here we show that there are nearly flat convex caps that have no unzipping. A convex cap is a \"top\" portion of a convex polyhedron; it has a boundary, i.e., it is not closed by a base.", "subjects": "Computational Geometry (cs.CG); Discrete Mathematics (cs.DM)", "title": "Un-unzippable Convex Caps", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.986571747626947, "lm_q2_score": 0.8128673201042492, "lm_q1q2_score": 0.8019519325840821 }
https://arxiv.org/abs/2005.07846
Jordan--Landau theorem for matrices over finite fields
Given a positive integer $r$ and a prime power $q$, we estimate the probability that the characteristic polynomial $f_{A}(t)$ of a random matrix $A$ in $\mathrm{GL}_{n}(\mathbb{F}_{q})$ is square-free with $r$ (monic) irreducible factors when $n$ is large. We also estimate the analogous probability that $f_{A}(t)$ has $r$ irreducible factors counting with multiplicity. In either case, the main term $(\log n)^{r-1}((r-1)!n)^{-1}$ and the error term $O((\log n)^{r-2}n^{-1})$, whose implied constant only depends on $r$ but not on $q$ nor $n$, coincide with the probability that a random permutation on $n$ letters is a product of $r$ disjoint cycles. The main ingredient of our proof is a recursion argument due to S. D. Cohen, which was previously used to estimate the probability that a random degree $n$ monic polynomial in $\mathbb{F}_{q}[t]$ is square-free with $r$ irreducible factors and the analogous probability that the polynomial has $r$ irreducible factors counting with multiplicity. We obtain our result by carefully modifying Cohen's recursion argument in the matrix setting, using Reiner's theorem that counts the number of $n \times n$ matrices with a fixed characteristic polynomial over $\mathbb{F}_{q}$.
\section{Introduction} \hspace{3mm} The purpose of this paper is to give some concrete arithmetic applications of the following fact: for any finitely many distinct $d_{1}, \dots, d_{r} \in \bZ_{\geq 1}$, the distribution of numbers of degree $d_{1}, \dots, d_{r}$ irreducible factors of the characteristic polynomial $f_{A}(t)$ of a matrix $A$ chosen uniformly at random from the set $\Mat_{n}(\bF_{q})$ of $n \times n$ matrices over $\bF_{q}$ converges to the distribution of numbers of length $d_{1}, \dots, d_{r}$ cycles of a random $\sigma \in S_{n}$, as $q$ goes to infinity. \ \begin{prop}\label{main} Let $d_{1}, \dots, d_{r} \in \bZ_{\geq 1}$ be distinct and $k_{1}, \dots, k_{r} \in \bZ_{\geq 0}$ not necessarily distinct. We have $$\lim_{q \ra \infty}\Prob_{A \in \Mat_{n}(\bF_{q})} \left(\begin{array}{c} f_{A}(t) \text{ has } k_{j} \text{ degree } d_{j} \\ \text{irreducible factors for } 1 \leq j \leq r \\ \text{counting with multiplicity} \end{array}\right) = \Prob_{\sigma \in S_{n}}\left(\begin{array}{c} \sigma \text{ has } k_{j} \text{ disjoint} \\ d_{j}\text{-cycles for } 1 \leq j \leq r \end{array}\right).$$ \ where $f_{A}(t)$ is the characteristic polynomial of $A \in \Mat_{n}(\bF_{q})$.\footnote{The event on the right-hand side means that when we write $\sigma$ as a unique product of disjoint cycles, there are $k_{j}$ cycles of length $d_{j}$ in the product for $1 \leq j \leq r$. The special case $r = 1$ is indirectly mentioned in ``Conclusion'' section of \cite{Sto}.} \end{prop} \ \begin{rmk} After the previous version of this paper, we have been noted that Proposition \ref{main} is known due to Hansen and Schumutz (\cite{HS}, Theorem 1.1). Their result is for $\GL_{n}(\bF_{q})$ rather than $\Mat_{n}(\bF_{q})$, but these two are easily interchangeable. Indeed, the last line of their proof (but not the statement), together with a well-known fact regarding some statistics of degree $n$ monic polynomials in $\bF_{q}[t]$ for large $q$, implies Proposition \ref{main}. The proof of Hansen and Schumutz uses Stong's work \cite{Sto}, on which our proof also depends, and they even compute some asymptotic errors with respect to $q$ and $r$. Since our work only considers the limit in $q$, our proof is simpler, so we decided to keep it in the paper. \ \hspace{3mm} In the following subsections, we discuss two applications of Proposition \ref{main}. The first one, Theorem \ref{p-adic}, considers the distribution of the cokernel of a random $n \times n$ matrix over $\bZ_{p}$, the ring of $p$-adic integers, with respect to the Haar (probability) measure on $\Mat_{n}(\bZ_{p}) = {\bZ_{p}}^{n^{2}}$, for large $p$ and $n$. According to our best knowledge and the consultations with the experts we have talked to, Theorem \ref{p-adic} is new. The second one, Theorem \ref{JL}, is a matrix version of Landau's theorem in number theory that estimates the number of irreducible factors of a random characteristic polynomial for large $n$ when $q \ra \infty$. Just as Proposition \ref{main} can be deduced from the work of Hansen and Schumutz \cite{HS}, Theorem \ref{JL} can also be deduced from their work if we apply another result of S. Cohen in \cite{Coh}. The interesting difference in our exposition is that we instead used Jordan's result in \cite{Jor} about the statistics of a random permutation in $S_{n}$ for large $n$ rather than the statistics of degree $n$ monic polynomials in $\bF_{q}[t]$ for large $q$ and $n$ in order to deduce Theorem \ref{JL}. \end{rmk} \ \subsection{The distribution of cokernels of Haar $\bZ_{p}$-random matrices with large $p$} Let $\bZ_{p}$ be the ring of $p$-adic integers, namely the inverse limit of the system of projection maps $\cdots \tra \bZ/(p^{3}) \tra \bZ/(p^{2}) \tra \bZ/(p)$. \ \begin{thm}\label{p-adic} Fix any distinct $d_{1}, \dots, d_{r} \in \bZ_{\geq 1}$. For any $d \in \bZ_{\geq 1}$ denote by $\mc{M}(d, p)$ the set of all monic irreducible polynomials of degree $d$ in $\bF_{p}[t]$. Then the limit $$\lim_{p \ra \infty} \Prob_{A \in \Mat_{n}(\bZ_{p})} \left(\begin{array}{c} \coker(P(A)) = 0 \\ \text{ for all } \bar{P} \in \bigsqcup_{j=1}^{r}\mc{M}(p, d_{j}) \end{array}\right)$$ \ exists, where for $P(A)$, we use any lift of $\bar{P}(t) \in \bF_{p}[t]$ in $\bZ_{p}[t]$ under the reduction modulo $p$ and the probability is according to the Haar measure on $\Mat_{n}(\bZ_{p})$. Moreover, we have $$\lim_{n \ra \infty} \lim_{p \ra \infty} \Prob_{A \in \Mat_{n}(\bZ_{p})} \left(\begin{array}{c} \coker(P(A)) = 0 \\ \text{ for all } \bar{P} \in \bigsqcup_{j=1}^{r}\mc{M}(p, d_{j}) \end{array}\right) = e^{-1/d_{1}} \cdots e^{-1/d_{r}}.$$ \end{thm} \ \begin{rmk} Taking the two limits in the particular order matters in Theorem \ref{p-adic}. For instance, we are unsure whether we can change the order of these limits, which may lead to an interesting consequence regarding the Cohen-Lenstra measure introduced in \cite{CL}. It is worth noting that our proof can be easily modified so that one may replace $\bZ_{p}$ with $\bF_{q}\llb t \rrb$, reduce modulo $(t)$ in place of $p$, and let $q \ra \infty$ instead of $p \ra \infty$ in Theorem \ref{p-adic}, although we will just focus on $\bZ_{p}$. The reader does not need to have much expertise of the Haar measure on $\Mat_{n}(\bZ_{p})$ to follow our proof. The only property of the Haar measure on $\Mat_{n}(\bZ_{p})$ we will use is that each fiber under the mod $p$ projection map $\Mat_{n}(\bZ_{p}) \tra \Mat_{n}(\bF_{p})$ has the same measure, which is necessarily $1/p^{n^{2}}$. \end{rmk} \ \hspace{3mm} We conjecture that the large $p$ limit in Theorem \ref{p-adic} has to do with independent Poisson random variables with means $1/d_{1}, \dots, 1/d_{r}$ in more generality. However, we have little empirical evidence, so it would be interesting to see any more examples that either support or disprove this conjecture. \ \begin{conj} Fix any distinct $d_{1}, \dots, d_{r} \in \bZ_{\geq 1}$ and not necessarily distinct $k_{1}, \dots, k_{r} \in \bZ_{\geq 0}$. Then the limit $$\lim_{p \ra \infty} \Prob_{A \in \Mat_{n}(\bZ_{p})} \left(\begin{array}{c} |\coker(P_{j}(A))| = p^{d_{j}k_{j}} \\ \text{ for all } (\bar{P}_{1}, \dots, \bar{P}_{r}) \in \prod_{j=1}^{r}\mc{M}(p, d_{j}) \end{array}\right)$$ \ must exist. Moreover, we must have $$\lim_{n \ra \infty}\lim_{p \ra \infty} \Prob_{A \in \Mat_{n}(\bZ_{p})} \left(\begin{array}{c} |\coker(P_{j}(A))| = p^{d_{j}k_{j}} \\ \text{ for all } (\bar{P}_{1}, \dots, \bar{P}_{r}) \in \prod_{j=1}^{r}\mc{M}(p, d_{j}) \end{array}\right) = \frac{e^{-1/d_{1}}(1/d_{1})^{k_{1}}}{k_{1}!} \cdots \frac{e^{-1/d_{r}}(1/d_{r})^{k_{r}}}{k_{r}!}.$$ \end{conj} \ \subsection{Jordan-Landau theorem for matrices over finite fields} It is a folklore that the cycle decomposition of a random permutation is analogous to the prime decomposition of a random positive integer. For instance, Granville (Section 2.1 of \cite{Gra}) notes that given $k \in \bZ_{\geq 1}$, a theorem of Jordan (p.161 of \cite{Jor}) says $$\Prob_{\sigma \in S_{n}}(\sigma \text{ has } k \text{ disjoint cycles}) \sim \frac{(\log n)^{k-1}}{(k-1)! n}$$ \ for large $n \in \bZ_{\geq 1}$, while a theorem of Landau (p.211 of \cite{Lan} or Theorem 437 on p.491 of \cite{HW}) says \begin{align*} \Prob_{1 \leq N \leq x}\left(\begin{array}{c} N \text{ has } k \text{ distinct} \\ \text{prime factors} \end{array}\right) &\sim \Prob_{1 \leq N \leq x}\left(\begin{array}{c} N \text{ has } k \text{ prime factors} \\ \text{counting with multiplicity} \end{array}\right)\\ &\sim \frac{(\log \log x)^{k-1}}{(k-1)!\log x} \end{align*} \ for large $x \in \mathbb{R}_{\geq 1}$. The first probability is given uniformly at random from the set $S_{n}$ of permutations on $n$ letters. The second probability is given by choosing the integer $N$ uniformly at random from the set $\{1, 2, \dots, \lfloor x \rfloor\}$. We wrote ``$f(x) \sim g(x)$ for large $x$'' to mean that $f(x)/g(x) \ra 1$ as $x \ra \infty$. This is a generalization of the Prime Number Theorem (i.e., the case $k = 1$). \ \hspace{3mm} By the Prime Number Theorem, for large $x$, we may expect one prime in every interval of length $\log x$. For large $n$, we expect one cycle in $n$ permutations (since there are $(n-1)!$ $n$-cycles while there are far less cycles of shorter lengths), so Jordan's result is analogous to Landau's result. We take this analogy further. In particular, summing the identity given in Proposition \ref{main} over all tuples $(d_{1}, \dots, d_{r}) \in (\bZ_{\geq 1})^{r}$ and $(k_{1}, \dots, k_{r}) \in (\bZ_{\geq 0})^{r}$ such that \begin{itemize} \item $r \in \bZ_{\geq 0}$, \item $n = \sum_{j=1}^{r}k_{j}d_{j}$, and \item $k = \sum_{j=1}^{r} k_{j}$, \end{itemize} \ we have $$\lim_{q \ra \infty}\Prob_{A \in \Mat_{n}(\bF_{q})} \left(\begin{array}{c} f_{A}(t) \text{ has } k \text{ irreducible factors} \\ \text{counting with multiplicity} \end{array}\right) = \Prob_{\sigma \in S_{n}}(\sigma \text{ has } k \text{ disjoint cycles}),$$ \ which is asymptotically $(\log n)^{k-1}/((k-1)!n)$ for large $n$, as noted above. Using the facts that almost all matrices $A \in \Mat_{n}(\bF_{q})$ have square-free characteristic polynomials and almost all monic polynomials of degree $n$ are square-free when $q \ra \infty$ (see Section \ref{Landau}), we may obtain the following. \ \begin{thm}[Jordan-Landau theorem for $\bF_{q}$-matrices]\label{JL} Given $k \in \bZ_{\geq 1}$, for large $n \in \mathbb{Z}_{\geq 1}$, we have \begin{align*} \lim_{q \ra \infty}\Prob_{A \in \Mat_{n}(\bF_{q})} \left(\begin{array}{c} f_{A}(t) \text{ has } k \text{ distinct} \\ \text{irreducible factors} \end{array}\right) &= \lim_{q \ra \infty}\Prob_{A \in \Mat_{n}(\bF_{q})} \left(\begin{array}{c} f_{A}(t) \text{ has } k \text{ irreducible factors} \\ \text{counting with multiplicity} \end{array}\right) \\ &\sim \frac{(\log n)^{k-1}}{(k-1)! n}. \end{align*} \end{thm} \ \begin{rmk} There is an analogous result of Theorem \ref{JL}, where the sample space is the set $\bA^{n}(\bF_{q})$ of all degree $n$ monic polynomials in $\bF_{q}[t]$ due to S. Cohen \cite{Coh}. It says given $k \in \bZ_{\geq 1}$, for large $n \in \mathbb{Z}_{\geq 1}$, we have \begin{align*} \lim_{q \ra \infty}\Prob_{f \in \bA^{n}(\bF_{q})} \left(\begin{array}{c} f(t) \text{ has } k \text{ distinct} \\ \text{irreducible factors} \end{array}\right) &\sim \lim_{q \ra \infty}\Prob_{f \in \bA^{n}(\bF_{q})} \left(\begin{array}{c} f(t) \text{ has } k \text{ irreducible factors} \\ \text{counting with multiplicity} \end{array}\right) \\ &\sim \frac{(\log n)^{k-1}}{(k-1)! n}. \end{align*} \end{rmk} \ With these two results in mind, we note that the push-forward measure $\Mat_{n}(\bF_{q}) = \bA^{n^{2}}(\bF_{q}) \tra \bA^{n}(\bF_{q})$ given by taking the characteristic polynomial of a matrix does not change the probability of the event we compute in $\bA^{n}(\bF_{q})$, when we take $q \ra \infty$. This can be also taken as another example to be added in the list of similarly behaving combinatorial examples in \cite{ABT}, a survey paper by Arratia, Barbour, and Tavar\'e. \ \subsection{Fulman's generalization and further directions} Recently, we have also realized that Fulman generalized Proposition \ref{main} to certain linear algebraic groups $G$ over $\bF_{q}$, where $G = \GL_{n}$ is the desired special case for Proposition \ref{main} (\cite{Ful99}, Theorem 15), where $S_{n}$ occurs as the Weyl group of $G(\ol{\bF_{q}}) = \GL_{n}(\ol{\bF_{q}})$. Fulman's argument needs to assume that the probability that a random element in $G(\bF_{q}) \subset G(\ol{\bF_{q}})$ is regular semisimple converges to $1$, as $q \ra \infty$, which we previously thought as a mere heuristics. There is a statement in Fulman's paper (\cite{Ful99}, Theorem 5 and Theorem 7) that proves a similar property related to this hypothesis for $G = \GL_{n}$, but it takes $n \ra \infty$ before we can let $q \ra \infty$. However, for $g \in \GL_{n}(\bF_{q})$ saying that $g$ is regular semisimple turns out to mean that the characteristic polynomial of $g$ is square-free in $\bF_{q}[t]$, so using a similar argument as in Lemma \ref{sfmat}, we see that $G = \GL_{n}$ satisfies Fulman's hypothesis because the number of degree $n$ monic square-free polynomials $f(t) \in \bF_{q}[t]$ such that $f(0) \neq 0$ is precisely $$q^{n} - 2q^{n-1} + 2q^{n-2} - \cdots + (-1)^{n-1}2q + (-1)^{n},$$ \ which replaces Lemma \ref{sf} in the argument. We are not sure whether the above count is explicitly written in literature, but this is certainly well-known (e.g., more general phenomena are dealt in various works such as \cite{BE}, \cite{Kim}, and \cite{VW}). More specifically, one can obtain this count by counting the $\bF_{q}$-points of the unordered configuration space $\Conf^{n}(\bA^{1} \sm \{0\})$ of $n$ points on the affine line minus the origin over $\bF_{q}$. The generating function for such count $|\Conf^{n}(\bA^{1} \sm \{0\})(\bF_{q})|$ is given by $$\sum_{n=0}^{\infty} |\Conf^{n}(\bA^{1} \sm \{0\})(\bF_{q})| t^{n} = \frac{Z_{\bA^{1} \setminus \{0\}}(t)}{Z_{\bA^{1} \setminus \{0\}}(t^{2})} = \frac{1 - qt^{2}}{(1 + t)(1 - qt)},$$ \ where $Z_{X}(t)$ means the zeta series of an algebraic variety over $\bF_{q}$, and expanding the right-hand side, one may obtain the desired count. We are not sure whether this hypothesis may or may not be satisfied by other linear algebraic groups $G$ over $\bF_{q}$. The reader must note that the reasons that this works for $G = \GL_{n}$ are as follows: \be \item we are aware of the size of each fiber of the map $\Mat_{n}(\bF_{q}) = \bA^{n^{2}}(\bF_{q}) \rightarrow \bA^{n}(\bF_{q})$ given by taking the characteristic polynomials and \item the preimage of the set of degree $n$ monic polynomials with nonzero constant terms under this map is precisely $\GL_{n}(\bF_{q})$. \ee Nevertheless, it would be extremely interesting if one can identify which algebraic groups $G$, other than $\GL_{n}$, over $\bF_{q}$ produce the sets $G(\bF_{q})$ of $\bF_{q}$-points that can replace $\Mat_{n}(\bF_{q})$ with the Weyl groups of $G(\ol{\bF_{q}})$, replacing $S_{n}$ in Proposition \ref{main}. If such algebraic groups are given in a sequence $G_{n}$ in $n \in \mathbb{Z}_{\geq 1}$, then we can hope that studying the corresponding sequence $W_{n}$ of Weyl groups asymptotically in $n$ may produce analogous statements for Theorem \ref{p-adic} and Theorem \ref{JL}. \ \subsection{Organization for the rest} In Section \ref{Landau}, we will show that Proposition \ref{main} implies Theorem \ref{JL}. In Section \ref{Haar}, we will show that Proposition \ref{main} implies Theorem \ref{p-adic}. We will provide a crucial formula due to Stong in Section \ref{Stong}, but our proof is slightly different from the original one, as we use direct counting of Young diagrams. Stong's formula will be used in proving Proposition \ref{main} in Section \ref{mainproof}. Finally, in Section \ref{permproof}, we will provide another proof of Lemma \ref{perm}, an influential result of Shepp and Lloyd, which states that the number of length $d$ cycles of a random permutation in $S_{n}$ asymptotically follows the Poisson distribution with mean $1/d$ when $n$ is large, and the probability is given independently for finitely many different choices of $d$. It is interesting that our proof is entirely combinatorial, while the original proof was probabilistic. Since the result is quite popular in literature, our alternative proof of Lemma \ref{perm} may be known to experts, although we were unable to locate it. \ \subsection{Acknowledgment} We thank Yifeng Huang, Nathan Kaplan, and Jeff Lagarias for helpful conversations. G. Cheong was supported by NSF grant DMS-1162181 and the Korea Institute for Advanced Study for his visits to the institution regarding this research. M. Yu was supported by a KIAS Individual Grant (SP075201) via the Center for Mathematical Challenges at Korea Institute for Advanced Study. He was also supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. 2020R1C1C1A01007604). \ \section{Proposition \ref{main} implies Theorem \ref{JL}}\label{Landau} \hspace{3mm} We already know that Proposition \ref{main} implies that $$\lim_{q \ra \infty}\Prob_{A \in \Mat_{n}(\bF_{q})} \left(\begin{array}{c} f_{A}(t) \text{ has } k \text{ irreducible factors} \\ \text{counting with multiplicity} \end{array}\right) \sim \frac{(\log n)^{k-1}}{(k-1)! n}$$ \ for large $n$ in the statement of Theorem \ref{JL}. We need to justify the statement about distinct irreducible factors. We need two lemmas, well-known to the experts, to provide a complete exposition. \ \begin{lem}\label{sf} Let $\bA^{n}(\bF_{q})$ be the set of all monic polynomials of degree $n$ in $\bF_{q}[t]$. For $n \geq 2$, we have $$|\{f \in \bA^{n}(\bF_{q}): f(t) \text{ is square-free in } \bF_{q}[t]\}| = q^{n}(1 - q^{-1}).$$ \ In particular, we have $$\lim_{q \ra \infty}\Prob_{f \in \bA^{n}(\bF_{q})}(f(t) \text{ is square-free in } \bF_{q}[t]) = 1.$$ \end{lem} \ \begin{proof} This is a well-known fact in literature. For a proof, see Lemma 3.4 of \cite{CWZ}. A more combinatorial proof using partitions is also available (e.g., Proposition 5.9 in \cite{VW} with $a = 2$). \end{proof} \ \begin{lem}\label{sfmat} For any $n \in \bZ_{\geq 1}$, we have $$\lim_{q \ra \infty}\Prob_{A \in \Mat_{n}(\bF_{q})}(f_{A}(t) \text{ is square-free in } \bF_{q}[t]) = 1.$$ \end{lem} \ \begin{proof} Fix any monic square-free polynomial $$f(t) = P_{1}(t) \cdots P_{r}(t),$$ \ where $P_{i}(t)$ are distinct monic irreducible polynomials in $\bF_{q}[t]$. By Theorem 2 in \cite{Rei}, the number of $A \in \Mat_{n}(\bF_{q})$ such that $f_{A}(t) = f(t)$ is $$q^{n^{2} - n} \cdot \frac{(1 - q^{-1})(1 - q^{-2}) \cdots (1 - q^{-n})}{(1 - q^{-\deg(P_{1})}) (1 - q^{-\deg(P_{2})}) \cdots (1 - q^{-\deg(P_{r})})} \geq q^{n^{2} - n} \cdot \frac{(1 - q^{-1})(1 - q^{-2}) \cdots (1 - q^{-n})}{(1 - q^{-n})^{r}}.$$ \ For $n = 1$, the map $A \mapsto f_{A}(t)$ is bijective, and thus letting $q \ra \infty$, we are done. Foe $n \geq 2$, by Lemma \ref{sf}, we have $$\Prob_{A \in \Mat_{n}(\bF_{q})}(f_{A}(t) \text{ is square-free in } \bF_{q}[t]) \geq (1 - q^{-1}) \cdot \frac{(1 - q^{-1})(1 - q^{-2}) \cdots (1 - q^{-n})}{(1 - q^{-n})^{r}},$$ \ so letting $q \ra \infty$ gives the result. \end{proof} \ \begin{proof}[Proof that Proposition \ref{main} implies Theorem \ref{JL}] Again, we only need to show the first identity in the statement. By Lemma \ref{sfmat}, we have \begin{align*} \lim_{q \ra \infty}\Prob_{A \in \Mat_{n}(\bF_{q})} \left(\begin{array}{c} f_{A}(t) \text{ has } k \text{ distinct} \\ \text{irreducible factors} \end{array}\right) &= \lim_{q \ra \infty}\Prob_{A \in \Mat_{n}(\bF_{q})} \left(\begin{array}{c} f_{A}(t) \text{ is square-free in } \bF_{q}[t] \text{ and}\\ \text{has } k \text{ irreducible factors} \\ \text{counting with multiplicity} \end{array}\right) \\ &= \lim_{q \ra \infty}\Prob_{A \in \Mat_{n}(\bF_{q})} \left(\begin{array}{c} f_{A}(t) \text{ has } k \text{ irreducible factors} \\ \text{counting with multiplicity} \end{array}\right), \end{align*} \ because the number of monic non-square-free polynomials of degree $n$ is negligible as $q \ra \infty$. \end{proof} \ \section{Proposition \ref{main} implies Theorem \ref{p-adic}}\label{Haar} \subsection{Useful lemmas} The following result, due to Shepp and Lloyd, will be crucial in applying Proposition \ref{main} to obtain Theorem \ref{p-adic}. In \cite{SL}, Shepp and Lloyd obtained the result by showing that the the characteristic function of the distribution of the numbers of cycles of fixed lengths $d_{1}, \dots, d_{r}$ of a random permutation converges to the characteristic function of the distribution given by the independent Poisson random variables with the means $1/d_{1}, \dots, 1/d_{r}$ and then applying L\'evy's theorem. We will provide our own combinatorial proof of this result in Section \ref{permproof}. \ \begin{notn} For any permutation $\sigma \in S_{n}$, we denote by $m_{d}(\sigma)$ the number of $d$-cycles in the cycle decomposition of $\sigma$. \end{notn} \ \begin{lem}[cf. p.342 in \cite{SL}]\label{perm} Fix $r \in \bZ_{\geq 0}$. Given distinct $d_{1}, \dots, d_{r} \in \bZ_{\geq 1}$ and not necessarily distinct $k_{1}, \dots, k_{r} \in \bZ_{\geq 0}$, we have $$\lim_{n \ra \infty}\Prob_{\sigma \in S_{n}}(m_{d_{j}}(\sigma) = k_{j} \text{ for } 1 \leq j \leq r) = \frac{e^{-1/d_{1}}(1/d_{1})^{k_{1}}}{k_{1}!} \cdots \frac{e^{-1/d_{r}}(1/d_{r})^{k_{r}}}{k_{r}!},$$ \ which means that the number of cycles of length $d_{1}, \dots, d_{r}$ of a random permutation of $n$ letters are asymptotically given by independent Poisson random variables with means $1/d_{1}, \dots, 1/d_{r}$ when $n$ is large. \end{lem} \ \hspace{3mm} We need the following lemma to deduce Theorem \ref{p-adic} from Proposition \ref{main}. This lemma contains some ideas used for proving Theorem C of \cite{CH}, but the proof here is simpler, since we only need a special case of their argument. \ \begin{lem}\label{red} Let $P_{1}, \dots, P_{r} \in \bZ_{p}[t]$ be any monic polynomials whose images in $\bF_{p}[t]$ under the mod $p$ reduction are distinct irreducible polynomials. Then $$\Prob_{A \in \Mat_{n}(\bZ_{p})}\left(\begin{array}{c} \coker(P_{j}(A)) = 0 \\ \text{ for } 1 \leq j \leq r \end{array}\right) = \Prob_{A \in \Mat_{n}(\bF_{p})}\left(\begin{array}{c} A[\bar{P}_{j}^{\infty}] = 0 \\ \text{ for } 1 \leq j \leq r \end{array}\right),$$ \ where $A[\bar{P}_{j}^{\infty}]$ means the $\bar{P}_{j}$-part of the $\bF_{p}[t]$-module structure given by the matrix action $A \acts \bF_{p}^{n}$. \end{lem} \begin{proof} First, note that for any $A \in \Mat_{n}(\bF_{p})$, we have $\dim_{\bF_{p}}\ker(\bar{P}_{j}(A)) = \dim_{\bF_{p}}\coker(\bar{P}_{j}(A))$ and $\ker(\bar{P}_{j}(A)) = A[\bar{P}_{j}^{\infty}]/\bar{P}_{j}(t)A[\bar{P}_{j}^{\infty}]$, as $\bF_{p}[t]$-modules where the action of $t$ is given by multiplying the matrix $A$ on the left. Hence, the following are equivalent: \ \begin{itemize} \item $A[\bar{P}_{j}^{\infty}] = 0$; \item $\ker(\bar{P}_{j}(A)) = 0$; \item $\coker(\bar{P}_{j}(A)) = 0$, \end{itemize} \ so it is enough to show that $$\Prob_{A \in \Mat_{n}(\bZ_{p})}\left(\begin{array}{c} \coker(P_{j}(A)) = 0 \\ \text{ for } 1 \leq j \leq r \end{array}\right) = \Prob_{A \in \Mat_{n}(\bF_{p})}\left(\begin{array}{c} \coker(\bar{P}_{j}(A)) = 0 \\ \text{ for } 1 \leq j \leq r \end{array}\right).$$ \ For any $B \in \Mat_{n}(\bZ_{p})$, the right exactness of $(-) \otimes_{\bZ_{p}} \bF_{p}$ implies that $$\coker(B) \otimes_{\bZ_{p}} \bF_{p} \simeq \coker(\bar{B}),$$ where $\bar{B} \in \Mat_{n}(\bF_{p})$ is the reduction of $B$ mod $p$. Hence, by Nakayama's lemma, we have $\coker(B) = 0$ if and only if $\coker(\bar{B}) = 0$. Thus, if we consider the mod $p$ reduction map $\pi : \Mat_{n}(\bZ_{p}) \tra \Mat_{n}(\bF_{p})$, we have $$\pi^{-1}\left\{\begin{array}{c} A \in \Mat_{n}(\bF_{p}) : \coker(\bar{P}_{j}(A)) = 0 \\ \text{ for } 1 \leq j \leq r \end{array}\right\} = \left\{\begin{array}{c} A \in \Mat_{n}(\bZ_{p}) : \coker(P_{j}(A)) = 0 \\ \text{ for } 1 \leq j \leq r \end{array}\right\}.$$ \ Since each fiber under $\pi$ has the same measure $1/p^{n^{2}}$, this finishes the proof. \end{proof} \ \subsection{Proof of Theorem \ref{p-adic} given Proposition \ref{main}} We now prove Theorem \ref{p-adic}, assuming Proposition \ref{main}. Proposition \ref{main} will be proven in Section \ref{mainproof}. \ \begin{proof}[Proof of Theorem \ref{p-adic} given Proposition \ref{main}] Let $\mc{M}(q, d)$ be the set of degree $d$ monic irreducible polynomials in $\bF_{q}[t]$ so that $M(q, d) = |\mc{M}(q, d)|$. By Lemma \ref{red}, we have \begin{align*}\Prob_{A \in \Mat_{n}(\bZ_{p})} \left(\begin{array}{c} \coker(P(A)) = 0 \\ \text{ for all } P \in \bigsqcup_{j=1}^{r}\mc{M}(p, d_{j}) \end{array}\right) &= \Prob_{A \in \Mat_{n}(\bF_{p})} \left(\begin{array}{c} A[\bar{P}_{j}^{\infty}] = 0 \\ \text{ for all } P \in \bigsqcup_{j=1}^{r}\mc{M}(p, d_{j}) \end{array}\right) \\ &= \Prob_{A \in \Mat_{n}(\bF_{p})} \left(\begin{array}{c} f_{A}(t) \text{ has no irreducible factors of} \\ \text{degrees } d_{1}, \dots, d_{r} \end{array}\right). \end{align*} \ Hence, we may apply Proposition \ref{main} and Lemma \ref{perm} to finish the proof. \end{proof} \ \section{Stong's formula}\label{Stong} \subsection{Set-up} Let $R$ be a PID, and say $\mf{m}$ is a maximal ideal of $R$ such that $R/\mf{m} = \bF_{q}$. Denote by $\mc{P}$ the set of all partitions including the empty partition $\es$. Then any finite length (or equivalently, finite size) $\mf{m}^{\infty}$-torsion $R$-module is isomorphic to $$H_{\mf{m},\ld}^{R} := R/\mf{m}^{\ld_{1}} \op \cdots \op R/\mf{m}^{\ld_{l}},$$ \ for a unique partition $\ld = (\ld_{1}, \dots, \ld_{l}) \in \mc{P}$. We will always assume $\ld_{1} \geq \cdots \geq \ld_{l} > 0$, and we write $|\ld| := \ld_{1} + \cdots + \ld_{l}$. The case $l = 0$ will correspond to the empty partition $\ld = \es$. Note that $$|H_{\mf{m},\ld}^{R}| = q^{\ld_{1} + \cdots + \ld_{l}} = q^{|\ld|}.$$ \ We will write $H_{\mf{m},\ld} := H_{\mf{m},\ld}^{R}$ if the meaning of $R$ is evident from the context. Denote by $m_{d}(\ld)$ the number of parts of size $d$ in $\ld$, and define $$n(\ld) := 0 \cdot \ld_{1} + 1 \cdot \ld_{2} + 2 \cdot \ld_{3} + \cdots + (l-1) \cdot \ld_{l}.$$ \ \hspace{3mm} The number $|\Aut_{R}(H_{\mf{m},\ld}^{R})|$ of automorphisms of $H_{\mf{m},\ld}^{R}$ can be computed by noting the fact that $$|\Aut_{R}(H_{\mf{m},\ld}^{R})| = |\Aut_{R_{\mf{m}}}(H_{\mf{m}R_{\mf{m}},\ld}^{R_{\mf{m}}})|$$ \ and the following formula due to Macdonald. \ \begin{lem}[(1.6) on p.181 in \cite{Mac}]\label{Mac} Let $(R, \mf{m})$ be a DVR (discrete valuation ring) with the finite residue field $R/\mf{m} = \bF_{q}$. Then we have $$|\Aut_{R}(H_{\mf{m},\ld})| = q^{|\ld|+2n(\ld)} \prod_{d = 1}^{\infty}\prod_{i=1}^{m_{d}(\ld)}(1 - q^{-i}).$$ \end{lem} \ \hspace{3mm} To our best knowledge, the following result is due to Stong in \cite{Sto}, but we rephrase Stong's result in a more general setting and provide a slightly different proof. We will essentially go through some key ideas in Lemma 6 in \cite{Sto}, which relies on the Fine-Herstein theorem (i.e., the number of $n \times n$ nilpotent matrices over $\bF_{q}$ is $q^{n(n-1)}$), but unlike the reference, we will avoid the use of M\"obius inversion along with exponentiation, logarithm, differentiation and integration of power series by counting partitions (i.e., Young diagrams) instead. \ \begin{lem}[cf. Proposition 19 in \cite{Sto}]\label{key} Let $R$ be any Dedekind domain with a maximal ideal $\mf{m}$ such that $R/\mf{m} = \bF_{q}$. Then $$\sum_{M \in \Mod_{R_{\mf{m}}}^{<\infty}} \frac{u^{q\dim_{\bF_q}(M)}}{|\Aut_{R}(M)|} = \prod_{i=1}^{\infty}\frac{1}{1 - q^{-i}u^{q}},$$ \ where $u$ is a complex variable with $|u| < q^{1/q}$. Equivalently, we have $$\sum_{\ld \in \mc{P}} \frac{y^{|\ld|}}{|\Aut_{R}(H_{\mf{m},\ld})|} = \prod_{i=1}^{\infty}\frac{1}{1 - q^{-i}y},$$ \ where $y$ is a complex variable with $|y| < q$. \end{lem} \begin{proof} As explained in the beginning of this section, finite $\mf{m}^{\infty}$-torsion $R$-modules are parametrized by partitions. Hence, taking $y = u^{q}$, we see that the two given statements are equivalent. We will prove the second statement. Applying Lemma \ref{Mac}, this reduces our problem to the case where $R = \bF_{q}[t]$ and $\mf{m} = (t)$ by replacing $R_{\mf{m}}$ with $\bF_{q}[t]_{(t)}$ or $\bF_{q}\llb t \rrb$. This reduction lets us rewrite the left-hand side of the desired identity as $$\sum_{n = 0}^{\infty}\sum_{\ld \vdash n}\frac{y^{n}}{|\Stab_{\GL_{n}(\bF_{q})}(J_{0,\ld})|},$$ \ where $J_{0,\ld}$ is the Jordan canonical form given by the $0$-Jordan blocks of whose sizes are equal to the parts of the partition $\ld$, and the action $\GL_{n}(\bF_{q}) \acts \Mat_{n}(\bF_{q})$ is given by the conjugation. By the orbit-stabilizer theorem, we have $$\frac{1}{|\Stab_{\GL_{n}(\bF_{q})}(J_{0,\ld})|} = \frac{|\GL_{n}(\bF_{q}) \cdot J_{0,\ld}|}{|\GL_{n}(\bF_{q})|}.$$ \ Recall that $n \times n$ matrices all of whose eigenvalues are $0$ are precisely nilpotent matrices. Thus, we have \begin{align*} \sum_{n = 0}^{\infty}\sum_{\ld \vdash n}\frac{y^{n}}{|\Stab_{\GL_{n}(\bF_{q})}(J_{0,\ld})|} &= \sum_{n = 0}^{\infty}\frac{|\Nil_{n}(\bF_{q})|y^{n}}{|\GL_{n}(\bF_{q})|} \\ &= 1 + \sum_{n = 1}^{\infty}\frac{q^{n(n-1)}y^{n}}{(q^{n} - 1)(q^{n} - q) \cdots (q^{n} - q^{n-1})} \\ &= 1 + \sum_{n = 1}^{\infty}\frac{(q^{-1}y)^{n}}{(1 - q^{-1})(1 - q^{-2}) \cdots (1 - q^{-n})} \\ &= \sum_{n = 0}^{\infty}(q^{-1}y)^{n} \left(\sum_{j_{1}=0}^{\infty} q^{-j_{1}}\right) \left(\sum_{j_{2}=0}^{\infty} q^{-2j_{2}}\right) \cdots \left(\sum_{j_{n}=0}^{\infty} q^{-nj_{n}}\right) \\ &= \sum_{n = 0}^{\infty}(q^{-1}y)^{n} \sum_{j_{1}=0}^{\infty} \sum_{j_{2}=0}^{\infty} \cdots \sum_{j_{n}=0}^{\infty} q^{-j_{1} - 2j_{2} - \cdots - nj_{n}} \\ &= \prod_{i=0}^{\infty} (1 + q^{-i}(q^{-1}y) + q^{-2i}(q^{-1}y)^{2} + \cdots) \\ &= \prod_{i=0}^{\infty} \frac{1}{1 - q^{-(i+1)}y} \\ &= \prod_{i=1}^{\infty} \frac{1}{1 - q^{-i}y}, \end{align*} \ where for the second equality we used the elementary identity $|\GL_{n}(\bF_{q})| = (q^{n} - 1)(q^{n} - q) \cdots (q^{n} - q^{n-1})$, and Fine-Herstein theorem (e.g., Theorem 1 of \cite{FH}), which gives the number $|\Nil_{n}(\bF_{q})|$ of nilpotent matrices in $\Mat_{n}(\bF_{q})$: $$|\Nil_{n}(\bF_{q})| = q^{n(n-1)}.$$ \ For the sixth equality, we have used the fact that the coefficient of $Y^{n}$ of the following product $$\prod_{i=0}^{\infty}(1 + X^{i}Y + X^{2i}Y^{2} + \cdots)$$ \ is equal to $$\sum_{j_{1}=0}^{\infty} \sum_{j_{2}=0}^{\infty} \cdots \sum_{j_{n}=0}^{\infty} X^{j_{1} + 2j_{2} + \cdots + nj_{n}}.$$ \ where $X, Y$ are considered to be complex numbers varying within the open unit disk centered at $0$ (i.e., $|X|, |Y| < 1$) so that we can take $X = q^{-1}$ and $Y = q^{-1}y$ with $|y| < q$ in our proof for the sixth equality in the chain of equalities above. Indeed, when we expand the given product, we have \begin{align*} \prod_{i=0}^{\infty}(1 + X^{i}Y + X^{2i}Y^{2} + \cdots) &= \sum_{ m_{0}, m_{1}, m_{2}, \dots \geq 0} X^{0 \cdot m_{0} + 1 \cdot m_{1} + 2 \cdot m_{2} + \cdots} Y^{m_{0} + m_{1} + m_{2} + \cdots}\\ &= \sum_{n=0}^{\infty} Y^{n} \sum_{\substack{ m_{0}, m_{1}, m_{2}, \dots \geq 0 \\ m_{0} + m_{1} + m_{2} + \cdots = n}} X^{m_{1} + 2m_{2} + \cdots}, \end{align*} \ so it is enough to show that $$\sum_{\substack{ m_{0}, m_{1}, m_{2}, \dots \geq 0 \\ m_{0} + m_{1} + m_{2} + \cdots = n}} X^{m_{1} + 2m_{2} + \cdots} = \sum_{j_{1}=0}^{\infty} \sum_{j_{2}=0}^{\infty} \cdots \sum_{j_{n}=0}^{\infty} X^{j_{1} + 2j_{2} + \cdots + nj_{n}}.$$ \ Note that we have a bijection $$\{(m_{0}, m_{1}, m_{2}, \dots) \in \bZ_{\geq 0}^{\infty} : m_{0} + m_{1} + m_{2} + \cdots = n \} \lra \{(m_{1}, m_{2}, \dots) \in \bZ_{\geq 0}^{\infty} : m_{1} + m_{2} + \cdots \leq n \}.$$ \ given by $(m_{0}, m_{1}, m_{2}, \dots) \mapsto (m_{1}, m_{2}, \dots)$. This reduces our problem to the following: $$\sum_{\substack{ m_{1}, m_{2}, \dots \geq 0 \\ m_{1} + m_{2} + \cdots \leq n}} X^{m_{1} + 2m_{2} + \cdots} = \sum_{j_{1}=0}^{\infty} \sum_{j_{2}=0}^{\infty} \cdots \sum_{j_{n}=0}^{\infty} X^{j_{1} + 2j_{2} + \cdots + nj_{n}}.$$ \ If $n = 0$, both sides are $1$, so let $n \geq 1$. Let $A_{n}$ be the set of partitions whose number of parts counting with multiplicity (i.e., the row lengths of Young diagrams) is $\leq n$ and $B_{n}$ the set of partitions whose parts (i.e., the column lengths of Young diagrams) are $\leq n$. Then we have a bijection $A_{n} \lra B_{n}$ given by taking conjugation, so in particular, we have $$\sum_{\ld \in A_{n}}X^{|\ld|} = \sum_{\ld \in B_{n}}X^{|\ld|}.$$ \ Now, noting that $$\sum_{\ld \in A_{n}}X^{|\ld|} = \sum_{\substack{ m_{1}, m_{2}, \dots \geq 0 \\ m_{1} + m_{2} + \cdots \leq n}} X^{m_{1} + 2m_{2} + \cdots}$$ \ and $$\sum_{\ld \in B_{n}}X^{|\ld|} = \sum_{j_{1}=0}^{\infty} \sum_{j_{2}=0}^{\infty} \cdots \sum_{j_{n}=0}^{\infty} X^{j_{1} + 2j_{2} + \cdots + nj_{n}},$$ \ we finish the proof. \end{proof} \ \section{Proof of Proposition \ref{main}}\label{mainproof} \hspace{3mm} For our proof of Proposition \ref{main}, we need to analyze polynomials that encode some information about random matrices in $\Mat_{n}(\bF_{q})$ and random permutations in $S_{n}$. Such a polynomial is called a ``cycle index''. \ \subsection{Cycle index} Given any permutation group $G \leqslant S_{n}$, we define the \textbf{cycle index} of $G$ as the following polynomial: $$\mc{Z}(G, \bs{x}) := \frac{1}{|G|}\sum_{g \in G}x_{1}^{m_{1}(g)} \cdots x_{n}^{m_{n}(g)},$$ \ where we recall that $m_{d}(g)$ means the number of $d$-cycles in the cycle decomposition of $g$. Cycle indices were used to solve various enumeration problems, and to our best knowledge, these were invented independently by Redfield \cite{Red} and P\'olya \cite{Pol}. (P\'olya's paper is in German, but there is an English translation \cite{PR} by Read). We will use the cycle index $\mc{Z}(S_{n}, \bs{x})$ of the full symmetric group $S_{n}$. Note that for any $\sigma, \tau \in S_{n}$, we note that $\sigma$ and $\tau$ are conjugate to each other in $S_{n}$ if and only if $x_{1}^{m_{1}(\sigma)} \cdots x_{n}^{m_{n}(\sigma)} = x_{1}^{m_{1}(\tau)} \cdots x_{n}^{m_{n}(\tau)}$, so $\mc{Z}(S_{n}, \bs{x})$ captures some information about the conjugate action $S_{n} \acts S_{n}$. \ \hspace{3mm} We will also make use of the following polynomial: $$\mc{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x}) := \frac{1}{|\GL_{n}(\bF_{q})|}\sum_{A \in \Mat_{n}(\bF_{q})}\prod_{P \in |\bA^{1}_{\bF_{q}}|}x_{P, \mu_{P}(A)},$$ \ where $\mu_{P}(A) = (\mu_{1}, \dots, \mu_{l})$ is the partition defined by the $P$-part of the $\bF_{q}[t]$-module structure defined by the matrix multiplication $A \acts \bF_{q}^{n}$: $$A[P^{\infty}] \simeq \bF_{q}[t]/(P)^{\mu_{1}} \op \cdots \op \bF_{q}[t]/(P)^{\mu_{l}}.$$ \ We denoted by $|\bA^{1}_{\bF_{q}}|$ the set of all monic irreducible polynomials in $\bF_{q}[t]$. For $P \in |\bA^{1}_{\bF_{q}}|$ and a nonempty partition $\ld$, the notation $x_{P,\ld}$ means a formal variable associated to the pair $(P, \ld)$. For the empty partition, we define $x_{P,\es} := 1$. We call $\mc{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})$ the \textbf{cycle index} of the conjugate action $\GL_{n}(\bF_{q}) \acts \Mat_{n}(\bF_{q})$. The terminology makes sense because each monomial $\prod_{P \in |\bA^{1}_{\bF_{q}}|}x_{P, \mu_{P}(A)}$ is characterized by an orbit of such action. That is, two matrices $A$ and $B$ are in the same orbit if and only if $\prod_{P \in |\bA^{1}_{\bF_{q}}|}x_{P, \mu_{P}(A)} = \prod_{P \in |\bA^{1}_{\bF_{q}}|}x_{P, \mu_{P}(B)}$. The notion of $\mc{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})$ is introduced by Stong \cite{Sto}, generalizing a similar definition for the conjugate action $\GL_{n}(\bF_{q}) \acts \GL_{n}(\bF_{q})$ introduced by Kung \cite{Kun}. We will use the following factorization results for the generating functions for the cycle indicies. \ \begin{lem}[p.14 of \cite{PR}]\label{fac1} In the ring $\bQ [\bs{x}] \llb u \rrb$ of formal power series in $u$ over the polynomial ring $\bQ [\bs{x}] = \bQ[x_{1},x_{2}, \dots]$, we have $$\sum_{n=0}^{\infty}\mc{Z}(S_{n}, \bs{x}) u^{n} = \prod_{d=1}^{\infty}e^{x_{d}u^{d}/d}.$$ \end{lem} \begin{proof} Given $\ld \vdash n$ (i.e., a partition of $n$), the number of permutations in $S_{n}$ with cycle type $\ld$ is $$\frac{n!}{m_{1}(\ld)!1^{m_{1}(\ld)} \cdots m_{n}(\ld)!n^{m_{n}(\ld)}},$$ \ so $$\mc{Z}(S_{n}, \bs{x}) = \sum_{\ld \vdash n}\frac{x_1^{m_{1}(\ld)} \cdots x_{n}^{m_{n}(\ld)}}{m_{1}(\ld)!1^{m_{1}(\ld)} \cdots m_{n}(\ld)!n^{m_{n}(\ld)}}.$$ \ Thus, we have \begin{align*} \sum_{n=0}^{\infty}\mc{Z}(S_{n}, \bs{x})u^{n} &= \sum_{n=0}^{\infty}\sum_{\ld \vdash n}\frac{x_1^{m_{1}(\ld)} \cdots x_{n}^{m_{n}(\ld)} u^{1 \cdot m_{1}(\ld)} \cdots u^{n \cdot m_{n}(\ld)}}{m_{1}(\ld)!1^{m_{1}(\ld)} \cdots m_{n}(\ld)!n^{m_{n}(\ld)}} \\ &= \sum_{\ld \in \mc{P}} \prod_{d=1}^{\infty} \frac{ (x_{d}u^{d}/d)^{m_{d}(\ld)} }{m_{d}(\ld)!} \\ &= \prod_{d=1}^{\infty}\sum_{m=0}^{\infty}\frac{(x_{d}u^{d}/d)^{m}}{m!} \\ &= \prod_{d=1}^{\infty}e^{x_{d}u^{d}/d}, \end{align*} \ as desired. \end{proof} \ \begin{lem}[Lemma 1 (2) in \cite{Sto}]\label{fac2} In the ring $\bQ [\bs{x}] \llb u \rrb$ of formal power series in $u$ over the polynomial ring $\bQ [\bs{x}] = \bQ[x_{P,\ld}]_{P \in |\bA^{1}_{\bF_{q}}|, \ld \in \mc{P} \sm \{\es\}}$, we have $$\sum_{n=0}^{\infty}\mc{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})u^n = \prod_{P \in |\bA^{1}_{\bF_{q}}|} \sum_{\ld \in \mc{P}} \frac{x_{P,\ld}u^{|\ld|\deg(P)}}{|\Aut_{\bF_{q}[t]}(H_{P, \ld})|},$$ \ where $H_{P, \ld} := H_{(P),\ld}^{\bF_{q}[t]}$, following the notation defined in the beginning of Section \ref{Stong}. \end{lem} \ \begin{rmk} Lemma \ref{fac2} says that on the left-hand side, the coefficient of $u^{n}$ is given by $$\sum_{|\ld^{(1)}|\deg(P_{1}) + \cdots + |\ld^{(s)}|\deg(P_{s}) = n}\frac{x_{P_{1}, \ld^{(1)}} \cdots x_{P_{s}, \ld^{(s)}}}{|\Aut_{\bF_{q}[t]}(H_{P_{1}, \ld^{(1)}})| \cdots |\Aut_{\bF_{q}[t]}(H_{P_{s}, \ld^{(s)}})|},$$ \ where the sum is over distinct $P_{1}, \dots, P_{s} \in |\bA^{1}_{\bF_{q}}|$ and not necessarily distinct $\ld^{(1)}, \dots, \ld^{(s)}$ with the stipulated condition. This is a finite sum, so it makes sense to evaluate any complex numbers in the variables $x_{P, \ld}$ for $\ld \neq \es$ of the infinite product in the statement of Lemma \ref{fac2} and get an identity in $\bC \llb u \rrb$. (Recall that $x_{P, \es} = 1$.) \end{rmk} \ \subsection{Connection between two cycle indices} We define $\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})$ to be the polynomial given by specializing $x_{P,\ld} = x_{\deg(P)}^{|\ld|}$ in the cycle index $\mc{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})$, which also makes sense for the empty partition because $x_{P,\es} = 1 = x_{\deg(P)}^{0}$. More explicitly, we have \begin{align*} \bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x}) &= \frac{1}{|\GL_{n}(\bF_{q})|}\sum_{A \in \Mat_{n}(\bF_{q})}\prod_{P \in |\bA^{1}_{\bF_{q}}|}x_{\deg(P)}^{|\mu_{P}(A)|} \\ &= \frac{1}{|\GL_{n}(\bF_{q})|}\sum_{A \in \Mat_{n}(\bF_{q})}x_{1}^{m_{1}(A)} \cdots x_{n}^{m_{n}(A)}, \end{align*} \ where $m_{d}(A)$ is the number of degree $d$ monic irreducible polynomials, counting with multiplicity, in $\bF_{q}[t]$ dividing the characteristic polynomial $f_{A}(t)$ of $A$. Note that only $x_{1}, \dots, x_{n}$ will occur in $\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q})$ because the degree of $f_A(t)$ is $n$ for $A \in \Mat_{n}(\bF_{q})$, so $f_A(t)$ is not divisible by any irreducible polynomial of degree $> n$. In the concluding remark of \cite{Sto}, Stong essentially observed that there is a close relationship between $\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})$ and $\mc{Z}(S_{n}, \bs{x})$ when $q$ is large. We rigorously formulate what he might have meant. \ \begin{lem}[cf. ``Conclusion'' in \cite{Sto}]\label{conv} We have $$\lim_{q \ra \infty}\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x}) = \mc{Z}(S_{n}, \bs{x}),$$ \ meaning that the coefficients of $\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})$ converge to the coefficients of $\mc{Z}(S_{n}, \bs{x})$ as $q \ra \infty$. \end{lem} \begin{proof} Let $x_{d} \in \bC$ such that $|x_{d}| \leq 1$ for any $d \in \bZ_{\geq 1}$. Applying Lemma \ref{fac2} and then Lemma \ref{key}, we have \begin{align*} \sum_{n=0}^{\infty}\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})u^{n} &= \prod_{P \in |\bA^{1}_{\bF_{q}}|} \sum_{\ld \in \mc{P}} \frac{(x_{\deg(P)}u^{\deg(P)})^{|\ld|}}{|\Aut_{\bF_{q}[t]}(H_{P, \ld})|} \\ &= \prod_{P \in |\bA^{1}_{\bF_{q}}|} \prod_{i=1}^{\infty}(1 - x_{\deg(P)}(q^{-i}u)^{\deg(P)})^{-1} \\ &= \prod_{d=1}^{\infty}\prod_{i=1}^{\infty}(1 - x_{d}(q^{-i}u)^{d})^{-M(q,d)}, \end{align*} \ for $|u| < 1$, where $M(q, d)$ is the number of monic irreducible polynomials in $\bF_{q}[t]$ with degree $d$. Since $|x_{d}| \leq 1$ for all $d \geq 1$ so that we have $$|\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})|^{1/n} \leq \left(\prod_{i=1}^{n}\frac{1}{1 - q^{-i}}\right)^{1/n} \leq \frac{1}{1 - q^{-1}},$$ \ the radius of convergence of the power series $\sum_{n=0}^{\infty}\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})u^{n}$ in $u$ is at least $1 - q^{-1} > 0$. Since our statement only involves finitely many $x_{d}$, we may set all but finitely many of them to be $0$, say $x_{d} = 0$ for all $d > m$ for fixed $m \in \bZ_{\geq 1}$. This lets us have $$\sum_{n=0}^{\infty}\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})u^{n} = \prod_{d=1}^{m}\prod_{i=1}^{\infty}(1 - x_{d}(q^{-i}u)^{d})^{-M(q,d)}.$$ \ In what follows, we will use the fact that $$\lim_{q \ra \infty}\frac{M(q,d)}{q^{d}/d} = 1,$$ \ for any $d \geq 1$, which can be found as Theorem 2.2 in \cite{Ros}. (Note that for $M(q, 1) = q$, so we do not need to take the limit to see this for $d = 1$.) Now, for $|u| < 1 - q^{-1}$, we have \begin{align*} \lim_{q \ra \infty}\sum_{n=0}^{\infty}\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})u^{n} &= \prod_{d=1}^{m}\lim_{q \ra \infty}\prod_{i=1}^{\infty}(1 - x_{d}(q^{-i}u)^{d})^{-M(q,d)} \\ &= \prod_{d=1}^{m}\lim_{q \ra \infty}\prod_{i=1}^{\infty}\left((1 - x_{d}(q^{-i}u)^{d})^{-q^{d}/d}\right)^{M(q,d)d/q^{d}} \\ &= \prod_{d=1}^{m}\prod_{i=1}^{\infty}\lim_{q \ra \infty}\left((1 - x_{d}(q^{-i}u)^{d})^{-q^{d}/d}\right)^{M(q,d)d/q^{d}} \\ &= \prod_{d=1}^{\infty}\lim_{q \ra \infty}(1 - x_{d}(q^{-1}u)^{d})^{-q^{d}/d} \\ &= \prod_{d=1}^{\infty}\left(\lim_{q \ra \infty}\left(1 - \frac{x_{d}u^{d}}{q^{d}}\right)^{- q^{d}/(x_{d}u^{d})}\right)^{(x_{d}u^{d})/d} \\ &= \prod_{d=1}^{\infty}e^{x_{d}u^{d}/d}. \end{align*} \ Applying Lemma \ref{fac1} to the last expression, we see that $$\lim_{q \ra \infty}f_{q}(u) = f(u),$$ \ where $$f_{q}(u) := \sum_{n=0}^{\infty}\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})u^{n}$$ \ and $$f(u) := \sum_{n=0}^{\infty}\mc{Z}(S_{n}, \bs{x})u^{n}$$ \ are holomorphic functions in $u$ defined for $|u| < 1-q^{-1}$. Take any $0 < \epsilon < 1-q^{-1}$ and write $C_{\eps} := \{u \in \bC : |u| = \eps\}$. By the Cauchy integral formula, we have $$\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x}) = \frac{1}{2\pi i}\varointctrclockwise_{z \in C_{\eps}}\frac{f_{q}(z)}{z^{n+1}}dz$$ \ and $$\mc{Z}(S_{n}, \bs{x}) = \frac{1}{2\pi i}\varointctrclockwise_{z \in C_{\eps}}\frac{f(z)}{z^{n+1}}dz.$$ \ Thus, by the Dominated Convergence Theorem, we have \begin{align*} \lim_{q \ra\infty} \bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x}) &= \lim_{q \ra \infty} \frac{1}{2\pi i}\varointctrclockwise_{z \in C_{\eps}}\frac{f_{q}(z)}{z^{n+1}}dz \\ &= \frac{1}{2\pi i}\varointctrclockwise_{z \in C_{\eps}}\frac{f(z)}{z^{n+1}}dz \\ &= \mc{Z}(S_{n}, \bs{x}). \end{align*} \ Since $x_{1}, \dots, x_{m}$ are arbitrary with the restriction $|x_{d}| \leq 1$, this is enough to prove the desired statement about the convergence of the coefficients in $x_{1}, \dots, x_{m}$ by the Cauchy integral formula for several variables (e.g., Theorem 2.1.3 of \cite{Fie}) with the Dominate Convergence Theorem. \end{proof} \ \subsection{Proof of Proposition \ref{main}} Again, Lemma \ref{conv} says that the coefficients of the polynomial $$\bs{Z}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x}) = \frac{1}{|\GL_{n}(\bF_{q})|}\sum_{A \in \Mat_{n}(\bF_{q})}x_{1}^{m_{1}(A)} \cdots x_{n}^{m_{n}(A)} \in \bQ[x_{1}, \dots, x_{n}]$$ \ converge to the coefficients of the polynomial $$\mc{Z}(S_{n}, \bs{x}) = \frac{1}{|S_{n}|}\sum_{\sigma \in S_{n}}x_{1}^{m_{1}(\sigma)} \cdots x_{n}^{m_{n}(\sigma)} \in \bQ[x_{1}, \dots, x_{n}]$$ \ as $q \ra \infty$, where the convergences are happening in $\bR$ or $\bC$, although the limits still lie in $\bQ$. This implies that, setting $x_{d,0} := 1$ and letting $x_{d,m}$ be formal for any $d, m \in \bZ_{\geq 1}$, the coefficients of $$\hat{\bs{Z}}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x}) := \frac{1}{|\GL_{n}(\bF_{q})|}\sum_{A \in \Mat_{n}(\bF_{q})}x_{1, m_{1}(A)} \cdots x_{n, m_{n}(A)} \in \bQ \left[\begin{array}{c} x_{1,1}, \dots, x_{1,n}, \\ \cdots, \cdots, \cdots \\ x_{n,1}, \dots, x_{n,n} \end{array}\right]$$ \ converge to the coefficients of $$\hat{\mc{Z}}(S_{n}, \bs{x}) := \frac{1}{|S_{n}|}\sum_{\sigma \in S_{n}}x_{1, m_{1}(\sigma)} \cdots x_{n, m_{n}(\sigma)} \in \bQ \left[\begin{array}{c} x_{1,1}, \dots, x_{1,n}, \\ \cdots, \cdots, \cdots \\ x_{n,1}, \dots, x_{n,n} \end{array}\right].$$ \ To refer back to this fact, we record this as follows. \ \begin{lem}\label{conv'} As $q \ra \infty$, the coefficients of the polynomial $\hat{\bs{Z}}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x})$ converge to those of $\hat{\mc{Z}}(S_{n}, \bs{x})$. In particular, for any evaluation of the variables $x_{d,m}$ in $\bC$, we have $$\lim_{q \ra \infty}\hat{\bs{Z}}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x}) = \hat{\mc{Z}}(S_{n}, \bs{x}).$$ \end{lem} \ \hspace{3mm} We are now ready to prove Proposition \ref{main}: \begin{proof}[Proof of Proposition \ref{main}] Recall the statements of Proposition \ref{main}, since we will use the notations in them. We will also use the notations given in this section. If we evaluate \begin{itemize} \item $x_{d_{1},k_{1}} = \cdots = x_{d_{r},k_{r}} = 1$, \item $x_{d_{j},m} = 0$ for all $m \neq 0, k_{j}$ for $1 \leq j \leq r$, and \item $x_{d,m} = 1$ for all the other cases from above, \end{itemize} \ then we have $$\hat{\bs{Z}}([\Mat_{n}/\GL_{n}](\bF_{q}), \bs{x}) = \frac{\left|\left\{\begin{array}{c} A \in \Mat_{n}(\bF_{q}) : f_{A}(t) \text{ has } k_{j} \text{ irreducible} \\ \text{factors of degree } d_{j} \text{ for } 1 \leq j \leq r \end{array} \right\}\right|}{|\GL_{n}(\bF_{q})|},$$ \ where we count the factors with multiplicity, and $$\hat{\mc{Z}}(S_{n}, \bs{x}) = \frac{\left|\left\{\begin{array}{c} \sigma \in S_{n} : \sigma \text{ has } k_{j} \text{ cycles} \\ \text{of length } d_{j} \text{ for } 1 \leq j \leq r \end{array} \right\}\right|}{|S_{n}|}.$$ \ Thus, applying Lemma \ref{conv'} and noting that $$\lim_{q \ra \infty}\frac{|\Mat_{n}(\bF_{q})|}{|\GL_{n}(\bF_{q})|} = \lim_{q \ra \infty}(1 - q^{-1})^{-1} \cdots (1 - q^{-n})^{-1} = 1,$$ \ we have $$\lim_{q \ra \infty}\Prob_{A \in \Mat_{n}(\bF_{q})}\left(\begin{array}{c} A \in \Mat_{n}(\bF_{q}) : f_{A}(t) \text{ has } k_{j} \\ \text{irreducible factors of} \\ \text{degree } d_{j} \text{ for } 1 \leq j \leq r \end{array} \right) = \Prob_{\sigma \in S_{n}}\left(\begin{array}{c} \sigma \in S_{n} : \sigma \text{ has } k_{j} \text{ cycles} \\ \text{of length } d_{j} \text{ for } 1 \leq j \leq r \end{array} \right),$$ \ as desired. \end{proof} \ \section{Combinatorial proof of Lemma \ref{perm}}\label{permproof} \hspace{3mm} In this section, we give a proof of Lemma \ref{perm}, a result due to Shepp and Lloyd in \cite{SL}, originally proven by computing the characteristic functions of the distributions. Unlike their proof, we will directly compute the desired probability with a more combinatorial method. Most of our argument will be encoded in the following lemma. \ \begin{lem}\label{issue} Take \begin{itemize} \item $x_{d_{j},m} = 0$ with $1 \leq j \leq s$ and $m \neq 0$, \item $x_{d_{j},m} = 0$ with $s + 1 \leq j \leq r$ and $m \neq 0, k_{j}$, and \item $x_{d,m} = 1$ for any other ones not on the above list. \end{itemize} Then $$\sum_{n=0}^{\infty} \hat{\mc{Z}}(S_{n}, \bs{x})u^{n} = \left( \frac{e^{-u^{d_1}/d_{1}} \cdots e^{-u^{d_r}/d_{r}}}{1 - u} \right) \cdot \left( \prod_{j=s+1}^{r} \left(1 + \frac{(u^{d_j}/d_{j})^{k_{j}}}{k_{j}!} \right) \right) \in \bC \llb u \rrb.$$ \end{lem} \ \begin{rmk}\label{Tau} Before proving Lemma \ref{issue}, we first see how it implies Lemma \ref{perm}. We will make use of the following useful observation: for any $f(u) = c_{0} + c_{1}u + c_{2}u^{2} + \cdots \in \bC \llb u \rrb$, if $f(1) = c_{0} + c_{1} + c_{2} + \cdots$ exists, then the limit of the coefficient sequence of the power series $$\frac{f(u)}{1 - u} = b_{0} + b_{1}u + b_{2}u^{2} + \cdots$$ \ is given by $$\lim_{n \ra \infty} b_{n} = f(1),$$ \ because $(1 - u)^{-1} = 1 + u + u^{2} + \cdots$ so that $b_{n} = c_{0} + c_{1} + \cdots + c_{n}$, which is the coefficient of $u^{n}$ for the power series $f(u)(1 - u)^{-1}$. \end{rmk} \ \begin{proof}[Proof of Lemma \ref{perm}] Using Lemma \ref{issue} with $s = r$, we have $$\sum_{n=0}^{\infty}\Prob_{\sigma \in S_{n}}\left( m_{d_{j}}(\sigma) = 0 \text{ for } 1 \leq j \leq r \right) u^{n} = \frac{e^{-u^{d_1}/d_{1}} \cdots e^{-u^{d_r}/d_{r}}}{1 - u}.$$ \ Using Remark \ref{Tau}, this implies that $$\lim_{n \ra \infty}\Prob_{\sigma \in S_{n}}\left(m_{d_{j}}(\sigma) = 0 \text{ for } 1 \leq j \leq r\right) = e^{-1/d_{1}} \cdots e^{-1/d_{r}},$$ \ as claimed. For the general case, we work with the induction on the number of nonzero integers among $k_{1}, \dots, k_{r}$. The base case is when such number is $0$, which is exactly what we have proved above. Suppose that the result is true when such number is $0, 1, \dots, s-1$, where $s - 1 \geq 0$. Then we assume that there are $s$ nonzero elements among $k_{1}, \dots, k_{r}$, so permuting them if necessary, say $$k_{1} = \cdots = k_{s} = 0,$$ \ while $$k_{s+1}, \dots, k_{r} \neq 0.$$ \ Applying Lemma \ref{issue}, we get \begin{align*} \sum_{n=0}^{\infty}&\Prob_{\sigma \in S_{n}}\left( \begin{array}{c} m_{d_{j}}(\sigma) = 0 \text{ for } 1 \leq j \leq s, \text{ and } \\ m_{d_{s+1}}(\sigma) \in \{0, k_{s+1}\}, \dots, m_{d_{r}}(\sigma) \in \{0, k_{r}\} \end{array} \right) u^{n} \\ &= \left( \frac{e^{-u^{d_1}/d_{1}} \cdots e^{-u^{d_r}/d_{r}}}{1 - u} \right) \cdot \left( \prod_{j=s+1}^{r} \left(1 + \frac{(u^{d_j}/d_{j})^{k_{j}}}{k_{j}!} \right) \right), \end{align*} \ so \begin{align*} \sum_{\eps_{s+1} \in \{0, k_{s+1}\}} \cdots \sum_{\eps_{r} \in \{0, k_{r}\}} & \lim_{n \ra \infty}\Prob_{\sigma \in S_{n}}\left( \begin{array}{c} m_{d_{1}}(\sigma) = \cdots = m_{d_{s}}(\sigma) = 0, \\ m_{d_{s+1}}(\sigma) = \eps_{s+1}, \dots, m_{d_{r}}(\sigma) = \eps_{r} \end{array} \right) \\ &= e^{-1/d_{1}} \cdots e^{-1/d_{r}} \prod_{j=s+1}^{r} \left(1 + \frac{(1/d_{j})^{k_{j}}}{k_{j}!} \right) \end{align*} \ Applying the induction hypothesis after expanding the right-hand side of the above identity implies that $$\Prob_{\sigma \in S_{n}}\left( \begin{array}{c} m_{d_{1}}(\sigma) = \cdots = m_{d_{s}}(\sigma) = 0, \\ m_{d_{s+1}}(\sigma) = k_{s+1}, \dots, m_{d_{r}}(\sigma) = k_{r} \end{array} \right) = e^{-1/d_{1}} \cdots e^{-1/d_{r}} \frac{(1/d_{s+1})^{k_{s+1}}}{k_{s+1}!} \cdots \frac{(1/d_{r})^{k_{r}}}{k_{r}!},$$ \ which finishes the proof. \end{proof} \ \hspace{3mm} Finally, we provide a proof of Lemma \ref{issue}. \ \begin{proof}[Proof of Lemma \ref{issue}] For now, let us not evaluate any variables in $\bs{x} = (x_{d,m})$. Recall that $$\hat{\mc{Z}}(S_{n}, \bs{x}) = \frac{1}{|S_{n}|}\sum_{\sigma \in S_{n}}x_{1, m_{1}(\sigma)} \cdots x_{n, m_{n}(\sigma)},$$ \ where $x_{d,0} = 1$ and $x_{d,m}$ are defined to be formal variables for $d, m \geq 1$. We observe that $$\hat{\mc{Z}}(S_{n}, \bs{x}) = \sum_{\ld \vdash n}\frac{x_{1,m_{1}(\ld)} \cdots x_{n,m_{n}(\ld)}}{m_{1}(\ld)!1^{m_{1}(\ld)} \cdots m_{n}(\ld)!n^{m_{n}(\ld)}},$$ \ so $$\sum_{n=0}^{\infty}\hat{\mc{Z}}(S_{n}, \bs{x})u^{n} = \prod_{d=1}^{\infty}\sum_{m=0}^{\infty}\frac{x_{d,m}u^{dm}}{m!d^{m}} = \prod_{d=1}^{\infty}\sum_{m=0}^{\infty}\frac{x_{d,m}(u^d/d)^{m}}{m!}.$$ \ Note that taking all $x_{d,m} = 1$ in the identity, we get $$\frac{1}{1 - u} = 1 + u + u^{2} + \cdots + = \prod_{d=1}^{\infty}\sum_{m=0}^{\infty}\frac{(u^d/d)^{m}}{m!}.$$ \ Hence, with the given evaluation in the variables $\bs{x} = (x_{d,m})$, we get \begin{align*} \sum_{n=0}^{\infty}\hat{\mc{Z}}(S_{n}, \bs{x}) u^{n} &= \left( \prod_{d=1}^{\infty}\sum_{m=0}^{\infty}\frac{(u^d/d)^{m}}{m!} \right) \cdot \left( \prod_{j=1}^{r} \left( \sum_{m=0}^{\infty}\frac{(u^{d_j}/d_{j})^{m}}{m!} \right)^{-1} \right) \cdot \left( \prod_{j=s+1}^{r} \left(1 + \frac{(u^{d_j}/d_{j})^{k_{j}}}{k_{j}!} \right) \right) \\ &= \left( \frac{e^{-u^{d_1}/d_{1}} \cdots e^{-u^{d_r}/d_{r}}}{1 - u} \right) \cdot \left( \prod_{j=s+1}^{r} \left(1 + \frac{(u^{d_j}/d_{j})^{k_{j}}}{k_{j}!} \right) \right). \end{align*} \ Since the identities makes sense as those of holomorphic functions in $u \in \bC$ with $|u| < 1$, this finishes the proof. \end{proof} \ \newpage
{ "timestamp": "2020-05-19T02:06:15", "yymm": "2005", "arxiv_id": "2005.07846", "language": "en", "url": "https://arxiv.org/abs/2005.07846", "abstract": "Given a positive integer $r$ and a prime power $q$, we estimate the probability that the characteristic polynomial $f_{A}(t)$ of a random matrix $A$ in $\\mathrm{GL}_{n}(\\mathbb{F}_{q})$ is square-free with $r$ (monic) irreducible factors when $n$ is large. We also estimate the analogous probability that $f_{A}(t)$ has $r$ irreducible factors counting with multiplicity. In either case, the main term $(\\log n)^{r-1}((r-1)!n)^{-1}$ and the error term $O((\\log n)^{r-2}n^{-1})$, whose implied constant only depends on $r$ but not on $q$ nor $n$, coincide with the probability that a random permutation on $n$ letters is a product of $r$ disjoint cycles. The main ingredient of our proof is a recursion argument due to S. D. Cohen, which was previously used to estimate the probability that a random degree $n$ monic polynomial in $\\mathbb{F}_{q}[t]$ is square-free with $r$ irreducible factors and the analogous probability that the polynomial has $r$ irreducible factors counting with multiplicity. We obtain our result by carefully modifying Cohen's recursion argument in the matrix setting, using Reiner's theorem that counts the number of $n \\times n$ matrices with a fixed characteristic polynomial over $\\mathbb{F}_{q}$.", "subjects": "Combinatorics (math.CO); Number Theory (math.NT)", "title": "Jordan--Landau theorem for matrices over finite fields", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9865717464424892, "lm_q2_score": 0.8128673178375734, "lm_q1q2_score": 0.8019519293850368 }
https://arxiv.org/abs/2208.08506
On $d$-permutations and Pattern Avoidance Classes
Bonichon and Morel first introduced $d$-permutations in their study of multidimensional permutations. Such permutations are represented by their diagrams on $[n]^d$ such that there exists exactly one point per hyperplane $x_i$ that satisfies $x_i= j$ for $i \in [d]$ and $j \in [n]$. Bonichon and Morel previously enumerated $3$-permutations avoiding small patterns, and we extend their results by first proving four conjectures, which exhaustively enumerate $3$-permutations avoiding any two fixed patterns of size $3$. We further provide a enumerative result relating $3$-permutation avoidance classes with their respective recurrence relations. In particular, we show a recurrence relation for $3$-permutations avoiding the patterns $132$ and $213$, which contributes a new sequence to the OEIS database. We then extend our results to completely enumerate $3$-permutations avoiding three patterns of size $3$.
\section{Introduction} Starting with Knuth's \cite{knuth1973art} work on permutations in 1973, the field of pattern avoidance has been well-studied in enumerative combinatorics. Simion and Schmidt first considered pattern avoidance in their work on enumerating permutation avoidance classes in 1985 \cite{simion1985restricted}. Pattern avoidance can be defined as follows: \begin{definition} Let $\sigma \in S_{n}$ and $\pi \in S_{k}$, where $k \leq n$. We say that the permutation $\sigma$ \emph{contains} the pattern $\pi$ if there exists indices $c_1 < \dots < c_k$ such that $\sigma(c_1) \cdots \sigma(c_k)$ is order-isomorphic to $\pi$. We say a permutation \emph{avoids} a pattern if it does not contain it. \end{definition} It is well-known that permutations avoiding certain patterns are in bijection with other combinatorial objects, such as Dyck paths \cite{krattenthaler2001permutations, reifegerste2003diagram}, maximal chains of lattices \cite{simion1985restricted}, and the Catalan and Schr{\"o}der numbers \cite{west1995generating}. In their work, Simion and Schmidt \cite{simion1985restricted} completely enumerated permutations avoiding any single pattern, two patterns, or three patterns of size $3$, paving the path for more work in the field of pattern avoidance. More recently, Bonichon and Morel \cite{bonichon2022baxter} defined a multidimensional generalization of a permutation, called a $d$-permutation, which resembles the structure of a $(d-1)$-tuple of permutations. Tuples of permutations have been studied before \cite{gunby2019asymptotics, aldred2005permuting}, but $d$-permutations have not been thoroughly studied yet, mainly appearing in a few papers related to separable permutations \cite{gunby2019asymptotics, asinowski2010separable}. In particular, Asinoski and Mansour \cite{asinowski2010separable} presented a generalization of separable permutations that are similar to $d$-permutations and characterized these generalized permutations with sets of forbidden patterns. The study of pattern-avoidance classes of permutations has received much attention and permutations avoiding sets of small patterns have been exhaustively enumerated \cite{simion1985restricted, mansour2020enumeration, knuth1973art}. However, $d$-permutations introduced by Bonichon and Morel are slightly different than the one introduced by Asinoski and Mansour \cite{asinowski2010separable} and coincide with the classical permutation for $d=2$. Similar to the enumeration Simion and Schmidt \cite{simion1985restricted} did in 1985, Bonichon and Morel \cite{bonichon2022baxter} started the enumeration of $d$-permutations avoiding small patterns and made many conjectures regarding the enumeration of $3$-permutations avoiding sets of two patterns. We present two main classes of results regarding the enumeration of $3$-permutation avoiding small patterns. We first completely enumerate $3$-permutations avoiding classes of two patterns and prove their respective recurrence relations, solving the conjectures presented by Bonichon and Morel \cite{bonichon2022baxter}. Further, we derive a recurrence relation for $3$-permutations avoiding $132$ and $213$, which does not appear on the OEIS database \cite{oeis} and Bonichon and Morel were unable to make a conjecture about. We then further initiate and completely enumerate $3$-permutations avoiding classes of three patterns, similar to Simion and Schmidt \cite{simion1985restricted}'s results in 1985. This paper is organized as follows. In Section 2, we introduce preliminary definitions and notation. In Section 3, we completely enumerate sequences of $3$-permutations avoiding two patterns of size 3 and prove four conjectures of Bonichon and Morel \cite{bonichon2022baxter}. In addition, we prove a recurrence relation for an avoidance class that does not appear on the OEIS database \cite{oeis}, completing our enumeration. In Section 4, we extend our enumeration to $3$-permutations avoiding three patterns of size 3 and prove recurrence relations for their avoidance classes. We conclude with open problems in Section 5. \section{Preliminaries} \label{sec:preliminaries} Let $S_n$ denote the set of permutations of $[n] = \{ 1,2, \dots, n \}$. Note that we can represent each permutation $\sigma \in S_n$ as a sequence $\sigma(1) \cdots \sigma(n)$. Further, let $\mathrm{Id}_n$ denote the identity permutation $12 \cdots n$ of size $n$ and given a permutation $\sigma \in S_n$, let $\rev(\sigma)$ denote the reverse permutation $\sigma(n) \sigma(n-1) \cdots \sigma(1)$. We further say that a sequence $w$ is \emph{consecutively increasing} (respectively \emph{decreasing}) if for every index $i$, $w(i+1) = w(i)+1$ (respectively $w(i+1) = w(i)-1$). For a sequence $w = w(1) \cdots w(n)$ with distinct real values, the \emph{standardization} of $w$ is the unique permutation with the same relative order. Note that once standardized, a consecutively-increasing sequence is the identity permutation and a consecutively-decreasing sequence is the reverse identity permutation. Moreover, we say that in a permutation $\sigma$, the elements $\sigma(i)$ and $\sigma(i+1)$ are \emph{adjacent} to each other. More specifically, $\sigma(i)$ is \emph{left-adjacent} to $\sigma(i+1)$ and similarly, the element $\sigma(i+1)$ is $\emph{right-adjacent}$ to $\sigma(i)$. The following definitions in this section were introduced in \cite{bonichon2022baxter}. \begin{definition} A \emph{$d$-permutation} $\boldsymbol{\sigma} := (\sigma_1 , \dots, \sigma_{d-1})$ of size $n$ is a tuple of permutations, each of size $n$. Let $S_{n}^{d-1}$ denote the set of $d$-permutations of size $n$. We say that $d$ is the \textit{dimension} of $\boldsymbol{\sigma}$. Moreover, the \textit{diagram} of $\sigma$ is the set of points $(i, \sigma_1(i) ,\dots, \sigma_{d-1}(i))$ for all $i \in [n]$. \end{definition} Note that the identity permutation is implicitly included in the diagram of a $d$-permutation, which justifies why a $d$-permutation is a $(d-1)$-tuple of permutations. For a $d$-permutation $\boldsymbol{\sigma} = (\sigma_1 ,\dots, \sigma_{d-1}),$ let $\boldsymbol{\Bar{\sigma}} = (\mathrm{Id}_n, \sigma_1, \dots, \sigma_{d-1}).$ Further, with this definition, it is natural to consider the projections of the diagram of a $d$-permutation, which is useful in defining the notion of pattern avoidance for $d$-permutations. \begin{definition} Given $d' \in \mathbb N$ and $\boldsymbol{i} = i_1, \dots, i_{d'} \in [d]^{d'}$, the \emph{projection on $\boldsymbol{i}$} of some $d$-permutation $\boldsymbol{\sigma}$ is the $d'$-permutation $\mathrm{proj}_{\boldsymbol{i}}(\boldsymbol{\sigma}) = (\boldsymbol{\Bar{\sigma}}_{i_2} \circ \boldsymbol{\Bar{\sigma}}_{i_1}^{-1}, \dots, \boldsymbol{\Bar{\sigma}}_{i_{d'}} \circ \boldsymbol{\Bar{\sigma}}_{i_1}^{-1} ).$ \end{definition} We say that a projection is \emph{direct} if $i_1 < \dots < i_{d'}$ and \emph{indirect} otherwise. \begin{remark} There are only three direct projections of dimension $2$ of a $3$-permutation $\boldsymbol{\sigma} = (\sigma, \sigma')$. Namely, they are $\sigma$, $\sigma'$, and $\sigma' \circ \sigma^{-1}$. \end{remark} In the remainder of the section, we use the projection of a $3$-permutation $\boldsymbol{\sigma} = (\sigma, \sigma')$ to refer to the projection $\sigma' \circ \sigma^{-1}$. Using direct projections, Bonichon and Morel \cite{bonichon2022baxter} introduced the following definition of pattern avoidance, which coincides with the existing concept of pattern avoidance for regular permutations. \begin{definition} Let $\boldsymbol{\sigma} = (\sigma_1 ,\dots, \sigma_{d-1}) \in S_{n}^{d-1}$ and $\boldsymbol{\pi} = (\pi_1, \dots, \pi_{d'-1}) \in S_{k}^{d'-1}$, where $k \leq n$. We say that the $d$-permutation $\boldsymbol{\sigma}$ \emph{contains} the pattern $\boldsymbol{\pi}$ if there exists a direct projection $\boldsymbol{\sigma'}$ of dimension $d'$ and indices $c_1 < \dots < c_k$ such that $\boldsymbol{\sigma'}_i(c_1) \cdots \boldsymbol{\sigma'}_i(c_k)$ is order-isomorphic to $\pi_i$ for all $i$. We say a $d$-permutation \emph{avoids} a pattern if it does not contain it. \end{definition} Given $m$ patterns $\boldsymbol{\pi_1}, \dots, \boldsymbol{\pi_m} \in S^{d'-1}_{n'}$, we write $S_{n}^{d-1}(\boldsymbol{\pi_1}, \dots, \boldsymbol{\pi_m})$ to mean the set of $d$-permutations of size $n$ that simultaneously avoid $\boldsymbol{\pi_1}, \dots, \boldsymbol{\pi_m}$. Bonichon and Morel \cite{bonichon2022baxter} also noted symmetries on $d$-permutations that correspond to symmetries on the $d$-dimensional cube. In particular, these symmetries are counted by signed permutation matrices of dimension $d$. Such a signed permutation matrix is a square matrix with entries consisting of $-1, 0$, or $1$ such that each row and column contain exactly one nonzero element. We call $\textit{d-Sym}$ the set of such signed permutation matrices of size $d$. This allows us to extend the well-known definitions of Wilf-equivalence and trivial Wilf-equivalence to higher dimensions. \begin{definition} We say that two sets of patterns $\boldsymbol{\pi_1}, \dots, \boldsymbol{\pi_k}$ and $\boldsymbol{\tau_1}, \dots, \boldsymbol{\tau_\ell}$ are \emph{d-Wilf-equivalent} if $|S_{n}^{d-1}(\boldsymbol{\pi_1}, \dots, \boldsymbol{\pi_k})| = |S_n^{d-1}(\boldsymbol{\tau_1}, \dots, \boldsymbol{\tau_\ell})|$. Moreover, these patterns are \emph{trivially d-Wilf-equivalent} if there exists a symmetry $s \in \textit{d-Sym}$ that maps $S_{n}^{d-1}(\boldsymbol{\pi_1}, \dots, \boldsymbol{\pi_k})$ to $S_n^{d-1}(\boldsymbol{\tau_1}, \dots, \boldsymbol{\tau_\ell})$ bijectively. \end{definition} \section{Enumeration of Pattern Avoidance Classes of at most size 2} \label{sec:enumeration} Bonichon and Morel \cite{bonichon2022baxter} proposed the problem of enumerating sequences of $3$-permutations avoiding at most two patterns of size 2 or 3. They provided Table \ref{double avoidance}, conjecturing the recurrences in the last four rows and leaving the remainder as open problems. \begin{table}[htp] \centering \begin{tabular}{|c | c | c | c | c|} \hline Patterns & \#TWE & Sequence & OEIS Sequence & Comment \\ [0.5ex] \hline\hline 12 & 1 & $1,0,0,0,0,\dots$ & & \cite{bonichon2022baxter} \\ \hline 21 & 1 & $1,1,1,1,1,\dots$ & & \cite{bonichon2022baxter} \\ \hline 123 & 1 & $1,4,20,100,410,1224,2232, \dots$ & & Not in OEIS \\ \hline 132 & 2 & $1,4,21,116,646,3596,19981, \dots$ & & Not in OEIS \\ \hline 231 & 2 & $1,4,21,123,767,4994,35584, \dots$ & & Not in OEIS \\ \hline 321 & 1 & $1,4,21,128,850,5956,43235, \dots$ & & Not in OEIS \\ \hline $123, 132$ & 2 & $1,4,8,8,0,0,0, \dots$ & & Terminates after $n=4$ \\ \hline $123, 231$ & 2 & $1,4,9,6,0,0,0, \dots$ & & Terminates after $n=4$ \\ \hline $123, 321$ & 1 & $1,4,8,0,0,0,0, \dots$ & & Terminates after $n=3$ \\ \hline $132, 213$ & 1 & $1,4,12,28,58,114,220, \dots$ & & Theorem \ref{132,213} \\ \hline $132, 231$ & 4 & $1,4,12,32,80,192,448, \dots$ & \href{http://oeis.org/A001787}{A001787} & Theorem \ref{132,231} \\ \hline $132, 321$ & 2 & $1,4,12,27,51,86,134, \dots$ & \href{http://oeis.org/A047732}{A047732} & Theorem \ref{132,321} \\ \hline $231, 312$ & 1 & $1,4,10,28,76,208,568, \dots$ & \href{http://oeis.org/A026150}{A026150} & Theorem \ref{231,312} \\ \hline $231, 321$ & 2 & $1,4,12,36,108,324,972, \dots$ & \href{http://oeis.org/A003946}{A003946} & Theorem \ref{231,321} \\ \hline \end{tabular} \caption{Sequences of $3$-permutations avoiding at most two patterns of size 2 or 3. The second column indicates the number of trivially Wilf-equivalent patterns.} \label{double avoidance} \end{table} In all of the following theorems, we take constructive approaches to prove recurrence relations. Given an element $\boldsymbol{\sigma}$ in $S_n^2(\pi_1, \pi_2)$, we attempt to construct elements in $S_{n+1}^2(\pi_1, \pi_2)$ via inserting a maximal element $n+1$ into the permutations in $\boldsymbol{\sigma}$. Note that if a permutation $\sigma \in S_n$ contains a pattern $\pi$, then adding a maximal element $n+1$ anywhere into $\sigma$ still contains $\pi$. Similarly, if a permutation $\sigma \in S_n$ avoids a pattern $\pi$, then removing the maximal element $n$ from $\sigma$ will still avoid $\pi$. \begin{theorem}\label{132,231} Let $a_n = |S_n^2(132,231)|$. Then $a_n$ satisfies the recurrence relation $a_{n+1} = 2a_n + 2^{n}$ with initial term $a_1 =1$, which corresponds with OEIS sequence \href{http://oeis.org/A001787}{A001787}. \end{theorem} \begin{proof} Given any $\boldsymbol{\sigma} = (\sigma, \sigma') \in S^2_n(132,231)$, we construct an element of $S^2_{n+1}(132,231)$ by inserting a maximal element $n+1$ in both $\sigma$ and $\sigma'$. To avoid both $132$ and $231$, the maximal element $n+1$ must be inserted into either the beginning or end of $\sigma$ and $\sigma'$; otherwise if there are elements on both sides of $n+1$, then there must be either an occurrence of $132$ or $231$. Appending a maximal element $n+1$ onto the left of both $\sigma$ and $\sigma'$ or onto the right of both $\sigma$ and $\sigma'$ also avoids $132$ and $231$. In other words, $(\sigma (n+1), \sigma'(n+1))$ and $((n+1)\sigma, (n+1)\sigma')$ both still avoid $132$ and $231$. This contributes $2a_{n}$ different $3$-permutations in $S_{n+1}^2(132,231)$. We now make the following claims: \begin{claim} The $3$-permutation $(\sigma(n+1), (n+1)\sigma')$ avoids $132$ and $231$ if and only if $\sigma$ is $\mathrm{Id}_n$ and $\sigma' \in S_n^1(132,231)$. \end{claim} \begin{proof} For the forwards direction, suppose that $(\sigma(n+1), (n+1)\sigma')$ avoids $132$ and $231$. Now writing the projection $((n+1)\sigma') \circ (\sigma(n+1))^{-1} = (\sigma_L (n+1) \sigma_R)$ for some subpermutations $\sigma_L$ and $\sigma_R$, note that $\sigma_R$ is nonempty, and hence using the reasoning mentioned above, $\sigma_L$ is empty. Otherwise, $(\sigma_L (n+1) \sigma_R)$ contains an occurrence of either $132$ or $231$. And thus $\sigma$ begins with the minimal element $1$. But because $\sigma$ is forced to avoid the $132$ pattern, it is forced to be consecutive and hence is the identity permutation. For the backwards direction, both $(\sigma(n+1))$ and $((n+1)\sigma')$ still avoid $132$ and $231$. Further, the projection $((n+1)\sigma') \circ (\sigma(n+1))^{-1}$ evaluates to $(n+1)\sigma'$, which also still avoids $132$ and $231$. \end{proof} \begin{claim} The $3$-permutation $((n+1)\sigma, \sigma'(n+1))$ avoids $132$ and $231$ if and only if $\sigma$ is $\mathrm{rev}(\mathrm{Id}_n)$ and $\sigma' \in S_n^1(132,231)$. \end{claim} \begin{proof} For the forwards direction, we write the projection $(\sigma'(n+1))\circ((n+1)\sigma)^{-1}$ in the form of $\sigma_L (n+1) \sigma_R$. As above, $\sigma_R$ is nonempty and hence, $\sigma_L$ must be empty to avoid the patterns $132$ and $231$. So we conclude that $\sigma$ must end with a minimal element $1$. And because $\sigma$ must avoid the $231$ permutation, it is forced to be consecutively decreasing and hence is $\mathrm{rev}(\mathrm{Id}_n)$. For the backwards direction, $((n+1)\sigma)$ and $(\sigma'(n+1))$ both still avoid $132$ and $231$. The projection $(\sigma'(n+1))\circ ((n+1)\sigma)^{-1}$ evaluates to $(n+1)\mathrm{rev}(\sigma')$. Because $132$ and $231$ are reverses of each other, $\mathrm{rev}(\sigma')$ still avoids $132$ and $231$, and thus $(n+1)\mathrm{rev}(\sigma')$ avoids these patterns as well. \end{proof} Thus we have shown that given any $3$-permutation $\boldsymbol{\sigma} = (\sigma, \sigma') \in S^2_n(132,231)$, we can construct two elements in $S^2_{n+1}(132,231)$; furthermore, we can construct two additional elements in $S^2_{n+1}(132,231)$ if and only if $\sigma' \in S^1_n(132,231)$ and $\sigma$ is $\Id_n$ or $\rev(\Id_n)$. Simion and Schmidt \cite{simion1985restricted} have shown that $|S^1_n(132,231)| = 2^{n-1}$. In the cases where $\sigma$ is $\Id_n$ or $\rev(\Id_n)$, it follows that $\boldsymbol{\sigma}$ avoids $132$ and $231$ if and only if $\sigma'$ avoids these patterns, and hence it follows that \begin{align*} a_{n+1} = 2a_n + 2^n. & \qedhere \end{align*} \end{proof} \begin{theorem}\label{132,321} Let $a_n = |S^2_n(132,321)|$. Then $a_n$ follows the recurrence relation $a_{n+1} = a_n +n(n+2)$ with initial term $a_1 = 1$, which corresponds with the OEIS sequence \href{http://oeis.org/A047732}{A047732}. \end{theorem} \begin{proof} Let us write $\boldsymbol{\sigma} = (\sigma, \sigma') \in S^2_n(132,321)$ in the form $(\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$. We construct an element of $S^2_{n+1}(132,321)$ by inserting a maximal element $n+1$ in both $\sigma$ and $\sigma'$. Inserting $n+1$ onto the end of $\sigma$ and $\sigma'$ always constructs a $132$-avoiding and $321$-avoiding $3$-permutation, and thus contributes $a_n$ different $3$-permutations to $S^2_{n+1}(132,321)$. We also have the following three cases: \begin{enumerate} \item $\sigma_R$ and $\sigma_R'$ are both nonempty and $\sigma_L$, $\sigma_L'$, $\sigma_R$, and $\sigma_R'$ are all consecutively increasing. Moreover, every element of $\sigma_L$ and $\sigma_L'$ is greater than every element of $\sigma_R$ and $\sigma_R'$, respectively. \item Exactly one of $\sigma_R$, $\sigma_R'$ is empty and the other is of the form $\sigma_L n \sigma_R$, where $\sigma_L$ and $\sigma_R$ are consecutively increasing and every element of $\sigma_L$ is greater than every element of $\sigma_R$. \item Both $\sigma_R$, $\sigma_R'$ are empty. \end{enumerate} First we show that when $\boldsymbol{\sigma}$ is none of these cases, inserting a maximal element $n+1$ into $\boldsymbol{\sigma}$ cannot avoid these patterns. So let $\boldsymbol{\sigma} = (\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$, where without loss of generality, $\sigma_L$ is nonincreasing. Every element of $\sigma_L$ and $\sigma_L'$ still must be greater than every element of $\sigma_R$ and $\sigma_R'$, respectively; otherwise, they would contain an occurrence of $132$. So because $\sigma_L$ is not consecutively increasing, there is an occurrence of $\ell (\ell-c)$ in $\sigma_L$. If $\sigma_R$ contains elements in the interval $(\ell-c, n)$, then $\boldsymbol{\sigma}$ would contain an occurrence of $132$. Similarly, if $\sigma_R$ contains elements in the interval $(1, \ell-c)$, then $\boldsymbol{\sigma}$ would contain an occurrence of $321$. So $\sigma_R$ is empty. Inserting $n+1$ to the left of $\ell$ gives an occurrence of $321$. And inserting $n+1$ to the right of $\ell$ gives an occurrence of $132$. So nothing outside these cases avoids $132$ and $321$. Now we present each case: \begin{enumerate} \item We claim that only $(\sigma_L n (n+1) \sigma_R, \sigma_L' n (n+1) \sigma_R')$ avoids $132$ and $321$. Because $\sigma_L$ and $\sigma_R$ are consecutive and $\sigma_R$ must start with $1$, then in the projection $(\sigma_L' n (n+1) \sigma_R') \circ (\sigma_L n (n+1) \sigma_R)^{-1}$, either $n+1$ is right-adjacent to $n$, or the composition begins with $n+1$. In the former case, this projection avoids $132$ and $321$, and hence $\boldsymbol{\sigma}$ avoids these patterns too. In the latter case, the projection is of the form $(n+1) \sigma_R' \sigma_L' n$. But $\sigma_R' \sigma_L' n$ is strictly increasing, so the projection also avoids $132$ and $321$. Therefore the $3$-permutation $(\sigma_L n (n+1) \sigma_R, \sigma_L' n (n+1) \sigma_R')$ avoids these patterns too. Now we show that inserting $n+1$ into $\boldsymbol{\sigma}$ anywhere else cannot result in an element in $S_{n+1}^2(132,321)$. In particular, we show that we are forced to insert $n+1$ at the end of $\sigma$ and $\sigma'$ or directly after $n$ in these two permutations. Otherwise, because $\sigma_L$, $\sigma_L'$, $\sigma_R$, and $\sigma_R'$ are all consecutively increasing, $\sigma$ or $\sigma'$ would contain $132$. Now it is sufficient to show $(\sigma_L n \sigma_R (n+1), \sigma_L' n (n+1) \sigma_R')$ and $(\sigma_L n (n+1) \sigma_R, \sigma_L' n \sigma_R' (n+1))$ do not avoid $132$ and $321$. To see the former, we take the projection $\rho$ and note that it must take one of the following three forms: \begin{enumerate} \item $\rev(\Id_{n+1})$. \item $(\pi_1 c \pi_2 m \pi_3 (c+1))$, where $m >c$ for some $m$ and $c$. \item $(\pi (n+1)c)$, where $c$ is the maximum element in $\sigma_R'$ and $\pi$ is a subpermutation. \end{enumerate} In Case (a), $\rho$ contains $321$, in Case (b), $\rho$ contains $132$, and in Case (c), $\rho$ contains $132$. Similar reasoning shows that $(\sigma_L n (n+1) \sigma_R, \sigma_L' n \sigma_R' (n+1))$ does not avoid $132$ and $321$. There are $(n-1)$ ways to choose $\sigma_L n \sigma_R$ and $\sigma_L' n \sigma_R'$, so this case contributes $(n-1)^2$ distinct $3$-permutations to $S^2_{n+1}(132,321)$. \item Without loss of generality, let $\sigma_R$ be empty. Then we claim that only the $3$-permutations $((n+1)\sigma_L n, \sigma_L' n (n+1) \sigma_R')$ and $(\sigma_L n(n+1), \sigma_L' n (n+1) \sigma_R')$ avoid $132$ and $321$. Checking that both of these $3$-permutations avoid $132$ and $321$ uses a similar argument to the previous case. Now we show that inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ cannot avoid the patterns $132$ and $321$. In particular, we must insert $n+1$ into the beginning or end of $\sigma$ and either right-adjacent to $n$ or at the end in $\sigma'$. So it is sufficient to show that $((n+1)\sigma_L n, \sigma_L' n \sigma_R' (n+1))$ does not avoid $132$ and $321$. Now taking the projection gives us a permutation of the form $\pi (n+1) c$, where $c$ is the first element of $\sigma_L'$ and $\pi$ is some subpermutation. Because $\sigma_R'$ is nonempty, $\pi$ contains elements in $\sigma_R'$, and this composition contains an instance of $132$. A similar argument holds for when $\sigma_R'$ is empty and $\sigma_R$ is nonempty. This case contributes $4(n-1)$ total $3$-permutations to $S^2_{n+1}(132,321)$. \item Then $(\sigma_L n (n+1), (n+1) \sigma_L' n)$, $((n+1)\sigma_L n, (n+1) \sigma_L' n)$, and $((n+1)\sigma_L n, \sigma_L' n (n+1))$ all avoid $132$ and $321$. Checking that these avoid $132$ and $321$ follow a similar reasoning to Case 1. Because we are forced to insert the maximal element $n+1$ to the beginning or end of $\sigma$ and $\sigma'$, any other insertion would not avoid $132$ and $321$. This case contributes $3$ new elements in $S^2_{n+1}(132,321)$. \end{enumerate} And hence we conclude that \begin{align*} a_{n+1} &= a_n + (n-1)^2 + 4(n-1) +3 \\ &= a_n + n(n+2). & \qedhere \end{align*} \end{proof} \begin{theorem}\label{231,312} Let $a_n = |S^2_n(231,312)|$. Then $a_n$ follows the recurrence relation $a_{n+1} = 2a_n + 2a_{n-1}$ with initial terms $a_1 = 1$ and $a_2 =4$, which corresponds to the OEIS sequence \href{http://oeis.org/A026150}{A026150}. \end{theorem} \begin{proof} Let $\boldsymbol{\sigma} = (\sigma, \sigma') \in S_n^2(231,312)$ and write $\boldsymbol{\sigma}$ in the form $(\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$. Note that each element of $\sigma_L$ and $\sigma_L'$ are less than each element of $\sigma_R$ and $\sigma_R'$, respectively. Further, $\sigma_R$ and $\sigma_R'$ have to be consecutively decreasing. If $\sigma_R$ is nonempty, $n-1$ must be right-adjacent to $n$ in $\sigma$ to avoid instances of $231$ and $312$. We then have the following cases, where $\sigma_R$ and $\sigma_R'$ are nonempty: \begin{enumerate} \item $\boldsymbol{\sigma}$ is of the form $(\sigma_L n, \sigma_L' n)$. \item $\boldsymbol{\sigma}$ is of the form $(\sigma_L (n-1) n \sigma_R, \sigma_L' (n-1) n \sigma_R')$. \item $\boldsymbol{\sigma}$ is of the form $(\sigma_L (n-1) n \sigma_R, \sigma_L' n (n-1) \sigma_R')$. \item $\boldsymbol{\sigma}$ is of the form $(\sigma_L n (n-1) \sigma_R, \sigma_L' (n-1) n \sigma_R')$. \item $\boldsymbol{\sigma}$ is of the form $(\sigma_L n (n-1) \sigma_R, \sigma_L' n (n-1) \sigma_R')$. \end{enumerate} Now we present each case: \begin{enumerate} \item $(\sigma, \sigma') = (\sigma_L n, \sigma_L' n)$. The maximal element $n+1$ must be inserted adjacent to $n$ in both $\sigma$ and $\sigma'$. If not, then there would be an occurrence of $312$. Thus the following $3$-permutations all avoid $231$ and $312$: $(\sigma_L n (n+1), \sigma_L' n (n+1)), (\sigma_L n (n+1), \sigma_L' (n+1)n), (\sigma_L (n+1) n, \sigma_L' n (n+1)),$ and $(\sigma_L (n+1)n, \sigma_L' (n+1)n)$. So each instance of $\boldsymbol{\sigma}$ in this case contributes $4$ new $3$-permutations that avoid $231$ and $312$. \item $(\sigma, \sigma')= (\sigma_L (n-1) n \sigma_R, \sigma_L' (n-1) n \sigma_R')$. Then both $\sigma_R$ and $\sigma_R'$ are empty (or else there would be an occurrence of $231$). And this is counted in Case 1. \item $(\sigma, \sigma')= (\sigma_L (n-1) n \sigma_R, \sigma_L' n (n-1) \sigma_R')$. Then $\sigma_R$ must be empty and $n (n-1) \sigma_R'$ must be consecutively decreasing. Appending the maximal element $n+1$ onto the end of $\sigma$ and $\sigma'$ also avoids $231$ and $312$. In other words, $(\sigma_L (n-1) n (n+1), \sigma_L' n (n-1) \sigma_R' (n+1))$ avoids $231$ and $312$. In addition, the $3$-permutation $(\sigma_L (n-1) n (n+1), \sigma_L' (n+1) n (n-1) \sigma_R')$ also avoids $231$ and $312$. To see this, we first evaluate the projection of $(\sigma_L (n-1) n, \sigma_L' n (n-1) \sigma_R')$. As shown in Figure 1, we can subdivide $\sigma_L$ into $\pi_L$ and $\pi_R$, where $|\pi_L| = |\sigma_L'|$. For the projection to avoid $312$, $\pi_R (n-1) n$ must be strictly increasing because $n (n-1) \sigma_R'$ is consecutively decreasing. \begin{figure}[htp] \centering \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (2,0) coordinate (x2) --++ (0.6,0) coordinate (x3) --++ (0.8,0) coordinate (x4) --++ (0.8,0) coordinate (x5) --++ (2,0) coordinate (x6) --++ (0.8,0) coordinate (x7) --++ (0.8,0) coordinate (x8); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L2-|x1) -- (L2-|x6) node[above,pos=0.5] {$\sigma_L$}; \node at (L2-|x7) {$n-1$}; \node at (L2-|x8) {$n$}; \draw[|-|] (L3-|x1) -- (L3-|x2) node[above,pos=0.5] {$\sigma_L'$}; \node at (L3-|x3) {$n$}; \node at (L3-|x4) {$n-1$}; \draw[|-|] (L3-|x5) -- (L3-|x8) node[above,pos=0.5] {$\sigma_R'$}; \draw[|-|,dashed] (L1-|x1) -- (L1-|x2) node[above,pos=0.5] {$\pi_L$}; \draw[|-|,dashed] (L1-|x6) -- (L1-|x2) node[above,pos=0.5] {$\pi_R$}; \end{tikzpicture} \caption{The two-line notation used to evaluate $(\sigma_L' n (n-1) \sigma_R') \circ (\sigma_L (n-1)n)^{-1}$. The second line represents the first permutation in the $3$-permutation and the last line represents the second permutation in the $3$-permutation.} \end{figure} This projection is of the form $(\sigma_L' \circ \pi_L^{-1}) n (n-1) \sigma_R'$, which still avoids $231$ and $312$. A similar argument shows that $(\sigma_L (n-1) n (n+1), \sigma_L' (n+1) n (n-1) \sigma_R')$ also avoids $231$ and $312$ by noting that its projection is of the form $(\sigma_L' \circ \pi_L^{-1})(n+1)n (n-1) \sigma_R'$, which still avoids these patterns. Now we show that inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ cannot produce a $3$-permutation that avoids $231$ and $312$. We must insert $n+1$ adjacent to $n$ in $\sigma$. If not, then inserting $n+1$ anywhere to the left of $n-1$ contains an occurrence of $312$. Similarly, we must insert $n+1$ left-adjacent to $n$ or at the end in $\sigma'$. Inserting $n+1$ anywhere to the right of $n-1$ contains an occurrence of $231$. We first show that $(\sigma_L (n-1) (n+1) n, \sigma_L' (n+1) n (n-1) \sigma_R')$ cannot avoid these patterns. As discussed above, we can subdivide $\sigma_L$ into $\pi_L$ and $\pi_R$, where $\pi_L$ is the same size as $\sigma_L'$ and $\pi_R(n-1)$ is consecutively increasing. \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (2,0) coordinate (x2) --++ (0.8,0) coordinate (x3) --++ (0.8,0) coordinate (x4) --++ (0.8,0) coordinate (x5) --++ (0.8,0) coordinate (x6) --++ (0.8,0) coordinate (x7) --++ (0.8,0) coordinate (x8) --++ (1.2,0) coordinate (x9) --++ (0.8,0) coordinate (xx); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L2-|x1) -- (L2-|x7) node[above,pos=0.5] {$\sigma_L$}; \node at (L2-|x8) {$n-1$}; \node at (L2-|x9) {$n+1$}; \node at (L2-|xx) {$n$}; \draw[|-|] (L3-|x1) -- (L3-|x2) node[above,pos=0.5] {$\sigma_L'$}; \node at (L3-|x5) {$n-1$}; \node at (L3-|x4) {$n$}; \node at (L3-|x3) {$n+1$}; \draw[|-|] (L3-|x6) -- (L3-|xx) node[above,pos=0.5] {$\sigma_R'$}; \draw[|-|,dashed] (L1-|x1) -- (L1-|x2) node[above,pos=0.5] {$\pi_L$}; \draw[|-|,dashed] (L1-|x7) -- (L1-|x2) node[above,pos=0.5] {$\pi_R$}; \end{tikzpicture} \end{center} Then the projection is of the form $(\sigma_L' \circ \pi_L^{-1}) \pi (r+2) r (r+1)$, where $r$ is the minimal element of $\sigma_R'$ and $\pi$ is a subpermutation. This contains an occurrence of $312$. A similar calculation shows that $(\sigma_L (n-1) (n+1) n, \sigma_L' n (n-1) \sigma_R' (n+1))$ cannot avoid $231$ and $312$ either, because the projection is of the form $\pi (r+1) (n+1) r$ for a subpermutation $\pi$, which contains an occurrence of $231$. Hence each instance of $\boldsymbol{\sigma}$ in this case contributes $2$ new elements in $S_{n+1}^2(231,312)$. \item $(\sigma, \sigma')= (\sigma_L n (n-1) \sigma_R, \sigma_L' (n-1) n \sigma_R')$. Similar to the previous case, $n(n-1)\sigma_R$ must be consecutively decreasing and $\sigma_R'$ must be empty. As in the previous cases, $(\sigma_L n (n-1) \sigma_R (n+1), \sigma_L' (n-1) n (n+1))$ avoids $231$ and $312$. Moreover, $(\sigma_L (n+1) n (n-1) \sigma_R, \sigma_L' (n-1) n (n+1))$ also avoids these patterns. To see this, we first evaluate the projection of $(\sigma_L n (n-1) \sigma_R, \sigma_L' (n-1) n)$. \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (2,0) coordinate (x2) --++ (0.6,0) coordinate (x3) --++ (1.0,0) coordinate (x4) --++ (0.8,0) coordinate (x5) --++ (2,0) coordinate (x6) --++ (0.8,0) coordinate (x7) --++ (0.8,0) coordinate (xx); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L1-|x1) -- (L1-|x2) node[above,pos=0.5] {$\sigma_L$}; \node at (L1-|x3) {$n$}; \node at (L1-|x4) {$n-1$}; \draw[|-|] (L1-|x5) -- (L1-|xx) node[above,pos=0.5] {$\sigma_R$}; \draw[|-|] (L2-|x1) -- (L2-|x6) node[above,pos=0.5] {$\sigma_L'$}; \node at (L2-|x7) {$n-1$}; \node at (L2-|xx) {$n$}; \draw[|-|,dashed] (L3-|x1) -- (L3-|x2) node[above,pos=0.5] {$\pi_L$}; \draw[|-|,dashed] (L3-|x6) -- (L3-|x2) node[above,pos=0.5] {$\pi_R$}; \node at (L3-|x4) { }; \end{tikzpicture} \end{center} Because the projection of $(\sigma_L n (n-1) \sigma_R, \sigma_L' (n-1) n)$ is of the form $(\pi_L \circ \sigma_L^{-1}) n (n-1) \mathrm{rev}(\pi_R)$, we conclude that $n(n-1)\mathrm{rev}(\pi_R)$ must be consecutively decreasing, since this projection must avoid $231$ and $312$. We evaluate the projection of $(\sigma_L (n+1) n (n-1) \sigma_R, \sigma_L' (n-1) n (n+1))$. \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (2,0) coordinate (x2) --++ (0.8,0) coordinate (x3) --++ (1.0,0) coordinate (x4) --++ (1.0,0) coordinate (x5) --++ (0.8,0) coordinate (x6) --++ (0.6,0) coordinate (x7) --++ (0.8,0) coordinate (x8) --++ (1.0,0) coordinate (x9) --++ (1.0,0) coordinate (xx); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L1-|x1) -- (L1-|x2) node[above,pos=0.5] {$\sigma_L$}; \node at (L1-|x3) {$n+1$}; \node at (L1-|x4) {$n$}; \node at (L1-|x5) {$n-1$}; \draw[|-|] (L1-|x6) -- (L1-|xx) node[above,pos=0.5] {$\sigma_R$}; \draw[|-|] (L2-|x1) -- (L2-|x7) node[above,pos=0.5] {$\sigma_L'$}; \node at (L2-|x8) {$n-1$}; \node at (L2-|x9) {$n$}; \node at (L2-|xx) {$n+1$}; \draw[|-|,dashed] (L3-|x1) -- (L3-|x2) node[above,pos=0.5] {$\pi_L$}; \draw[|-|,dashed] (L3-|x7) -- (L3-|x2) node[above,pos=0.5] {$\pi_R$}; \node at (L3-|x8) { }; \end{tikzpicture} \end{center} Then the projection of the $3$-permutation $(\sigma_L (n+1) n (n-1) \sigma_R, \sigma_L' (n-1) n (n+1))$ is of the form $(\pi_L \circ \sigma_L^{-1}) (n+1) n (n-1) \mathrm{rev}(\pi_R)$, which also avoids $231$ and $312$. Now we show that inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ cannot produce a $3$-permutation that avoids $231$ and $312$. Using similar logic to the previous case, it is sufficient to show that $(\sigma_L (n+1) n (n-1) \sigma_R, \sigma_L' (n-1) (n+1) n)$ and $(\sigma_L n (n-1) \sigma_R (n+1), \sigma_L' (n-1) (n+1) n)$ do not avoid $231$ and $312$. For the former $3$-permutation, we take the projection: \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (2,0) coordinate (x2) --++ (0.8,0) coordinate (x3) --++ (1.0,0) coordinate (x4) --++ (1.0,0) coordinate (x5) --++ (0.8,0) coordinate (x6) --++ (0.6,0) coordinate (x7) --++ (0.8,0) coordinate (x8) --++ (1.2,0) coordinate (x9) --++ (1.0,0) coordinate (xx); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L1-|x1) -- (L1-|x2) node[above,pos=0.5] {$\sigma_L$}; \node at (L1-|x3) {$n+1$}; \node at (L1-|x4) {$n$}; \node at (L1-|x5) {$n-1$}; \draw[|-|] (L1-|x6) -- (L1-|xx) node[above,pos=0.5] {$\sigma_R$}; \draw[|-|] (L2-|x1) -- (L2-|x7) node[above,pos=0.5] {$\sigma_L'$}; \node at (L2-|x8) {$n-1$}; \node at (L2-|x9) {$n+1$}; \node at (L2-|xx) {$n$}; \draw[|-|,dashed] (L3-|x1) -- (L3-|x2) node[above,pos=0.5] {$\pi_L$}; \draw[|-|,dashed] (L3-|x7) -- (L3-|x2) node[above,pos=0.5] {$\pi_R$}; \node at (L3-|x8) { }; \end{tikzpicture} \end{center} The projection is of the form $(\pi_L \circ \sigma_L^{-1}) n (n+1) (n-1) \mathrm{rev}(\pi_R)$, which contains an occurrence of $231$. For the latter $3$-permutation, a similar argument shows that this projection contains an occurrence of $312$. Hence each instance of $\boldsymbol{\sigma}$ in this case contributes $2$ new elements in $S_{n+1}^2(231,312)$. \item $(\sigma, \sigma')= (\sigma_L n (n-1) \sigma_R, \sigma_L' n (n-1) \sigma_R')$. Then $n(n-1)\sigma_R$ and $n(n-1)\sigma_R'$ must be consecutively decreasing. We claim $|\sigma_R| = |\sigma_R'|$. For the sake of contradiction, suppose that $|\sigma_R'|>|\sigma_R|$. Then the $3$-permutations are: \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (2,0) coordinate (x2) --++ (0.6,0) coordinate (x3) --++ (1,0) coordinate (x4) --++ (0.8,0) coordinate (x5) --++ (2,0) coordinate (x6) --++ (0.6,0) coordinate (x7) --++ (1,0) coordinate (x8) --++ (0.8,0) coordinate (x9) --++ (2.5,0) coordinate (xx); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L1-|x1) -- (L1-|x6) node[above,pos=0.5] {$\sigma_L$}; \node at (L1-|x7) {$n$}; \node at (L1-|x8) {$n-1$}; \draw[|-|] (L1-|x9) -- (L1-|xx) node[above,pos=0.5] {$\sigma_R$}; \draw[|-|] (L2-|x1) -- (L2-|x2) node[above,pos=0.5] {$\sigma_L'$}; \node at (L2-|x3) {$n$}; \node at (L2-|x4) {$n-1$}; \draw[|-|] (L2-|x5) -- (L2-|xx) node[above,pos=0.5] {$\sigma_R'$}; \end{tikzpicture} \end{center} The projection is of the form $\pi_1 n \pi_2 r \pi_3 (r+c)$, where $r$ is the minimal element of $\sigma_R'$, $c$ is some positive integer, and $\pi_1$, $\pi_2$, $\pi_3$ are subpermutations. And hence the projection contains $312$, a contradiction. A similar argument holds for $|\sigma_R'|<|\sigma_R|$, so the two permutations must be the same size. Moreover, because both $n(n-1)\sigma_R$ and $n(n-1)\sigma_R'$ are consecutively decreasing, then $\sigma_R = \sigma_R'$. We immediately see that $(\sigma_L n (n-1) \sigma_R (n+1), \sigma_L' n (n-1) \sigma_R' (n+1))$ is in $S_{n+1}^2 (231,312)$. Moreover, $(\sigma_L (n+1) n (n-1) \sigma_R, \sigma_L' (n+1) n (n-1) \sigma_R')$ also avoids $231$ and $312$, because the projection is of the form $(\sigma_L' \circ \sigma_L^{-1}) (\sigma_R' \circ \sigma_R^{-1}) (n-1) n (n+1)$. Now we show that inserting the maximal element $n+1$ anywhere else cannot avoid $231$ and $312$. In fact, $n+1$ can only be inserted either at the end of $\sigma$ and $\sigma'$ or left-adjacent to $n$. If $n+1$ is inserted anywhere in $\sigma_L$ or $\sigma_L'$, then there would be an occurrence of $312$. If $n+1$ is inserted anywhere to the right of $n$ and not at the end of the permutation, then there would be an occurrence of $231$. So we show that the $3$-permutations $(\sigma_L (n+1) n (n-1) \sigma_R, \sigma_L' n (n-1) \sigma_R' (n+1))$ and $(\sigma_L n (n-1) \sigma_R (n+1), \sigma_L' (n+1) n (n-1) \sigma_R')$ cannot avoid $231$ and $312$. For the first $3$-permutation, the projection looks like \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (2,0) coordinate (x2) --++ (0.8,0) coordinate (x3) --++ (0.8,0) coordinate (x4) --++ (0.8,0) coordinate (x5) --++ (0.8,0) coordinate (x6) --++ (0.8,0) coordinate (x7) --++ (0.8,0) coordinate (xx); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L1-|x1) -- (L1-|x2) node[above,pos=0.5] {$\sigma_L$}; \node at (L1-|x5) {$n-1$}; \node at (L1-|x4) {$n$}; \node at (L1-|x3) {$n+1$}; \draw[|-|] (L1-|x6) -- (L1-|xx) node[above,pos=0.5] {$\sigma_R$}; \draw[|-|] (L2-|x1) -- (L2-|x2) node[above,pos=0.5] {$\sigma_L'$}; \node at (L2-|x4) {$n-1$}; \node at (L2-|x3) {$n$}; \node at (L2-|xx) {$n+1$}; \draw[|-|] (L2-|x5) -- (L2-|x7) node[above,pos=0.5] {$\sigma_R'$}; \end{tikzpicture} \end{center} Evaluating the projection gives the form $(\sigma_L' \circ \sigma_L^{-1})(n+1) \pi (n-1)(n)$ for a subpermutation $\pi$, which contains $312$. A similar argument shows that for the second $3$-permutation, the projection is of the form $(\sigma_L' \circ \sigma_L^{-1}) \pi (n-1)n(n+1)r$, where $r$ is the minimal element of $\sigma_R'$ and $\pi$ is a subpermutation. This contains $231$. So each instance of $\boldsymbol{\sigma}$ in this case contributes $2$ new elements in $S_{n+1}^2(231,312)$. \end{enumerate} Now we show that $3$-permutations avoiding $231$ and $312$ must be one of the forms above. We have only one form to consider, where exactly one of $\sigma$, $\sigma'$ is empty. Let $\sigma_R$ be empty and $\sigma_R'$ be nonempty. In particular, $(\sigma, \sigma')= (\sigma_L n, \sigma_L' n \sigma_R')$. Now $n-1$ must be adjacent to $n$ in $\sigma'$. If $\sigma' = \sigma_L' (n-1) n \sigma_R'$, then $\sigma_R'$ must be empty to avoid an occurrence of $231$. Then Case 1 covers this. If $\sigma' = \sigma_L' n (n-1) \sigma_R'$, then we show that $n-1$ is adjacent to $n$ in $\sigma$. So suppose, for the sake of contradiction, that this is not the case. First, $n(n-1)\sigma_R'$ must be consecutively decreasing. Taking the projection $\sigma' \circ \sigma^{-1}$, we conclude that it is of the form $\pi_L n \pi_R k r$, where $k \neq r+1$ and $r$ is the minimal element in $\sigma_R'$. Now we consider where the element $r+1$ is in the permutation. If $r+1$ is in $\pi_L$, then there is an occurrence of $231$. If $r+1$ is in $\pi_R$, then if $k>r+1$, then there is an occurrence of $231$, and if $k<r+1$, then there is an occurrence of $312$. Hence $n-1$ must be adjacent to $n$ in $\sigma$, and a similar argument from Case 3 covers this case. A similar argument also holds for when the case when $\sigma_R$ is nonempty and $\sigma_R'$ is empty. So we see that for every $3$-permutation $\boldsymbol{\sigma} = (\sigma, \sigma')$ in $S_n^2(231,312)$, inserting a maximal element $n+1$ onto the end of both $\sigma$ and $\sigma'$ always yields a $3$-permutation in $S_{n+1}^2(231,312)$; moreover, inserting a maximal element such that the parities of the two largest elements in $\sigma$ and $\sigma'$ are preserved also always yields another $3$-permutation. This contributes $2a_n$ different $3$-permutations to $S_{n+1}^2(231,312)$. In the case that $\boldsymbol{\sigma}$ is in the form in Case 1 (where $\sigma$, $\sigma'$ each end with the maximal element $n$), each $\boldsymbol{\sigma}$ can construct two elements in $S_{n+1}^2(231,312)$ in addition to the elements generated above, so this case constructs an additional $2a_{n-1}$ elements in $S_{n+1}^2(231,312)$. Hence \begin{align*} a_{n+1} = 2a_n + 2a_{n-1}. & \qedhere \end{align*} \end{proof} \begin{theorem}\label{231,321} Let $a_n = |S^2_n(231,321)|$. Then $a_n$ follows the formula $a_{n+1} = 4 \cdot 3^{n-1}$ (where $a_1 = 1$), which corresponds to the OEIS sequence \href{http://oeis.org/A003946}{A003946}. \end{theorem} \begin{proof} Let $\boldsymbol{\sigma} = (\sigma, \sigma') \in S_n^2(231,321)$ and let $\boldsymbol{\sigma}$ be of the form $(\sigma_L 1 \sigma_R, \sigma_L' 1 \sigma_R')$. Note that $\sigma_L$ and $\sigma_L'$ either contain one element or are empty. We insert a minimal element $0$ to the permutation and standardize the new permutation. The element $0$ must be inserted adjacent to $1$ or in the front of both $\sigma$ and $\sigma'$. We then have the following cases: \begin{enumerate} \item $(\sigma, \sigma') = (1 \sigma_R, 1 \sigma_R')$. We can see that $(0 \ 1 \ \sigma_R, 0 \ 1 \ \sigma_R')$, $(0 \ 1 \ \sigma_R, 1 \ 0 \ \sigma_R')$, $(1 \ 0 \ \sigma_R, 0 \ 1 \ \sigma_R')$, and $(1 \ 0 \ \sigma_R, 1 \ 0 \ \sigma_R')$ all avoid $231$ and $321$. So a $3$-permutation $\boldsymbol{\sigma}$ in this case constructs $4$ distinct $3$-permutations in $S_{n+1}^2 (231,321)$. \item $(\sigma, \sigma') = (1 \sigma_R, \ell 1 \sigma_R')$ for some integer $\ell$. Similar to the previous case, we see that $(0 \ 1 \ \sigma_R, 0 \ell 1 \sigma_R')$, $(0 \ 1 \ \sigma_R, \ell 0 \ 1 \ \sigma_R')$, $(1 \ 0 \ \sigma_R, 0 \ell 1 \sigma_R')$, and $(1 \ 0 \ \sigma_R, \ell 0 \ 1 \ \sigma_R')$ all avoid $231$ and $321$. A $3$-permutation $\boldsymbol{\sigma}$ in this case also constructs $4$ distinct $3$-permutations in $S_{n+1}^2 (231,321)$. \item $(\sigma, \sigma') = (\ell 1 \sigma_R, 1 \sigma_R')$ for some integer $\ell$. Appending a minimal element $0$ onto the front of $\sigma$ and $\sigma'$ still avoids $231$ and $321$. In particular, $(0 \ell 1 \sigma_R, 0 \ 1 \ \sigma_R')$, as well as $(\ell 0 \ 1 \ \sigma_R, 1 \ 0 \ \sigma_R')$, avoids these patterns. Now we show that inserting $0$ anywhere else cannot avoid these patterns. In particular, note that $(0 \ell 1 \sigma_R, 1 \ 0 \ \sigma_R')$ and $(\ell 0 \ 1 \ \sigma_R, 0 \ 1 \ \sigma_R')$ cannot avoid $231$ and $321$. For both $3$-permutations, the projection is of the form $1 \pi_L 0 \pi_R$ for subpermutations $\pi_L$ and $\pi_R$ (where $\pi_L$ is nonempty). This contains an instance of $231$. Hence $3$-permutations $\boldsymbol{\sigma}$ in this case constructs $2$ different elements in $S_{n+1}^2 (231,321)$. \item $(\sigma, \sigma') = (\ell_L 1 \sigma_R, \ell_R' 1 \sigma_R')$ for integers $\ell_L$, $\ell_R'$. As in the previous cases, note that $(0 \ell_L 1 \sigma_R, 0 \ell_R' 1 \sigma_R')$, as well as $(\ell_L 0 \ 1 \ \sigma_R, \ell_R' 0 \ 1 \ \sigma_R')$, avoids $231$ and $321$. Now we show that inserting $0$ anywhere else in $\boldsymbol{\sigma}$ cannot avoid $231$ and $321$. In particular, we show that $(0 \ell_L 1 \sigma_R, \ell_R' 0 \ 1 \ \sigma_R')$ and $(\ell_L 0 \ 1 \ \sigma_R, 0 \ell_R' 1 \sigma_R')$ cannot avoid $231$ and $321$. For both $3$-permutations, the projection is $\ell_R' 1 \pi_R$ for some subpermutation $\pi_R$, which contains an instance of $321$ since $\pi_R$ must contain the element $0$. And hence $3$-permutations $\boldsymbol{\sigma}$ in this case constructs $2$ distinct elements in $S_{n+1}^2 (231,321)$. \end{enumerate} We claim that in $S_n^2(231,321)$, exactly half of the elements $\boldsymbol{\sigma} = (\sigma, \sigma')$ satisfy $\sigma(1) = 1$ after standardization. The base case can be seen in $S_2^2(231,321)$. Then for our inductive step let us assume that this is the case for $S_n^2(231,321)$. We wish to show that this is true for $S_{n+1}^2(231,321)$. In each case above, exactly half of the $3$-permutations constructed have the property $\sigma(1)= 1$ and the other half satisfy $\sigma(1) \neq 1$ after standardization, so via induction, exactly half of the elements in $S_n^2(231,321)$ satisfy $\sigma(1) = 2$. Then when $\sigma(1) = 1$, we are in Case 1 or Case 2, which contribute $4$ elements in $S_{n+1}^2(231,321)$. When $\sigma(1) \neq 1$, we are in Case 3 or Case 4, which contribute $2$ elements in $S_{n+1}^2(231,321)$. Thus we conclude that $$a_{n+1} = \frac{a_n}{2}\cdot 4 + \frac{a_n}{2} \cdot 2 = 3a_n.$$ We can see that $a_2 = 4$, so we conclude that \begin{align*} a_{n+1} = 4 \cdot 3^{n-1}. & \qedhere \end{align*} \end{proof} This allows us to prove all the conjectures Bonichon and Morel \cite{bonichon2022baxter} have made in regard to $3$-permutations avoiding two patterns of size $3$. However, there is one class of $3$-permutations that have yet to be classified, which we now enumerate. We begin with a well-known lemma. \begin{lemma}\label{inverselemma} Let $\sigma$ be a permutation and $\pi$ be an involution. Then $\sigma$ avoids $\pi$ if and only if $\sigma^{-1}$ avoids $\pi$. \end{lemma} Because $132$ and $213$ are both involutions, $\sigma$ avoids $132$ if and only if $\sigma^{-1}$ avoids $132$. The same reasoning holds for the pattern $213$. \begin{theorem}\label{132,213} Let $a_n = |S_n^2(132,213)|$. Then $a_n$ follows the recurrence relation $$a_{n+1} = a_n + 3 \cdot 2^{n-1} +2(n-1)$$ with the initial term $a_1 = 1$. \end{theorem} \begin{proof} Let $\boldsymbol{\sigma} = (\sigma, \sigma') \in S_n^2(132,213)$ and let $\boldsymbol{\sigma}$ be of the form $(\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$. Note that $\sigma_L n$ and $\sigma_L' n$ are increasing; otherwise we would contain an occurrence of $213$. Moreover, they must be consecutively increasing; if not we would have an occurrence of $132$. Adding a maximal element $n+1$ right-adjacent to $n$ in both $\sigma$ and $\sigma'$ always produces a $3$-permutation in $S_{n+1}^2(132,213)$. To see this, suppose that $|\sigma_L| > |\sigma_L'|$. Then the projection $\sigma' \circ \sigma^{-1}$ would look like \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (2,0) coordinate (x2) --++ (0.6,0) coordinate (x3) --++ (0.6,0) coordinate (x4) --++ (0.8,0) coordinate (x5) --++ (2,0) coordinate (x6) --++ (0.6,0) coordinate (x7) --++ (0.6,0) coordinate (x8) --++ (0.8,0) coordinate (x9) --++ (2.5,0) coordinate (xx); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L1-|x1) -- (L1-|x6) node[above,pos=0.5] {$\sigma_L$}; \node at (L1-|x7) {$n$}; \draw[|-|] (L1-|x8) -- (L1-|xx) node[above,pos=0.5] {$\sigma_R$}; \draw[|-|] (L2-|x1) -- (L2-|x2) node[above,pos=0.5] {$\sigma_L'$}; \node at (L2-|x3) {$n$}; \draw[|-|] (L2-|x4) -- (L2-|xx) node[above,pos=0.5] {$\sigma_R'$}; \draw[|-|,dashed] (L3-|x1) -- (L3-|x4) node[above,pos=0.5] {$\pi_L$}; \draw[|-|,dashed] (L3-|xx) -- (L3-|x8) node[above,pos=0.5] {$\pi_R$}; \draw[|-|,dashed] (L3-|x4) -- (L3-|x8) node[above,pos=0.5] {$\pi_M$}; \end{tikzpicture} \end{center} This has the form $(\pi_R \circ \sigma_R^{-1}) \pi_L \pi_M$, where $\pi_L$ is consecutively increasing and ends with $n$. This must avoid $132$ and $213$. Now consider $(\sigma_1, \sigma_2) = (\sigma_L n (n+1) \sigma_R, \sigma_L' n (n+1) \sigma_R')$. The projection would look like \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (2,0) coordinate (x2) --++ (0.6,0) coordinate (x3) --++ (1,0) coordinate (x4) --++ (0.8,0) coordinate (x5) --++ (2,0) coordinate (x6) --++ (0.6,0) coordinate (x7) --++ (1,0) coordinate (x8) --++ (0.8,0) coordinate (x9) --++ (2.5,0) coordinate (xx); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L1-|x1) -- (L1-|x6) node[above,pos=0.5] {$\sigma_L$}; \node at (L1-|x7) {$n$}; \node at (L1-|x8) {$n+1$}; \draw[|-|] (L1-|x9) -- (L1-|xx) node[above,pos=0.5] {$\sigma_R$}; \draw[|-|] (L2-|x1) -- (L2-|x2) node[above,pos=0.5] {$\sigma_L'$}; \node at (L2-|x3) {$n$}; \node at (L2-|x4) {$n+1$}; \draw[|-|] (L2-|x5) -- (L2-|xx) node[above,pos=0.5] {$\sigma_R'$}; \draw[|-|,dashed] (L3-|x1) -- (L3-|x4) node[above,pos=0.5] {$\pi_L$}; \draw[|-|,dashed] (L3-|xx) -- (L3-|x9) node[above,pos=0.5] {$\pi_R$}; \draw[|-|,dashed] (L3-|x4) -- (L3-|x9) node[above,pos=0.5] {$\pi_M$}; \end{tikzpicture} \end{center} This has the form $(\pi_R \circ \sigma_R^{-1}) \pi_L (n+1) \pi_M$, which still avoids $132$ and $213$. The case where $|\sigma_L| = |\sigma_L'|$ follows as well. To see when $|\sigma_L| < |\sigma_L'|$, we utilize Lemma 3.5. We've showed that if $\sigma$, $\sigma'$, and $\sigma' \circ \sigma^{-1}$ avoid $132$ and $213$ and $|\sigma_L| > |\sigma_L'|$, then $\sigma_2 \circ \sigma_1^{-1}$ also avoids $132$ and $213$. So it follows that if $\sigma'$, $\sigma$, and $\sigma \circ \sigma'^{-1}$ avoid $132$ and $213$ and $|\sigma_L'| > |\sigma_L|$, then $\sigma_1 \circ \sigma_2^{-1}$ also avoids $132$ and $213$. But Lemma 3.5 shows that $\sigma_1 \circ \sigma_2^{-1}$ avoiding $132$ and $213$ implies that $\sigma_2 \circ \sigma_1^{-1}$ also avoids $132$ and $213$. Hence appending a maximal element $n+1$ right-adjacent to $n$ in both $\sigma$ and $\sigma'$ contributes $a_{n}$ elements in $S^2_{n+1}(132,213).$ In addition, we have the following cases: \begin{enumerate} \item $\sigma = \sigma'$. Then $((n+1) \sigma, (n+1) \sigma')$ also avoids $132$ and $213$. Now we show that inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ does not avoid $132$ and $213$. Note that $n+1$ must be inserted either at the beginning or right-adjacent to $n$ in $\sigma$ and $\sigma'$. If we insert $n+1$ to the left of $n$ (but not at the beginning), then we contain an instance of $132$. Similarly, if we insert $n+1$ to the right of $n$ (but not adjacent to $n$), then we contain an instance of $213$. Then we show that $((n+1)\sigma, \sigma_L' n (n+1) \sigma_R')$ and $(\sigma_L n (n+1) \sigma_R, (n+1) \sigma')$ cannot avoid $132$ or $213$. To see that the first $3$-permutation cannot avoid $132$ or $213$, its projection is of the form $1\pi(n+1)\ell$, where $\ell$ is the first element in $\sigma_L'$ and $\pi$ is a subpermutation. This contains an occurrence of $132$. A similar argument shows that the projection of the latter $3$-permutation also contains $213$. The exception is when $\sigma = \sigma' = \Id_n$, in which case $((n+1)\Id_n, \Id_n (n+1))$ and $(\Id_n (n+1), (n+1) \Id_n)$ both avoid $132$ and $213$. Since Simion and Schmidt \cite{simion1985restricted} showed there are $2^{n-1}$ possible permutations that avoid $132$ and $213$ with size $n$, this contributes an additional $2^{n-1} + 2$ elements in $S_{n+1}^2(132,213)$. \item $\sigma = \Id_n$ and $\sigma' \neq \Id_n$. We note that $(\sigma (n+1), (n+1) \sigma')$ avoids $132$ and $213$. In the special case where $\sigma_R'$ is consecutively increasing, then $((n+1) \sigma, \sigma_L' n (n+1) \sigma_R')$ is also an element in $S_{n+1}^2(132,213)$. Now we show that inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ cannot avoid $132$ and $213$. We first show that $(\sigma (n+1), \sigma' (n+1))$ and $((n+1)\sigma, (n+1) \sigma')$ cannot avoid $132$ and $213$. Taking the projection of these $3$-permutations evaluates to $\sigma' (n+1)$, which contains an occurrence of $213$ because $\sigma'$ is not the identity and hence must contain an occurrence of $21$. Now we show that $((n+1) \sigma, \sigma' (n+1))$ cannot avoid $132$ and $213$. Let us write $\sigma' = \ell \pi_2$. Note that $\ell>1$, because if $\sigma'$ started with $1$, then it would become the identity. Taking the projection gives us the form $\pi_2 (n+1) \ell$, which contains an occurrence of $132$. Now let $\sigma_R'$ be nonincreasing. We wish to show that $((n+1) \sigma, \sigma_L' n (n+1) \sigma_R')$ cannot avoid $132$ and $213$. Because $\sigma_R'$ contains an instance of $21$ and every element in $\sigma_R'$ is smaller than every element in $\sigma_L'$, taking the projection gives an occurrence of $213$. Simion and Schmidt \cite{simion1985restricted} has shown that there are $2^{n-1}$ different $\sigma'$ that avoid $132$ and $213$, so $(\sigma (n+1), (n+1) \sigma')$ contributes $2^{n-1}-1$ different elements to $S_{n+1}^2(132,213)$. Moreover, the special case $((n+1) \sigma, \sigma_L' n (n+1) \sigma_R')$ contributes $n-1$ elements to $S_{n+1}^2(132,213)$. \item $\sigma \neq \Id_n$ and $\sigma' = \Id_n$. This case also contributes $2^{n-1}+n-2$ elements in $S_{n+1}^2(132,213)$. This is a consequence of Lemma 3.5 and the reasoning discussed above. \end{enumerate} Now we show that nothing else can contribute to $S_{n+1}^2(132,213)$. Let $\boldsymbol{\sigma} = (\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$ and assume that $\sigma_R$ and $\sigma_R'$ are nonempty and that $\sigma \neq \sigma'$. This implies that $\sigma$ and $\sigma'$ cannot be the identity permutation. Inserting $n+1$ at the beginning of $\sigma$ and $\sigma'$ gives the projection $(\sigma' \circ \sigma^{-1}) (n+1)$. And because $\sigma \neq \sigma'$, then $\sigma' \circ \sigma^{-1}$ cannot be the identity and hence contains an occurrence of $21$. So the projection contains an occurrence of $213$. We show that $((n+1) \sigma, \sigma_L' n (n+1) \sigma_R')$ cannot avoid $132$ and $213$ either. To see this, let $|\sigma_L| \geq |\sigma_L'|$. Then we evaluate the projection. \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (1.1,0) coordinate (x2) --++ (2,0) coordinate (x3) --++ (0.6,0) coordinate (x4) --++ (0.8,0) coordinate (x5) --++ (0.8,0) coordinate (x6) --++ (0.6,0) coordinate (x7) --++ (0.6,0) coordinate (x8) --++ (0.6,0) coordinate (x9) --++ (2.5,0) coordinate (xx); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L1-|x2) -- (L1-|x7) node[above,pos=0.5] {$\sigma_L$}; \node at (L1-|x8) {$n$}; \node at (L1-|x1) { \ \ \ \ \ $n+1$}; \draw[|-|] (L1-|x9) -- (L1-|xx) node[above,pos=0.5] {$\sigma_R$}; \draw[|-|] (L2-|x1) -- (L2-|x3) node[above,pos=0.5] {$\sigma_L'$}; \node at (L2-|x4) {$n$}; \node at (L2-|x5) {$n+1$}; \draw[|-|] (L2-|x6) -- (L2-|xx) node[above,pos=0.5] {$\sigma_R'$}; \end{tikzpicture} \end{center} This is of the form $r \pi_L (n+1) \pi_R \ell$, where $\ell$ is an element in $\sigma_L'$ (or $n$ if $\sigma_L'$ is empty), $r$ is an element in $\sigma_R'$, and $\pi_L$ and $\pi_R$ are subpermutations. Because elements in $\sigma_L'$ are greater than elements in $\sigma_R'$, then the projection contains an occurrence of $132$. Lemma 3.5 concludes that when $|\sigma_L| \leq |\sigma_L'|$, the projection cannot avoid $132$ either. Finally we show that $( \sigma_L n (n+1) \sigma_R, (n+1) \sigma')$ cannot avoid $132$ and $213$. Let $|\sigma_L| \geq |\sigma_L'|$. Then we evaluate the projection. \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (1.1,0) coordinate (x2) --++ (2,0) coordinate (x3) --++ (0.6,0) coordinate (x4) --++ (0.6,0) coordinate (x5) --++ (2,0) coordinate (x6) --++ (0.6,0) coordinate (x7) --++ (1,0) coordinate (x8) --++ (0.8,0) coordinate (x9) --++ (2.5,0) coordinate (xx); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L1-|x1) -- (L1-|x6) node[above,pos=0.5] {$\sigma_L$}; \node at (L1-|x7) {$n$}; \node at (L1-|x8) {$n+1$}; \draw[|-|] (L1-|x9) -- (L1-|xx) node[above,pos=0.5] {$\sigma_R$}; \draw[|-|] (L2-|x2) -- (L2-|x3) node[above,pos=0.5] {$\sigma_L'$}; \node at (L2-|x4) {$n$}; \node at (L2-|x1) {\ \ \ \ \ $n+1$}; \draw[|-|] (L2-|x5) -- (L2-|xx) node[above,pos=0.5] {$\sigma_R'$}; \end{tikzpicture} \end{center} This is of the form $(\sigma_R' \circ \sigma_R^{-1}) (n+1) \pi_L n \pi_R$, where $\pi_L$ and $\pi_R$ are subpermutations. Because $\sigma_R' \circ \sigma_R^{-1}$ must contain the minimal element $1$, the projection contains an occurrence of $132$, and Lemma 3.5 concludes that when $|\sigma_L| \leq |\sigma_L'|$, this result holds as well. And because nothing else can contribute elements in $S_{n+1}^2(132,213)$, we conclude that \begin{align*} a_{n+1} = a_n + 3 \cdot 2^{n-1} + 2(n-1). & \qedhere \end{align*} \end{proof} These theorems allow us to enumerate all $3$-permutations avoiding two patterns of size $3$ that correspond to existing OEIS sequences. Moreover, since the sequence in Theorem \ref{132,213} does not yet correspond to a sequence on the OEIS database \cite{oeis}, it allows the complete classification and enumeration of all $3$-permutations avoiding two patterns of size $3$. \section{Enumeration of Pattern Avoidance Classes of size 3} Having enumerated all $3$-permutations avoiding two patterns, we now turn our attention to enumerating $3$-permutations avoiding three patterns, as Simion and Schmidt \cite{simion1985restricted} have done with classic permutations. In Table \ref{triple avoidance}, we extend Bonichon and Morel's \cite{bonichon2022baxter} conjectures to $3$-permutations avoiding three patterns of size $3$. \begin{table}[h] \centering \begin{tabular}{|c | c | c | c | c|} \hline Patterns & \#TWE & Sequence & OEIS Sequence & Comment \\ [0.5ex] \hline\hline $123,132,213$ & 3 & $1,4,2,0,0, \dots$ & & Terminates after $n=3$ \\ \hline $123,132,231$ & 4 & $1,4,3,0,0,\dots$ & & Terminates after $n=3$ \\ \hline $123,231,312$ & 1 & $1,4,0,0,0, \dots$ & & Terminates after $n=2$ \\ \hline $123,231,321$ & 2 & $1,4,3,0,0, \dots$ & & Terminates after $n=3$ \\ \hline $132,213,312$ & 2 & $1,4,6,8,10, \dots$ & \href{http://oeis.org/A005843}{A005843} & Theorem \ref{132,213,312} \\ \hline $132,213,321$ & 1 & $1,4,9,16,25, \dots$ & \href{http://oeis.org/A000290}{A000290} & Theorem \ref{132,213,321} \\ \hline $132,231,312$ & 2 & $1,4,7,10,13, \dots$ & \href{http://oeis.org/A016777}{A016777} & Theorem \ref{132,231,312} \\ \hline $213, 231, 321$ & 4 & $1,4,6,8,10, \dots$ & \href{http://oeis.org/A005843}{A005843} & Theorem \ref{213,231,321} \\ \hline $231,312,321$ & 1 & $1,4,7,19,40, \dots$ & \href{http://oeis.org/A006130}{A006130} & Theorem \ref{231,312,321} \\ \hline \end{tabular} \caption{Sequences of $3$-permutations avoiding three permutations of size $3$. The second column indicates the number of trivially Wilf-equivalent patterns.} \label{triple avoidance} \end{table} \begin{theorem}\label{132,213,312} Let $a_n = |S_n^2(132,213,312)|$. Then $a_{n+1}$ follows the formula $a_{n+1} = 2(n+1)$ for $n>0$ (with initial term $a_1=1$). \end{theorem} \begin{proof} Let $\boldsymbol{\sigma} = (\sigma, \sigma') \in S_n^2(132,213,312)$. Let $\boldsymbol{\sigma}$ be of the form $(\sigma_L n \sigma_R, \sigma_L' n \sigma_R').$ Note that $\sigma_R$ and $\sigma_R'$ have to be either empty or consecutively decreasing, and similarly, $\sigma_L$ and $\sigma_L'$ have to be either empty or consecutively increasing. Moreover, every element in $\sigma_L$ and $\sigma_L'$ must be larger than every element in $\sigma_R$ and $\sigma_R'$, respectively. If not, there would be an occurrence of $132$. We have the following cases: \begin{enumerate} \item $\sigma_L$, $\sigma_R$ are nonempty. If $\sigma_L'$ and $\sigma_R'$ are nonempty, consider $\boldsymbol{\sigma} = (\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$ in $S_{n}^2 (132,213,312)$. Note that inserting $n+1$ right-adjacent to $n$ in both $\sigma$ and $\sigma'$ will avoid $132$, $231$, and $312$. In particular, $(\sigma_L n (n+1) \sigma_R, \sigma_L' n (n+1) \sigma_R')$ avoids $132$, $213$, and $312$. Now we show that inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ does not yield an element of $S_{n+1}^2 (132,213,312)$. Consider $\sigma_L n \sigma_R$. We cannot insert $n+1$ in the beginning of this permutation, or else there would be an instance of $312$. Further, we cannot insert $n+1$ anywhere to the left of $n$, or else there would be an instance of $132$. There would also be an occurrence of $213$ if $n+1$ is inserted anywhere to the right of $n$ that is not adjacent to $n$. Hence $n+1$ is forced to be right-adjacent to $n$ in $\sigma$. The same conclusion follows for $\sigma'$. If $\sigma_L'$ is empty, then $\sigma' = \rev(\Id_n)$. The projection $\sigma' \circ \sigma^{-1}$ contains an occurrence of $132$, and hence this case is impossible. Similarly, if $\sigma_R'$ is empty, the projection $\sigma' \circ \sigma^{-1}$ contains an occurrence of $312$, and this case is also impossible. Therefore every element in this case contributes $1$ element in $S_{n}^2 (132,213,312)$. \item $\sigma_L$ is empty. If both $\sigma_L'$ and $\sigma_R'$ are nonempty, then $\boldsymbol{\sigma} = (\rev(\Id_n), \sigma_L' n \sigma_R')$. Taking the projection gives an instance of $132$ because every element in $\sigma_L'$ is larger than every element in $\sigma_R'$. And hence this is not a valid element in $S_{n}^2 (132,213,312)$, so this case is impossible. If $\sigma_L'$ is empty, then we conclude that $(\sigma, \sigma') = (\rev(\Id_n), \rev(\Id_n))$. Following similar logic to the previous case, $n+1$ must be inserted adjacent to $n$ in both $\sigma$ and $\sigma'$ to avoid $312$ and $213$. Note that $((n+1)\rev(\Id_n), (n+1)\rev(\Id_n))$ avoids $132$, $213$, and $312$. However, the projections of the $3$-permutations $((n+1)\rev(\Id_n), (n (n+1) \rev(\Id_{n-1}))$, $(n(n+1)\rev(\Id_{n-1}), (n+1)\rev(\Id_n))$, and $(n(n+1)\rev(\Id_{n-1}), n(n+1)\rev(\Id_{n-1}))$ all cannot avoid $132$, $213$, and $312$, so $(\rev(\Id_n), \rev(\Id_n))$ contributes one additional element to $S_{n+1}^2 (132,213,312)$, in addition to inserting $n+1$ right-adjacent to $n$ in both $\sigma$ and $\sigma'$ as discussed in the previous case. If $\sigma_R'$ is empty, then $(\sigma, \sigma') = (\rev(\Id_n), \Id_n)$. The element $n+1$ must be inserted adjacent to $n$ in $\sigma$, while $n+1$ must be inserted at the end of $\sigma'$. We can see that $((n+1)\rev(\Id_n), \Id_n (n+1))$ is an element of $S_{n+1}^2 (132,213,312)$ and furthermore, $(n (n+1) \rev(\Id_{n-1}), \Id_n (n+1))$ is not an element, because the projection $\sigma' \circ \sigma^{-1}$ contains an instance of $312$. Hence each element in this case contributes $1$ element towards $S_{n+1}^2 (132,213,312)$, with the exception of $(\rev(\Id_n), \rev(\Id_n))$, which contributes $2$ elements towards $S_{n+1}^2 (132,213,312)$. \item $\sigma_R$ is empty. If $\sigma_L'$ is nonempty, then $\boldsymbol{\sigma} = (\Id_n, \sigma_L' n \sigma_R')$. Then $n+1$ is forced to be right-adjacent to $n$ for both $\sigma$ and $\sigma'$, which contributes $1$ element to $S_{n+1}^2 (132,213,312)$. If $\sigma_L'$ is empty, then $(\sigma, \sigma') = (\Id_n, \rev(\Id_n))$. Note that $(\Id_n(n+1), (n+1)\rev(\Id_n))$ avoids $132$, $213$, and $312$, so $(\Id_n, \rev(\Id_n))$ contributes one additional element to $S_{n+1}^2 (132,213,312)$ in addition to inserting $n+1$ right-adjacent to $n$ in both $\sigma$ and $\sigma'$. Hence each element in this case contributes $1$ element towards $S_{n+1}^2 (132,213,312)$, with the exception of $(\Id_n, \rev(\Id_n))$, which contributes $2$ elements towards $S_{n+1}^2 (132,213,312)$. \end{enumerate} Inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ cannot provide an element in $S_{n+1}^2 (132,213,312)$, and hence $a_{n+1} = a_n +2$. We have the base case $a_2 = 4$, thus \begin{align*} a_{n+1} = 2(n+1). & \qedhere \end{align*} \end{proof} \begin{theorem}\label{132,213,321} Let $a_n = |S_n^2(132,213,321)|$. Then $a_{n+1}$ follows the formula $a_{n+1} = (n+1)^2$. \end{theorem} \begin{proof} Let $\boldsymbol{\sigma} = (\sigma, \sigma') \in S_n^2(132,213,321)$. Write $(\sigma, \sigma')$ as $(\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$. Using a similar reasoning discussed in Theorem \ref{132,213,312}, note that $\sigma_L$, $\sigma_L'$, $\sigma_R$, and $\sigma_R'$ are consecutively increasing. Moreover, every element in $\sigma_L$ and $\sigma_L'$ is larger than every element in $\sigma_R$ and $\sigma_R'$, respectively. Similar to the reasoning in Theorem \ref{132,213,312}, $(\sigma_L n (n+1) \sigma_R, \sigma_L' n (n+1) \sigma_R')$ is in $S_{n+1}^2(132,213,321)$. This contributes $a_n$ different $3$-permutations to $S_{n+1}^2(132,213,321)$. We also have the following cases: \begin{enumerate} \item $\sigma_R$ is empty and $\sigma_R'$ is nonempty. Note that this implies that $\sigma = \Id_n$ and $\sigma' \neq \Id_n$. Then $((n+1) \Id_n, \sigma_L' n (n+1) \sigma_R')$ avoids $132$, $213$, and $312$. Inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ cannot avoid $132$, $213$, and $312$. In $\sigma$, we must insert $n+1$ at the beginning of the permutation, or else we contain an instance of $132$. Similarly, we must insert $n+1$ right-adjacent to $n$ in $\sigma'$. If $n+1$ is left of $n$, then $\sigma'$ contains an instance of $321$. If $n+1$ is right of $n$ but not adjacent, then $\sigma'$ contains an instance of $213$. And hence $((n+1) \Id_n, \sigma_L' n (n+1) \sigma_R')$ is the only $3$-permutation we can construct in $S_{n+1}^2(132,213,321)$. Hence this case contributes $n-1$ elements to $S_{n+1}^2(132,213,321)$. \item $\sigma_R$ is nonempty and $\sigma_R'$ is empty. This implies that $\sigma' = \Id_n$ and $\sigma \neq \Id_n$. Note that $(\sigma_L n (n+1) \sigma_R, (n+1) \Id_n)$ belongs to $S_{n+1}^2(132,213,321)$. Using a similar argument as in Case 1, inserting $n+1$ in $\boldsymbol{\sigma}$ anywhere else does not avoid $132$, $213$, and $312$. Hence this case contributes $n-1$ different $3$-permutations to $S_{n+1}^2(132,213,321)$. \item Both $\sigma_R$, $\sigma_R'$ are empty. This implies that $\sigma = \sigma' = \Id_n$. We see that $((n+1)\Id_n,\Id_n (n+1))$, $((n+1)\Id_n, (n+1)\Id_n)$, and $(\Id_n (n+1), (n+1) \Id_n)$ all avoid $132$, $213$, and $312$. And using the same reasoning as in Case 1 shows that inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ cannot avoid these patterns, and hence this case contributes $3$ elements to $S_{n+1}^2(132,213,321)$. \end{enumerate} Lastly when $\sigma$ and $\sigma'$ are not the identity permutation, then $\sigma_R$ and $\sigma_R'$ are both nonempty, and the same argument in Case 1 shows that inserting $n+1$ anywhere not right-adjacent to $n$ in $\sigma$ and $\sigma'$ cannot avoid $132$, $213$, and $312$. Hence no other insertions of $n+1$ in $\boldsymbol{\sigma}$ produce an element in $S_{n+1}^2(132,213,321)$, and so $$a_{n+1} = a_n + 2n + 1.$$ The base case is $a_1 = 1$, so we conclude that \begin{align*} a_{n+1} = (n+1)^2. & \qedhere \end{align*} \end{proof} \begin{theorem}\label{132,231,312} Let $a_n = |S_n^2(132,231,312)|$. Then $a_n$ follows the recurrence relation $a_{n+1} = a_n+3$ with initial term $a_1 = 1$. \end{theorem} \begin{proof} Let $\boldsymbol{\sigma} = (\sigma, \sigma') \in S_n^2(132,231,312)$. Write $(\sigma, \sigma')$ as $(\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$. Note that $n \sigma_R$ and $n \sigma_R'$ must be consecutively decreasing to avoid $312$ and $132$. We insert a maximal element $n+1$ into $\sigma$ and $\sigma'$ to count how many elements in $S_{n+1}^2(132,231,312)$ there are. Note that $(\sigma (n+1), \sigma' (n+1))$ avoids $132$, $231$, and $312$. So this contributes $a_n$ different $3$-permutations to $S_{n+1}^2(132,231,312)$. We have the following additional cases: \begin{enumerate} \item $\sigma = \rev(\Id_n)$ and $\sigma' \neq \rev(\Id_n)$. This forces $\sigma'$ to be the identity. Then $((n+1)\rev(\Id_n), \sigma' (n+1))$ avoids $132$, $231$, and $312$. Now inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ cannot avoid these patterns. Namely, if $n+1$ is inserted anywhere not in the beginning or end of $\sigma$, there is an occurrence of $231$. Moreover, inserting $n+1$ into the beginning of $\sigma'$ contains $312$. If $n+1$ is inserted anywhere not in the beginning or end of $\sigma'$, there is an occurrence of either $231$ or $132$. Hence $((n+1)\rev(\Id_n), \sigma' (n+1))$ is the only element we can construct in $S_{n+1}^2(132,231,312)$ in this case. And this case contributes one element towards $S_{n+1}^2(132,231,312)$. \item $\sigma' = \rev(\Id_n)$ and $\sigma \neq \rev(\Id_n)$. Then similar to Case 1, $(\sigma (n+1), (n+1) \rev(\Id_n))$ avoids $132$, $231$, and $312$, and inserting $n+1$ anywhere else into this $3$-permutation cannot construct a $3$-permutation in $S_{n+1}^2(132,231,312)$. Hence this case contributes one element towards $S_{n+1}^2(132,231,312)$. \item $\sigma = \sigma' =\rev(\Id_n)$. Note that $((n+1) \rev(\Id_n), (n+1) \rev(\Id_n))$ works. Now we show that no other insertions of $n+1$ into this $3$-permutation avoids the patterns $132$, $231$, and $312$. The projection of $( \rev(\Id_n) (n+1), (n+1) \rev(\Id_n))$ contains an occurrence of $231$ and the projection of $((n+1) \rev(\Id_n), \rev(\Id_n) (n+1))$ contains an occurrence of $312$, and hence this case contributes one element towards $S_{n+1}^2(132,231,312)$. \end{enumerate} Inserting $n+1$ into $(\sigma, \sigma') = (\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$ anywhere else cannot avoid $132$, $231$, and $312$, where $\sigma, \sigma' \neq \rev(\Id_n)$. This implies that $\sigma_L$ and $\sigma_L'$ are nonempty. Inserting $n+1$ left-adjacent to $n$ contains $132$ and inserting $n+1$ anywhere to the left of this contains $312$. Further, inserting $n+1$ anywhere to the right of $n$ (but not at the end of the permutation) contains $231$. Hence we must insert $n+1$ at the end of the permutation, and no other insertions of $n+1$ in $\boldsymbol{\sigma}$ avoid $132$, $231$, and $312$. Thus $$a_{n+1} = a_n+3.$$ Because our base case is $a_1 = 1$, this is equivalent to $a_{n+1} = 3n+1$. \end{proof} \begin{theorem}\label{213,231,321} Let $a_n = |S_n^2(213, 231, 321)|$. Then $a_{n+1}$ follows the formula $a_{n+1} = 2(n+1)$ for $n>0$ (with initial term $a_1 = 1$). \end{theorem} \begin{proof} Let $\boldsymbol{\sigma} = (\sigma, \sigma') \in S_n^2(213,231,321)$. Writing $(\sigma, \sigma')$ as $(\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$, note that $\sigma_L$, $\sigma_L'$, $\sigma_R$, and $\sigma_R'$ are all consecutively increasing or empty. We insert a maximal element $n+1$ to $\sigma$ and $\sigma'$ in an attempt to construct an element in $S_{n+1}^2(213,231,321)$. If $\sigma_R$ and $\sigma_R'$ are nonempty, we cannot construct an element of $S_{n+1}^2(213,231,321)$ via insertion because inserting $n+1$ to the left of $n$ contains $321$, inserting $n+1$ right-adjacent to $n$ contains $231$, and inserting $n+1$ anywhere else contains $213$. Then it is enough to consider $(\sigma, \sigma') = (\Id_n, \Id_n)$. We have two cases: \begin{enumerate} \item We insert $n+1$ to the end of $\sigma$. Then we can insert $n+1$ anywhere in $\sigma'$ and the resulting $3$-permutation is an element of $S_{n+1}^2(213,231,321)$. So this case contributes $n+1$ different elements to $S_{n+1}^2(213,231,321)$. \item We do not insert $n+1$ to the end of $\sigma$. Note that inserting $n+1$ into the same position in $\sigma$ and $\sigma'$ avoids $213$, $231$, and $321$. Further, $(\Id_{n-1} (n+1)n, \Id_{n} (n+1))$ also avoids these patterns. Inserting $n+1$ anywhere else contains one of these patterns because the resulting projection contains either $321$ or $231$, and hence this case contributes $n+1$ different $3$-permutations to $S_{n+1}^2(213,231,321)$. \end{enumerate} And hence \begin{align*} a_{n+1} = 2(n+1). & \qedhere \end{align*} \end{proof} \begin{theorem}\label{231,312,321} Let $a_n = |S_n^2(231,312,321)|$. Then $a_n$ follows the recurrence relation $$a_{n+1} = a_n + 3a_{n-1}$$ with initial terms $a_1 = 1$ and $a_2 = 4$. \end{theorem} \begin{proof} Let $\boldsymbol{\sigma} = (\sigma, \sigma') \in S_n^2(231,312,321)$. Write $(\sigma, \sigma')$ as $(\sigma_L n \sigma_R, \sigma_L' n \sigma_R')$. Note that $(\sigma (n+1), \sigma' (n+1))$ is an element of $S_{n+1}^2(231,312,321)$. This contributes $a_n$ different $3$-permutations towards $S_{n+1}^2(231,312,321)$. We consider the following additional case: when $\sigma_R$ and $\sigma_R'$ are empty. Then $(\sigma, \sigma') = (\sigma_L n, \sigma_L' n)$, and thus $(\sigma_L (n+1)n, \sigma_L' n(n+1))$, $(\sigma_L (n+1)n, \sigma_L' (n+1)n)$, and $(\sigma_L n(n+1), \sigma_L' (n+1)n)$ are all elements in $S_{n+1}^2(231,312,321)$. Inserting $n+1$ anywhere else cannot avoid these patterns, because inserting $n+1$ anywhere non-adjacent to $n$ contains $312$. And hence this case contributes $3a_{n-1}$ distinct $3$-permutations to $S_{n+1}^2(231,312,321)$. Now when either $\sigma_R$ and $\sigma_R'$ are nonempty, we show that inserting $n+1$ anywhere but the end of the $3$-permutation cannot avoid $231$, $312$, and $321$. Let $\sigma_R$ be nonempty. Then we must insert $n+1$ at the end of $\sigma$; otherwise, inserting $n+1$ to the right of $n$ contains $231$, inserting left-adjacent to $n$ contains $321$, and inserting to the left of $n$ contains $312$. And we evaluate the projection $(\sigma_L n \sigma_R (n+1), \sigma_L' (n+1) n)$: \begin{center} \begin{tikzpicture}[coo/.style={coordinate}] \path (0,0) coordinate (x1) --++ (2,0) coordinate (x2) --++ (0.6,0) coordinate (x3) --++ (0.6,0) coordinate (x4) --++ (0.6,0) coordinate (x5) --++ (0.8,0) coordinate (x6) --++ (0.8,0) coordinate (x7); \foreach \i in {1,2,3} \coordinate[] (L\i) at (0,-\i); \draw[|-|] (L1-|x1) -- (L1-|x2) node[above,pos=0.5] {$\sigma_L$}; \node at (L1-|x3) {$n$}; \node at (L1-|x7) {$n+1$}; \draw[|-|] (L1-|x4) -- (L1-|x6) node[above,pos=0.5] {$\sigma_R$}; \draw[|-|] (L2-|x1) -- (L2-|x5) node[above,pos=0.5] {$\sigma_L'$}; \node at (L2-|x7) {$n$}; \node at (L2-|x6) {$n+1$}; \end{tikzpicture} \end{center} Since $\sigma_R$ is nonempty, this contains an instance of $312$. The case where $\sigma_R'$ is nonempty is similar. Inserting $n+1$ anywhere else in $\boldsymbol{\sigma}$ cannot produce an element in $S_{n+1}^2(231,312,321)$, and hence \begin{align*} a_{n+1} = a_n + 3a_{n-1}. & \qedhere \end{align*} \end{proof} \section{Final Remarks and Open Problems} In this paper, we completely enumerated $3$-permutations avoiding two patterns of size $3$ and three patterns of size $3$. The theorems in this paper prove all the conjectures by Bonichon and Morel \cite{bonichon2022baxter} regarding $3$-permutations avoiding two patterns of size $3$ and extend their conjectures to classify $3$-permutations avoiding all classes of three patterns of size $3$. We conclude with the following open problems. \begin{problem} Enumerate $3$-permutations avoiding one pattern of size $3$ or one pattern of size $4$. \end{problem} Although this paper has shown connections between $3$-permutations avoiding two patterns of size $3$ and their recurrence relations, there are no existing OEIS sequences \cite{oeis} that correspond with the number of $3$-permutations avoiding one pattern of size $3$ or $3$ permutations avoiding one pattern of size $4$. In a similar vein, enumeration of $d$-permutations with dimension greater than $3$ remains an open problem. Bonichon and Morel \cite{bonichon2022baxter} also introduced other tables enumerating $3$-permutations avoiding other combinations of patterns, such as avoiding patterns with dimension $3$ or avoiding exactly one permutation of size 2 and dimension 3 and exactly one permutation of size 3 and dimension 2. We present a few of these as future directions to continue. \begin{conjecture}[Bonichon and Morel \cite{bonichon2022baxter}] The $3$-permutations avoiding the $3$-patterns $(12,12)$ and $(231,312)$ are enumerated by the OEIS sequence \href{http://oeis.org/A295928}{A295928}. \end{conjecture} In addition, Table \ref{future problems} by Bonichon and Morel \cite{bonichon2022baxter} presents $3$-permutations avoiding a permutation of size 2 and dimension 3 as well as a pattern of size 3 and dimension 2. Many of these sequences also correspond to existing sequences on the OEIS database \cite{oeis}, but little research has been done to enumerate such sequences of $3$-permutations. It would be interesting to prove these recurrences that result from $3$-permutations avoiding patterns with different dimensions. \begin{table}[htp] \centering \begin{tabular}{|c | c | c | c | c|} \hline Patterns & \#TWE & Sequence & OEIS Sequence & Comment \\ [0.5ex] \hline\hline $123, (12,12)$ & 1 & $1,3,14,70,288,822,1260, \dots$ & & Not in OEIS \\ \hline $123, (12,21)$ & 3 & $1,3,6,6,0,0,0, \dots$ & & Terminates after $n=4$ \\ \hline $132, (12,12)$ & 2 & $1,3,11,41,153,573,2157, \dots$ & \href{http://oeis.org/A281593}{A281593?} & \\ \hline $132, (12,21)$ & 6 & $1,3,11,43,173,707,2917, \dots$ & \href{http://oeis.org/A026671}{A026671?} & \\ \hline $231, (12,12)$ & 2 & $1,3,9,26,72,192,496, \dots$ & \href{http://oeis.org/A072863}{A072863?} & \\ \hline $231, (12,21)$ & 4 & $1,3,11,44,186,818,3706, \dots$ & & Not in OEIS \\ \hline $231, (21,12)$ & 2 & $1,3,12,55,273,1428,7752, \dots$ & \href{http://oeis.org/A001764}{A001764?} & \\ \hline $321, (12,12)$ & 1 & $1,3,2,0,0,0,0, \dots$ & & Terminates after $n=3$ \\ \hline $321, (12,21)$ & 3 & $1,3,11,47,221,1113,5903, \dots$ & \href{http://oeis.org/A217216}{A217216?} & \\ \hline \end{tabular} \caption{Sequences of $3$-permutations avoiding one pattern of size 3 and dimension 2 and one pattern of size 2 and dimension 3. The ``?" after the OEIS sequences mean that the sequences match on the first few terms and Bonichon and Morel \cite{bonichon2022baxter} conjectured that they are the same. The second column indicates the number of trivially Wilf-equivalent patterns.} \label{future problems} \end{table} We notice that the sequence \href{http://oeis.org/A001787}{A001787} in Theorem \ref{132,231} counts the number of $132$-avoiding permutations of length $n+2$ with exactly one occurrence of a $123$-pattern and the number of Dyck $(n+2)$-paths with exactly one valley at height $1$ and no higher valley \cite{oeis}. In this spirit, we propose the following problem: \begin{problem} Find combinatorial bijections to explain the relationships between the $3$-permutation avoidance classes found in this paper and their recurrence relations. \end{problem} In general, the problem of enumerating $d$-permutations avoiding sets of small patterns is widely open. Since several of these enumeration sequences correspond to sequences on the OEIS database \cite{oeis}, there are certainly interesting combinatorial properties of these $3$-permutation avoidance classes, and there are several bijections to find that explain these sequences. \section*{Acknowledgements} This research was conducted at the 2022 University of Minnesota Duluth REU and is supported by Jane Street Capital, the NSA (grant number H98230-22-1-0015), the NSF (grant number DMS-2052036), and the Harvard College Research Program. The author is indebted to Joe Gallian for his dedication and organizing the University of Minnesota Duluth REU. Lastly, a special thanks to Joe Gallian, Amanda Burcroff, Maya Sankar, and Andrew Kwon for their invaluable feedback and advice on this paper. \newpage
{ "timestamp": "2022-08-26T02:03:14", "yymm": "2208", "arxiv_id": "2208.08506", "language": "en", "url": "https://arxiv.org/abs/2208.08506", "abstract": "Bonichon and Morel first introduced $d$-permutations in their study of multidimensional permutations. Such permutations are represented by their diagrams on $[n]^d$ such that there exists exactly one point per hyperplane $x_i$ that satisfies $x_i= j$ for $i \\in [d]$ and $j \\in [n]$. Bonichon and Morel previously enumerated $3$-permutations avoiding small patterns, and we extend their results by first proving four conjectures, which exhaustively enumerate $3$-permutations avoiding any two fixed patterns of size $3$. We further provide a enumerative result relating $3$-permutation avoidance classes with their respective recurrence relations. In particular, we show a recurrence relation for $3$-permutations avoiding the patterns $132$ and $213$, which contributes a new sequence to the OEIS database. We then extend our results to completely enumerate $3$-permutations avoiding three patterns of size $3$.", "subjects": "Combinatorics (math.CO)", "title": "On $d$-permutations and Pattern Avoidance Classes", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.98657174604767, "lm_q2_score": 0.8128673133042217, "lm_q1q2_score": 0.8019519245916243 }
https://arxiv.org/abs/1310.7984
Polarization of Koszul cycles with applications to powers of edge ideals of whisker graphs
In this paper, we introduced the polarization of Koszul cycles and use it to study the depth function of powers of edge ideals of whisker graphs.
\section*{Introduction} Polarization is a technique to deform an arbitrary monomial ideal $I$ in a polynomial ring $S$ into a squarefree monomial ideal $I^\wp$ in a larger polynomial ring $S^\wp$ such that $S/I$ is a quotient of $S^\wp/ I^\wp$ modulo a regular sequence of linear forms. The polarized ideal $I^\wp$ has a nice property that it has the same graded Betti numbers as $I$. Therefore, many questions regarding monomial ideals can be reduced to the study of squarefree monomial ideals. The fact that $I$ and $I^\wp$ has same graded Betti numbers implies that the corresponding Koszul homology modules of the ideal and its polarization have the same vector-space dimension. Therefore, it is natural to ask whether cycles whose homology classes form a basis of the Koszul homology of $I$ can be naturally lifted to cycles representing a basis for the Koszul homology of $I^\wp$. In Theorem~\ref{main}, it is shown that this is indeed the case. In his book \cite[Proposition 6.3.2]{V}, Villarreal uses polarization to give a simple proof of the fact that the edge ideal of a whisker graph is Cohen-Macaulay. Given a finite simple graph $G$ on the vertex set $V(G)= \{x_1, \ldots, x_n\}$ and the edge set $E(G)$. One defines whisker graph $G^*$ of $G$ to be the graph with vertex set $\{x_1, \ldots, x_n, y_1, \ldots, y_n\}$ and edge set $E(G) \cup \{ \{x_i, y_i\} : \; i=1, \ldots, n\}$. By using the results of Section~\ref{polarization}, one easily sees that the homology classes of the cycles \begin{eqnarray}\label{basis1} x_{i_1} \ldots x_{i_k} e_{j_1} \wedge e_{j_{n-k}}\wedge f_{i_1} \wedge \ldots \wedge f_{i_k} \end{eqnarray} with ${\mathcal S} = \{i_1, \ldots, i_k\}$ a maximal independent set of $G$ and $\{j_1, \ldots, j_{n-k}\}=V(G) \setminus {\mathcal S}$, form a basis of the Koszul homology $H_n(x_1, \ldots, x_n, y_1, \ldots, y_n ; S^*/I(G^*))$. Here $e_1, \ldots, e_n, f_1, \ldots, f_n$ is a $S^*$-basis of free module $K_1(x_1, \ldots, x_n, y_1, \ldots, y_n; S^*/I(G^*))$ with $\partial (e_i) = x_i$ and $\partial (f_j)= y_j$. A basis cycle as described in (\ref{basis1}) is used in Section~\ref{whisker graphs} in the study of the powers of edge ideals of whisker graphs. The homological and algebraic behavior of powers of an ideal has been subject of many research papers in recent years. In particular, the nature of the depth function $f(k)= \depth(S/I^k)$ of a graded ideal $I$ in a polynomial ring $S$ is still quite mysterious. While it is known by a classical result of Brodmann \cite{Br1} that $f(k)$ for $k \gg 0$ is constant, the behavior of $f(k)$ is not so well understood for initial values of $k$. In \cite{HH1}, it is shown that any non-decreasing bounded integer function $f(k)$ is the depth function of a suitable monomial ideal and it is conjectured that $f(k)$ can be any convergent nonnegative integer valued function. In support of this conjecture, it was shown in \cite{BHH} that $f(k)$ may have arbitrarily many local maxima. On the other hand, it seems that the depth function for the edge ideals behave more tamely. In particular, it is expected that the depth function of an edge ideal is a non-increasing function. This is indicated by the fact that edge ideals satisfy the persistence property for the associated prime ideals of their powers, as shown in \cite{CMS}. Interesting lower bounds for the depth function of an edge ideal have been achieved by Morey \cite{M}. On the other hand, even for simple graphs like a line graph or a cycle, the precise depth function is unknown. In this paper we give an upper bound for the depth function for any connected whisker graph. In fact we show in Theorem~\ref{whisker} that for any connected graph $G$ on the vertex set $[n]$, we have $\depth (S^*/I(G^*)^k) \leq n-k+1$ for $k= 1, \ldots , n$. It can be shown by examples that this upper bound is no longer valid if we drop the assumption that $G$ is connected. For connected graph this upper bound is obtained by constructing suitable non-vanishing homology classes for the Koszul homology of the powers of $I(G^*)$. The cycles representing these non-vanishing homology classes are obtained as products of certain 1-cycles and an $(n-1)$-cycle which is defined via an independent set of $G$. For showing that the homology of this product of cycles in the corresponding homology group is non-vanishing, we use a combinational fact proved in Proposition~\ref{gamma} which says that any connected graph admits a friendly independent set in the sense as described in this proposition. By using results from \cite{CMS} and \cite {EH2}, we show in Corollary~\ref{limit} that $\depth(S^*/I(G^*)^k) =1$ for $k\geq n$ if $G$ is bipartite and $\depth(S^*/I(G^*)^k) =0$ for $k\geq n$ if $G$ is non-bipartite. The upper bound for the depth of the powers of a whisker graph given by our Theorem~\ref{whisker} is not always attained. The simplest examples for such case are the whisker graphs of a $3$-cycle or $4$-cycle. On the other hand, Villarreal \cite[Proposition 6.3.7]{V}, has shown that $\depth(S^*/I(G^*)^2) \geq n-1$ if $G$ is tree (or even a forest) on the vertex set $[n]$. In Theorem~\ref{tree}, we extend the result of Villarreal and show that for any forest $G$ one has $\depth(S^*/I(G^*)^k) \geq n-k+1$ for $k=1, \ldots , n$. Together with Theorem~\ref{whisker} we conclude that for any tree $G$ we have $\depth(S^*/I(G^*)^k) = n-k+1$ for $k=1, \ldots, n$. \section{Polarization of Koszul cycles} \label{polarization} Let $K$ be a field and $I\subset S=K[x_1,\ldots,_n]$ a monomial ideal in the polynomial ring $S$. We denote as usual by $G(I)$ the unique minimal set of monomial generators of $I$. If $u=x_1^{a_1}\cdots x_n ^{a_n}$ is a monomial, we call ${\bold a}=(a_1,\ldots,a_n)$ the multi-degree of $u$ and set $\deg_{x_i}u=a_i$ for all $i$. Let $c_i=\max\{\deg_{x_i} u\:\; u\in G(I)\}$ for $i=1,\ldots,n$, and let $S^{\wp}$ be the polynomial ring over $K$ in the variables $x_{ij}$, $i=1,\ldots,n$, $j=1,\ldots.c_i$. The {\em polarization} of $I$ is the squarefree monomial $I^\wp\subset S^\wp$ generated by the monomials $u^\wp$ with $u\in G(I)$ where for $u=x_1^{a_1}\cdots x_n^{a_n}$ one sets \[ u^\wp=\prod_{i=1,\ldots,n}\prod_{j=1,\ldots,a_i} x_{ij}. \] We extend this polarization operation to elements in the Koszul complex. Let $K({\bold x};I)$ be the Koszul complex of the sequence ${\bold x}=x_1,\ldots,x_n$ with values in $I$. Recall that $K_i({\bold x})=\bigwedge^iF$ where $F=\Dirsum_j^nSe_j$ and where $\partial e_j=x_j$ for $j=1,\ldots,n$, and that $K({\bold x};I)=K({\bold x})\tensor I$. Thus an element of $K_i({\bold x};I)$ is of the form \[ \sum_{J}f_Je_J, \] where the sum is taken over all ordered sets $J=\{j_1<j_2<\cdots <j_i\}$ of cardinality $i$, where $f_J\in I$ and where $e_J=e_{j_1}\wedge e_{j_2}\wedge \cdots \wedge e_{j_i}$. Next we consider the Koszul complex $K({\bold x}^\wp; I^\wp)$. Here ${\bold x}^\wp$ is the sequence \[ x_{11},x_{12},\ldots,x_{1c_1},x_{21},\ldots, x_{2c_2},\ldots, x_{n1},\ldots, x_{nc_n}, \] and $K_i({\bold x}^\wp)=\bigwedge^i G$ where $G=\Dirsum_{i=1,\ldots,n}\Dirsum_{j=1,\ldots,c_i}S^{\wp}e_{ij}$. We call an element $u_Je_J$ a {\em monomial} of $K({\bold x};I)$ if $u_J$ is a monomial. We set \[ \deg_{x_i}(u_J e_J)= \deg_{x_i}u_J + \delta_j , \] where \[ \delta_j= \left\{ \begin{array}{ll} 1, & \;\textnormal{if $j\in J$}, \\ 0, & \;\text{otherwise.} \end{array} \right. \] and call \[ \deg(u_J e_J) = (\deg_{x_1}(u_J e_J), \ldots, \deg_{x_n} (u_J e_J) ) \] the multi-degree of $u_Je_J$. For any monomial $u_J e_J$ of multi-degree $\leq {\bold c}$ (componentwise) where ${\bold c}= (c_1, \ldots, c_n)$, we define the polarization of $u_Je_J$ to be the monomial \[ (u_Je_J)^\wp=u_J^\wp e_{j_1a_{j_1}+1}\wedge e_{j_2a_{j_2}+1}\wedge \cdots \wedge e_{j_ia_{j_i}+1}, \] in $K({\bold x}^{\wp};I^\wp)$, where $J=\{j_1<j_2<\cdots <j_i\}$, and $a_i = \deg_{x_i} u_J$. We extend this polarization operator to an arbitrary multi-homogeneous element $f=\sum_J\lambda_Ju_Je_J$, $\lambda_J\in K$, of multi-degree $\leq {\bold c}$ , by setting \[ f^\wp= \sum_J\lambda_J(u_Je_J)^\wp. \] If follows from \cite[Theorem 3.1]{BHbook} that any non-vanishing homology class of $H_i({\bold x};I)$ can be represented by a multi-homogeneous cycle $z=\sum_J\lambda_Ju_Je_J$ in $K_i({\bold x};I)$ with $\deg z \leq {\bold c}$. Thus the polarization of such cycles is defined. \medskip The following example demonstrate the polarization of cycles: let $I=(x_1^2 x_2, x_1 x_2^2)$. Then $z= x_1x_2^2 e_1 - x_1^2x_2 e_2$ is a cycle in $K_1(x_1, x_2;I)$, and $z^\wp = x_{11}x_{21} x_{22} e_{12} - x_{11}x_{12}x_{21}e_{22}$. \medskip With the notation introduced, we have \begin{Theorem} \label{main} Let $I \subset S=K[x_1, \ldots, x_n]$ be a monomial ideal and let ${\bold c}=(c_1, \ldots, c_n)$ be the integer vector with $c_i=\max\{\deg_{x_i} u\:\; u\in G(I)\}$ for $i=1, \ldots, n$. Let $z_1, \ldots, z_r$ be multi-homogeneous cycles with multi-degree $\leq {\bold c}$, whose homology classes form a $K$-basis of $H_i({\bold x};I)$. Then the homology classes of the cycles $z_1^\wp, \ldots, z_r^\wp$ form a $K$-basis of $H_i({\bold x}^\wp; I^\wp)$. \end{Theorem} The theorem will be a consequence of the following \begin{Lemma} \label{comparison} Let $M$ be a finitely graded $S$-module, and assume that $x_1$ is a non zero-divisor modulo $M$. Then there is a natural isomorphism \[ \varphi: H_{i}(x_1,\ldots,x_n;M)\to H_i(x_2,\ldots,x_n;\bar{M}), \] where $\bar{M}$ is the $\bar{S} = S/x_1S$-module $M/x_1M$. This isomorphism is given as follows: let $z \in Z_i(x_1,\ldots,x_n;M)$ and write $z=e_1\wedge z_0+z_1$ with $z_1\in K_i(x_2,\ldots,x_n;M)$. Then the homology class $[z]\in H_{i}(x_1,\ldots,x_n;M)$ is mapped to $[\bar{z}_1]\in H_{i}(x_2,\ldots,x_n;\bar{M})$, where $\bar{z}_1$ is obtained from $z_1$ by taking the residue classes of the coefficients of $z_1$ modulo $x_1$. \end{Lemma} \begin{proof} Observe that $\bar{z_1}$ is indeed a cycle in $K(x_2, \ldots x_n; \bar{M})$, because $0= x_1 z_0 - e_1 \wedge \partial z_0 + \partial z_1$. From this equation it follows that $e_1 \wedge \partial z_0 =0$ and hence $\partial \bar{z_1} =0$. Next we show that $\varphi$ is well defined. Let $z$ be as in the statement and let $w=z+\partial b$ where $b \in K_{i+1}(x_1, \ldots, x_n;M)$. Let $b= e_1 \wedge b_0 + b_1$ with $b_1 \in K_{i+1}(x_2, \ldots, x_n; M)$. Then $w= e_1\wedge w_0 + w_1 $ where $w_1 = z_1 + x_1 b_0 + \partial b_1$. We have to show that $[\bar{w}_1] = [\bar{z}_1]$. But this is obvious, because $\bar{w}_1 = \bar{z}_1 + \partial \bar{b}_1$, so that $\bar{w}_1$ and $\bar{z}_1$ differ only by a boundary in $K_{i}(x_2, \ldots, x_n;\bar{M})$. Since $H_i (x_1, \ldots, x_n ; M) \iso \Tor_i^S (K;M)$ and $H_i (x_2, \dots, x_n; \bar{M}) \iso \Tor_i^{\bar{S}} (K, \bar{M})$, we conclude that $\dim_K H_i (x_1, \ldots, x_n ; M) = \dim_K H_i (x_2, \dots, x_n; \bar{M})$. Indeed, since $x_1$ is a non-zero on $M$, the graded minimal free resolution of $\bar{M}$ is obtained from the graded minimal free resolution of $M$ be reduction modulo $x_1$. This implies that $\dim_K \Tor_i^S (K;M) = \dim_K \Tor_i^{\bar{S}} (K, \bar{M})$. Hence, in order to prove that $\varphi$ is an isomorphism, it suffices to show that $\varphi$ is surjective. Let $[v] \in H_i (x_2, \ldots, x_n ; \bar{M})$. There exists $z_1 \in K_i (x_2, \ldots, x_n; M)$ with $\bar{z_1} = v$. It follows that $\partial z_1 = -x_1 z_0$ for some $z_0 \in K_{i-1} (x_2, \ldots, x_n;M)$. Since $0= \partial^2 z_1 = -x_1 \partial z_0$, we see that $\partial z_0 =0$. Now we set $z= e_1 \wedge z_0 + z_1$. Then $z$ is a cycle and $\varphi [z] = [v] $. \end{proof} \begin{proof}[Proof of Theorem~\ref{main}] Fix an integer $1 \leq i \leq n$. For each $u \in G(I)$ we define \[ u'= \left\{ \begin{array}{ll} (u/x_i)y , & \;\textnormal{if $x_i^2 | u$}, \\ u, & \;\text{otherwise.} \end{array} \right. \] The element $u'$ is called the {\em $1$-step polarization} of $u$ with respect to the variable $x_i$, and the ideal $I' = (\{u'| u \in G(I)\})$ is called a 1-step polarization of $I$. Obviously, the (complete) polarization of $I$ can be obtained by a sequence of 1-step polarization. Let $I'$ be the 1-step polarization of $I$ with respect to $x_i$. Without loss of generality, we may assume that $i=1$. We consider the Koszul complex $K(y,x_1, \ldots, x_n; I')= (\bigwedge H) \tensor I'$ where $H$ is the free $S[y]$-module with basis $f, e_1 \ldots, e_n$ and where $\partial f = y$ and $\partial e_j = x_j$ for $j=1, \ldots, n$. Let $z= \sum_{J} \lambda_J u_J e_J $ be a multi-homogenous cycle of $K_i(x_1, \ldots, x_n;I)$ with $\deg z \leq {\bold c}$ whose homology class is non-zero. We set $z' = \sum_{J} \lambda_J (u_J e_J)'$, where \[ (u_J e_J)'= \left\{ \begin{array}{ll} u_J e_J , & \;\textnormal{if $x_1 \nmid u_J$}, \\ u_{J}'e_J, & \;\text{if $x_1|u_J$ and $1 \notin J$}, \\ u_{J}e_{J}', & \;\text{if $x_1|u_J$ and $1 \in J$}. \end{array} \right. \] Here $e_{J}'$ is obtained from $e_J$ by replacing the factor $e_1$ in $e_J$ by $f$. \medskip As an example we consider again the cycle $z= x_1x_2^2 e_1 - x_1^2x_2 e_2$ in $K_1(x_1,x_2;I)$ where $I=(x_1^2 x_2, x_1x_2^2)$. Then $z' = x_1 x_2^2 f - x_1 y x_2 e_2$. \medskip We claim that $z'$ is a cycle in $K_i(y, x_1, \ldots, x_n; I')$, and that the map \[ H_i (y, x_1, \ldots, x_n;I') \rightarrow H_i( x_1, \ldots, x_n; I), \quad [z'] \mapsto [z] \] is an isomorphism. From this claim the theorem follows by induction on the number of 1-step polarization which are required to obtain the polarization $I^{\wp}$ of $I$. \medskip Proof of the claim: we first show that $z'$ is a cycle. We first discuss the case when $\deg_{x_1} z \leq 1$. By the definition of $(u_J e_J)'$, we have $(u_J e_J)' = u_J e_J$, for all $J$. It shows that $z=z'$ and hence $z'$ is a cycle. Now we discuss the case when $\deg_{x_1}z > 1$. Let $z=e_1\wedge z_0+z_1$ and $z'= f \wedge z'_0 +z'_1$ with $z_1\in K_i(x_2,\ldots,x_n;I)$ and $z'_1\in K_i(x_1,\ldots,x_n;I')$. Moreover, $z_0 = \sum_{1 \in J} \lambda_J u_J e_{J \setminus \{1\}}$ and $z_1 = \sum_{1 \notin J} \lambda_J u_J e_J$. From the definition of $z'$ we see that $z'_0 = z_0$ and $z'_1 = \sum_{1\notin J} \lambda _J u'_J e_J$ where $u'_J = y u_J / x_1 $. It implies that $z'_1 = (y/x_1 ) z_1$. By applying $\partial$ on $z'$, we get $\partial (z') = y z_0 + \partial (z'_1) = y z_0 + (y/x_1) \partial (z_1)$. It shows that $x_1\partial (z') = y \partial (z) = 0$. Hence $\partial (z') = 0$. We first observe that $y-x_1$ is a non-zero divisor on $S[y]/I'$ and that $I' / (y-x_1)I' = I$. Therefore, by Lemma~\ref{comparison}, there exists the isomorphism $\varphi : H_i(y, x_1, \ldots, x_n) = H_i(y-x_1, x_1, \ldots, x_n; I') \rightarrow H_i (x_1, \ldots, x_n; I)$. Thus it remains to be shown that $\varphi([z']) = [z]$. Applying the Lemma~\ref{comparison}, we write $z'= (f-e_1) \wedge w_0 +w_1$. By definition, \begin{eqnarray*} z'&=& \sum_{x_1 \nmid u_J} u_Je_J + \sum_{x_1 | u_J, 1 \notin J} u'_Je_J + \sum_{x_1|u_J, 1 \in J} u_J f \wedge e_{J \setminus \{1\}} \\ &=& \sum_{x_1 \nmid u_J} u_Je_J + \sum_{x_1 | u_J, 1 \notin J} u'_Je_J + \sum_{x_1|u_J, 1 \notin J} u_J (f-e_1) \wedge e_{J \setminus \{1\}} + \sum_{x_1|u_J, 1 \in J} u_J e_{J}. \end{eqnarray*} Therefore, \begin{eqnarray*} w_1&=& \sum_{x_1 \nmid u_J} u_Je_J + \sum_{x_1 | u_J, 1 \notin J} u'_Je_J + \sum_{x_1|u_J, 1 \in J} u_J e_{J}. \end{eqnarray*} From this it follows that $\bar{w}_1 = z$, which by the definition of $\varphi$ implies that $\varphi ([z']) = [z]$, as desired. \end{proof} \begin{Corollary}\label{polarize} Let $I\subset S$ be a monomial ideal as in Theorem~\ref{main}. Let $z_1, \ldots, z_r$ be multi-homogeneous cycles with multi-degree $\leq {\bold c}$, whose homology classes form a $K$-basis of $H_i({\bold x};S/I)$ for $i \geq 1$. Then the homology classes of the cycles $z_1^\wp, \ldots, z_r^\wp$ form a $K$-basis of $H_i({\bold x}^\wp; S^\wp/I^\wp)$. \end{Corollary} \begin{proof} We notice that for $i \geq 1$ there is an isomorphism $\varphi: H_i({\bold x}^\wp;S^\wp/I^\wp) \rightarrow H_{i+1}({\bold x}^\wp;I^\wp)$ with $\varphi ([z]) = [\partial(w)]$ and $w \in K({\bold x}^\wp; S^\wp)$ such that $z=w+I^\wp K({\bold x}^\wp;S^\wp)$. Since $\partial(f^\wp) = (\partial(f))^\wp$ for any multi-homogenous element $f \in K({\bold x}; S^\wp)$ with $\deg f \leq {\bold c}$, the desired conclusion follows. \end{proof} As an example for the polarization of Koszul cycles, we consider whisker graphs. Let $G$ be a finite simple graph on the vertex set $[n]=\{1, \ldots, n\}$. The {\em whisker graph} $G^*$ of $G$ is the graph with the vertex set $V(G^*)=\{1, \ldots, n\} \cup \{1', \ldots, n'\}$ and the edge set $E(G^*)=E(G) \cup \{ \{1, 1'\}, \{2, 2'\}, \ldots, \{n, n'\} \}$. \medskip Figure~\ref{one} displays the whisker graph of the graph $G$ with edges $\{1,2\}, \{2,3\}, \{3,4\}$ and $\{4,2\}$. \begin{figure}[hbt] \begin{center} \label{one} \psset{unit=0.6cm}\begin{pspicture}(0.5,1)(4.5,5) \pspolygon(2,2)(3,3.71)(4,2)\psline(3,3.71)(3,5.2)\psline(0.6,1.1)(2,2)\psline(4,2)(4,3.49) \psline(2,2)(2,3.49) \psline(0.6,1.1)(0.6,2.59) \rput(0.6,2.59){$\bullet$} \rput(2,3.49){$\bullet$}\rput(4,3.49){$\bullet$}\rput(2,2){$\bullet$}\rput(3,3.71){$\bullet$}\rput(4,2){$\bullet$}\rput(3,5.2){$\bullet$}\rput(0.6,1.1){$\bullet$}\rput(0.6,3.1){$1'$}\rput(2,4){$2'$}\rput(4.25,4){$3'$}\rput(2,1.5){$2$}\rput(3.25,4){$4$} \rput(4,1.5){$3$}\rput(3.1,5.7){$4'$}\rput(0.6,0.6){$1$} \end{pspicture} \end{center} \caption{} \label{example}\end{figure} Let $K$ be a field. The edge ideal $I(G)$ of $G$ is the monomial ideal in $S=K[x_1, \ldots, x_n]$ generated by the monomials $x_ix_j$ with $\{i,j\} \in E(G)$. We consider the edge ideal $I(G^*)$ of the whisker graph $G^*$ of $G$ as the monomial ideal in $S^*=K[x_1, \ldots, x_n, y_1, \ldots, y_n]$ with $I(G^*) = I_G + (\{ x_k y_k | k \in [n] \})$. Next, we let $J(G) = ( I(G) , x_1^2 , \ldots, x_n^2 )$. Then, obviously, $I(G^*) = J(G)^{\wp}$, where for simplicity we set $x_i = x_{i1} , y_i = x_{i2}$, for $i= 1, \ldots, n$. For the polarized Koszul complex of $K(x_1, \ldots, x_n; I(G))$ we use the notation $e_i = e_{i1}$ and $f_i = e_{i2}$. Given a cycle $\sum_J \lambda_J u_J e_J \in K(x_1, \ldots, x_n; J(G))$ representing a non-zero homology class, the polarized cycle in $K(x_1, \ldots, x_n, y_1, \ldots, y_n ; I(G^*))$ is given as $\sum_J \lambda_j u_j e_{J'}$ where $e_{J'} $ is obtained from $e_J$ by replacing $e_{j}$ for $j \in J$ by $f_j$ if $x_j |u$. Note that $H_n(x_1, \ldots, x_n; J(G))$ is minimally generated by the homology classes $[u e_1\wedge \ldots \wedge e_n]$ with $u = x_{i_1}\ldots x_{i_k}$ where $\{i_1, \ldots, i_k\}$ is a maximal independent set of $G$. Recall that a subset ${\mathcal S} \subset V(G)$ is an {\em independent set} of $G$ if $\{i,j\} \notin E(G)$ for all $i,j \in {\mathcal S}$. The set ${\mathcal S}$ is called a maximal independent set if ${\mathcal S} \cup \{k\}$ is not independent for all $k \notin V(G) \setminus {\mathcal S}$. It follows from Corollary \ref{polarize}, that the elements \begin{eqnarray}\label{basis} x_{i_1} \ldots x_{i_k} e_{j_1} \wedge e_{j_{n-k}}\wedge f_{i_1} \wedge \ldots \wedge f_{i_k} \end{eqnarray} form a basis of $H_n(x_1, \ldots, x_n, y_i, \ldots, y_n ; S^*/I(G^*))$ where ${\mathcal S} = \{i_1, \ldots, i_k\}$ is a maximal independent set of $G$ and $\{j_1, \ldots, j_{n-k}\}=V(G) \setminus {\mathcal S}$. \section{Powers of whisker graphs}\label{whisker graphs} In this section, we want to study the powers of whisker graphs. For the formulation of the main result we introduce the following concept. Let $G$ be a finite simple graph on $[n]$, and let ${\mathcal S}$ be a maximal independent subset of $V(G)$. We define the graph $\Gamma_{{\mathcal S}}(G)$ with vertex $V(\Gamma_{{\mathcal S}}(G)) = {\mathcal S}$ and $\{i,j\} \in E(\Gamma_{{\mathcal S}}(G))$ if and only if there exists $k \in V(G) \setminus {\mathcal S}$ such that $\{i,k\}, \{j,k\} \in E(G)$. \begin{Proposition}\label{gamma} Let $G$ be a finite simple connected graph. Then there exists an independent set ${\mathcal S}$ such that $\Gamma_{{\mathcal S}}(G)$ is connected. \end{Proposition} \begin{proof} Let $\Delta(G)$ be the clique complex of $G$ with cliques $F_1, \ldots, F_r$. We are going to construct the independent set ${\mathcal S}$ of $G$ as follows. Let $v_1 \in V(F_1)$. We may assume that $v_1 \in V(F_i)$ for $i = 1, \ldots, t$ and $v_1 \notin V(F_i)$ for $i>t$. If $t=r$, then we are done. Assume that $t <r$. Since $G$ is connected, there exists $F_i$ with $i>t$, say $F_{t+1}$, such that $V(F_{t+1}) \cap V(F_j) \neq \emptyset $ for some $j \leq t$. Since $F_{t+1}$ is a maximal clique, $V(F_{t+1}) \not\subset \bigcup_{i=1}^t V(F_i) $ because otherwise $v_1 \in V(F_{t+1})$, a contradiction. Hence, we may choose $v_2 \in V(F_{t+1} )\setminus \bigcup_{i=1}^t V(F_i) $. We may assume that $v_2 \in V(F_i)$ for $i=t+1, \ldots, s$ and does not belong to any other clique. If $s=r$, then $\Gamma_{{\mathcal S}}(G) $ is a line graph with vertex set $ \{v_1, v_2\}$. Indeed, $\{v_1, v_2\} \notin E(G)$ because the set of neighbors of $v_1$ is equal to $\bigcup_{i=1}^t F_i$ and $v_2 \notin \bigcup_{i=1}^t F_i $. On the other hand, if $k \in V(F_{t+1}) \cap V(F_j)$. then $\{v_1,k\}, \{v_2,k\} \in E(G)$. Therefore, $\{v_1, v_2\} \in E(\Gamma_{{\mathcal S}}(G))$. Consider all $F_j$ for $j >s$ such that $V(F_j) \subset \bigcup_{i=1}^t V(F_i)$. We may assume that it is the case for $F_{s+1}, \ldots, F_k$. If $k=r$, then $\{v_1, v_2\}$ is an independent set for $G$, and we are done. If $k<r$, then since $G$ is connected there exists a clique $F_i$, say $F_{k+1}$, such that $V(F_{k+1}) \cap V(F_j) \neq \emptyset $ for some $j<k$ and $V(F_{k+1}) \not\subset \bigcup_{i=1}^s V(F_i) (= \bigcup_{i=1}^k V(F_i) ) $. We choose $v_3 \in V(F_{k+1}) \setminus \bigcup_{i=1}^s V(F_i)$. If $j<t$ then $\{v_1, v_3\} $ will be an edge of $\Gamma_{{\mathcal S}}(G)$, and if $t+1\leq j\leq s$, then $\{v_2,v_3\}$ will be an edge of $\Gamma_{{\mathcal S}}(G)$. Proceeding this way, we obtain the desired independent set ${\mathcal S}$ of $G$ such that $\Gamma_{{\mathcal S}}(G)$ is connected. \end{proof} We call an independent set ${\mathcal S}$ of $G$ {\em friendly} if it satisfies the condition that $\Gamma_{{\mathcal S}}(G)$ is connected. For example, if we consider the line graph $L$ on vertex set $[4]$ with edges $\{\{1,2\}, \{2,3\}, \{3,4\}\}$. Then ${\mathcal S}= \{1,3\}$ is a friendly independent set of $L$ while $\{1,4\}$ is not a friendly independent set of $L$. \begin{Theorem}\label{whisker} Let $G$ be a finite simple connected graph on vertex set $[n]$, and $G^*$ be the whisker graph of $G$. Furthermore, let $I(G^*)\subset S^*=K[x_1, \ldots, x_n, y_1, \ldots, y_n]$ be the edge ideal of $G^*$. Then \[ \depth (S^*/I(G^*)^k) \leq n-k+1 , \text{\; for \;} k= 1, \ldots, n. \] \end{Theorem} \begin{proof} Let $M$ be an $S^*$-module and consider the Koszul complex \[ K(M) = K(x_1, \ldots, x_n, y_1, \ldots, y_n;M) \] with $K_1(M) = \Dirsum_{i=1}^n M e_i \dirsum \Dirsum_{j=1}^n M f_j$ and $\partial e_i = x_i$ and $\partial f_j = y_j$, for all $i,j$. We first show that \[ H_{2n-2} (I(G^*)^n) \neq 0. \] This will imply that $\depth (S^*/ I(G^*)^n) \leq 1$. To see that the above Koszul homology does not vanish, we proceed as follows. By Proposition~\ref{gamma} we may choose a friendly independent set ${\mathcal S}$ of $G$ with $|{\mathcal S}| =s$. Since $\Gamma_{{\mathcal S}}(G)$ is connected, there exists a spanning tree $T$ of $\Gamma_{{\mathcal S}}(G)$ with $s-1$ edges $\alpha_1, \ldots, \alpha_{s-1}$. We may assume that $\alpha_1, \ldots, \alpha_{s-1}$ is a leaf order for $T$. In other words, the following conditions are satisfied: (i) $\alpha_1$ has a free vertex in $T$, (ii) for each $j>1$, $ \alpha_j \cap \alpha_i \neq \emptyset$ for some $i<j$ and $\alpha_j$ has a free vertex in the tree $T_j= \alpha_1, \ldots, \alpha_{j}$. Now, we label the vertices of $T$ inductively as follows: $1$ is the free vertex of $\alpha_1$ in $T_1$ and the other vertex in $T_1$ is given the label $2$. Suppose, the labeling of $T_{j-1}$ is defined. Then we give the new vertex of $T_j$, the label $j+1$. Then $\alpha_1=\{1,2\}$ and for each $j>1$, $\alpha_j=\{i_j,j+1\}$, where $\{i_j\}= \alpha_j \cap \alpha_i$. The following Figure~\ref{two}, gives an example of such a labeling. \begin{figure}[hbt]\begin{center}\label{two} \psset{unit=0.6cm}\begin{pspicture}(0.5,1.5)(4.5,3)\psline(2,2)(4,2)\psline(0,1)(2,2) \psline(0,3)(2,2)\psline(4,2)(6,1)\psline(4,2)(6,3)\rput(2,2){$\bullet$}\rput(4,2){$\bullet$}\rput(0,1){$\bullet$} \rput(0,3){$\bullet$}\rput(6,1){$\bullet$}\rput(6,3){$\bullet$}\rput(2,1.5){$2$}\rput(4,1.5){$4$}\rput(-0.4,0.9){$1$} \rput(-0.4,3.1){$3$}\rput(6.4,0.9){$5$}\rput(6.4,3.1){$6$}\rput(3,1.6){$\alpha_3$}\rput(1.3,1.1){$\alpha_1$} \rput(1.3,2.9){$\alpha_2$}\rput(4.9,2.9){$\alpha_4$}\rput(5,1.1){$\alpha_5$} \end{pspicture} \end{center} \caption{} \label{example}\end{figure} According to our labeling of $T$, we have ${\mathcal S}= \{1, \ldots, s\}$. By definition of $\Gamma_{{\mathcal S}}(G)$, there exists for each edge $\alpha_j=\{i_j,j+1\} \in E(T)$, a vertex $v_j \in \{s+1, \ldots, n\}$ such that $\{i_j,v_{j}\}, \{v_j,j+1\} \in E(G)$. Then $z_j= x_{i_j}x_{v_j} e_{j+1} - x_{j+1} x_{v_j} e_{i_j}$ is a cycle belonging to $Z_1(I(G^*))$. Furthermore, for each $k \in \{s+1, \ldots, n\}$, we choose $j_k \in {\mathcal S}$ such that $\{k, j_k\} \in E(G)$. Then $z_k = x_k x_{j_k} f_k - x_k y_k e_{j_k} $ is a cycle belonging to $Z_1(I(G^*))$. This gives $n-s$ such cycles. Let \[ c= \prod _{i =1}^s x_i e_{s+1} \wedge \ldots \wedge e_n \wedge f_1 \wedge \ldots \wedge f_s. \] Note that by (\ref{basis}), $c$ is a cycle in $Z_n(S^*/I(G^*))$ whose homology class $[c]$ in $H_n(S^*/I(G^*))$ is non-zero. In particular, $[\partial (c)]$ is non-zero homology class in $H_{n-1}(I(G^*))$. Let \[ a=c \wedge z_1\wedge \ldots \wedge z_{s-1} \wedge z_{s+1} \wedge \ldots \wedge z_n. \] Observe that $a \in K_{2n-1}(I(G^*)^{n-1})$. We set $z= \partial (a)$. Then $z \in Z_{2n-2} (I(G^*)^n)$. Indeed, $z = \partial (c) \wedge z_1\wedge \ldots \wedge z_{s-1} \wedge z_{s+1} \wedge \ldots \wedge z_n$, and it has coefficients in $I(G^*)^n$ because $\partial (c)$ and each $z_i$ has coefficients in $I(G^*)$. We claim that $[z]$ is a non-zero homology class in $H_{2n-2} (I(G^*)^n)$. To prove the claim, we show that $z$ is not a boundary, that is, there does not exist any $b \in K_{2n-1}(I(G^*)^n)$ such that $z= \partial b$. On contrary, assume that such $b$ exists. Then $\partial (b) = \partial (a) =z$ implies $\partial ( a-b) = 0$ which gives $a-b \in Z_{2n-1}(I(G^*)^{n-1} )$. Then, there exists $b' \in K_{2n}(S^*)$ such that $\partial (b')= a-b$ where $b' = v e_1 \wedge \ldots \wedge e_n \wedge f_1 \wedge \ldots \wedge f_n$ and $v$ is a monomial in $S^*$ because all cycles under consideration are ${\NZQ Z}^{2n}$-graded. Note that \[ w= ( \prod _{i=1}^s x_i ) (\prod_{k=s+1}^n x_k x_{j_k} ) (\prod_{j=1}^{s-1} x_{i_j} x_{v_j}) e_2 \wedge \cdots \wedge e_{n} \wedge f_1 \wedge \cdots \wedge f_n \] with $i_j, j_k \in {\mathcal S}$, $k, v_j \in \{s+1, \ldots, n\}$ is a non-zero term of $a$ and it is not cancelled by any other term of $a$ because the product $ e_2 \wedge \cdots \wedge e_n \wedge f_1 \wedge \cdots \wedge f_n$ appears only once in the expansion of $a$. To see this, consider \begin{eqnarray}\label{c} c \wedge z_{s+1} \wedge \cdots \wedge z_n = ( \prod _{i=1}^s x_i ) (\prod_{k=s+1}^n x_k x_{j_k} ) e_{s+1} \wedge \cdots \wedge e_{n} \wedge f_1 \wedge \cdots \wedge f_n. \end{eqnarray} Therefore, it follows that the term $w$ appears only once in the expansion of $a$ if $e_2 \wedge \cdots \wedge e_{s-1}$ appears only once in the expansion of $z_1 \wedge \cdots \wedge z_{s-1}$. Now to see this, we write $z_j=g_j - h_j$, where $g_j=x_{i_j}x_{v_j} e_{j+1} $ and $h_j=x_{j+1} x_{v_j} e_{i_j}$ for $j= 1, \ldots, s-1$. Note that for $1 \leq t \leq s-1$ the wedge product $z_1\wedge \cdots \wedge z_{t}$ is a linear combination of $g_{i_1}\wedge \cdots \wedge g_{i_k} \wedge h_{j_1}\wedge \cdots \wedge h_{j_t}$ with $\{i_1, \ldots, i_k\}\cup\{ j_1, \ldots, j_t\}= \{1, \ldots, t\}$. We prove by induction on $t$ that among these terms $g_1 \wedge \cdots \wedge g_t$ is the only term that does not contain $e_1$. For $t=1$, the assertion is trivial. Now let $t>1$ and assume that th sonly term that does not contain $e_1$ is $g_1 \wedge \cdots \wedge g_{t-1}$. Then, the only terms of $z_1 \wedge \cdots \wedge z_t$ which do not contain $e_1$ are either $g_1 \wedge \cdots \wedge g_{t}$ or $g_1 \wedge \cdots \wedge g_{t-1} \wedge h_t$. However, by the definition of the cycles $z_j$ given in terms of the tree $T$ it follows that $i_t = \{2, \ldots, t-1\}$. Therefore, $g_1 \wedge \cdots \wedge g_{t-1} \wedge h_t=0$. \medskip Next, we show that $a \in K_{2n-1}(I(G^*)^{n-1}) \setminus K_{2n-1} (I(G^*)^n)$. For this it suffice to show that $w \in K_{2n-1}(I(G^*)^{n-1}) \setminus K_{2n-1} (I(G^*)^n)$, because $w$ is a non-zero term of $a$ which does not cancel against any other term in $a$, as we have just seen. In fact, $ ( \prod _{i=1}^s x_i ) (\prod_{k=s+1}^n x_k x_{j_k} ) (\prod_{j=1}^{s-1} x_{i_j} x_{v_j}) $ which is coefficient of $w$, there are $n-1$ terms with indices in $\{s+1, \ldots, n\}$ and $n+s-1$ terms with indices in ${\mathcal S}=\{1, \ldots, s\}$. Since ${\mathcal S}$ is a maximal independent set, this implies that $w$ contains a product of exactly $n-1$ generator of $I(G^*)$. Since all coefficients of $b=a-\partial(b')$ are in $I(G^*)^n$ and the term $w$ which appears in the expansion of $a$ does not have coefficient in $I(G^*)^{n}$, $w$ must be cancelled by some term of $\partial (b')$. This gives \[ v x_1 e_2 \wedge \cdots \wedge e_{n} f_1 \wedge \ldots \wedge f_n = ( \prod _{i=1}^s x_i ) (\prod_{k=s+1}^n x_k x_{j_k} ) (\prod_{j=1}^{s-1} x_{i_j} x_{v_j}) e_2 \wedge \cdots \wedge e_{n} \wedge f_1 \wedge \cdots \wedge f_n, \] which implies \[ v= ( \prod _{i=2}^s x_i ) (\prod_{k=s+1}^n x_k x_{j_k} ) (\prod_{j=1}^{s-1} x_{i_j} x_{v_j}) \in I(G^*)^{n-1}. \] The coefficient of the term $v y_n e_1 \wedge \ldots \wedge e_n f_1 \wedge \cdots \wedge f_{n-1} $ which appears in the expansion of $\partial(b')$ does not belong to $I(G^*)^{n}$ because $x_n$ is the only neighbor of $y_n$. Also the term $v y_n e_1 \wedge \ldots \wedge e_n f_1 \wedge \cdots \wedge f_{n-1} $ is not cancelled by any of the terms of $a$ because from (\ref{c}) we can see that all terms of $a$ contain the wedge product $f_1\wedge \cdots \wedge f_n$ as a factor. Hence, our assumption that $z$ is a boundary leads us to contradiction. For simplicity of notation, we set $z'_i=z_i$ for $i = 1, \ldots, s-1$ and $z'_i=z_{i+1}$ for $i=s+1, \ldots, n-1$. Note that $\partial(c)\wedge z'_1 \wedge \cdots \wedge z'_{k-1} \in Z_{n+k-2}(I(G^*)^k)$. We claim that this cycle is not a boundary in $K(I(G^*)^k)$. This then implies that $\depth (S^*/I(G^*)^k) \leq n+k-1$ since $H_{n+k-1}(S^*/I(G^*)^k) \iso H_{n+k-2} (I(G^*)^k) \neq 0$. In order to prove the claim, assume that there exists $b \in K_{n+k-1}(I(G^*)^k)$ such that $\partial (b) = \partial (c) \wedge z'_1 \wedge \cdots \wedge z'_{k-1}$. Let $b'=b \wedge z'_k \wedge \cdots \wedge z'_{n-1} $. Then $b' \in K_{2n-1}(I(G^*)^n)$ and $\partial(b') = \partial (b) \wedge z'_k \wedge \cdots \wedge z'_{n-1} = z$, a contradiction. \end{proof} Our hypothesis of Theorem~\ref{whisker} which requires that $G$ is connected is needed. For example, if we take the disconnected graph $G$ on vertex $[4]$ with edge $\{\{1,2\}, \{3,4\}\}$, then $\depth(S^*/I(G^*)^4) = 2$. \begin{Remark}{\em Let $I$ be an arbitrary monomial ideal in $K[x_1, \ldots, x_n, y_1, \ldots, y_n]$. In \cite[Theorem 3.3]{HQ}, it is shown that $\depth (S/I^k) \leq 2n-k+1$ for $k=1, \ldots, r$, where $r< 2n$ is a number depending on $I$. Comparing this result with our Theorem~\ref{whisker}, where $I$ is the edge ideal of a whisker graph, our bound is about half of the bound which is valid for general monomial ideals.} \end{Remark} \begin{Corollary}\label{limit} Let $G$ be a finite simple connected graph on vertex set $[n]$, $G^*$ be the whisker graph of $G$, and $I(G^*)\subset S^*$ be the edge ideal of $G^*$. If $G$ is bipartite, then $\depth (S^*/I(G^*)^k) =1$ for all $k \geq n$, and if $G$ is non-bipartite, then $\depth (S^*/I(G^*)^k) =0$ for all $k\geq n$. \end{Corollary} \begin{proof} Suppose first that $G$ is bipartite. Then $G^*$ is bipartite as well. It follows from \cite[Theorem 5.9]{SVV} that the $\depth (S^*/I(G^*)^k )\geq 1$ for all $k$. Thus our Theorem~\ref{whisker} implies that $\depth(S^*/I(G^*)^n)=1$. On the other hand, since the Rees ring $R(I(G^*))$ of $I(G^*)$ is Cohen-Macaulay (see for example \cite[Corollary 5.20]{EH}), the result of Eisenbud and Huneke \cite[Proposition 3.3]{EH2} yields the desired conclusion. Now let $G$ be a non-bipartite graph. It follows from \cite[Corollary 4.3]{CMS}, applied to our case, that $\Ass (S^*/I(G^*)^k) = \Ass(S^*/I(G^*)^{n})$ for all $k \geq n$. On the other hand, since $G$ is non-bipartite, we know by \cite[Corollary 3.4]{CMS} that $\depth (S^*/I(G^*)^k) =0$ for $k \gg 0$. This implies that $\depth (S^*/I(G^*)^k) =0$ for all $k\geq n$. \end{proof} In general the upper bounds for the depth of the powers of the edge ideal of a whisker graph given in Theorem~\ref{whisker} are not attained. For example, if $G$ is a 3-cycle then $\depth(S^*/I(G^*))=3 $ and $\depth(S^*/I(G^*)^k)=0$ for all $k \geq 2$. Even if $G$ is bipartite, this bound is not attained. For example, if $G$ is a 4-cycle, then $\depth(S^*/I(G^*))=4$, $\depth(S^*/I(G^*)^2)=3$ and $\depth(S^*/I(G^*)^k) =1$ for $k \geq 3$. On the other hand, Villarreal showed \cite[Proposition 6.3.7]{V}, that if $G$ is a forest then $\depth(S^*/I(G^*)^2) \geq n-1$. Together with our Theorem~\ref{whisker} it follows that $\depth(S^*/I(G^*)^2) = n-1$. By using the arguments applied in Villarreal's proof, we now show more generally \begin{Theorem} \label{tree} \label{whiskertree} Let $G$ be a forest on $n$ vertices and let \[ I=I(G^*)\subset S^*=K[x_1,\dots,x_n,y_1,\ldots,y_n] \] be the edge ideal of $G^*$. Then \[ \depth(S^*/I^k) \geq n-k+1 \quad \text{for $k=1,\ldots,n$}. \] \end{Theorem} \begin{proof} We show this by induction on $k + n$. If $k+n=1$, then either $k=1$ or $n=1$. If $k=1$, then $\depth(S^*/I^k)=n$ since $I$ is a Cohen-Macaulay of height $n$ and for $n=1$ the assertion is trivial. Let $x_n$ be a free vertex of the forest $G$ with $\{x_{n-1}, x_n\} \in E(G)$. Following the notation used in the proof \cite[Proposition 6.3.7]{V}, we denote by $J$ the ideal which is obtained by $I$ by substituting $x_n=0$ and by $L$ the ideal which obtained from $J$ by substituting $x_{n-1} = 0$. Furthermore, we set $K=(J^k, x_{n-1}x_n, x_ny_n)$. Since $(J^k, x_{n-1}x_n):x_n = (L^k, x_{n-1})$, we obtain the exact sequence \begin{eqnarray*}\label{depth1} 0 \rightarrow S^*/ (L^k, x_{n-1}) \rightarrow S^* / ( {J^k, x_{n-1}x_n} )\rightarrow S^* / (J^k, x_n) \rightarrow 0 . \end{eqnarray*} Since $J$ is edge of a whisker forest on $n-1$ vertices and $L$ is the edge ideal of the whisker forest on $n-2$ vertices, our induction hypothesis implies that \[ \depth (S^*/ (L^k, x_{n-1})) \geq n-k+2 \text{ and } \depth (S^* / (J^k, x_n)) \geq n-k+1. \] This implies that \begin{eqnarray}\label{depth2} \depth (S^* / (J^k, x_{n-1}x_n )) \geq n-k+1. \end{eqnarray} Since $(J^k, x_{n-1}x_n): x_ny_n = (L^k, x_{n-1})$, we obtain the exact sequence \begin{eqnarray}\label{depth3} 0 \rightarrow S^*/ (L^k, x_{n-1}) \rightarrow S^* / ( {J^k, x_{n-1}x_n} )\rightarrow S^* / K \rightarrow 0 . \end{eqnarray} From (\ref{depth2}) and (\ref{depth3}), we obtain \begin{eqnarray}\label{depth4} \depth (S^* / K) \geq n-k+1. \end{eqnarray} Note that $(I^k, x_n y_n) = (J, x_{n-1}x_n)^k + (x_n y_n)$. Therefore, $(I^k , x_n y_n) : x_{n-1} x_n = (J, x_{n-1} x_n)^k : x_{n-1}x_n + (y_n)$. Since $(J, x_{n-1} x_n)$ is the edge ideal of the graph for which $\{n-1, n\}$ is an edge with free vertex $n$, it follows by result of Morey \cite[Lemma 2.10]{M} that $(J, x_{n-1} x_n)^k : x_{n-1}x_n = (J, x_{n-1} x_n)^{k-1}$. Therefore, altogether we have that $(I^k, x_ny_n): x_{n-1}x_n = (J, x_{n-1}x_n)^{k-1} + (y_n)$. Thus, we obtain the exact sequence \begin{eqnarray}\label{depth5} 0 \rightarrow S^*/ ((J, x_{n-1x_n})^{k-1} + (y_n) ) \rightarrow S^* / ( {I^k, x_n y_n} )\rightarrow S^* / K \rightarrow 0 . \end{eqnarray} We claim that for $k=2, \ldots, n$ we have $\depth (S^*/ (J, x_{n-1}x_n)^{k-1} ) \geq n-k+2$. Therefore, (\ref{depth4}) and (\ref{depth5}) implies \begin{eqnarray}\label{depth6} \depth (S^* / ({I^k, x_n y_n}) ) \geq n-k+1. \end{eqnarray} By using (\ref{depth6}) and our induction hypothesis, the exact sequence \begin{eqnarray*} 0 \rightarrow S^*/ I^{k-1} \rightarrow S^* / I^k \rightarrow S^* / (I^k, x_n y_n) \rightarrow 0 . \end{eqnarray*} yields that $\depth (S^*/ I^k) \geq n-k+1$ and proves our theorem. It remains to prove the claim. For that we use induction on $k$. For $k=2$, this inequality is shown in the proof of \cite[Proposition 6.3.7]{V}. Suppose that $k>2$. Since $(J, x_{n-1}x_n)$ is the edge ideal of a tree with free vertex $n$, we may apply \cite[Lemma 2.10]{M}, and obtain the exact sequence \begin{eqnarray*}\label{depth7} 0 \rightarrow S^*/ (J, x_{n-1x_n})^{k-2} \rightarrow S^* / ( J, x_{n-1} x_n )^{k-1} \rightarrow S^* / (J^{k-1}, x_{n-1}x_n) \rightarrow 0 . \end{eqnarray*} By our induction hypothesis, $\depth ( S^*/ (J, x_{n-1x_n})^{k-2} ) \geq n-k+3$ and (\ref{depth2}) applied for $k-1$ yields $\depth ( S^*/ (J, x_{n-1x_n})^{k-2} ) \geq n-k+2$. Therefore, it follows that $\depth (S^* / ( J, x_{n-1} x_n )^{k-1} ) \geq n-k+2$, as desired. \end{proof} By combining Theorem~\ref{whisker} and Theorem~\ref{tree}, we obtain \begin{Corollary} Let $G$ be a tree. Then \[ \depth(S^*/I(G^*)^k) = n-k+1 \quad \text{for $k=1,\ldots,n$}. \] \end{Corollary} More generally, we expect that if $G$ is a forest with $m$ connected components, then \[ \depth(S^*/I(G^*)^k) = \left\{ \begin{array}{ll} n-k+1, & \text{if $k\leq n-m+1$}, \\ m, &\text{if $k \geq n-m+1$}. \end{array} \right. \]
{ "timestamp": "2013-10-31T01:01:43", "yymm": "1310", "arxiv_id": "1310.7984", "language": "en", "url": "https://arxiv.org/abs/1310.7984", "abstract": "In this paper, we introduced the polarization of Koszul cycles and use it to study the depth function of powers of edge ideals of whisker graphs.", "subjects": "Commutative Algebra (math.AC)", "title": "Polarization of Koszul cycles with applications to powers of edge ideals of whisker graphs", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9808759593358227, "lm_q2_score": 0.8175744828610095, "lm_q1q2_score": 0.8019391552047819 }
https://arxiv.org/abs/2211.03218
Projection error-based guaranteed L2 error bounds for finite element approximations of Laplace eigenfunctions
For conforming finite element approximations of the Laplacian eigenfunctions, a fully computable guaranteed error bound in the $L^2$ norm sense is proposed. The bound is based on the a priori error estimate for the Galerkin projection of the conforming finite element method, and has an optimal speed of convergence for the eigenfunctions with the worst regularity. The resulting error estimate bounds the distance of spaces of exact and approximate eigenfunctions and, hence, is robust even in the case of multiple and tightly clustered eigenvalues. The accuracy of the proposed bound is illustrated by numerical examples. The demonstration code is available at https://ganjin.online/xfliu/EigenfunctionEstimation4FEM .
\section{Introduction} Deriving guaranteed error bounds for approximate eigenfunctions of the Laplace operator is a challenging task due to possible ill-posedness of eigenfunctions. In the case of multiple and tightly clustered eigenvalues, the corresponding exact \LIU{eigenfunctions} are sensitive even to small perturbations of the problem and may change abruptly. Any accurate error bound has to take into account this sensitivity. Therefore, the recent results \cite{CanDusMadStaVoh2017,CanDusMadStaVoh2018,CanDusMadStaVoh2020,LiuVej2022} consider an arbitrary cluster of eigenvalues and the space generated by corresponding eigenfunctions. The resulting error bounds then estimate a distance between the eigenfunction spaces associated to exact and approximate eigenvalues. The particular distances between spaces are naturally based either on the energy or $L^2$ norm. Interestingly, the $L^2$ norm bounds provided by Algorithm I of \cite{LiuVej2022}, which is based on the Rayleigh quotients of the approximate eigenfunctions, are considerably less accurate than the corresponding bounds in the energy norm. For the linear FEM solutions to the Laplacian eigenvalue problems, Algorithm I of \cite{LiuVej2022} provides the error bounds in the energy norm that exhibit the optimal rate of convergence. However, the $L^2$ error bounds often converge sub-optimally and are considerably less accurate. It is worth pointing out that, the residual error-based Algorithm II of \cite{LiuVej2022} provides an optimal $L^2$ norm bound through the re-constructed flux and the Prager--Synge method; see the computation results for the L-shaped domain in \S \ref{sec:l-shape}. \medskip In this paper, we bound the $L^2$ error by utilizing the {\em a priori} error estimation proposed {in} \cite{LiuOis2013} for the boundary value problems, and the resulting bound achieves the optimal rate of convergence and more accurate numerical results. For eigenfunction space $E$ associated to eigenvalues in a cluster $\mathcal{C}$ with eigenvalues $\{\lambda_n, \cdots, \lambda_N\}$, we obtain the following bounds in Lemma \ref{thm:u_and_pi_k_u_new_ver} and Theorem \ref{th:main_new_}: \begin{equation*} \delta_b(E,E_h) \le (1+\beta) \max_{u \in E, \|\nabla u\|=1} \|u - P_h u\| \le (1+\beta) {\lambda_N} C_h^2~. \end{equation*} Here, $\delta_b(E,E_h)$ is the $L^2$ norm directed distance to measure the distance between $E$ and its approximate eigenspace $E_h$; $\beta$ is a quantity related to the cluster width and the gap between the cluster $\mathcal{C}$ and the rest of the spectrum; $C_h$ is a quantity with an explicitly known or computable value that comes from the {\em a priori} error estimation for the projection operator $P_h$. The proposed estimate of quantity $\beta$ can be regarded as an improvement of \cite{CarGed2011}; see the comparison in Remark \ref{remark:comparison_carsten_gedick}. Also, the obtained explicit bounds are consistent with the standard qualitative analysis for the eigenfunction approximations; see the discussion in Remark \ref{remark:convergence-rate}. \medskip To evaluate the bounds on eigenfunctions, suitable lower and upper bounds on eigenvalues are required. In this paper, we assume that sufficiently accurate two-sided bounds on eigenvalues are available, although we admit that computing guaranteed eigenvalue bounds, especially from below, is not a simple task. We use the recent method \cite{Liu2015} based on the explicitly know interpolation constant for the Crouzeix--Raviart finite element method; see also, \cite{LiuOis2013,CarGal2014,CarGed2014}. This method provides lower bounds on eigenvalues and we further use the Lehmann--Goerisch method \cite{Lehmann1949,Lehmann1950,GoeHau1985} to compute their high-precision improvements. Let us note that there is a vast literature on error estimates for symmetric elliptic eigenvalue problems. Classical works \cite{Chatelin1983,BabOsb:1991,Boffi:2010} provide the fundamental theories. Many existing {\em a posteriori} error bounds on eigenvalues contain unknown constants or are valid asymptotically; see, e.g., \cite{DurGasPad1999,ArmDur2004,Yang2010,MehMie2011,DarDurPad2012,GiaHal2012,JiaCheXie2013,HuHuaLin2014}. In the last years, several results providing fully computable and guaranteed a posteriori error estimates for eigenvalues appeared; see \cite{CarGal2014,CarGed2014,Liu2015,LiuOis2013,SebVej2014,Vejchodsky2018b,Vejchodsky2018,carstensen2021direct}. These estimates contain no unknown constants and bound eigenvalues on all meshes, not only asymptotically. In particular, the general framework proposed in \cite{Liu2015} was applied to the Stokes eigenvalue problem \cite{Xie2LIU-2018}, Steklov eigenvalue problem \cite{you-xie-liu-2019}, and biharmonic operators related to the quadratic interpolation error constants \cite{liu-you:2018,LiaoYuLiu2019}. The series of papers \cite{CanDusMadStaVoh2017,CanDusMadStaVoh2018,CanDusMadStaVoh2020} provides guaranteed, robust, and optimally convergent {\em a posteriori} bounds for eigenvalues and even for corresponding eigenfunctions for both conforming and nonconforming approximations. The last paper in the series solves the difficult case of multiple and tightly clustered eigenvalues. The recent work \cite{LiuVej2022} proposes two algorithms to handle multiple and tightly clustered eigenvalues as well and provides alternative guaranteed and fully computable error bounds for eigenfunctions. Particularly, the residual error-based Algorithm II of \cite{LiuVej2022} provides high-precision bounds by successfully extending the Davis--Kahan theorem to weakly formulated eigenvalue problems. \medskip The rest of the paper is organized as follows. Section~\ref{se:eigenproblem} briefly recalls the Laplace eigenvalue problem, its discretization by the finite element method, and division of the spectrum into clusters. Section~\ref{se:optimalL2} derives a project error-based bound for finite element eigenfunctions in the $L^2$ {sense}. Section~\ref{se:numex} presents the results of two numerical examples and Section~\ref{se:conclusions} draws the conclusions. Below is url of the online demonstration: \begin{center} \url{https://ganjin.online/xfliu/EigenfunctionEstimation4FEM} \end{center} \section{Laplace eigenvalue problem} \label{se:eigenproblem} Let us considers the Laplace eigenvalue problem to find eigenvalues $\lambda_i \in \mathbb{R}$ and corresponding eigenfunctions $u_i \neq 0$ such that \begin{equation} \label{eq:modpro} -\Delta u_i = \lambda_i u_i \quad\text{in }\Omega, \qquad u_i = 0 \quad\text{on }\partial\Omega, \end{equation} where $\Omega \subset \mathbb{R}^d$ is a bounded, Lipschitz $d$-dimensional domain. The weak formulation of this eigenvalue problem reads: find $\lambda_i \in \mathbb{R}$ and $u_i \in H^1_0(\Omega) \setminus \{0\}$ such that \begin{equation} \label{eq:weakf} (\nabla u_i, \nabla v) = \lambda_i (u_i,v) \quad \forall v \in H^1_0(\Omega), \end{equation} where $H^1_0(\Omega)$ is the usual Sobolev space of square integrable functions with the square integrable gradients and with zero traces on the boundary $\partial\Omega$; and $(\cdot,\cdot)$ stands for the $L^2(\Omega)$ inner product. The Laplace eigenvalue problem is well studied in \cite{BabOsb:1991,Boffi:2010}. There exists a countable sequence of eigenvalues $$ 0 < \lambda_1 \leq \lambda_2 \leq \cdots, $$ where we repeat each eigenvalue according to its multiplicity. The corresponding eigenfunctions $u_i \in H^1_0(\Omega)$ are assumed to be normalized such that $$ (u_i,u_j) = \delta_{ij}, \quad i,j = 1,2, \dots. $$ We discretize problem \eqref{eq:weakf} by the standard conforming finite element method. For simplicity, we assume $\Omega$ to be a polytope. We consider the usual conforming simplicial mesh $\mathcal{T}_h$ in $\Omega$ and define the finite element space $V_h$ of piece-wise polynomial and continuous functions over the mesh $\mathcal{T}_h$ satisfying the Dirichlet boundary conditions as $$ V_h =\{v_h \in H^1_0(\Omega) : {v_h} |_K \in \mathbb{P}_p(K) \text{ for all } K \in \mathcal{T}_h \}, $$ where $\mathbb{P}_p(K)$ stands for the space of polynomials of degree at most $p$ defined in $K$. The finite element eigenvalue problem reads: find $\lambda_{h,i}\in\mathbb{R}$ and $u_{h,i} \in V_h\setminus\{0\}$ such that \begin{equation} \label{eq:eig_pro_with_fem} (\nabla u_{h,i,} \nabla v_h) = \lambda_{h,i} (u_{h,i}, v_h)\quad \forall v_h \in V_h, \end{equation} where $i=1,2,\dots,\operatorname{dim} (V_h)$. Discrete eigenfunctions are assumed to be normalized such that $(u_{h,i},u_{h,j})=\delta_{ij}$ and $(\nabla u_{h,i},\nabla u_{h,j})=\lambda_{h,i} \delta_{ij}$. As we mentioned in the introduction, {we will formulate the $L^2$ error bound on eigenfunctions for clusters of eigenvalues.} For the purpose of the theory, the splitting of the spectrum into clusters can be arbitrary. Let $n_k$ and $N_k$ stand for indices of the first and the last eigenvalue in the $k$th cluster; see Figure~\ref{fi:clusters}. {Note that eigenvalues in a cluster need not equal to each other}. We consider the $k$th cluster to be of interest, and set $n=n_k$ and $N=N_k$ to simplify the notation. Let $E_k$ be the space of exact eigenfunctions associated to $k$th cluster of eigenvalues: $$ E_k = \operatorname{span}\{ u_{n}, u_{n+1}, \dots, u_{N} \} ~. $$ Similarly, finite element approximations $u_{h,i} \in H^1_0(\Omega)$ of exact eigenfunctions $u_i$, {for $i=n,n+1,\dots,N$}, form the corresponding approximate space: $$ E_{h,k} = \operatorname{span}\{ u_{h,n}, u_{h,n+1}, \dots, u_{h, N} \}~. $$ \begin{figure}[t] \begin{tikzpicture}[scale=1] \newcommand{0.1}{0.1} \newcommand{\tick}[1]{\draw [semithick] (#1,-0.1)--(#1,0.1);} \draw [semithick] (0,0)--(6.2,0); \draw [thick,dotted] (6.2+0.2,0)--(6.2+1.3,0); \draw [thick] (6.2+1.5,0)--(6.2+7,0); \tick{0.5}\node [below] at (0.5,-0.1) {$0$}; \tick{1.7}\node [above] at (1.7,0.1) {$\lambda_{n_1}$}; \tick{1.9} \tick{2.5}\node [above] at (2.3,0.1) {$\cdots$}; \tick{2} \tick{2.7} \node [above] at (2.9,0.1) {$\lambda_{N_1}$}; \node [below] at (2.25,-0.1) {$1$st cluster}; \tick{4}\node [above] at (4,0.1) {$\lambda_{n_2}$}; \tick{4.25} \tick{4.6}\node [above] at (4.7,0.1) {$\cdots$}; \tick{5.1} \tick{5.2} \node [above] at (5.3,0.1) {$\lambda_{N_2}$}; \node [below] at (4.75,-0.1) {$2$nd cluster}; \tick{9.5}\node [above] at (9,0.1) {$(\lambda_n:=)\lambda_{n_k}$}; \tick{9.65} \tick{10} \tick{10.1} \tick{10.35} \tick{10.5} \tick{10.7} \tick{10.8} \tick{11}\node [above] at (11.75,0.1) {$\lambda_{N_k}(=:\lambda_{N})$}; \node [below] at (10.5,-0.1) {$k$th cluster}; \end{tikzpicture} \caption{Clusters of eigenvalues on the real axis.} \label{fi:clusters} \end{figure} Denoting by $\|\cdot\|$ the $L^2(\Omega)$ norm, the directed distances of spaces measured in the energy and $L^2$ norms are defined as follows. \begin{equation} \label{eq:Delta} \delta_a(E_k, E_{h,k}) = \max_{\substack{v \in E_k\\ \|\nabla v \|=1}} \min_{ v_h \in E_{h,k}} \|\nabla v - \nabla v_h \| ,\quad \delta_b(E_k, E_{h,k}) = \max_{\substack{v \in E_k\\ \| v \|=1}} \min_{ v_h \in E_{h,k}} \| v- v_h \|. \end{equation} For reader's convenience and for the later reference, we recall the recent error bounds from \cite{LiuVej2022}. Take $\rho$ such that $\lambda_n < \rho \leq \lambda_{N+1}$, then { \begin{align} \label{eq:Deltaest} \delta_a^2(E_k,E_{h,k}) &\leq {\frac{\rho (\hat\lambda^{(k)}_N - \lambda_n) + \lambda_n \hat\lambda^{(k)}_N \vartheta^{(k)}}{\hat\lambda^{(k)}_N(\rho - \lambda_n)}} \quad\text{and} \\ \label{eq:deltaest} \delta_b^2(E_k,E_{h,k}) &\leq {\frac{\hat\lambda^{(k)}_N - \lambda_n + \theta^{(k)}}{\rho - \lambda_n}}, \end{align} where \begin{align*} \hat\lambda^{(k)}_N &= \max_{v_h \in E_{h,k}} \frac{\norm{\nabla v_h}^2}{\norm{v_h}^2}, \quad \vartheta^{(k)} = \sum_{\ell=1}^{k-1} \left( \frac{\rho}{ \lambda_{n_\ell} } - 1 \right) \left[ \hat{\zeta}(E_{h,\ell},E_{h,k}) + \delta_a(E_\ell,E_{h,\ell}) \right]^2, \\ \theta^{(k)} &= \sum_{\ell=1}^{k-1} \left(\rho - \lambda_{n_\ell}\right) \left[ \hat{\varepsilon}(E_{h,\ell},E_{h,k}) + \delta_b(E_\ell,E_{h,\ell}) \right]^2. \end{align*} Note that quantities \begin{equation*} \hat{\zeta}(E_{h,\ell},E_{h,k}) = \max_{\substack{v\in E_{h,\ell}\\ \|\nabla v\|=1}} \max_{\substack{w\in E_{h,k}\\ \|\nabla w\|=1}} (\nabla v, \nabla w), \quad \hat{\varepsilon}(E_{h,\ell},E_{h,k}) = \max_{\substack{v\in E_{h,\ell}\\ \| v\|=1}} \max_{\substack{w\in E_{h,k}\\ \|w\|=1}} ( v , w ) \end{equation*} }% measure the non-orthogonality between spaces of approximate eigenfunctions for the previous clusters and can be easily computed by using \cite[Lemma~2]{LiuVej2022}. {Further} {note that in \cite{LiuVej2022}, the approximate eigenfunction $\{u_{h,i}\}$ are considered as arbitrary and the orthogonality of $\{u_{h,i}\}$ is not required.} \section{Projection error-based estimate in the $L^2$ norm} \label{se:optimalL2} The result of \cite[Theorem 8.1]{Boffi:2010}, and the explicitly known value of the constant in the \emph{a priori} error estimate for the energy projection \cite{LiuOis2013} enable us to mimic this approach for the eigenvalue problem and derive an optimal order convergent guaranteed and fully computable upper bound on the directed distance of the exact and approximate spaces of eigenfunctions measured in the $L^2(\Omega)$-norm. First, we mention that $u_{h,i}$ is not available in practical computation, in general, because it is a result of a generalized matrix eigenvalue solver polluted typically by rounding errors and truncation errors of iterative algorithms. In principle, {we could apply the interval arithmetic to have a rigorous representation of $u_{h,i}$,} {but such} argument would make the paper lengthy and not easy to read. Therefore, we concentrate here on a theoretical analysis of the discretization error $(u_{h,i}-u_i)$, {where $u_{h,i}$ is the exact solution of the discrete problem \eqref{eq:eig_pro_with_fem}.} For the reader's convenience, we recall several results about the \emph{a priori} error estimates for finite element solutions of the Poisson equation. These \emph{a priori} error estimates will play an important role in subsequent error bounds for eigenfunctions. Given $f\in L^2(\Omega)$, let $u\in H_0^1(\Omega)$ be the weak solution of the Poisson problem satisfying $$ (\nabla u, \nabla v) = (f,v) \quad \forall v \in H^1_0(\Omega). $$ The corresponding Galerkin approximation $u_h \in V_h(\subset H^1_0(\Omega))$ is determined by the identity $$ (\nabla u_h, \nabla v_h) = (f,v_h) \quad \forall v_h \in V_h. $$ The energy projector $P_h : H_0^1(\Omega) \rightarrow V_h$ is defined by $(\nabla (u - P_h u), \nabla v_h) = 0$ for all $v_h \in V_h$. Clearly, $u_h = P_h u$. In \cite{LiuOis2013}, Liu proposed the following constructive \emph{a priori} error estimate with a computable constant $C_h$: \begin{equation} \label{eq:Ch} \|\nabla(u-P_h u) \| \le C_h \|f\|, \quad \| u - P_h u \| \le C_h \| \nabla(u-P_h u) \| \le C_h^2 \|f\| \end{equation} and the following lower eigenvalue bounds: \begin{equation} \lambda_{k} \ge \frac{\lambda_{h,k}}{1+C_h^2 \lambda_{h,k}}~\quad (k=1, 2,\cdots, \operatorname{dim}(V_h)). \end{equation} \iffalse \cred{Do we need the first inequality? It is included in the second one. If you think it can be deleted, please, remove it. If not, keep it.} \cblue{It depends. For some non-conforming FEM space $V_h$, we can evaluate $C_h$ in $ \| u - P_h u \| \le C_h \| \nabla(u-P_h u) \|$ directly and without additional cost, compared with the hypercircle method which has to solve a side problem.} \cred{So, would you like to keep it as it is? (It is fine with me.)} \cblue{Yes.leave it unchanged. There is no pace to discuss the non-conforming case. } \fi In case of non-convex domains, the value of $C_h$ can be computed by solving a dual saddle-point problem based on the hypercircle method; see \cite[Sections 3.2--3.3]{LiuOis2013}. In case of convex domains, the value of $C_h$ can be easily computed by considering the Lagrange interpolation error constant; see \cite[Theorem~3.1]{LiuOis2013}. The specific value of $C_h$ is provided below in Section~\ref{se:numex} for the considered examples. Throughout this section, we consider an arbitrary cluster of eigenvalues $\lambda_n, \lambda_{n+1}, \dots, \lambda_N$. We denote by $\mathcal{C} = \{n, n + 1, \dots, N\}$ the set of indices of eigenvalues in this cluster and by $|\mathcal{C}| = N - n + 1$ their number. Spaces of exact and finite element eigenfunctions corresponding to this clusters are $E = \operatorname{span}\{ u_i : i \in \mathcal{C} \}$ and $E_h = \operatorname{span}\{ u_{h,i} : i \in \mathcal{C} \}$, respectively. It is also assumed that \begin{equation} \label{eq:no_overlap_of_eigs} \lambda_{h,n-1} < \lambda_n, \quad \lambda_N < \lambda_{h,N+1}. \end{equation} Such an assumption makes it possible to define the following quantities: \begin{equation*} \tau = \max_{j \in \mathcal{C}}\max_{i {\in}\mathcal{I} \setminus \mathcal{C}}\frac{\lambda_j}{|\lambda_{h,i}-\lambda_j|} ,\quad \tau_h = \max_{j \in \mathcal{C}}\max_{i {\in}\mathcal{I} \setminus \mathcal{C}}\frac{\lambda_{h,i}}{|\lambda_{h,i}-\lambda_j|}, \end{equation*} where $\mathcal{I} = \{1,2,\dots,\operatorname{dim} (V_h)\}$ stands for the set of all indices. These quantities extend the one in \cite[pages 53, 57]{Boffi:2010} and have their origin in \cite{raviart1983introduction}. The application of the quantity $\tau $ can be found in \cite[Prop.~3.1]{CarGed2011}. The result in Lemma~\ref{thm:u_and_pi_k_u} can be regarded an improvement of the one of \cite{CarGed2011}. To derive the projection error-based upper bound on the directed distance of the exact and approximate spaces of eigenfunctions measured in the $L^2(\Omega)$-norm by applying estimates \eqref{eq:Ch}, we need to bound the error of the $L^2(\Omega)$ orthogonal projection $\Pi^\mathcal{C}_h: H^1_0(\Omega) \rightarrow E_h$ by the error of the energy projection $P_h: H^1_0(\Omega) \rightarrow V_h$. To achieve this goal, we first introduce several quantities and two auxiliary lemmas. \medskip Let us introduce the unit ball $E^B := \{ u \in E : \|u\|=1 \}$. For the given cluster of eigenfunction, we introduce $\beta$ as the optimal (minimal) quantity that makes the inequality \begin{equation} \label{def:beta} \|(I-\Pi^\mathcal{C}_h) P_h u\| \le \beta \max_{\substack{v \in E^B}} \|v - P_h v\| \quad \mbox{hold for any } u\in E^B ~, \end{equation} and aim to obtain an upper bound of $\beta$. In case $\|u - P_h u\|=0$ for all $u\in E^B$, it is natural to define $\beta=0$. Given $u =\sum_{j\in \mathcal{C}} c_j u_j \in E^B$, let $\kappa:E^B \to E^B$ be the mapping such that \begin{equation} \label{eq:def-kappa-h} \kappa u = \overline{\lambda}^{-1} \sum_{j\in \mathcal{C}} {c_j \lambda_j} u_j,\quad \text{where}\quad {\overline{\lambda}^2} =\sum_{j\in \mathcal{C}}c_j^2 \lambda_j^2 . \end{equation} It is easy to see that $\kappa:E^B \to E^B$ is bijective. We set $\overline u = \kappa u$, define the relative width $\xi := (\lambda_N - \lambda_n)/\lambda_n$ of the eigenvalue cluster of interest, and note that the following estimate holds: $$ \|u-\overline{u}\|^2 = \sum_{j\in \mathcal{C} } c_j^2 (1-\lambda_j/\overline{\lambda})^2 \le \xi^2. $$ \begin{lemma}\label{thm:u_and_pi_k_u} Given an arbitrary clusters of eigenvalues, the quantity $\beta$ satisfies \begin{equation} \label{eq:est_of_beta_k} \beta \leq \tau |\mathcal{C}|^{1/2}. \end{equation} Further, if \begin{equation} \label{eq:condition_for_lemma} \tau_h\xi < 1 - |\mathcal{C}|^{-1/2} \end{equation} then \begin{equation} \label{eq:est_of_beta_k_sharper} \beta \leq \frac{\tau}{1-\tau_h\xi} ~. \end{equation} \end{lemma} \begin{remark} Note that if condition \eqref{eq:condition_for_lemma} is satisfied than the estimate \eqref{eq:est_of_beta_k_sharper} is always sharper then the bound \eqref{eq:est_of_beta_k}. Further, a smaller relative width of the cluster $\xi$ leads to a sharper upper bound \eqref{eq:est_of_beta_k_sharper}. In the extreme case of a multiple eigenvalue such that $\lambda_n=\lambda_N$, we have $\xi = 0$ and hence $\beta \le \tau$. \end{remark} \begin{proof} {First, for $u_j \in E^B$ as an eigenfunction, let us apply the standard argument (see, e.g., \cite{Boffi:2010}) to show that $\|(I - \Pi^\mathcal{C}_h) P_h u_j\| \le \tau \| (I - P_h) u_j\|$. } Note that $$ (I - \Pi^\mathcal{C}_h) P_h u_j = \sum_{i \in \mathcal{I} \setminus \mathcal{C}} (P_h u_j, u_{h,i}) u_{h,i} \in V_h, $$ which leads to \begin{equation} \label{eq:IPik} \|(I - \Pi^\mathcal{C}_h ) P_h u_j\|^2 = \sum_{i \in \mathcal{I} \setminus \mathcal{C}} (P_h u_j, u_{h,i})^2\:. \end{equation} In equality \begin{equation} \label{eq:equality_in_lemma} \lambda_{h,i} (P_h u_j , u_{h,i}) = (\nabla P_h u_j, \nabla u_{h,i}) = (\nabla u_j, \nabla u_{h,i}) = \lambda_j (u_j,u_{h,i}), \end{equation} we subtract $\lambda_j (P_h u_j,u_{h,i})$ on both sides and obtain $$ (P_h u_j, u_{h,i}) = \frac{\lambda_j}{\lambda_{h,i} -\lambda_j } (u_j - P_h u_j,u_{h,i}). $$ Summation over $i \in \mathcal{I}\setminus\mathcal{C}$ gives $$ \sum_{i \in \mathcal{I} \setminus \mathcal{C}} (P_h u_j, u_{h,i})^2 \le \tau^2 \sum_{i \in \mathcal{I} \setminus \mathcal{C}} (u_j - P_h u_j,u_{h,i})^2 \le \tau^2 \|u_j - P_h u_j\|^2, $$ where the last inequality follows form the identity $\sum_{i \in \mathcal{I}} (u_j - P_h u_j,u_{h,i})^2 = \| \Pi_h(u_j - P_h u_j) \|^2$ with $\Pi_h : H^1_0(\Omega) \rightarrow V_h$ denoting the $L^2(\Omega)$ orthogonal projector. Using this in \eqref{eq:IPik}, we finally derive \begin{equation} \label{eq:est_for_single_uj} \|(I - \Pi^\mathcal{C}_h) P_h u_j\| \le \tau \| (I - P_h) u_j\|. \end{equation} Next, we consider any $u \in E^B$ and express it in the form $u =\sum_{j\in \mathcal{C}} c_j u_j$ with $\sum_{j\in\mathcal{C}} c_j^2 = 1$. Denoting the linear operator $(I - \Pi^\mathcal{C}_h) P_h $ by $L$, the estimate \eqref{eq:est_for_single_uj} leads to $$ \|L u \|^2 = \left\| \sum_{j \in \mathcal{C}} c_j L u_j \right\|^2 \le \sum_{j \in \mathcal{C}} \|Lu_j\|^2 \le \tau^2 \sum_{j \in \mathcal{C}} \| (I - P_h) u_j\|^2. $$ Thus, we can estimate $\|(I - \Pi^\mathcal{C}_h) P_h u\|$ as \begin{equation} \label{eq:lemma_estimate_for_general_u_ver0} \|(I - \Pi^\mathcal{C}_h) P_h u \| \le \tau |\mathcal{C}|^{1/2} \max_{u \in E^B} \|u - P_h u\| \end{equation} and statement \eqref{eq:est_of_beta_k} easily follows. Finally, we consider the case when the condition \eqref{eq:condition_for_lemma} holds true. Given $u\in E^B$, we take $\overline{u}=\kappa u$ and $\overline{\lambda}$ as defined in \eqref{eq:def-kappa-h}. Using inequality \eqref{eq:equality_in_lemma}, we obtain for $u$ the identity $$ \lambda_{h,i} (P_h u , u_{h,i}) = \overline{\lambda} (\overline{u},u_{h,i}). $$ Subtracting $\overline{\lambda} (P_h \overline{u},u_{h,i})$ on both sides, we derive $$ \left( \lambda_{h,i} P_h (u - \overline{u}) + (\lambda_{h,i} -\overline{\lambda}) P_h \overline{u} , u_{h,i} \right) = \overline{\lambda} (\overline{u} -P_h \overline{u} ,u_{h,i}). $$ Thus, $$ (P_h \overline{u}, u_{h,i}) = \frac{\overline{\lambda}}{\lambda_{h,i} -\overline{\lambda} } ( \overline{u} - P_h \overline{u},u_{h,i}) - \frac{\lambda_{h,i}}{\lambda_{h,i} -\overline{\lambda} } (P_h (u - \overline{u}), u_{h,i}). $$ Since function $g(t)=t/ |\lambda_{h,i}-t|$ satisfies $g(t) \le \max(g(\lambda_n), g(\lambda_N))$ for all $t \in [\lambda_n, \lambda_N]$ and function $g_h(t)=\lambda_{h,i}/|\lambda_{h,i}-t|$ is bounded in the same way, we have $$ |(P_h \overline{u}, u_{h,i})| \leq \tau |\left(( I - P_h) \overline{u},u_{h,i}\right)| + \tau_h |(P_h (u - \overline{u}), u_{h,i})|. $$ Now, considering these inequalities for $i\in\mathcal{I}\setminus\mathcal{C}$, using the geometric inequality\footnote{ Given vectors $a=(a_1, \cdots, a_n), b=(b_1, \cdots, b_n), c=(c_1, \cdots, c_n)$ with $a_i,b_i,c_i>0$ and $a_i\leq b_i+c_i$, then their Euclidean norms satisfy $\|a\| \le \|b\|+\|c\|.$ } and the general fact that $\sum_{i\in \mathcal{I}\setminus\mathcal{C}} (\varphi,u_{h,i})^2 = \| (I - \Pi^\mathcal{C}_h) \Pi_h \varphi \|^2$ for any $\varphi \in H^1_0(\Omega)$ with $\Pi_h: H^1_0(\Omega) \rightarrow V_h$ being the $L^2(\Omega)$ orthogonal projector, we derive the bound \begin{equation} \label{eq:sub_inequality_1} \|(I-\Pi^\mathcal{C}_h)P_h {\overline{u}}\| \le \tau \|\overline{u}-P_h \overline{u}\| + \tau_h \|(I-\Pi^\mathcal{C}_h)P_h (u - \overline{u})\|. \end{equation} Since $\|u-\overline{u}\| \le \xi $, the definition of $\beta$ gives \begin{equation} \label{eq:sub_inequality_2} \|(I-\Pi^\mathcal{C}_h) P_h (u-\overline{u})\| \le \xi \beta \max_{u \in E^B} \|u - P_h u\|. \end{equation} Inequalities \eqref{eq:sub_inequality_1} and \eqref{eq:sub_inequality_2} lead to the relation \begin{equation} \label{eq:lemma_estimate_for_general_u_2} \|(I - \Pi^\mathcal{C}_h) P_h \overline{u} \| \le \left( \tau + \tau_h \xi \beta \right) \max_{u \in E^B} \|u - P_h u\|. \end{equation} Since $\overline u = \kappa u$ and $\kappa: E^B \to E^B$ is a bijection, we have $$ \max_{u\in E^B} \|(I - \Pi^\mathcal{C}_h) P_h \kappa u \| = \max_{u\in E^B} \|(I - \Pi^\mathcal{C}_h) P_h u \|. $$ Consequently, from the bound \eqref{eq:lemma_estimate_for_general_u_2} and the definition of $\beta$, we obtain $$ \beta \le \tau + \tau_h \xi \beta. $$ Since condition \eqref{eq:condition_for_lemma} implies $1-\tau_h\xi>0$, the estimate \eqref{eq:est_of_beta_k_sharper} follows. \end{proof} In next lemma, we show the relation between $\delta_b(E,E_h)$ and the projection error using the quantity $\beta$. \begin{lemma}\label{thm:u_and_pi_k_u_new_ver} For the given cluster of eigenvalues, the following estimate holds: \begin{equation} \label{eq:thm_u_and_pi_k_u_new_} \delta_b(E,E_h) = \max_{u\in E^B} \|u-\Pi^\mathcal{C}_h u\| \le (1+\beta) \max_{u \in E^B} \|u - P_h u\|. \end{equation} \end{lemma} \begin{proof} For any $u\in E^B$, since $\Pi^\mathcal{C}_h u$ provides the best approximation of $u$ under the $L^2$ norm in $E_h$, we have \begin{equation} \label{eq:lem_triangle_inequality_new_} \|u - \Pi^\mathcal{C}_h u \| \le \|u - \Pi^\mathcal{C}_h P_h u \| \le \|u - P_h u \| + \|P_h u-\Pi^\mathcal{C}_h P_h u \| . \end{equation} Using the definition of the quantity $\beta$, we easily draw the conclusion. \end{proof} \begin{remark} \label{remark:comparison_carsten_gedick} In Proposition 3.2 of \cite{CarGed2011}, with $m:=|\mathcal{C}|=N-n+1$, the following result is obtained: For $u_{h,j} $ in $E$ as an eigenfunction associated to $\lambda_{h,j}$, $$ \min_{u\in E} \|u_{h,j}-u\| \le \sqrt{2} m(2m+1)(1+\tau) \max_{u_j\in E^B} \|u_j-P_h u_j\|~. $$ If such a result is applied to $\delta_b(E,E_h)$, one can obtain the following estimate. \begin{equation} \label{eq:overestimation} \delta_b(E,E_h) \le \sqrt{2} m(2m+1)(1+\tau) \max_{u \in E^B} \|u - P_h u\|~. \end{equation} This bound is larger than the result in Lemma~\ref{thm:u_and_pi_k_u_new_ver}. Particularly, if the eigenvalue cluster is tight, i.e., $\xi\approx 0$, we have $\beta\approx \tau$ and the bound \eqref{eq:overestimation} is overestimated by the factor $\sqrt{2} m(2m+1)$. \end{remark} Bounding $\beta$ by Lemma~\ref{thm:u_and_pi_k_u}, the following theorem presents the main result. \begin{theorem}\label{th:main_new_} Let $\beta$ be the quantity defined in \eqref{def:beta}. For an arbitrary cluster of eigenvalues, the following estimate holds: \begin{equation} \label{eq:l2_norm_optimal} \delta_b(E,E_h) \le (1+\beta) {\lambda_N} C_h^2~. \end{equation} \end{theorem} \begin{proof} Given $u\in E_k$, $u = \sum_{j\in \mathcal{C}}c_j u_j$, let $\overline{u}=\kappa u$ and $\overline{\lambda}^2 = \sum_{j\in \mathcal{C}}c_j^2 \lambda_j^2$. Then the identity $$ (\nabla u, \nabla v)= (\overline{\lambda}\overline{u}, v), \quad \forall v\in V, $$ holds and the {\em a priori} error estimate \eqref{eq:Ch} with $f=\overline{\lambda}\overline{u}$ yields \begin{equation} \label{eq:local_est_1_new} \|u-P_h u\| \le C_h^2 \| \overline{\lambda}\overline{u}\| \le C_h^2\lambda_N . \end{equation} Definition of $\delta_b(E,E_h)$ and bounds \eqref{eq:thm_u_and_pi_k_u_new_} and \eqref{eq:local_est_1_new} give \begin{equation*} \label{eq:est2_new} \delta_b(E,E_h) = \max_{u\in E^B} \|u-\Pi^\mathcal{C}_h u\| \leq (1+\beta) \max_{u \in E^B} \| u - P_h u \| \le (1+\beta) C_h^2 \lambda_N. \end{equation*} \end{proof} \begin{remark} \label{remark:convergence-rate} The result \cite{LiuOis2013} shows how to compute the quantity $C_h$. For convex domains, the solution of the Poisson problem has the regularity $u\in H^2(\Omega)$ and, consequently, we have $C_h=O(h)$ via the Lagrange interpolation error estimation. For non-convex domains, the solution $u$ belongs to $H^{1+\alpha}(\Omega)$, where $\alpha \in (0,1]$ depends on the angles of re-entrant non-convex corners. In this case, the value of $C_h$ is evaluated by the hypercircle method using the Raviart--Thomas FEM, and it is expected that $C_h=O(h^{\alpha})$. As {it} is pointed out in \cite[Theorem 9.13]{Boffi:2010}, the FEM solutions approximate the eigenfunction independently. That is, even for non-convex domains, if an eigenfunction has the $H^2$-regularity, then the FEM approximation to such an eigenfunction has the $O(h^2)$ convergence rate under $L^2$ norm. Since the projection error in the estimation \eqref{eq:thm_u_and_pi_k_u_new_} is restricted to the function in $E$ for the specified eigenvalue cluster, the estimation of Lemma \ref{thm:u_and_pi_k_u_new_ver} is consistent with the theoretical analysis of \cite{Boffi:2010}. The proposed estimation \eqref{eq:l2_norm_optimal} using $C_h$ has a defect that, in case of non-convex domains, for an eigenfunction with a better regularity, the proposed bound still keeps the degenerated convergence rate, which is because the {\em a priori} error estimation is considering the worst case for the projection error. If the regularity for eigenfunction in $E$ is known, then the estimation in Theorem \ref{th:main_new_} can also be improved since the estimation only depends on the projection error for eigenfunctions in the specified cluster. For example, in the case of an L-shaped domain of \S \ref{sec:l-shape}, the eigenfunction $u=\sin(\pi x)\sin (\pi y)$ associated to $\lambda_3=2\pi^2$ has the $H^2$-regularity, thus one can take $C_h<0.493h$ (where $h$ is the largest leg length for right triangles in the triangulation) for FEM approximation using triangulation with right triangles. \end{remark} \begin{remark} Theorem 3 in \cite{LiuVej2022} provides the following estimate: \begin{equation} \label{eq:energy_by_L2} \delta_a^2(E,E_h) \leq {2 - 2 \lambda_n \left( \frac{1 - \delta_b^2(E,E_h) }{\lambda_N \lambda_{h,N}} \right)^{1/2}} ~, \end{equation} where the energy error $\delta_a(E,E_h)$ {is bounded by} the $L^2$ error $\delta_b(E,E_h)$. However, bound \eqref{eq:energy_by_L2} is not optimal for clusters of a positive widths, i.e., $\lambda_n<\lambda_{N}$. On the other hand, for clusters consisting of a simple or a multiple eigenvalue, we have $\lambda_n = \lambda_N$ and the bound \eqref{eq:energy_by_L2} has the optimal speed of convergence. Indeed, in this case, it can be easily shown that the right-hand side of \eqref{eq:energy_by_L2} is dominated by $|\lambda_{h,N} - \lambda_N|$ and the other terms, including $\delta_b^2(E,E_h)$ are of higher order. {Consequently, bound \eqref{eq:energy_by_L2} combined with \eqref{eq:l2_norm_optimal} provides a guaranteed and fully computable error bound in the energy norm with the optimal speed of convergence for a cluster consisting of only one simple or multiple eigenvalue.} \end{remark} \iffalse \begin{remark}\label{re:re2} Theorem 8.1 of \cite{Boffi:2010} proves the estimate $$ \Delta(E_{h,K}, E_K) \le C(K) \sup_{v\in E_1\cup \cdots \cup E_{K}, \|v\|=1} \|\nabla(v-P_h v)\|, $$ where we use the notation of the current paper. Although the explicit bound on $\Delta(E_{h,K}, E_K)$ is not given in \cite{Boffi:2010}, we believe that using the constant $C_h$ from \eqref{eq:Ch}, we can provide an explicit bound on $C(K)$ and $\|\nabla(v-P_h v)\|$ and, thus, an estimate for $\Delta(E_{h,K}, E_K)$. In our future work, we will derive this estimate and compare it with bounds derived in the current paper. [DO WE WANT TO PROMISE THIS?] \cred{[What is your idea?]} \LIU{[In Corollary 9.11 of \cite{Boffi:2010}, there is a stronger result. But be careful that the assumption there is that the cluster width is zero. In case of zero width of cluster, our result in \eqref{eq:energy_by_L2} already proves this property.]} \LIU{[As a conclusion, there is no direct evidence or idea that we can obtain the ``optimal" explicit estimation of $\delta_a$ for a cluster with different eigenvalues. So, the remark shall be further weakened or removed.]} \cred{[OK, please, remove this remark completely. We do not need it for the current paper at all.]} \end{remark} \fi \section{Numerical examples} \label{se:numex} This section provides numerical examples to illustrate the accuracy of proposed bounds on the directed distances of spaces of exact and approximate eigenfunctions. The first example is the Laplace eigenvalue problem \eqref{eq:modpro} in the unit square domain for which {the exact eigenvalues and eigenfunctions are well known.} The second example is the same problem considered {in a non-convex L-shaped domain where eigenfunctions may have singularities at the re-entrant corner.} Both examples are computed in the floating point arithmetic and the influence of rounding errors is not taken into account {for simplicity.} However, if needed, mathematically rigorous estimates could be obtained by employing the interval arithmetic \cite{moore2009introduction}. \subsection{The unit square domain} \label{sec:unit_square} Consider the Laplace eigenvalue problem with homogeneous Dirichlet boundary conditions in the unit square $\Omega=(0,1)^2$: find eigenvalues $\lambda_i \in \mathbb{R}$ and corresponding eigenfunctions $u_i \neq 0$ such that \begin{equation} \label{eq:modpro} -\Delta u_i = \lambda_i u_i \quad\text{in }\Omega; \qquad u_i = 0 \quad\text{on }\partial\Omega. \end{equation} \begin{table}[ht] \caption{\label{ta:square_domain_clusters}The four leading clusters for the unit square.} \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline \rule[-2mm]{0cm}{6mm}{} Cluster & 1 & 2 & 3 & 4 \\ \hline \rule[-2mm]{0cm}{6mm}{} Eigenvalues & $\lambda_1 = 2\pi^2$ & $\lambda_2 = \lambda_3 = 5\pi^2$ & $\lambda_4 = 8\pi^2$ & $\lambda_5 = \lambda_6 = 10\pi^2$ \\ \hline \end{tabular} \end{center} \end{table} The exact eigenpairs are known analytically to be $$ \lambda_{ij} = (i^2+j^2)\pi^2,\quad u_{ij}=\sin(i\pi x) \sin(j\pi y), \quad i,j=1,2,3, \dots. $$ These eigenvalues are either simple or multiple and we cluster them according to the multiplicity. The first four clusters are listed in Table~\ref{ta:square_domain_clusters}. Since the exact eigenvalues are known, we use them to evaluate error bounds. {To compute bounds \eqref{eq:Deltaest} and \eqref{eq:deltaest} for the cluster $\{\lambda_{n}, \lambda_{n+1}, \cdots, \lambda_{N}\}$, we choose $\rho = \lambda_{N+1}$. } \begin{figure}[ht] \begin{center} \includegraphics[scale=0.25]{square_uniform.eps} \end{center} \caption{\label{fig:uniform_mesh_square} The uniform mesh with $h=1/4$ for the unit square.% } \label{fi:squaremesh} \end{figure} {Problem \eqref{eq:modpro}} is discretized by the conforming finite element method using piecewise linear functions. The finite element mesh $\mathcal{T}_h$ is chosen as the uniform triangulation consisting of isosceles right triangles; {see Figure~\ref{fi:squaremesh} for an illustration. } The projection error constant can be easily obtained through the interpolation error constant {as} $C_h\le h/0.493$. { For each cluster, we compute bounds on $\delta_b(E_k,E_{h,k})$ and $\delta_a(E_k,E_{h,k})$ by the estimate \eqref{eq:l2_norm_optimal} from Theorem~\ref{th:main_new_} and its combination with the relation \eqref{eq:energy_by_L2}, respectively. We then compare these results with the bounds \eqref{eq:Deltaest} and \eqref{eq:deltaest} computed by Algorithm I of \cite{LiuVej2022}. } The convergence behavior of computed bounds for the four leading clusters is shown in Figure ~\ref{fig:unit-square-l2} and \ref{fig:unit-square-h1}. The results confirm the expected optimal convergence rate $O(h^2)$ of the estimate \eqref{eq:l2_norm_optimal} for {$\delta_b(E_k,E_{h,k})$}, and the sub-optimal rate $O(h)$ from Algorithm I of \cite{LiuVej2022}. The estimate by Algorithm II of \cite{LiuVej2022} can provide impressively sharp bounds and the optimal convergence rate for the error of approximate eigenfunctions under both $L^2$ and $H^1$ norms. Since such an approach needs more {effort to post-process the approximate eigenfunction, reconstruct the flux, and estimate the} residual error of the eigenfunction approximation, the comparison with Algorithm II of \cite{LiuVej2022} is omitted here. \begin{figure}[htp] \centering \includegraphics[width=\textwidth]{Unitsquare_L2.eps} \caption{\label{fig:unit-square-l2}Bounds on the $L^2(\Omega)$ distances of spaces of eigenfunctions {$\delta_b(E_k,E_{h,k})$} for the square domain and four leading clusters of eigenvalues {$k=1,2,3,4$.} } \end{figure} \begin{figure}[htp] \centering \includegraphics[width=\textwidth]{Unitsquare_H1.eps} \caption{\label{fig:unit-square-h1} Bounds on the energy distances of spaces of eigenfunctions {$\delta_a(E_k,E_{h,k})$} for the square domain and the four leading clusters {$k=1,2,3,4$}. } \end{figure} \subsection{The L-shaped domain} \label{sec:l-shape} We consider the Laplace eigenvalue problem \eqref{eq:modpro} in the L-shaped domain $\Omega = (-1,1)^2\setminus(-1,0]^2$ to present the standard example with singularities of eigenfunctions and also to demonstrate the versatility of the proposed method. We solve this problem by using the classical linear conforming finite element space over a uniform mesh. \begin{figure}[ht] \centering \includegraphics[scale=0.4]{L_shape_mesh.eps} \caption{L-shaped domain and the initial mesh } \label{fig:l_shaped_domain} \end{figure} Since the exact eigenvalues are not known, the eigenvalue bounds are evaluated by using two-sided bounds on eigenvalues, which were computed in \cite{liu2014high} and we list them in Table~\ref{tab:l_shaped_eig_lower_bound}. The first four eigenvalues are simple and form trivial clusters. The values of the projection error constants are obtained by applying the hypercircle method proposed in \cite{LiuOis2013}; see Table \ref{table:lshape-projection-error-constant}. \begin{table}[ht] \centering \caption{Lower bounds on the leading eigenvalues for the L-shaped domain.} \begin{tabular}{|c|c|c|c|c|} \hline \rule[-2mm]{0mm}{6mm}{} $\lambda_1$ & $\lambda_2$ &$\lambda_3$ & $\lambda_4$ & $\lambda_5$ \\ \hline \rule[-2mm]{0mm}{6mm}{} $9.6397_{1}^{3}$ & $15.1972_{5}^{6}$ & $19.7392_0^1$ & $29.5214_7^9$ & $31.9126_2^4$ \\ \hline \end{tabular} \label{tab:l_shaped_eig_lower_bound} \end{table} \begin{table}[h] \centering \caption{\label{table:lshape-projection-error-constant}Projection error constants} \begin{tabular}{|c|c|c|c|c|} \hline \rule[-2mm]{0mm}{6mm}{} $h$ & $1/32$ &$1/64$ & $1/128$ & $1/256$ \\ \hline \rule[-2mm]{0mm}{6mm}{} $C_h$ & $0.0359$ & $0.0218$ & $0.0134$ & $0.00832$ \\ \hline \end{tabular} \end{table} The initial finite element mesh is displayed in Figure~\ref{fig:l_shaped_domain}. First, we apply the bounds \eqref{eq:Deltaest} and \eqref{eq:deltaest} to the four leading eigenvalue clusters. Since the exact error $\delta_b$ cannot be evaluated directly, we apply the residual error-based estimation, i.e., Algorithm II of \cite{LiuVej2022} to obtain a sharp bound of $\delta_b$. Numerical evaluation of such a bound implies that $\delta_b$ has the convergence rate as $O(h^{3/2})$ for the first cluster and $O(h^{2})$ for the rest $3$ clusters. Figure~\ref{fig:lshape-l2} shows the bounds on the $L^2(\Omega)$ distance $\delta_b$. Figure~\ref{fig:lshape-h1} compares the bounds on the energy distance $\delta_a$. The results confirm that the newly proposed estimate of $\delta_b$ based on the projection error estimate, namely the estimate \eqref{eq:l2_norm_optimal} in Theorem~\ref{th:main_new_}, provides improved convergence rates in comparison with the bound \eqref{eq:deltaest}. \begin{figure}[htp] \centering \includegraphics[width=\textwidth]{LShape_L2_with_REE.eps} \caption{\label{fig:lshape-l2}Bounds on the $L^2(\Omega)$ distances of spaces of eigenfunctions {$\delta_b(E_k,E_{h,k})$} for the L-shaped domain and four leading clusters {$k=1,2,3,4$}. } \end{figure} \begin{figure}[htp] \centering \includegraphics[width=\textwidth]{LShape_H1_N.eps} \caption{\label{fig:lshape-h1}Bounds on the energy distances of spaces of eigenfunctions {$\delta_a(E_k,E_{h,k})$} for the L-shaped domain and the four leading clusters {$k=1,2,3,4$}. } \end{figure} \section{Conclusions} \label{se:conclusions} For finite element eigenfunctions, we derived a projection error-based bound on the $L^2$ distance $\delta_b$ by employing the explicitly known value of the constant {$C_h$} in the \emph{a priori} error estimate for the energy projection. The obtained optimal estimate of $\delta_b$ can be further utilized to improve the bound for the energy distance $\delta_a$. The derived bound is fully computable and guaranteed. \bibliographystyle{amsplain}
{ "timestamp": "2022-11-08T02:16:37", "yymm": "2211", "arxiv_id": "2211.03218", "language": "en", "url": "https://arxiv.org/abs/2211.03218", "abstract": "For conforming finite element approximations of the Laplacian eigenfunctions, a fully computable guaranteed error bound in the $L^2$ norm sense is proposed. The bound is based on the a priori error estimate for the Galerkin projection of the conforming finite element method, and has an optimal speed of convergence for the eigenfunctions with the worst regularity. The resulting error estimate bounds the distance of spaces of exact and approximate eigenfunctions and, hence, is robust even in the case of multiple and tightly clustered eigenvalues. The accuracy of the proposed bound is illustrated by numerical examples. The demonstration code is available at https://ganjin.online/xfliu/EigenfunctionEstimation4FEM .", "subjects": "Numerical Analysis (math.NA)", "title": "Projection error-based guaranteed L2 error bounds for finite element approximations of Laplace eigenfunctions", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9835969689263265, "lm_q2_score": 0.8152324915965392, "lm_q1q2_score": 0.801860207704613 }
https://arxiv.org/abs/1712.04993
On the Alexander polynomial and the signature invariant of two-bridge knots
Fox conjectured the Alexander polynomial of an alternating knot is trapezoidal, i.e. the coefficients first increase, then stabilize and finally decrease in a symmetric way. Recently, Hirasawa and Murasugi further conjectured a relation between the number of the stable coefficients in the Alexander polynomial and the signature invariant. In this paper we prove the Hirasawa-Murasugi conjecture for two-bridge knots.
\section{Introduction} A knot is said to be alternating if it admits a diagram in which the crossings alternate between over- and underpasses. In 1962, Fox posed the following conjecture concerning a curious behavior of the Alexander polynomial of an alternating knot. \begin{conj}[\cite{Fox62}]\label{Foxconj} Let $K$ be an alternating knot with the Alexander polynomial $\Delta_K(t)=\Sigma_{j=0}^{2n}(-1)^ja_jt^{2n-j}$, $a_j>0$. Then $$a_0<a_1<\cdots<a_{n-m-1}<a_{n-m}=\cdots=a_{n+m}>a_{n+m+1}>\cdots>a_{2n}.$$ \end{conj} Polynomials satisfying the above condition are called trapezoidal, so this conjecture is known as Fox's trapezoidal conjecture. This conjecture remains open.It is, however, supported by the verification on several classes of alternating knots. The case of two-bridge knots is confirmed by Hartley \cite{Har79}. More generally Murasugi proved it for alternating algebraic knots \cite{Mur85}. The case of genus two alternating knots has also been verified by Ozsv\'ath and Szab\'o using Heegaard Floer homology \cite{OS03}, and by Jong via a combinatorial method \cite{Jon09}. Recently, Hirasawa and Murasugi showed that the conjecture holds for alternating stable knots \cite{HM13}. Moreover, in this case they observed that the signature of such knots are zero, and $m=0$ in Conjecture \ref{Foxconj}. Therefore, this progress led them to pose the following strengthened conjecture. \begin{conj}[\cite{HM13}]\label{HMconjecture} Let $K$ be an alternating knot, whose signature $|\sigma(K)|=2k$ and the Alexander polynomial $\Delta_K(t)=\Sigma_{j=0}^{2n}(-1)^ja_jt^{2n-j}$, $a_j>0$. Then $$a_0<a_1<\cdots<a_{n-m-1}<a_{n-m}=\cdots=a_{n+m}>a_{n+m+1}>\cdots>a_{2n},$$ moreover, $m\leq k$. \end{conj} We provide some evidence supporting this conjecture in this paper. We first observe that the case of genus two knots can be confirmed easily by using a result of Ozsv\'ath and Szab\'o in \cite{OS03}, or by Jong's inequalities in \cite{Jon10}, which is pointed out to the author by Kunio Murasugi. \begin{thm}\label{genus two case} If $K$ is an alternating knot of genus two, then it satisfies the statement of Conjecture \ref{HMconjecture}. \end{thm} \begin{proof} Note since the trapezoidal conjecture is true for genus two alternating knots, only $m\leq k$ are left to be verified. If $|\sigma(K)|=4=2g(K)$, then Conjecture \ref{HMconjecture} is obviously true since the degree the symmetric Alexander polynomial is less than or equal to $g(K)$. If $|\sigma(K)|=2$, Corollary 1.6 of \cite{OS03} or Theorem 1.6 of \cite{Jon10} implies $a_1\geq 2a_0+1$, hence the conjecture. If $\sigma(K)=0$, Corollary 1.6 of \cite{OS03} or Theorem 1.6 of \cite{Jon10} implies $a_1\geq 2a_2$, and $\Delta_K(1)=1$ implies $a_0=1+2a_1-2a_2$, therefore $a_0>a_1>a_2$. \end{proof} Our main result confirms the Hirasawa-Murasugi conjecture for two-bridge knots. \begin{thm}\label{main} Conjecture \ref{HMconjecture} is true for two-bridge knots. \end{thm} The proof of this theorem is given in Section 3. For the strategy of the proof, we extend Hartley's induction argument in \cite{Har79}. Hartley's induction utilizes extended digrams of two-bridge knots to compute their Alexander polynomials, and for our purpose we further implement Shinohara's algorithm in the induction to compute the signature invariant \cite{Shi76}.\\ \noindent\textbf{Organization.} In Section 2 we recall the preliminaries, which includes computing the Alexander polynomial using extended diagrams and Shinohara's result on the signature of two-bridge knots. Section 3 is devoted for the induction argument: after some further technical preparation for the induction in Subsection 3.1 and Subsection 3.2, the key parts of the proof are carried out in three parallel steps in Subsection 3.3-3.5.\\ \noindent\textbf{Acknowledgment:} I thank Stephan Burton, Matt Hedden, Effie Kalfagianni and Christine Ruey Shan Lee for their interest, and Matt Hedden again for his help on improving the exposition of this work. The author is grateful to Kunio Murasugi for pointing out a mistake in an earlier version of this paper, and suggesting a correction for the proof of Theorem \ref{genus two case}. \section{Preliminaries} For the reader's convenience, we recall some preliminaries regarding two-bridge knots (and links) in this section. As it will be clear, all the two-bridge links we consider will come with a preferred orientation, so this allows us to talk about the signature of a two-bridge link without ambiguity. This section has three parts, consisting of the Schubert normal form, extended diagrams and Shinohara's method for computing the signature invariant. In particular, we shall see both the Alexander polynomial and the signature of a two-bridge link can be read off from its extended diagram. \textbf{Convention.\ }For the ease of terminology, by the term two-bridge knot we often include the case of links and this shall not cause any confusion. With this convention, Theorem \ref{main} can also be understood as: any two-bridge link with the preferred orientation specified below satisfies Conjecture 1 (see Theorem \ref{reformulation of the main theorem} for a precise reformulation). \subsection{Two-bridge knots and their Schubert normal forms} A two-bridge knot is one that admits a bridge-presentation with two overarcs and two underarcs. Every two-bridge knot can be presented in its \emph{Schubert normal form}. More concretely, given a pair of coprime numbers $(p,q)$ such that $q$ is odd and $2p>q>0$, we may construct a two-bridge knot via following procedure. Firstly we draw two overarcs, placed horizontally on the same level, on which we mark $p+1$ points equidistantly, numbered from $0$ to $p$ with $0$ at the end near the center (see Fig.\ \ref{overarcs}). Then an underarc begins by spiralling out clockwisely from one of the $0$'s, passing under the two overarcs alternatively through the mark points $q$, $2q$, ... When reaching the outside, it makes a turn with a radius within $q/2$, and then spirals counterclockwisely, again passes through the overarcs alternatively under mark points with a $q$-unit difference. This process is repeated until the underarc reaches the tail of some overarc (Fig.\ \ref{(4,3) in Schubert normal form}). The other underarc is drew symmetrically. The so obtained diagram is called the Schubert normal form of the two-bridge knot of type $(p,q)$. Throughout, we orient thus obtained knots (or links) by requiring the orientation of overarcs to be center pointing. \begin{figure}[htb] \begin{minipage}[t]{0.5\linewidth} \centering{ \fontsize{0.5cm}{2em} \resizebox{65mm}{!}{\input{coordinates.pdf_tex}} \caption{$p=4$} \label{overarcs} } \end{minipage}% \begin{minipage}[t]{0.5\linewidth} \centering{ \fontsize{0.5cm}{2em} \resizebox{65mm}{!}{\input{sample1.pdf_tex}} \caption{$(4,3)$ with one underarc} \label{(4,3) in Schubert normal form} } \end{minipage}% \end{figure} \subsection{Extended diagrams and the Alexander polynomial} Closely related to the Schubert normal form of a two-bridge knot is its \emph{extended diagram}, introduced by Hartley \cite{Har79}. The extended diagram is obtained by unwinding the Schubert normal form: instead of drawing two overarcs horizontally, we draw a number of parallel overarcs, placed vertically, each one is marked off by numbers from $0$ to $p$ from the bottom to the top (see Fig.\ \ref{verticals}). Then starting from $0$ of one overarc, we let the underarc proceed from left to right if it were going clockwisely in the Schubert normal form, and going from right ot left if it were spiralling counterclockwisely (Fig.\ \ref{extended diagram for (4,3)}). The main advantage of the extended diagram presentation is that one could read off the (reduced) Alexander polynomial of the corresponding knot directly. To be precise, denote the overarcs which are hit by the underarc by $W_i$, with $i$ goes from $0$ to some number $l$ from left to right, and let $\alpha_i$ be the number of segments joining the $W_i$ and the $W_{i+1}$. By applying Fox calculus to the knot group presentation coming from the Schubert normal form, Hartley proved \begin{thm}[\cite{Har79}] $\Delta(t)=\Sigma_{i=0}^{l-1}(-1)^i\alpha_i t^i$. \end{thm} For example, the two bridge link of type $(4,3)$ shown in Fig. \ref{extended diagram for (4,3)} has $\Delta(t)=2-2t$. More technical results regarding extended diagrams will be needed for the proof of our main theorem, however, we defer that to Section 3 for the ease of reading. \begin{figure}[htb] \begin{minipage}[t]{0.5\linewidth} \centering{ \fontsize{0.5cm}{2em} \resizebox{63mm}{!}{\input{verticals.pdf_tex}} \caption{overacrs in extended diagram} \label{verticals} } \end{minipage}% \begin{minipage}[!htb]{0.5\linewidth} \centering{ \fontsize{0.5cm}{2em} \resizebox{40mm}{!}{\input{sample2.pdf_tex}} \caption{Extended diagram for $(4,3)$: $\alpha_0=2$, $\alpha_1=2$, so the $(4,3)$-link has $\Delta(t)=2-2t$.} \label{extended diagram for (4,3)} } \end{minipage} \end{figure} \subsection{The signature of two-bridge knots} Shinohara gave a convenient way of computing the signature invariant of a two-bridge knot from its Schubert normal form \cite{Shi76}. Keep in mind that the knots (especially links) that we consider are oriented. We have \begin{thm}[\cite{Shi76}]\label{Shinohara} For a two-bridge knot $K$ of type $(p,q)$, its signature $\sigma(K)$ equals the algebraic sum of the signed crossings of one underarc with the overarcs in its Schubert normal form. \end{thm} In view of the relation between the Schubert normal form and the extended diagram, we may say $\sigma$ equals the algebraic sum of the signed crossings of an underarc with the overarcs in the extended diagram, with the overarcs oriented as downward pointing. For exmaple, the two-bridge knot of type $(4,3)$ has signature $1$ (Fig.\ \ref{(4,3) in Schubert normal form} and Fig.\ \ref{extended diagram for (4,3)}). \begin{rmk} Denote by $\sigma(p,q)$ the signature of the two-bridge knot of type $(p,q)$, one can deduct the following well-known formula from Theorem \ref{Shinohara}: $$\sigma(p,q)=\sum_{i=1}^{p-1}(-1)^{[\frac{iq}{p}]}.$$ \end{rmk} \section{Proof the main theorem} In this section we give a proof of the main theorem. Overall, the strategy is to carry out an induction on the pairs $(p,q)$, starting from $(1,1)$ using three types of moves $T_i$ that will be defined later, $i=1,2,3$. This is the approach Hartley took to prove the trapezoidal conjecture for two-bridge knots. To facilitate our proof, the first subsection recalls technical results regarding extended diagrams from \cite{Har79}. Then we move to examine the effect of each move $T_i$. Among these, the effect of the $T_1$ move is most subtle and requires a fair amount of technical care, hence the corresponding discussion will be postponed to the last subsection. \subsection{More technical preparations on extended diagrams} To carry out the induction, rather than restricting to a pair $(p,q)$ such that $2p>q>0$, Hartley introduced a bigger set consisting of the so-called admissible pairs. \begin{defn} A pair of postive integers $(p,q)$ is said to be admissible if $gcd(p,q)=1$, and $q$ is odd. \end{defn} Note given an admissible pair $(p,q)$, we can similarly associate to it an extended diagram. More concretely, first introduce grid lines which consist of infinitely many parallel vertical lines, $W_i$, placed equidistantly, with subindex ranging from $-\infty$ to $\infty$, from left to right. On each grid line, mark $p+q$ points from $-(q-1)/2$ to $p+(q-1)/2$ with higher points having higher index (See Fig.\ \ref{gridlines}). The segments between $0$ and $p$ serve as the overarcs. Secondly, denote the point labeled by $j$ on $W_i$ by $x_{ij}$, and for all $i$, join the following pair of points by pairwise disjoint simple arcs lying within the region bounded by $W_i$ and $W_{i+1}$ (See Fig.\ \ref{arcs between two grid lines}): \begin{itemize} \item $x_{ij}$ and $x_{i+1,j+q}$, where $-(q-1)/2\leq j \leq p-(q+1)/2$ \item $x_{i+1,j}$ and $x_{i+1,-j}$ (called bottom loops), $x_{i,p-j}$ and $x_{i,p+j}$ (called top loops), where $1\leq j\leq (q-1)/2$ \end{itemize} After this, the above simple arcs piece up to give infinitly many underarcs (See Fig.\ \ref{infinitly many underarcs}). Arbitrarily pick a single underarc, by which we call \textit{the principal underarc}. Reindex the overarcs if necessary, so that leftmost overarc hit by the principal underarc is $W_0$. If the rightmost overarc hit by the principal underarc is $W_l$, we call $l$ to be the \textit{length} of $(p,q)$ (See Fig.\ \ref{infinitly many underarcs}).\\ There are two important sequence associated to the extended diagram of $(p,q)$. The first one is the \textit{arc sequence} $\alpha_i$, which is the number of arcs connecting $W_i$ and $W_{i+1}$ and coincides with the coefficients of the Alexander polynomial. The second one is the so-called \textit{bottom sequence} $b_i$, which is equal to twice the number of bottom loops of the principal underarc at $W_i$, plus one if the principal underarc starts at $W_i$ (sometimes we may consider $b_k$ with $k>l$, in this case $b_k$ should be understood as $0$). For example, in Fig.\ \ref{infinitly many underarcs} where the extended diagram of $(4,3)$ is shown, we see $l=2$, $b_0=2$, $b_1=1$ and $b_2=0$. \begin{figure}[htb] \begin{minipage}[t]{0.5\linewidth} \centering{ \fontsize{1.5cm}{2em} \resizebox{45mm}{!}{\input{extendverticals.pdf_tex}} \caption{grid lines for $(4,3)$} \label{gridlines} } \end{minipage}% \begin{minipage}[!htb]{0.5\linewidth} \centering{ \fontsize{0.5cm}{2em} \resizebox{40mm}{!}{\input{sample3.pdf_tex}} \caption{arcs between two grid lines} \label{arcs between two grid lines} } \end{minipage} \end{figure} \begin{figure}[htb] \centering{ \fontsize{1cm}{2em} \resizebox{85mm}{!}{\input{sample4.pdf_tex}} \caption{Revisiting the extended diagram for $(4,3)$: the thickened (and green) underarc is the principle underarc; the length of $(4,3)$ is $2$.} \label{infinitly many underarcs} } \end{figure} Key technical results regarding $\alpha_i$ and $b_i$ are summarized below. \begin{prop}[\cite{Har79}] Let $(p,q)$, $\alpha_i$ and $b_i$ be as above. Then $$\alpha_i-\alpha_{i-1}=b_i-b_{l-i},\ 1\leq i \leq l.$$ Moreover, $\{b_i\}$ satisfies the following three so-called IH properties: \begin{itemize} \item[(IH1)] There is an interger $h$ satisfying $1\leq h \leq l$ and an integer $r\leq h$ such that $b_i=0$ when $i>h$, and $0\leq S_0< S_1<\cdots<S_r=S_{r+1}=\cdots=S_h$, where $S_{2j}=b_{h-j}$ and $S_{2j+1}=b_j$. \item[(IH2)] If $h^*\geq h$ and $2j\leq h^*$, then $b_j\geq b_{h^*-j}$. \item[(IH3)] If $0\leq i<j$ and $b_i=b_j$, then $b_i=b_k=b_j$ for all $k$ such that $i\leq k \leq j$. \end{itemize} \end{prop} Now let us introduce the $T_i$ moves that we promised in the beginning of this section. These are operations on the admissible pairs defined as: \begin{align*} &T_1:(p,q)\longmapsto (p+q,q)\\ &T_2:(p,q)\longmapsto (p,2p+q)\\ &T_3:(p,q)\longmapsto (p,2p-q), \end{align*} where $T_3$ is only defined when $p>q$, hence $T_3$ cannot be applied after $T_2$ or $T_3$, but only after $T_1$. One nice feature of the $T_i$ moves is the following. \begin{prop}[\cite{Har79}]\label{inductible lemma} Any admissible pair $(p,q)$ can be obtained from $(1,1)$ via applying a sequence of $T_i$, $i=1,2,3$. \end{prop} In fact, the IH properties are proved inductively using these $T_i$ moves, and they imply the trapezoidal conjecture for two-bridge knots. For our purpose, the following facts will be important. \begin{prop}[\cite{Har79}]\label{change of bottom sequence} Given an admissible pair $(p,q)$, let $l$, $b_i$ and $\alpha_i$ denote the length, bottom sequence and the arc sequence respectively. Then the length and bottom sequence of $T_i(p,q)$, $i=1,2,3$ are summarized in the following table \begin{center} \begin{tabular}{|c|c|c|} \hline & length $l'$ & bottom sequence $b_i'$ \\ \hline $T_1(p,q)$ & $l+1$ & $b_i$ \\ \hline $T_2(p,q)$ & $l$ & $2\alpha_i+b_i$ \\ \hline $T_3(p,q)$ & $l$ & $2\alpha_i-b_i$ \\ \hline \end{tabular} \end{center} \end{prop} \subsection{Reformulation of the main theorem} Notice that an admissible pair $(p,q)$ may give rise to a two-component link. Therefore during this process of induction, both the degree of the Alexander polynomial and the signature may change their parity, so we would like to adjust the statement of the Hirasawa-Murasugi conjecture to take care of this issue. \begin{thm}\label{reformulation of the main theorem} Let $(p,q)$ be an admissible pair, $\sigma$ be its signature, and $\Delta_K(t)=a_0-a_1t+\cdots+(-1)^{l-1}a_{l-1}t^{l-1}$ be its Alexander polynomial, where $a_i>0$, $i=0,...,l-1$. Then \begin{equation}\label{Fox inequality} a_0<a_1<\cdots<a_{i_0-1}=a_{i_0}=\cdots=a_{l-i_0}>a_{l-i_0+1}>\cdots>a_{l-1}. \end{equation} Moreover, \begin{equation}\label{HM inequality} \lfloor \frac{|\sigma|+1}{2}\rfloor \geq \lfloor \frac{l-2(i_0-1)}{2}\rfloor. \end{equation} \end{thm} Here $\lfloor \cdot \rfloor$ is understood as taking the maximal integer part. It is obvious that the above theorem implies Theorem \ref{main}. \begin{proof} Note Inequality (\ref{Fox inequality}) is already proved by Hartley, so in the rest of this section, we will focus on proving Inequality (\ref{HM inequality}). To do that, we begin by noticing it is obviously true for the pair $(1,1)$. In view of Prop.\ \ref{inductible lemma}, we just need to see that if a pair $(p,q)$ satisfies Theorem \ref{reformulation of the main theorem}, so is $T_i(p,q)$ for $i=1,2,3$. This is done in subsections 3.3-3.5. \end{proof} \begin{rmk} From now on, we call $\lfloor \frac{l-2(i_0-1)}{2}\rfloor$ the radius of the stable terms of the Alexander polynomial. Note that the case in which all coefficients are distinct could happen, and in that case, $i_0-1=l-i_0$, which implies the radius is zero. \end{rmk} \subsection{The effect of $T_2$ move} This subsection is devoted to proving the following statement. \begin{prop} If Theorem \ref{reformulation of the main theorem} is true for an admissible pair $(p,q)$, then it is true for $T_2(p,q)$. \end{prop} \begin{proof} Let $\alpha_i$, $b_i$, $l$ denote the number of connecting arcs, bottom sequence, and length for $(p,q)$, and let $\alpha_i'$, $b_i'$, $l'$ be the corresponding quantities for $T_2(p,q)=(p,2p+q)$. By Prop.\ \ref{change of bottom sequence} $b_i'=2\alpha_i+b_{l-i}$ and $l'=l$, hence we have $$\begin{aligned} \alpha_i'-\alpha_{i-1}' &= b_i'-b_{l'-i}'\\ &=(2\alpha_i+b_{l-i})-(2\alpha_{l-i}+b_{i})\\ &=2(\alpha_i-\alpha_{l-i})-(b_i-b_{l-i})\\ &=2(\alpha_i-\alpha_{i-1})-(b_i-b_{l-i})\\ &=2(\alpha_i-\alpha_{i-1})-(\alpha_i-\alpha_{i-1})\\ &=\alpha_i-\alpha_{i-1}, \end{aligned}$$ where in the $4th$ equality we used $\alpha_{l-i}=\alpha_{i-1}$ due to the symmetry of the Alexander polynomial. So the radius $m=\lfloor \frac{l-2(i_0-1)}{2}\rfloor$ does not change after the $T_2$ move. The signature invariant is also unchanged after the $T_2$ move. To see this, note $$\sigma(p,q)=\sum_{i=1}^{p-1}(-1)^{\lfloor \frac{iq}{p}\rfloor }=\sum_{i=1}^{p-1}(-1)^{\lfloor \frac{i(q+2p)}{p} \rfloor}=\sigma(p,2p+q).$$ Therefore, Theorem \ref{reformulation of the main theorem} is true for $T_2(p,q)$ provided it is true for $(p,q)$. \end{proof} \subsection{The effect of $T_3$ move} In this subsection we examine the effect of $T_3$. \begin{prop} If Theorem \ref{reformulation of the main theorem} is true for an admissible pair $(p,q)$, then it is true for $T_3(p,q)$. \end{prop} \begin{proof} Let $\alpha_i$, $b_i$, $l$ denote the number of connecting arcs, bottom sequence, and length for $(p,q)$, and let $\alpha_i'$, $b_i'$, $l'$ be the corresponding quantities for $T_3(p,q)=(p,2p-q)$. In this case, we have $b_i'=2\alpha_i-b_i$ and $l'=l$ by Prop.\ \ref{change of bottom sequence}. Therefore, $$\begin{aligned} \alpha_i'-\alpha_{i-1}' &= b_i'-b_{l'-i}'\\ &=(2\alpha_i-b_{i})-(2\alpha_{l-i}-b_{l-i})\\ &=2(\alpha_i-\alpha_{l-i})-(b_i-b_{l-i})\\ &=2(\alpha_i-\alpha_{i-1})-(b_i-b_{l-i})\\ &=2(\alpha_i-\alpha_{i-1})-(\alpha_i-\alpha_{i-1})\\ &=\alpha_i-\alpha_{i-1}\\ \end{aligned}$$ So the radius $m=\lfloor \frac{l-2(i_0-1)}{2}\rfloor$ does not change after the $T_3$ move. For the signature, we have $$\sigma(p,2p-q)=\Sigma_{i=1}^{p-1}(-1)^{\lfloor\frac{i(2p-q)}{p}\rfloor}=\Sigma_{i=1}^{p-1}(-1)^{\lfloor\frac{i(-q)}{p}\rfloor}=-\Sigma_{i=1}^{p-1}(-1)^{\lfloor\frac{iq}{p}\rfloor}=-\sigma(p,q).$$ Therefore, neither does $|\sigma|$ change after the $T_3$ move. Hence Theorem \ref{reformulation of the main theorem} is true for $T_3(p,q)$ provided it is true for $(p,q)$. \end{proof} \subsection{The effect of $T_1$ move} In this subsection we will discuss the effect of $T_1$. Note on the level of knots, $T_2$ preserves the knot, and $T_3$ changes the knot to its mirror, and that is the reason these two cases are relatively easier compared to case of $T_1$. The goal of this subsection is to prove \begin{prop}\label{T_1 is good} If Theorem \ref{reformulation of the main theorem} is true for an admissible pair $(p,q)$, then it is true for $T_1(p,q)$. \end{prop} The proof of this proposition will come at the end of this subsection, after investigating the effect of $T_1$ on the signature and the Alexander polynomial. First of all, we present the effect of $T_1$ on the signature. \begin{lem}\label{effect of T_1 on signature} $\sigma(T_1(p,q))-\sigma(p,q)=1$. \end{lem} \begin{figure} \begin{minipage}[t]{0.5\linewidth} \centering{ \fontsize{1cm}{2em} \resizebox{40mm}{!}{\input{beforeT1.pdf_tex}} \caption{Before $T_1$} \label{BeforeT1} } \end{minipage}% \begin{minipage}[t]{0.5\linewidth} \centering{ \fontsize{1cm}{2em} \resizebox{50mm}{!}{\input{afterT1.pdf_tex}} \caption{After $T_1$} \label{AfterT1} } \end{minipage} \end{figure} \begin{figure} \centering{ \fontsize{1cm}{2em} \resizebox{85mm}{!}{\input{signatureT1.pdf_tex}} \caption{The effect of $T_1$ on signature} \label{signatureT1} } \end{figure} \begin{proof} The effect of $T_1$ on the extended diagram is shown in Fig.\ \ref{BeforeT1} and Fig.\ \ref{AfterT1}. We describe the effect as sliding the bottom end of the parallel overarcs one unit to the right. First compare the new overarcs and the old ones pair by pair who share the same top, starting from right to left. We see that the crossings between an old overarc and the principle underarc has counterparts in the crossings between the new overarc and the principle underarc (the crossing which are circled in Fig.\ \ref{signatureT1}). Secondly, the presence of each bottom circle in the principle underarc creates two new crosing with the new overarcs (the crossings which are boxed in Fig.\ \ref{signatureT1}), yet these two crossings cancel each other algebraically. Finally, sliding the overarc on which the principle underarc starts creates a new crossing (the crossing marked by a triangle in Fig.\ \ref{signatureT1}), and this crossing has positive sign. The conclusion then follows from comparing the sum of the signed crossings in view of Theorem \ref{Shinohara}. \end{proof} Next we examine how the radius of the stable terms behave under the $T_1$ operation. We separate the discussion into two cases. First we have \begin{prop}\label{radius change if there were no stable term} Given an admissible pair $(p,q)$, if there are no stable terms in the corresponding Alexander polynomial, then there are exactly two stable terms in the Alexander polynomial corresponding to $T_1(p,q)$. \end{prop} \begin{proof} In this case, in view of symmetry of the coefficients, the degree of the Alexander polynomial must be even, and hence $l=l(p,q)$ is odd. Let $l=2k+1$ and the coefficients of the Alexander polynomial for $(p,q)$ be $\alpha_0$,...,$\alpha_{k-1}$, $\alpha_{k}$,$\alpha_{k+1}$,...,$\alpha_{2k}$. After the $T_1$ move, denote the coefficients by $\alpha_0'$,...,$\alpha_{k}'$,$\alpha_{k+1}'$,...,$\alpha_{2k+1}'$. We have \begin{displaymath} \alpha_{k+1}'-\alpha_{k}'=b_{k+1}'-b_{(2k+2)-(k+1)}'=b_{k+1}-b_{k+1}=0, \end{displaymath} and \begin{displaymath} \alpha_{k}'-\alpha_{k-1}'=b_{k}'-b_{(2k+2)-(k)}'=b_{k}-b_{k+2}\geq b_{k}-b_{k+1}=\alpha_{k}-\alpha_{k-1}>0 \end{displaymath} where we used $b_{k+1}\geq b_{k+2}$ due to the second IH property. So the only stable terms are $\alpha_{k+1}'$ and $\alpha_{k}'$. \end{proof} Secondly, when there were stable terms in the Alexander polynomial before we apply $T_1$, we have the following proposition. \begin{prop}\label{radius change when there are stable terms} Let $\alpha_i$ and $\alpha_i'$ be the coefficients of the Alexander polynomial corresponding to $(p,q)$ and $T_1(p,q)$ respectively. Denote by $l$ the length of $(p,q)$. Let $i_0$, $i_0'$ be integers such that $$\alpha_0<\alpha_1<\cdots<\alpha_{i_0-1}=\alpha_{i_0}=\cdots=\alpha_{l-i_0}>\alpha_{l-i_0+1}>\cdots>\alpha_{l-1},$$ and $$\alpha_0'<\alpha_1'<\cdots<\alpha_{i_0'-1}'=\alpha_{i_0'}'=\cdots=\alpha_{l+1-i_0'}'>\alpha_{l-i_0'+2}'>\cdots>\alpha_{l}'.$$ If $i_0-1<l-i_0$, then one of the following statement is true: \begin{enumerate} \item $i_0'=i_0+1$\\ \item $i_0'=i_0$ and $b_{i_0}=b_{i_0+1}=\cdots =b_l=0$. \end{enumerate} \end{prop} \begin{proof} Note that $\alpha_{i_0}-\alpha_{i_0-1}=b_{i_0}-b_{l-i_0}=0$ and hence $b_{i_0}=b_{i_0+1}=\cdots=b_{l-i_0}$ by the third IH property. If $l-i_0>i_0$, then $\alpha_{i_0+1}'-\alpha_{i_0}'=b'_{i_0+1}-b'_{l'-(i_0+1)}=b_{i_0+1}-b_{l-i_0}=0$. Therefore, $i_0'\leq i_0+1$. If $l-i_0=i_0$, then $i_0'\leq i_0+1$ by considering the degree and symmetry of the Alexander polynomial. We move to see $i_0'\geq i_0$. If $i_0'< i_0+1$, we have $\alpha_{i_0}'-\alpha_{i_0-1}'=b_{i_0}-b_{l-i_0+1}=0$, then by third IH property, $b_{i_0}=b_{i_0+1}=\cdots=b_{l-i_0+1}$. We continue the discussion in two cases. \textbf{(Case 1) }If there is $\alpha_{i_0-2}$ term, i.e.\ $i_0\geq 2$, then since $b_{i_0-1}-b_{l-i_0+1}=\alpha_{i_0-1}-\alpha_{i_0-2}>0$, we learn that $b_{i_0-1}>b_{i_0}=b_{i_0+1}=\cdots=b_{l-i_0+1}$. Then $\alpha_{i_0-1}'-\alpha_{i_0-2}'=b_{i_0-1}-b_{l-i_0+2}\geq b_{i_0-1}-b_{l-i_0+1}>0$. Here $b_{l-i_0+1}\geq b_{l-i_0+2}$ follows from the second IH property. So in this case $i_0'=i_0$. Now let $h$ be the integer in the first IH property. If $h<l-i_0+1$, then by the first IH property, $b_{l-i_0+1}=\cdots=b_l=0$ and therefore $b_{i_0}=b_{i_0+1}=\cdots=b_l=0$. We claim $h$ cannot be greater than or equal to $l-i_0+1$. If not, $h\geq l-i_0+1$, then for some $j$ we have $S_j=b_{l-i_0+1}$, then $S_{j+1}$ must be $b_k$ for some $k\leq i_0-1$; otherwise, we have $b_{i_0-1}\leq b_{l-i_0+1}$ in view of first IH property, but this contradicts our earlier observation that $b_{i_0-1}>b_{l-i_0+1}$. This understood, we further oberserve that since $S_{j+2}=b_{l-i_0}=b_{l-i_0+1}=S_j$ (the existence of $S_{j+2}$ follows from the assumption $i_0-1<l-i_0$), so we have $S_j=S_{j+1}$ by the first IH property. Therefore $b_k=b_{l-i_0+1}$, and since $k\leq i_0-1\leq l-i_0+1$ we have $b_{i_0-1}=b_{l-i_0+1}$ by the third IH property. However, this contradicts $b_{i_0-1}>b_{l-i_0+1}$, so $h$ cannot be greater than or equal to $l-i_0+1$. In summary, we have $i_0'=i_0$ and $b_{i_0}=b_{i_0+1}=\cdots=b_{l}=0$. \textbf{(Case 2) }If there is no $\alpha_{i_0-2}$ term, i.e.\ $i_0=1$, then $i_0'=1$ and $\alpha_0=\alpha_1=\cdots=\alpha_{l-1}$. Since $0=\alpha_1-\alpha_0=b_1-b_{l-1}$, by the third IH property, we have $b_1=b_2=\cdots=b_{l-1}$. Moreover, $\alpha_1'-\alpha_0'=b_1-b_l=0$ implies $b_1=\cdots=b_l$. If $b_l$ is not zero, then the first IH property implies $b_l\leq b_0\leq b_{l-1}=b_l$, and hence all the $b_i's$ are equal by the third IH property, but this is impossible since by definition one and only one of the $b_i$'s is odd and all the others are even. Therefore, we must have $b_1=\cdots=b_l=0$, with $b_0$ being the only nonzero term. In summary, after $T_1$, the starting index of the stable terms either stays or increases by 1, and moreover, when the starting index stays, more than half of the $b_i$ sequence are zero. \end{proof} To prove Prop.\ \ref{T_1 is good}, we need further control of the signature in the case when the starting index of the stable terms stays level. This is what the following proposition addresses. \begin{prop}\label{signature is positive when half of b_i vanish} Let $(p,q)$ be an admissible pair with $\{b_i\}$ such that \begin{displaymath} b_0, b_1,..., b_{i_0-1}>0, \end{displaymath} \begin{displaymath} b_{i_0}=\cdots =b_l=0 \end{displaymath} where $l=l(p,q)$ and $i_0\leq \lfloor \frac{l}{2}\rfloor$. Then $\sigma(p,q)\geq 0$. \end{prop} To prove the proposition, we need four lemmas. \begin{lem}\label{signature is less than the bredth of the alexander poly} For any admissible pair $(p,q)$, let $\sigma=\sigma(p,q)$ and $l=l(p,q)$, then $|\sigma|\leq l-1$. \end{lem} \begin{proof} Note that this statement is true for $(1,1)$. Note $T_2$ and $T_3$ does not change $|\sigma|$ or $l$, while $T_1$ increase $l$ by one and increase $|\sigma|$ at most by one. So inductively, we can show the statement is true for all admissible pairs in view of Prop.\ \ref{inductible lemma}. \end{proof} \begin{lem}\label{$T_2(p,q)$ has no zero terms in its bottom sequence} $T_2(p,q)$ has no zero terms in its bottom sequence. \end{lem} \begin{proof} Let ${b_i}$ stands for the bottom sequence for $(p,q)$, and ${b_i'}$ for $T_2(p,q)$, $l=l(p,q)=l'(T_2(p,q))=l'$. Then $b_i'=2\alpha_i+b_{l-i}\geq 2\alpha_i>0$ when $i\leq l-1$ and $b_l'=2\alpha_l+b_0=b_0>0$. \end{proof} \begin{lem}\label{effect of T3T1} Let $\{b_i'\}_{0 \leq i\leq l+1}$ be the bottom sequence for $T_3\circ T_1(p,q)$. Then the only zero term is $b_{l+1}'$. \end{lem} \begin{figure} \centering{ \fontsize{0.5cm}{2em} \resizebox{10cm}{!}{\input{T3T1.pdf_tex}} \caption{} \label{T3T1} } \end{figure} \begin{proof} Let $l=l(p,q)$, and for $T_1(p,q)$, we denote by \begin{displaymath} b_0, b_1, ..., b_l, 0 \end{displaymath} \begin{displaymath} \alpha_0, \alpha_1,...,\alpha_{l} \end{displaymath} the bottom sequence and connecting arc sequence. So $b'_{l+1}=2\alpha_{l+1}-b_{l+1}=0-0=0$. We claim that $b'_l=2\alpha_l-b_l>0$, hence by the first IH property we know that $b_{l+1}'$ is the only zero term. To see $b'_l>0$, note $T_1(p,q)=(p+q,q)=(p',q')$, so $p'>q'>0$. First, if $b_l$ is zero or odd, then we are done since $2\alpha_l$ will never be zero or odd. Second, if $b_l$ is positive and even, see Fig.\ref{T3T1}: the lower arc joining a bottom loop of $(p',q')$ at $W_l$ must hit $W_{l+1}$ since $p'>q'$ (see Fig.\ \ref{T3T1}.(b)); the upper arc joining the bottom loop cannot turn over $W_l$ (See Fig.\ \ref{T3T1}.(c)), for that would imply $p'\leq \lfloor q'/2\rfloor+\lfloor q'/2\rfloor\leq q'$; therefore, what left are three possiblities (see the second row of Fig.\ \ref{T3T1}) and in all three cases, the existence of one bottom circle gives rise to at least two connecting arcs between $W_l$ and $W_{l+1}$, hence $\alpha_l\geq b_l$, which implies $b'_l>0$. \end{proof} \begin{lem} Starting from $(1,1)$, to obtain an admissible pair $(p,q)$ with ${b_i}$ such that \begin{displaymath} b_0, b_1,..., b_{i_0-1}>0, \end{displaymath} \begin{displaymath} b_{i_0}=\cdots =b_l=0 \end{displaymath} and $i_0\leq \lfloor \frac{l}{2} \rfloor$, then there must be at least $(l-i_0)$ $T_1$'s successively in the end, i.e.\ $(p,q)=T_1^{l-i_0}(p',q')$ for some $(p',q')$. \end{lem} \begin{proof} Note that there are $l-i_0+1$ zero terms in the bottom sequence of $(p,q)$. The operator cannot end with $T_2$ in view of Lemma \ref{$T_2(p,q)$ has no zero terms in its bottom sequence}. If the operator end with $T_3$, then since by definition $T_3$ cannot be applied successively or after $T_2$, nor can it be applied to $(1,1)$ directly, there must be a $T_1$ before it and hence $l\geq 2$. In view of Lemma \ref{effect of T3T1}, the pair we get will have only one zero term in its bottom sequence, not satisfying the assumption that more than half of the $b_i$'s are zero. So $(p,q)=T_1^k(p',q')$ for some $k$. If there is a $T_2$ before $T_1^k$, then $k=(l-i_0+1)$. If there is $T_3$ before $T_1^k$ or $(p',q')=(1,1)$, then there is already a zero in the bottom sequence and hence $k=l-i_0$. \end{proof} Now we are ready to give a proof to Prop.\ \ref{signature is positive when half of b_i vanish}. \begin{proof}[Proof of Prop.\ \ref{signature is positive when half of b_i vanish}] By the lemma above, $(p,q)=T_1^{l-i_0}(p',q')$. Let $l=l(p,q)$, $l'=l'(p',q')$. Then $l'=l-(l-i_0)=i_0$ and $|\sigma'(p',q')|<i_0$ in view of Lemma \ref{signature is less than the bredth of the alexander poly}. So by Prop.\ \ref{change of bottom sequence} we have $\sigma(p,q)=\sigma'(p',q')+(l-i_0)\geq l-2i_0\geq 0$. \end{proof} Finally, with all these preparations we are ready to prove Prop.\ \ref{T_1 is good}, hence concluding the proof of the main theorem. \begin{proof}[Proof of Prop.\ \ref{T_1 is good}] Recall we want to prove Theorem \ref{reformulation of the main theorem} is true for $T_1(p,q)$ provided it is true for $(p,q)$. Let $l=l(p,q)$, $l'=l(T_1(p,q))=l+1$. We separate the discussion into two cases. \textbf{(Case 1) }If there are no stable terms in the Alexander polynomial corresponding to $(p,q)$, then by Prop.\ \ref{radius change if there were no stable term}, there are exactly two stable terms in the Alexander polynomial corresponding to $T_1(p,q)$, and hence the radius of stable terms is $1$. Recall in the proof of Prop.\ \ref{radius change if there were no stable term} we observed $l$ is odd in this case, which implies the length $l'$ for $T_1(p,q)$ is even. This in turn implies $T_1(p,q)$ corresponds to a two-component link, so $|\sigma(T_1(p,q))|\geq 1$ since it must be odd. Hence $\lfloor \frac{|\sigma|+1}{2} \rfloor \geq 1$. \textbf{(Case 2) }If there are some stable terms in the Alexander polynomial corresponding to $(p,q)$. Let $i_0$, $i_0'$ be as in Prop.\ \ref{radius change when there are stable terms}. When $l=2n+1$, $|\sigma|=2k$. By assumption $k\geq \lfloor\frac{l-2(i_0-1)}{2}\rfloor=n-i_0+1$. If $i_0'=i_0+1$, then $\lfloor \frac{l'-2(i_0'-1)}{2}\rfloor=n-i_0+1\leq k$, and $\lfloor \frac{|\sigma'|+1}{2}\rfloor \geq \lfloor \frac{(2k-1)+1}{2}\rfloor=k$. If $i_0'=i_0$, then $\lfloor \frac{l'-2(i_0'-1)}{2}\rfloor=n-i_0+2\leq k+1$, and by Prop.\ \ref{radius change when there are stable terms}, Prop.\ \ref{signature is positive when half of b_i vanish} and Lemma \ref{effect of T_1 on signature}, $\lfloor \frac{|\sigma'|+1}{2}\rfloor \geq k+1$. When $l=2n$, the argument is similar and hence omitted. \end{proof} \bibliographystyle{abbrv
{ "timestamp": "2018-01-12T02:01:07", "yymm": "1712", "arxiv_id": "1712.04993", "language": "en", "url": "https://arxiv.org/abs/1712.04993", "abstract": "Fox conjectured the Alexander polynomial of an alternating knot is trapezoidal, i.e. the coefficients first increase, then stabilize and finally decrease in a symmetric way. Recently, Hirasawa and Murasugi further conjectured a relation between the number of the stable coefficients in the Alexander polynomial and the signature invariant. In this paper we prove the Hirasawa-Murasugi conjecture for two-bridge knots.", "subjects": "Geometric Topology (math.GT)", "title": "On the Alexander polynomial and the signature invariant of two-bridge knots", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9835969670030071, "lm_q2_score": 0.8152324871074608, "lm_q1q2_score": 0.8018602017212165 }
https://arxiv.org/abs/1705.04271
Lifting in Besov Spaces
Let $\Omega$ be a smooth bounded domain in $\mathbb R^n$ and u be a measurable function on $\Omega$ such that $|u(x)|=1$ almost everywhere in $\Omega$. Assume that u belongs to the $B^s_{p,q}(\Omega)$ Besov space. We investigate whether there exists a real-valued function $\varphi \in B^s_{p,q}$ such that $u=e^{i\varphi}$. This extends the corresponding study in Sobolev spaces due to Bourgain, Brezis and the first author. The analysis of this lifting problem leads us to prove some interesting new properties of Besov spaces, in particular a non restriction property when $q>p$.
\section*{Big picture} \section{Introduction} ${}$ Let $\Omega\subset {\mathbb R}^n$ be a bounded simply connected domain and $u:\Omega\rightarrow {\mathbb S}^1$ a continuous (resp. $C^k$, $k\geq 1$) function. It is a well-known fact that there exists a continuous (resp. $C^k$) real-valued function $\varphi$ such that $u=e^{\imath\varphi}$. In other words, $u$ has a continuous (resp. $C^k$) lifting. \par \noindent The analogous problem when $u$ belongs to the fractional Sobolev space $W^{s,p}$, $s>0$, $1\le p<\infty$, received an complete answer in \cite{lss}. Let us briefly recall the results: \begin{enumerate} \item when $n=1$, $u$ has a lifting in $W^{s,p}$ for all $s>0$ and all $p\in [1,\infty)$, \item when $n\geq 2$ and $0<s<1$, $u$ has a lifting in $W^{s,p}$ if and only if $sp<1$ or $sp\geq n$, \item when $n\ge 2$ and $s\geq 1$, $u$ has a lifting in $W^{s,p}$ if and only if $sp\geq 2$. \end{enumerate} Further developments in the Sobolev context can be found in \cite{bethuelchiron,nguyenphase,mironescuphase,mironescucras2}. In the present paper, we address the corresponding question in the framework of Besov spaces. More specifically, given $s, p, q$ in suitable ranges defined later, we ask whether a map $u\in B^s_{p,q}(\Omega ; {\mathbb S}^1)$ can be lifted as $u=e^{\imath\varphi}$, with $\varphi\in B^s_{p,q}(\Omega ; \mathbb R)$. We say that $B^s_{p,q}$ has the lifting property if and only if the answer is positive. When dealing with circle-valued functions and their phases, it is natural to consider only maps in $L^1_{loc}$. This is why we assume that $s>0$,\footnote{ However, we will discuss an appropriate version of the lifting problem when $s\le 0$; see Remark \ref{aa1} and Case \ref{T} below. } and we take the exponents $p$ and $q$ in the classical range $p\in [1,\infty)$, $q\in [1,\infty]$.\footnote{ We discard the uninteresting case where $p=\infty$. In that case, maps in $B^s_{\infty,q}$ are continuous. Arguing as in Case \ref{tri} below, we obtain the existence of a $B^s_{\infty,q}$ phase for every $u\in B^s_{\infty,q}(\Omega ; {\mathbb S}^1)$.} Since Besov spaces are microscopic modifications of Sobolev (or Slobodeskii) spaces, one expects a global picture similar to the one described before for Sobolev spaces. The analysis in Besov spaces is indeed partly similar to the one in Sobolev spaces, as far as the results and the techniques are concerned. However, several difficulties occur and some cases still remain open. Thus, the analysis of the lifting problem leads us to prove several new properties for Besov spaces (in connection with restriction or absence of restriction properties, sums of integer valued functions which are constant, products of functions in Besov spaces, disintegration properties for the Jacobian), which are interesting in their own right. We also provide detailed arguments for classical properties (some embeddings, Poincar\'e inequalities) which could not be precisely located in the literature. \medskip Let us now describe more precisely our results and methods. When $sp>n$, functions in $B^s_{p,q}$ are continuous, which readily implies that $B^s_{p,q}$ has the lifting property (Case \ref{tri}). \medskip In the case where $sp<1$, we rely on a characterization of $B^s_{p,q}$ in terms of the Haar basis \cite[Th\'eor\`eme 5]{bourdaud}, to prove that $B^s_{p,q}$ has the lifting property (Case \ref{A}). \medskip Assume now that $0<s\leq 1$, $sp=n$ and $q<\infty$. Let $u\in B^s_{p,q}(\Omega ; {\mathbb S}^1)$ and let $F(x,\varepsilon):=u\ast\rho_{\varepsilon}$, where $\rho$ is a standard mollifier. Since $B^s_{p,q}\hookrightarrow \text{\rm VMO}$, for all $\varepsilon$ sufficiently small and all $x\in \Omega$ we have $1/2<\left\vert F(x,\varepsilon)\right\vert\le 1$. Writing $F(x,\varepsilon)/\left\vert F(x,\varepsilon)\right\vert=e^{\imath\psi_\varepsilon}$, where $\psi_\varepsilon$ is $C^{\infty}$, and relying on a slight modification of the trace theory for weighted Sobolev spaces developed in \cite{tracesoldnew}, we conclude, letting $\varepsilon$tend to $0$, that $u=e^{\imath\psi_0}$, where $\psi_0=\lim_{\varepsilon\to 0}\psi_\varepsilon\in B^s_{p,q}$, and therefore $B^s_{p,q}$ still has the lifting property (Case \ref{X}). \medskip Consider now the case where $s>1$ and $sp\geq 2$. Arguing as in \cite[Section 3]{lss}, it is easily seen that the lifting property for $B^s_{p,q}$ will follow from the following property: given $u\in B^s_{p,q}(\Omega; {\mathbb S}^1)$, if $F:=u\wedge\nabla u\in L^p(\Omega ; {\mathbb R}^n)$, then $(*)$ $\operatorname{curl} F=0$. The proof of $(*)$ is much more involved than the corresponding one for $W^{s,p}$ spaces \cite[Section 3]{lss}. It relies on a disintegration argument for the Jacobians, more generally applicable in $W^{1/p,p}$. This argument, in turn, relies on the fact that $\operatorname{curl} F=0$ when $u\in \text{\rm VMO}$ and $n=2$, and a slicing argument. In particular, we need a {\it restriction property for Besov spaces}, namely the fact that, for $s>0$, $1\le p<\infty$ and $1\le q\le p$, for all $f\in B^s_{p,q}$, the partial maps of $f$ still belong to $B^s_{p,p}$ (see Lemma \ref{oa1} below). Thus, we obtain that, when $s>1$ and $1\leq p<\infty$, $B^s_{p,q}$ does have the lifting property when [$1\le q<\infty$, $n=2$, and $sp=2$], or [$1\le q\le p$, $n\ge 3$, and $sp= 2$], or [$1\le q\le \infty$, $n\ge 2$, and $sp> 2$]. One can improve the conclusion of Lemma \ref{oa1} as follows. For $s>0$, $1\le p<\infty$ and $1\le q\le p$, for all $f\in B^s_{p,q}$, the partial maps of $f$ belong to $B^s_{p,q}$ (Proposition \ref{qh1}). We emphasize the fact that this type of property holds only under the crucial assumption $q\le p$. More precisely, if $q>p$ and $s>0$, then we exhibit a compactly supported function $f\in B^s_{p,q}({\mathbb R}^2)$ such that, for almost every $x\in (0,1)$, $f(x,\cdot)\notin B^s_{p,\infty}({\mathbb R})$ (Proposition \ref{l7.26}). This phenomenon, which has not been noticed before, shows a picture strikingly different from the one for $W^{s,p}$, and even more generally for Triebel-Lizorkin spaces \cite[Section 2.5.13]{triebel2}. \medskip Let us return to the case when $0<s<1$, $1\le p< \infty$ and $n\geq 2$. Assume now that [$1\le q<\infty$ and $1\leq sp <n$], or [$q=\infty$ and $1< sp <n$]. In this case, we show that $B^s_{p,q}$ does not have the lifting property. The argument uses embedding theorems and the following fact, for which we provide a proof: let $s_i>0$, $1\le p_i<\infty$, and [$s_jp_j=1$ and $1\le q_j<\infty$], or [$s_jp_j>1$ and $1\le q_j\le\infty$], $i=1,2$. Then, if $f_i\in B^{s_i}_{p_i,q_i}$ and $f_1+f_2$ only takes integer values, then the function $f_1+f_2$ is constant. \medskip Assume finally that $0<s<\infty$, $1\le p<\infty$, $n\ge 2$ and [$1\le q< \infty$ and $1\le sp<2$] or [$q=\infty$ and $1\le sp\le 2$]. In this case, $B^s_{p,q}$ does not have the lifting property either. We provide a counterexample of topological nature, inspired by \cite[Section 4]{lss}: namely, the function $\displaystyle u(x)=\frac{(x_1,x_2)}{\displaystyle\left(x_1^2+x_2^2\right)^{1/2}}$ belongs to $B^s_{p,q}$ but has no lifting in $B^s_{p,q}$. \medskip Contrary to the case of Sobolev spaces, some cases remain open. A first case occurs when $s>1$, $1\le p<\infty$, $p<q<\infty$, $n\ge 3$, and $sp=2$. In this situation, since the restriction property for $B^s_{p,q}$ does not hold, the argument sketched before does not work any longer and we do not know if $B^s_{p,q}$ has the lifting property. The case where $s=1$, $1\le p<\infty$, $n\ge 3$, and [$1\le q<\infty$ and $2\le p<n$] or [$q=\infty$ and $2< p\le n$] is also open (except when $s=1$ and $p=q=2$, since in this case, $B^1_{2,2}=W^{1,2}$ has the lifting property). This is related to the fact that it is not known whether the map $\varphi\mapsto e^{\imath\varphi}$ maps $B^1_{p,q}$ into itself. When $1\le p<\infty$, $s=1/p$ and $q=\infty$, we do not know if $B^{1/p}_{p,\infty}$ has the lifting property. In particular, it is unclear whether the Haar system provides a basis of $B^{1/p}_{p,\infty}$. The case where $q=\infty$, $n\le p<\infty$, $n\ge 3$ and $s=n/p$ is also open. Indeed, $B^s_{p,q}$ is not embedded into $\text{\rm VMO}$ in this case, and the argument briefly described above is not applicable any more. \medskip Let us summarize the main results of this paper concerning the lifting problem. We start with positive cases. \begin{theorem} \label{positive} Let $s>0$, $1\le p<\infty$, $1\le q\le \infty$. The lifting problem has a positive answer in the following cases: \begin{enumerate} \item $s>0$, $1\le q\le\infty$, and $sp>n$, \item $0<s<1$, $1\le q\le\infty$, and $sp<1$, \item $0<s\le 1$, $1\le q<\infty$, and $sp=n$, \item \begin{enumerate} \item $s>1$, $1\le q<\infty$, $n=2$, and $sp=2$, \item $s>1$, $1\le q\le p$, $n\ge 3$, and $sp= 2$, \item $s>1$, $1\le q\le \infty$, $n\ge 2$, and $sp> 2$. \end{enumerate} \end{enumerate} \end{theorem} \noindent The negative cases are as follows: \begin{theorem} \label{negative} Let $s>0$, $1\le p<\infty$, $1\le q\le \infty$. The lifting problem has a negative answer in the following cases: \begin{enumerate} \item \begin{enumerate} \item $0<s<1$, $1\le q<\infty$, $n\ge 2$, and $1\leq sp <n$, \item $0<s<1$, $q=\infty$, $n\ge 2$, and $1< sp <n$, \end{enumerate} \item \begin{enumerate} \item $0<s<\infty$, $1\le q<\infty$, $n\ge 2$, and $1\le sp<2$, \item $0<s<\infty$, $1\le p<\infty$, $q=\infty$, $n\ge 2$, and $1<sp\le 2$. \end{enumerate} \end{enumerate} \end{theorem} The paper is organized as follows. In Section \ref{fun}, we briefly recall the standard definition of Besov spaces and some classical characterizations of these spaces (by Littlewood-Paley theory and wavelets). In Section \ref{pos} we establish Theorem \ref{positive}, namely the cases where $B^s_{p,q}$ does have the lifting property, while Section \ref{neg} is devoted to negative cases (Theorem \ref{negative}). In Section \ref{ope}, we discuss the remaining cases, which are widely open. The final section gathers statements and proofs of various results on Besov spaces needed in the proofs of Theorems \ref{positive} and \ref{negative}. \subsubsection*{Acknowledgments} P. Mironescu thanks N. Badr, G. Bourdaud, P. Bousquet, A.C. Ponce and W. Sickel for useful discussions. He warmly thanks J. Kristensen for calling his attention to the reference \cite{uspenskii}. All the authors are supported by the ANR project \enquote{Harmonic Analysis at its Boundaries}, ANR-12-BS01-0013-03. P. Mironescu was also supported by the LABEX MILYON (ANR- 10-LABX-0070) of Universit\'e de Lyon, within the program \enquote{Investissements d'Avenir} (ANR-11-IDEX-0007) operated by the French National Research Agency (ANR). \section*{Notation, framework} \begin{enumerate} \item Most of our results are stated in a smooth bounded domain $\Omega\subset{\mathbb R}^n$. \item In few cases, proofs are simpler if we consider ${\mathbb Z}^n$-periodic maps $u: (0,1)^n\to{\mathbb S}^1$. In this case, we denote the corresponding function spaces $B^s_{p,q}({\mathbb T}^n ; {\mathbb S}^1)$, and the question is whether a map $u\in B^s_{p,q}({\mathbb T}^n ; {\mathbb S}^1)$ has a lifting $\varphi\in B^s_{p,q}((0,1)^n ; {\mathbb R})$. [Of course, $\varphi$ need not be, in general, ${\mathbb Z}^n$-periodic.] If such a $\varphi$ exists for every $u\in B^s_{p,q}({\mathbb T}^n ; {\mathbb S}^1)$, then $B^s_{p,q}({\mathbb T}^n ; {\mathbb S}^1)$ has the lifting property. However, in these results it is not crucial to work in ${\mathbb T}^n$. An inspection of the proofs shows that, with some extra work, we could take any smooth bounded domain. \item In the same vein, it is sometimes easier to work in $\Omega =(0,1)^n$ (with no periodicity assumption). \item Partial derivatives are denoted $\partial_j$, $\partial_j\partial_k$, and so on, or $\partial^\alpha$. \item $\wedge$ denotes vector product of complex numbers: $a\wedge b:=a_1b_2-a_2b_1$. Similarly, $u\wedge \nabla v:=u_1\nabla v_2-u_2\nabla v_1$. \item If $u:\Omega\to{\mathbb C}$ and if $\varpi$ is a $k$-form on $\Omega$ (with $k\in\llbracket 0, n-1\rrbracket$), then $\varpi\wedge (u\wedge\nabla u) $ denotes the $(k+1)$-form $\varpi\wedge(u_1d u_2-u_2du_1)$. \item We let ${\mathbb R}^n_+$ denote the open set ${\mathbb R}^{n-1}\times (0,\infty)$. \end{enumerate} \tableofcontents \section{Crash course on Besov spaces} \label{fun} ${}$ We briefly recall here the basic properties of the Besov spaces in ${\mathbb R}^n$, with special focus on the properties which will be instrumental for our purposes. For a complete treatment of these spaces, see \cite{triebel2,fjw,triebel3,runstsickel}. \par \subsection{Preliminaries} ${}$ In the sequel, ${\mathcal S}({\mathbb R}^n)$ is the usual Schwartz space of rapidly decreasing $C^{\infty}$ functions. Let ${\mathcal Z}({\mathbb R}^n)$ denote the subspace of ${\mathcal S}({\mathbb R}^n)$ consisting of functions $\varphi\in {\mathcal S}({\mathbb R}^n)$ such that $\partial^{\alpha}\varphi(0) = 0$ for every multi-index $\alpha\in {\mathbb N}^n$. Let ${\mathcal Z}^{\prime}({\mathbb R}^n)$ stand for the topological dual of ${\mathcal Z}({\mathbb R}^n)$. It is well-known \cite[Section 5.1.2]{triebel2} that ${\mathcal Z}^{\prime}({\mathbb R}^n)$ can be identified with the quotient space ${\mathcal S}^{\prime}({\mathbb R}^n)/{\mathcal P}({\mathbb R}^n)$, where ${\mathcal P}({\mathbb R}^n)$ denotes the space of all polynomials in ${\mathbb R}^n$. We denote by ${\mathcal F}$ the Fourier transform. For all sequence $(f_j)_{j\geq 0}$ of measurable functions on ${\mathbb R}^n$, we set \begin{equation*} \left\Vert (f_j)\right\Vert_{l^q(L^p)}:=\left(\sum_{j\geq 0} \left(\int_{{\mathbb R}^n} \left\vert f_j(x)\right\vert^pdx\right)^{q/p}\right)^{1/q}, \end{equation*} with the usual modification when $p=\infty$ and/or $q=\infty$. If $(f_j)$ is labelled by ${\mathbb Z}$, then $\left\Vert (f_j)\right\Vert_{l^q(L^p)}$ is defined analogously with $\sum_{j\geq 0}$ replaced by $\sum_{j\in {\mathbb Z}}$. Finally, we fix some notation for finite order differences. Let $\Omega\subset {\mathbb R}^n$ be a domain and let $f:\Omega\to{\mathbb R}$. For all integers $M\geq 0$, all $t>0$ and all $x, h\in {\mathbb R}^n$, set \begin{equation} \label{ia1} \Delta^M_hf(x)=\begin{cases}\displaystyle\sum_{l=0}^M {M \choose l} (-1)^{M-l} f(x+lh),&\text{if } x,\, x+h,\ldots,\, x+Mh\in \Omega\\ 0,&\text{otherwise} \end{cases}. \end{equation} \subsection{Definitions of Besov spaces} \label{mm7} ${}$ We first focus on inhomogeneous Besov spaces. Fix a sequence of functions $(\varphi_j)_{j\geq 0}\in {\mathcal S}({\mathbb R}^n)$ such that: \begin{itemize} \item[$1.$] $\displaystyle \mbox{supp }\varphi_0\subset B(0,2)$ and $\displaystyle \mbox{supp }\varphi_j\subset B(0,2^{j+1})\setminus B(0,2^{j-1}) \mbox{ for all }j\geq 1$. \item[$2.$] For all multi-index $\alpha\in {\mathbb N}^n$, there exists $c_{\alpha}>0$ such that $\displaystyle \left\vert D^{\alpha}\varphi_j(x)\right\vert\leq c_{\alpha}2^{-j\left\vert \alpha\right\vert}$, for all $x\in {\mathbb R}^n$ and all $j\geq 0$. \item[$3.$] For all $x\in {\mathbb R}^n$, it holds $ \sum_{j\geq 0} \varphi_j(x)=1$. \end{itemize} \begin{definition} [Definition of inhomogeneous Besov spaces] Let $s\in {\mathbb R}$, $1\leq p<\infty$ and $1\leq q\leq \infty$. Define $B^s_{p,q}({\mathbb R}^n)$ as the space of tempered distributions $f\in {\mathcal S}^{\prime}({\mathbb R}^n)$ such that \begin{equation*} \left\Vert f\right\Vert_{B^s_{p,q}({\mathbb R}^n)}:=\left\Vert \left(2^{sj} {\mathcal F}^{-1}\left(\varphi_j{\mathcal F}f(\cdot)\right)\right)\right\Vert_{l^q(L^p)}<\infty. \end{equation*} \end{definition} Recall \cite[Section 2.3.2, Proposition 1, p. 46]{triebel2} that $B^s_{p,q}({\mathbb R}^n)$ is a Banach space which does not depend on the choice of the sequence $(\varphi_j)_{j\geq 0}$, in the sense that two different choices for the sequence $(\varphi_j)_{j\geq 0}$ give rise to equivalent norms. Once the $\varphi_j$'s are fixed, we refer to the equality $f=\sum_j f_j$ in ${\mathcal S}'$ as the Littlewood-Paley decomposition of $f$. Let us now turn to the definition of homogeneous Besov spaces. Let $(\varphi_j)_{j\in {\mathbb Z}}$ be a sequence of functions satisfying: \begin{itemize} \item[$1.$] $\displaystyle \mbox{supp }\varphi_j\subset B(0,2^{j+1})\setminus B(0,2^{j-1}) \mbox{ for all }j\in {\mathbb Z}$. \item[$2.$] For all multi-index $\alpha\in {\mathbb N}^n$, there exists $c_{\alpha}>0$ such that$\displaystyle \left\vert D^{\alpha}\varphi_j(x)\right\vert\leq c_{\alpha}2^{-j\left\vert \alpha\right\vert} $, for all $x\in {\mathbb R}^n$ and all $j\in {\mathbb Z}$. \item[$3.$] For all $x\in {\mathbb R}^n\setminus \left\{0\right\}$, it holds $ \sum_{j\in {\mathbb Z}} \varphi_j(x)=1$. \end{itemize} \begin{definition} [Definition of homogeneous Besov spaces] Let $s\in {\mathbb R}$, $1\leq p<\infty$ and $1\leq q\leq \infty$. Define $\dot{B}^s_{p,q}({\mathbb R}^n)$ as the space of $f\in {\mathcal Z}^{\prime}({\mathbb R}^n)$ such that \begin{equation*} \left\vert f\right\vert_{B^s_{p,q}({\mathbb R}^n)}:=\left\Vert \left(2^{sj} {\mathcal F}^{-1}\left(\varphi_j{\mathcal F}f(\cdot)\right)\right)\right\Vert_{l^q(L^p)}<\infty. \end{equation*} \end{definition} Note that this definition makes sense since, for all polynomial $P$ and all $f\in {\mathcal S}^{\prime}({\mathbb R}^n)$, we have $\displaystyle \left\vert f\right\vert_{B^s_{p,q}({\mathbb R}^n)}=\left\vert f+P\right\vert_{B^s_{p,q}({\mathbb R}^n)}$. Again, $\dot{B}^s_{p,q}({\mathbb R}^n)$ is a Banach space which does not depend on the choice of the sequence $(\varphi_j)_{j\in {\mathbb Z}}$ \cite[Section 5.1.5, Theorem, p. 240]{triebel2}.\par \noindent For all $s>0$ and all $1\le p<\infty$, $1\leq q\leq \infty$, we have \cite[Section 2.3.3, Theorem]{triebel3}, \cite[Section 2.6.2, Proposition 3]{runstsickel} \begin{equation} \label{homoglp} B^s_{p,q}({\mathbb R}^n)=L^p({\mathbb R}^n)\cap \dot{B}^s_{p,q}({\mathbb R}^n)\mbox{ and }\left\Vert f\right\Vert_{B^s_{p,q}({\mathbb R}^n)}\sim \left\Vert f\right\Vert_{L^p({\mathbb R}^n)}+\left\vert f\right\vert_{B^s_{p,q}({\mathbb R}^n)}. \end{equation} Besov spaces on domains of ${\mathbb R}^n$ are defined as follows. \begin{definition} [Besov spaces on domains] Let $\Omega\subset {\mathbb R}^n$ be an open set. Then \begin{itemize} \item[$1.$] $\displaystyle B^s_{p,q}(\Omega):=\left\{f\in {\mathcal D}^{\prime}(\Omega);\mbox{ there exists }g\in B^s_{p,q}({\mathbb R}^n)\mbox{ such that } f=g\vert_{\Omega}\right\}$,\\ equipped with the norm \begin{equation*} \left\Vert f\right\Vert_{B^s_{p,q}(\Omega)}:=\inf \left\{\left\Vert g\right\Vert_{B^s_{p,q}({\mathbb R}^n)};\, g\vert_{\Omega}=f\right\}. \end{equation*} \item[$2.$] $\displaystyle \dot{B}^s_{p,q}(\Omega):=\left\{f\in {\mathcal D}^{\prime}(\Omega);\mbox{ there exists }g\in \dot{B}^s_{p,q}({\mathbb R}^n)\mbox{ such that } f=g\vert_{\Omega}\right\}$,\\ equipped with the semi-norm \begin{equation*} \left\Vert f\right\Vert_{\dot{B}^s_{p,q}(\Omega)}:=\inf \left\{\left\Vert g\right\Vert_{\dot{B}^s_{p,q}({\mathbb R}^n)};\, g\vert_{\Omega}=f\right\}. \end{equation*} \end{itemize} \end{definition} Local Besov spaces are defined in the usual way: $f\in B^s_{p,q}$ near a point $x$ if for some cutoff $\varphi$ which equals $1$ near $x$ we have $\varphi f\in B^s_{p,q}$. If $f$ belongs to $B^s_{p,q}$ near each point, then we write $f\in (B^s_{p,q})_{loc}$. The following is straightforward. \begin{lemma} \label{ka3} Let $f:\Omega\to{\mathbb R}$. If, for each $x\in\overline\Omega$, $f\in B^s_{p,q}(B(x,r)\cap \Omega)$ for some $r=r(x)>0$, then $f\in B^s_{p,q}$. \end{lemma} \subsection{Besov spaces on ${\mathbb T}^n$} \label{mm6} ${}$ Let $\varphi_0\in {\mathcal D}({\mathbb R}^n)$ be such that \begin{equation*} \varphi_0(x)=1\mbox{ for all } \left\vert x\right\vert<1\mbox{ and }\varphi_0(x)=0\mbox{ for all }\left\vert x\right\vert\geq \frac 32. \end{equation*} For all $k\geq 1$ and all $x\in {\mathbb R}^n$, define \begin{equation*} \varphi_k(x):=\varphi_0(2^{-k}x)-\varphi_0(2^{-k+1}x). \end{equation*} \begin{definition} \label{periodicbesov} Let $s\in {\mathbb R}$, $1\leq p<\infty$ and $1\le q\leq \infty$. Define $B^s_{p,q}({\mathbb T}^n)$ as the space of distributions $f\in {\mathcal D}^{\prime}({\mathbb T}^n)$ whose Fourier coefficients $(a_m)_{m\in{\mathbb Z}^n}$ satisfy \begin{equation*} \left\Vert f\right\Vert_{B^s_{p,q}({\mathbb T}^n)}:=\left(\sum_{j=0}^{\infty} 2^{jsq} \left\Vert x\mapsto \sum_{m\in {\mathbb Z}^n} a_m\varphi_j(2\pi m)e^{2\imath\pi m\cdot x}\right\Vert_{L^p({\mathbb T}^n)}^q\right)^{1/q}<\infty \end{equation*} (with the usual modification when $q=\infty$). Again, the choice of the system $(\varphi_j)_{j\ge 0}$ is irrelevant, and the equality $f=\sum f_j$, with $f_j:=\sum_{m} a_m\varphi_j(2\pi m)e^{2\imath\pi m\cdot x}$, is the Littlewood-Paley decomposition of $f$. \end{definition} Alternatively, we have $f\in B^s_{p,q}({\mathbb T}^n)$ if and only if $f$ can be identified with a ${\mathbb Z}^n$-periodic distribution in ${\mathbb R}^n$, still denoted $f$, which belongs to $(B^s_{p,q})_{loc}({\mathbb R}^n)$ \cite[Section 3.5.4, pp. 167-169]{schmeisser}. \subsection{Characterization by differences} \label{mm5} ${}$ Among the various characterizations of Besov spaces, we recall here the ones involving differences \cite[Section 5.2.3]{triebel2}, \cite[Theorem, p. 41]{runstsickel}, \cite[Section 1.11.9, Theorem 1.118, p. 74]{triebel06}. \begin{proposition} \label{p2.4} Let $s>0$, $1\leq p<\infty$ and $1\leq q\leq \infty$. Let $M>s$ be an integer. Then, with the usual modification when $q=\infty$: \begin{itemize} \item[$1.$] In the space $\dot B^s_{p,q}({\mathbb R}^n)$ we have the equivalence of semi-norms \begin{equation} \label{equivnormhomogrn} \begin{aligned} \left\vert f\right\vert_{B^s_{p,q}({\mathbb R}^n)}\sim &\left(\int_{{\mathbb R}^n} \left\vert h\right\vert^{-sq} \left\|\Delta_h^Mf\right\|_{L^p ({\mathbb R}^n)}^q\, \frac{dh}{\left\vert h\right\vert^n}\right)^{1/q}\\ \sim & \sum_{j=1}^n\left(\int_{{\mathbb R}} \left\vert h\right\vert^{-sq} \left\|\Delta_{h e_j}^Mf\right\|_{L^p ({\mathbb R}^n)}^q\, \frac{dh}{\left\vert h\right\vert}\right)^{1/q}. \end{aligned} \end{equation} \item[$2.$] The full $B^s_{p,q}$ norm satisfies, for all $\delta>0$, \begin{equation*} \left\Vert f\right\Vert_{B^s_{p,q}({\mathbb R}^n)}\sim \left\Vert f\right\Vert_{L^p({\mathbb R}^n)}+\left(\int_{\left\vert h\right\vert\le \delta} \left\vert h\right\vert^{-sq} \left\|\Delta_h^Mf\right\|_{L^p ({\mathbb R}^n)}^q\, \frac{dh}{\left\vert h\right\vert^n}\right)^{1/q}. \end{equation*} \end{itemize} \end{proposition} \subsection{Characterization by harmonic extensions} \label{chha} ${}$ In Section \ref{pos}, it will be convenient to work with extensions of maps in $B^{s}_{p,q}$. The connection between regularity of maps and the properties of their suitable extensions is a classical topic in the theory of function spaces. Here is a typical result in this direction. It characterizes $B^{s}_{p,q}$ by means of the harmonic extension \cite{triebelheat}, \cite[Section 2.12.2, Theorem, p. 184]{triebel2}. More specifically, if $f$ is measurable in ${\mathbb R}^n$ and $s\in (0,1)$, then we have \begin{equation} \label{besovnorm} \left\Vert f\right\Vert_{B^s_{p,q}({\mathbb R}^n)}\sim \left\Vert f\right\Vert_{L^p({\mathbb R}^n)}+ \left(\int_0^{\infty} t^{(1-s)q}\left\Vert \frac{\partial P_tf}{\partial t}(\cdot)\right\Vert_{L^p({\mathbb R}^n)}^q \frac{dt}t\right)^{1/q}, \end{equation} where $P_t$ stands for the Poisson semigroup generated by $-\Delta$, so that $(x,t)\mapsto P_tf(x)$, $t>0$, $x\in{\mathbb R}^n$, is the harmonic extension of $f$ to the upper-half space. Since when $p>1$ we have \begin{equation*}\left\Vert \frac{\partial P_tf}{\partial t}\right\Vert_{L^p({\mathbb R}^n)} =\left\Vert (-\Delta_x)^{1/2}P_tf\right\Vert_{L^p({\mathbb R}^n)}\sim \left\Vert \nabla_x P_tf\right\Vert_{L^p({\mathbb R}^n)}, \end{equation*} one also has, for $1<p<\infty$ and $1\le q\le \infty$, \begin{equation} \label{besovnormbis} \left\Vert f\right\Vert_{B^s_{p,q}({\mathbb R}^n)}\sim \left\Vert f\right\Vert_{L^p({\mathbb R}^n)}+ \left(\int_0^{\infty} t^{(1-s)q}\left\Vert \nabla P_tf(\cdot)\right\Vert_{L^p({\mathbb R}^n)}^q \frac{dt}t\right)^{1/q} \end{equation} (with the usual modification when $q=\infty$). The results in the literature are not suited to our context. We will need some variants of \eqref{besovnormbis}, which will be stated and proved in Section \ref{characext} below. \subsection{Lizorkin type characterizations} \label{mm1} ${}$ Such characterizations involve restrictions of the Fourier transform on cubes or corridors; see e.g. \cite[Section 2.5.4, pp. 85-86]{triebel2} or \cite[Section 3.5.3, pp. 166-167]{schmeisser}. The following special case \cite[Section 3.5.3, Theorem, p. 167]{schmeisser} will be useful later. \begin{proposition} \label{mm2} Let $s\in{\mathbb R}$, $1<p<\infty$ and $1\le q\le\infty$. Set $K_0:=\{0\}\subset{\mathbb Z}^n$ and, for $j\ge 1$, let $K_j:=\{ m\in{\mathbb Z}^n;\, 2^{j-1}\le |m|<2^j\}$.\footnote{ Here, $|m|:=\max_{l=1}^n|m_l|$.} Let $f\in{\cal D}'({\mathbb T}^n)$ have the Fourier series expansion $f=\sum_{m\in{\mathbb Z}^n} a_me^{2\imath\pi m\cdot x}$. We set $f_j:=\sum_{m\in K_j} a_me^{2\imath\pi m\cdot x}$. Then we have the norm equivalence \begin{equation*} \|f\|_{B^s_{p,q}({\mathbb T}^n)}\sim \left(\sum_{j=0}^{\infty} 2^{jsq} \left\Vert f_j\right\Vert_{L^p({\mathbb T}^n)}^q\right)^{1/q} \end{equation*} (with the usual modification when $q=\infty$). \end{proposition} \subsection{Characterization by the Haar system} \label{at7} ${}$ Besov spaces can also be described via the size of their wavelet coefficients. To illustrate this, we start with low smoothness Besov spaces, which can be described using the Haar basis. (The next section is devoted to smoother spaces and bases.) For the results of this section, see e.g. \cite[Corollary 5.3]{devorepopov}, \cite[Section 7]{bourdaud}, \cite[Theorem 1.58]{triebel06}, \cite[Theorem 2.21]{triebel10}. \par \noindent Let \begin{equation} \label{qa8} \psi_M(x):= \begin{cases} 1, & \text{if }0\leq x<1/2\\ -1, & \text{if }1/2\leq x\leq 1\\ 0, & \text{if }x\notin [0,1] \end{cases}, \text{ and }\psi_F(x):=\left\vert \psi_M(x)\right\vert. \end{equation} When $j\in{\mathbb N}$, we let \begin{equation} \label{qa1} G^j:=\begin{cases} \left\{F,M\right\}^{n},&\text{if }j=0\\ \left\{F,M\right\}^{n}\setminus\{ (F, F, \ldots, F)\},&\text{if }j>0 \end{cases}. \end{equation} For all $m\in {\mathbb Z}^n$, all $x\in {\mathbb R}^n$ and all $G\in \left\{F,M\right\}^{n}$, define \begin{equation} \label{qa2} \Psi_m^G(x):=\prod_{r=1}^n \psi_{G_r}(x_r-m_r). \end{equation} Finally, for all $m\in {\mathbb Z}^n$, all $j\in {\mathbb N}$, all $G\in G^j$ and all $x\in {\mathbb R}^n$, let \begin{equation} \label{qa3} \Psi_m^{j,G}(x):= \begin{cases} \displaystyle \Psi_m^G(x), & \text{if }j=0\\ \displaystyle 2^{nj/2}\Psi^G_m(2^jx), & \text{if }j\geq 1 \end{cases}. \end{equation} Recall that the family $(\Psi_m^{j,G})$, called the Haar system, is an orthonormal basis of $L^2({\mathbb R}^n)$ \cite[Proposition 1.53]{triebel06}. Moreover, we have the following result \cite[Theorem 2.21]{triebel10}. \begin{proposition} \label{at11} Let $s>0$, $1\leq p<\infty$, and $1\le q\le \infty$ be such that $sp<1$. Let $f\in {\mathcal S}^{\prime}({\mathbb R}^n$). Then $f\in B^s_{p,q}({\mathbb R}^n)$ if and only if there exists a sequence $\left(\mu^{j, G}_{m}\right)_{j\geq 0,\ G\in G^j,\ m\in {\mathbb Z}^n}$ such that \begin{equation} \label{qa4} \sum_{j=0}^{\infty}\ \sum_{G\in G^j} \left(\sum_{m\in {\mathbb Z}^n} \left\vert \mu^{j, G}_{m}\right\vert^p\right)^{q/p}<\infty \end{equation} (obvious modification when $q=\infty$) and \begin{equation} \label{decompof} f=\sum_{j=0}^{\infty}\ \sum_{G\in G^j} \sum_{m\in {\mathbb Z}^n} \mu^{j, G}_{m} 2^{-j\left(s-n/p\right)} 2^{-nj/2} \Psi^{j,G}_{m}. \end{equation} Here, the series in \eqref{decompof} converges unconditionally in $B^s_{p,q}({\mathbb R}^n)$ when $q<\infty$. Moreover, \begin{equation} \label{qa5} \left\Vert f\right\Vert_{B^s_{p,q}({\mathbb R}^n)}\sim \left(\sum_{j=0}^{\infty}\ \sum_{G\in G^j} \left(\sum_{m\in {\mathbb Z}^n} \left\vert \mu^{j, G}_{m}\right\vert^p\right)^{q/p}\right)^{1/q} \end{equation} (obvious modification when $q=\infty$). \end{proposition} Equivalently, Proposition \ref{at11} can be reformulated as follows. Consider the partition of ${\mathbb R}^n$ into standard dyadic cubes $Q$ of side $2^{-j}$. \footnote{ Thus the $Q$'s are of the form $Q=2^{-j}\prod_{k=1}^n[m_k,m_k+1)$, with $m_k\in{\mathbb Z}$.} For all $x\in{\mathbb R}^n$, denote by $Q_j(x)$ the unique dyadic cube of side $2^{-j}$ containing $x$. If $f\in L^1_{loc}({\mathbb R}^n)$, define $E_j(f)(x):=\fint_{Q_j(x)}f$ for all $j\geq 0$. We also set $E_{-1}(f):=0$. We have the following results (see \cite[Theorem 5 with $m=0$]{bourdaud} in ${\mathbb R}^n$; see also \cite[Appendix A]{lss} in the framework of Sobolev spaces on ${\mathbb T}^n$). \begin{proposition}\label{caracBesov} Let $s>0$, $1\leq p<\infty$, and $1\le q\le \infty$ be such that $sp<1$. Let $f\in L^1_{loc}({\mathbb R}^n$). Then \begin{equation*}\label{caracBesovBourdaud} \left\Vert f\right\Vert^q_{B^s_{p,q}({\mathbb R}^n)}\sim \sum_{j \ge 0} 2^{sjq}\|E_j(f)-E_{j-1}(f)\|_{L^p}^q \end{equation*} (obvious modification when $q=\infty$). \end{proposition} Similar results hold when ${\mathbb R}^n$ is replaced by $(0,1)^n$ or ${\mathbb T}^n$; it suffices to consider only dyadic cubes contained in $[0,1)^n$. \begin{corollary} \label{mq2} Let $s>0$, $1\leq p<\infty$, and $1\le q\le \infty$ be such that $sp<1$. Let $f\in L^1_{loc}({\mathbb R}^n$). Then \begin{equation*}\label{mq3} \left\Vert f\right\Vert^q_{B^s_{p,q}({\mathbb R}^n)}\sim \sum_{j \ge 0} 2^{sjq}\|f-E_{j}(f)\|_{L^p}^q \end{equation*} (obvious modification when $q=\infty$). Similar results hold when ${\mathbb R}^n$ is replaced by $(0,1)^n$ or ${\mathbb T}^n$. \end{corollary} \begin{corollary} \label{mp1} Let $s>0$, $1\leq p<\infty$, and $1\le q\le \infty$ be such that $sp<1$. Let $(\varphi_j)_{j\ge 0}$ be a sequence of functions on $(0,1)^n$ such that: for any $j$, $\varphi_j$ is constant on each dyadic cube $Q$ of size $2^{-j}$. Assume that $ \sum_{j \ge 1} 2^{sjq}\|\varphi_j-\varphi_{j-1}\|_{L^p}^q < \infty$. Then $(\varphi_j)$ converges in $L^p$ to some $\varphi \in B^s_{p,q}$, and we have \begin{equation*} \left\Vert \varphi \right\Vert_{B^s_{p,q}((0,1)^n)} \lesssim \left(\sum_{j \ge 0} 2^{sjq}\|\varphi_j-\varphi_{j-1}\|_{L^p}^q\right)^{1/q} \end{equation*} (with the convention $\varphi_{-1}:=0$ and with the usual modification when $q=\infty$). \end{corollary} In the framework of Sobolev spaces, Corollaries \ref{mq2} and \ref{mp1} are easy consequences of Proposition \ref{caracBesov}; see \cite[Appendix A, Theorem A.1]{lss} and \cite[Appendix A, Corollary A.1]{lss}. The arguments in \cite{lss} apply with no changes to Besov spaces. Details are left to the reader. \subsection{Characterization via smooth wavelets} \label{qa6} Proposition \ref{at11} has a counterpart when $sp\ge 1$; this requires smoother \enquote{mother wavelet} $\psi_M$ and \enquote{father wavelet} $\psi_F$. Given $\psi_F$ and $\psi_M$ two real functions, define $\psi_m^{j, G}$ as in \eqref{qa1}--\eqref{qa3}. Then \cite[Chapter 6]{meyer92}, \cite[Section 1.7.3]{triebel06} for every integer $k>0$ we may find some $\psi_F\in C^k_c({\mathbb R})$ and $\psi_M\in C^k_c({\mathbb R})$ such that the following result holds. \begin{proposition} \label{qb1} Let $s>0$, $1\leq p<\infty$, and $1\le q\le \infty$ be such that $s<k$. Let $f\in {\mathcal S}^{\prime}({\mathbb R}^n$). Then $f\in B^s_{p,q}({\mathbb R}^n)$ if and only if there exists a sequence $\left(\mu^{j, G}_{m}\right)_{j\geq 0,\ G\in G^j,\ m\in {\mathbb Z}^n}$ such that \begin{equation} \label{qb2} \sum_{j=0}^{\infty}\ \sum_{G\in G^j} \left(\sum_{m\in {\mathbb Z}^n} \left\vert \mu^{j, G}_{m}\right\vert^p\right)^{q/p}<\infty \end{equation} (obvious modification when $q=\infty$) and \begin{equation} \label{qb3} f=\sum_{j=0}^{\infty}\ \sum_{G\in G^j} \sum_{m\in {\mathbb Z}^n} \mu^{j, G}_{m} 2^{-j\left(s-n/p\right)} 2^{-nj/2} \Psi^{j,G}_{m}. \end{equation} Here, the series in \eqref{decompof} converges unconditionally in $B^s_{p,q}({\mathbb R}^n)$ when $q<\infty$. Moreover, \begin{equation} \label{qb4} \left\Vert f\right\Vert_{B^s_{p,q}({\mathbb R}^n)}\sim \left(\sum_{j=0}^{\infty}\ \sum_{G\in G^j} \left(\sum_{m\in {\mathbb Z}^n} \left\vert \mu^{j, G}_{m}\right\vert^p\right)^{q/p}\right)^{1/q} \end{equation} (obvious modification when $q=\infty$). \end{proposition} For further use, let us note that, if $f\in B^s_{p,q}({\mathbb R}^n)$ for some $s>0$, $1\le p<\infty$ and $1\le q\le\infty$, then we have \begin{equation} \label{qb40} \mu^{j, G}_{m}=\mu^{j, G}_{m}(f)=2^{j(s-n/p+n/2)}\, \int_{{\mathbb R}^n} f(x)\, \Psi^{j,G}_{m}(x)\, dx. \end{equation} This immediately leads to the following consequence of Proposition \ref{qb1}, the proof of which is left to the reader. \begin{corollary} \label{qb400} Let $s>0$, $1\le p<\infty$ and $1\le q\le\infty$ be such that $s<k$. Assume that $f\in L^p({\mathbb R}^n)$ is such that the coefficients $\mu^{j, G}_{m}$ given by \eqref{qb40} satisfy \begin{equation} \label{qb50} \sum_{j=0}^{\infty}\ \sum_{G\in G^j} \left(\sum_{m\in {\mathbb Z}^n} \left\vert \mu^{j, G}_{m}\right\vert^p\right)^{q/p}=\infty \end{equation} (obvious modification when $q=\infty$). Then $f\not\in B^s_{p,q}({\mathbb R}^n)$. \end{corollary} \subsection{Nikolski\u\i{} type decompositions} \label{mm8} ${}$ In practice, we often do not know the Littlewood-Paley decomposition of some given $f$, but only a Nikolski\u\i{} representation (or decomposition) of $f$. More specifically, set $\mathcal{C}_j:=B(0,2^{j+2})$, with $j\in{\mathbb N}$. Let $f^j\in\mathcal{S}'$ satisfy \begin{equation}\label{e230501} \operatorname{supp}{\mathcal F}f^j\subset\mathcal{C}_j,\ \forall\, j\in{\mathbb N},\ \text{and } f=\sum_jf^j\text{ in }{\mathcal S}'; \end{equation} the decomposition $f=\sum_j f^j$ is a Nikolski\u\i{} decomposition of $f$. Note that the Littlewood-Paley decomposition is a special Nikolski\u\i{} decomposition. We have the following result. \begin{proposition} \label{mm9} Let $s>0$, $1\le p<\infty$, $1\le q\le\infty$. Assume that \eqref{e230501} holds. Then we have \begin{equation} \label{28024} \Big\|\sum_{j}f^j\Big\|_{B^{s}_{p,q}}\lesssim\left(\sum_{j}2^{ sqj }\|f^j\|^q_{L^p}\right)^{1/q}, \end{equation} with the usual modification when $q=\infty$. \end{proposition} \noindent The above was proved in \cite[Lemma 1]{gnp} (see also \cite{yamazaki}) in the framework of Triebel-Lizorkin spaces $F^s_{p,q}$; the proof applies with no change to Besov spaces and will be omitted here. For related results in the framework of Besov spaces, see \cite[Section 2.5.2, pp. 79-80]{triebel2} and \cite[Section 2.3.2, Theorem, p. 105]{schmeisser}. \section{Positive cases} \label{pos} ${}$ We start with the trivial case. \begin{case} \label{tri} {\it Range.} $s>0$, $1\le p<\infty$, $1\le q\le\infty$, and $sp>n$. \smallskip \noindent {\it Conclusion.} $B^s_{p,q}(\Omega ; {\mathbb S}^1)$ does have the lifting property. \end{case} \begin{proof} Since $B^s_{p,q}(\Omega)\hookrightarrow C^0(\overline\Omega)$ (Lemma \ref{ka1}), we may write $u=e^{\imath\varphi}$, with $\varphi$ continuous. Locally, we have $\varphi=-\imath \ln u$, for some smooth determination $\ln$ of the complex logarithm. Then $\varphi$ belongs to $B^s_{p,q}$ locally in $\overline\Omega$ (Lemma \ref{ka2}), and thus globally (Lemma \ref{ka3}). \end{proof} \begin{case} \label{A} {\it Range.} $0<s<1$, $1\le p<\infty$, $1\le q\le\infty$, and $sp<1$. \smallskip \noindent {\it Conclusion.} $B^s_{p,q}(\Omega ; {\mathbb S}^1)$ does have the lifting property. \end{case} \begin{proof} The argument being essentially the one in \cite[Section 1]{lss}, we will be sketchy. Assume for simplicity that $\Omega=(0,1)^n$. Let $u \in B^s_{p,q}(\Omega ; {\mathbb S}^1)$. For all $j\in{\mathbb N}$, consider the function $U_j$ defined by \begin{equation*} U_j(x):= \begin{cases} \displaystyle E_j(u)(x)/|E_j(u)(x)|,&\mbox{if }E_j(u)(x) \neq 0\\ 1,&\mbox{if }E_j(u)(x) = 0 \end{cases}. \end{equation*} Since $E_j(u)\to u$ a.e., we find that $U_j \rightarrow u$ a.e. on $\Omega$. By induction on $j$, for all $j\in{\mathbb N}$ we construct a phase $\varphi_j$ of $U_j$, constant on each dyadic cube of size $2^{-j}$, and satisfying the inequality \begin{equation}\label{mq1} |\varphi_j -\varphi_{j-1}| \leq \pi |U_j-U_{j-1}| \quad\mbox{on }\Omega,\, \forall\, j\ge 1.\footnotemark \end{equation} \footnotetext{ Thus $\varphi_j$ is the phase of $U_j$ closest to $\varphi_{j-1}$.} As in \cite{lss}, \eqref{mq1} implies \begin{equation*} |\varphi_j -\varphi_{j-1}| \lesssim |u -E_j(u)|+ |u -E_{j-1}(u)|, \end{equation*} and thus, e.g. when $q<\infty$, we have \begin{equation*} \sum_{j \ge 1} 2^{sjq}\|\varphi_j-\varphi_{j-1}\|_{L^p}^q \lesssim \sum_{j \ge 0} 2^{sjq}\|u-E_{j}(u)\|_{L^p}^q. \end{equation*} Applying Corollaries \ref{mq2} and \ref{mp1}, we obtain that $\varphi_j \rightarrow \varphi $ in $L^p$ to some $\varphi \in B^s_{p,q}(\Omega ; {\mathbb R})$. Since $\varphi_j$ is a phase of $U_j$ and $U_j\to u$ a.e., we find that $\varphi$ is a phase of $u$. In addition, we have the control $\|\varphi\|_{B^s_{p,q}} \lesssim \|u\|_{B^s_{p,q}}$. \end{proof} \begin{case} \label{X} {\it Range.} $0<s<1$, $1\le p<\infty$, $1\le q<\infty$, and $sp=n$. \smallskip \noindent {\it Conclusion.} $B^s_{p,q}(\Omega ; {\mathbb S}^1)$ does have the lifting property. \end{case} \begin{proof} Here, it will be convenient to work with $\Omega={\mathbb T}^n$. Let $|\ |$ denote the sup norm in ${\mathbb R}^n$. Let $\rho\in C^\infty$ be a mollifier supported in $\{ |x|\le 1\}$ and set $F(x,\varepsilon):=u\ast\rho_\varepsilon(x)$, $x\in{\mathbb T}^n$, $\varepsilon>0$. Since $sp=n$, we have $u\in \text{\rm VMO}({\mathbb T}^n)$, by Lemma \ref{B-VMO}. Let us recall that, if $u\in\text{\rm VMO}({\mathbb T}^n ; {\mathbb S}^1)$ then, for some $\delta>0$ (depending on $u$) we have \cite[Remark 3, p. 207]{brezisnirenberg1} \begin{equation} \label{boundsv} \frac 12<\left\vert F(x,\varepsilon)\right\vert\le 1\mbox{ for all } x\in {\mathbb T}^n\mbox{ and all }\varepsilon\in (0,\delta).\footnotemark \end{equation} \footnotetext{ For an explicit calculation leading to \eqref{boundsv}, see e.g. \cite[p. 415]{surveypetru}.} Define \begin{equation*} w(x,\varepsilon):=\frac{F(x,\varepsilon)}{\left\vert F(x,\varepsilon)\right\vert}\mbox{ for all }x\in {\mathbb T}^n\mbox{ and all }\varepsilon\in (0,\delta). \end{equation*} Pick up a function $\psi\in C^{\infty}({\mathbb T}^n\times (0,\delta) ; {\mathbb R})$ such that $w=e^{\imath \psi}$. We note that for all $j\in\llbracket 1, n\rrbracket$ we have $ \nabla\psi=-\imath \overline{w}\nabla w $, and $ \partial_j |F|= |F|^{-1}(F\partial_j\overline{F}+\overline{F}\partial_jF)/2 $. Therefore, \eqref{boundsv} yields \begin{equation} \label{nablaw} \left\vert \nabla\psi\right\vert=\left\vert \nabla w\right\vert\lesssim \left\vert \nabla F\right\vert. \end{equation} In view of \eqref{nablaw} and estimate \eqref{cg1} in Lemma \ref{ab1}, we find that \begin{equation} \label{ka5} |u|_{B^{s,p}_q({\mathbb T}^n)}^q \gtrsim \displaystyle \int_0^\delta\varepsilon^{q-sq}\|(\nabla F)(\cdot,\varepsilon)\|_{L^p}^q\, \frac{d\varepsilon}{\varepsilon} \gtrsim \int_0^\delta\varepsilon^{q-sq}\|(\nabla \psi)(\cdot,\varepsilon)\|_{L^p}^q\, \frac{d\varepsilon}{\varepsilon}. \end{equation} Combining \eqref{ka5} with the conclusion of Lemma \ref{ab1}, we obtain that the phase $\psi$ has, on ${\mathbb T}^n$, a trace $\varphi\in B^s_{p,q}$, in the sense that the limit $\varphi:=\lim_{\varepsilon\to 0}\psi(\cdot,\varepsilon)$ exists in $B^s_{p,q}$. In particular (using Lemma \ref{kc2}), we have that $\psi(\cdot, \varepsilon_j)\to\varphi$ a.e. along some sequence $\varepsilon_j\to 0$; this leads to $w(\cdot, \varepsilon_j)=e^{\imath\psi(\cdot,\varepsilon_j)}\to e^{\imath\varphi}$ a.e. Since, on the other hand, we have $\lim_{\varepsilon\to 0}w(\cdot,\varepsilon)=u$ a.e., we find that $\varphi$ is a $B^s_{p,q}$ phase of $u$. \end{proof}\color{black} The next case is somewhat similar to Case \ref{X}, so that our argument is less detailed. \begin{case} \label{kc3} {\it Range.} $s=1$, $p=n$, $1\le q<\infty$. \smallskip \noindent {\it Conclusion.} $B^1_{n,q}(\Omega ; {\mathbb S}^1)$ does have the lifting property. \end{case} \begin{proof} We consider $\delta$, $w$ and $\psi$ as in Case \ref{X}. The analog of \eqref{nablaw} is the estimate \begin{equation} \label{kc4} |\partial_j\partial_k\psi|+|\nabla\psi|^2\lesssim |\partial_j\partial_kF|+|\nabla F|^2, \end{equation} which is a straightforward consequence of the identities \begin{equation*} \nabla\psi=-\imath \overline w\nabla w\text{ and }\partial_j\partial_k\psi=-\imath \overline w\partial_j\partial_kw+\imath w^2\partial_j w\partial_kw. \end{equation*} Combining \eqref{kc4} with the second part of Lemma \ref{kb2}, we obtain \begin{equation} \label{kg1a} |u|_{B^1_{n,q}}^q\gtrsim \int_0^\delta\varepsilon^q\left(\sum_{j, k=1}^n\left\|\partial_j\partial_k\psi(\cdot,\varepsilon)\right\|_{L^n}^q+\left\|\partial_\varepsilon\partial_\varepsilon\psi(\cdot,\varepsilon)\right\|_{L^n}^q+\|\nabla\psi(\cdot,\varepsilon)\|_{L^{2n}}^{2q}\right)\, \frac{d\varepsilon}{\varepsilon}. \end{equation} By \eqref{kg1a} and the first part of Lemma \ref{kb2}, we find that $\psi$ has a trace $\varphi:=\operatorname{tr} \psi \in B^1_{n,q}({\mathbb T}^n)$. Clearly, $\varphi$ is a $B^1_{n,q}$ phase of $u$. \end{proof} \begin{case} \label{Y} {\it Range.} $s>1$, $1\le p<\infty$, $1\le q<\infty$, $n=2$, and $sp=2$. Or $s>1$, $1\le p<\infty$, $1\le q\le p$, $n\ge 3$, and $sp= 2$. Or: $s>1$, $1\le p<\infty$, $1\le q\le \infty$, $n\ge 2$, and $sp> 2$. \smallskip \noindent {\it Conclusion.} $B^s_{p,q}(\Omega ; {\mathbb S}^1)$ does have the lifting property. \end{case} Note that, in the critical case where $sp=2$, our result is weaker in dimension $n\ge 3$ (when we ask $1\le q\le p$) than in dimension $2$ (when we merely ask $1\le q<\infty$). \begin{proof} The general strategy is the same as in \cite[Section 3, Proof of Theorem 3]{lss},\footnote{ See also \cite{carbou}.} but the key argument (validity of \eqref{at1} below) is much more involved in our case. It will be convenient to work in $\Omega={\mathbb T}^n$. Let $u\in B^s_{p, q}({\mathbb T}^n ; {\mathbb S}^1)$. Assume first that we do may write $u=e^{\imath\varphi}$, with $\varphi\in B^s_{p, q}((0,1)^n ; {\mathbb R})$. Then $u, \varphi\in W^{1,p}$ (Lemma \ref{kc2}). We are thus in position to apply chain's rule and infer that $\nabla u=\imath u\nabla \varphi$, and therefore \begin{equation} \label{at2} \nabla\varphi=\frac 1{\imath u}\nabla u=F,\ \text{with }F:=u\wedge\nabla u\in L^p({\mathbb T}^n ; {\mathbb R}^n). \end{equation} The assumptions on $s$, $p$, $q$ imply that $F\in B^{s-1}_{p, q}$ (Lemma \ref{at3}). We may now argue as follows. If $\varphi$ solves \eqref{at2}, then $\nabla\varphi\in B^{s-1}_{p, q}$, and thus $\varphi\in B^s_{p, q}$ (Lemma \ref{at4}). Next, since $u, \varphi\in W^{1,p}\cap L^{\infty}$, we find that \begin{equation*} \nabla(u\, e^{-\imath\varphi})=\nabla u\, e^{-\imath\varphi}-\imath u\, e^{-\imath\varphi}\nabla\varphi=\imath u\, e^{-\imath\varphi}(u\wedge\nabla u-\nabla\varphi)=0. \end{equation*} Thus $u\, e^{-\imath\varphi}$ is constant, and therefore $\varphi$ is, up to an appropriate additive constant, a $B^s_{p, q}$ phase of $u$. There is a flaw in the above. Indeed, \eqref{at2} need not have a solution. In ${\mathbb T}^n$, the necessary and sufficient conditions for the solvability of \eqref{at2} are\footnote{ This is easily seen by an inspection of the Fourier coefficients.} \begin{equation} \label{at5} \int_{{\mathbb T}^n}F=\widehat F(0)=0 \end{equation} and \begin{equation} \label{at1} \operatorname{curl} F=0. \end{equation} Clearly, \eqref{at5} holds.\footnote{ Expand $u\wedge\nabla u$ in Fourier series.} We complete Case \ref{Y} by noting that \eqref{at1} holds in the relevant range of $s$, $p$, $q$ and $n$ (Lemma \ref{at6}). \end{proof} \begin{remark} \label{aa1} We briefly discuss the lifting problem when $s\le 0$. For such $s$, distributions in $B^s_{p,q}$ need not be integrable functions, and thus the meaning of the equality $u=e^{\imath\varphi}$ is unclear. We therefore address the following reasonable version of the lifting problem: let $u:\Omega\to {\mathbb S}^1$ be a measurable function such that $u\in B^s_{p, q}(\Omega)$. Is there any $\varphi\in L^1_{loc}\cap B^s_{p,q}(\Omega ; {\mathbb R})$ such that $u=e^{\imath\varphi}$? Let us note that the answer is trivially positive when $s<0$, $1\le p<\infty$, $1\le q\le\infty$. Indeed, let $\varphi$ be any bounded measurable lifting of $u$. Then $\varphi\in B^s_{p, q}$, since $L^\infty\hookrightarrow B^s_{p,q}$ when $s<0$ (see Lemma \ref{ia2}). \end{remark} \section{Negative cases} \label{neg} \begin{case} \label{B} {\it Range.} $0<s<1$, $1\le p< \infty$, $1\le q<\infty$, $n\ge 2$, and $1\leq sp <n$. Or $0<s<1$, $1\le p< \infty$, $q=\infty$, $n\ge 2$, and $1< sp <n$. \smallskip \noindent {\it Conclusion.} $B^s_{p,q}(\Omega ; {\mathbb S}^1)$ does not have the lifting property. \end{case} \begin{proof} We want to show that there exists a function $u\in B^{s}_{p,q}$ such that $u\neq e^{{\imath} \varphi}$ for any $\varphi \in B^{s}_{p,q}$. For sufficiently small $\varepsilon>0$, set $ s_1:=s/(1-\varepsilon)$ and $p_1:=(1-\varepsilon)p$. By Lemma \ref{Besovemb}, we have $B^{s_1}_{p_1,q_1}\not\hookrightarrow B^s_{p,q}$ (for any $q_1$). We will use later this fact for $q_1:=(1-\varepsilon)q$. Let $\psi\in B^{s_1}_{p_1,q_1}\setminus B^s_{p,q}$ and set $u:=e^{{\imath} \psi}$. Then $u\in B^{s_1}_{p_1,q_1}\cap L^{\infty}$ (Lemma \ref{eipsi}) and thus $u\in B^{s}_{p,q}$ (Lemma \ref{gn}). We claim that there is no $\varphi\in B^s_{p,q}$ such that $u=e^{{\imath} \varphi}$. Argue by contradiction. Since $u=e^{\imath\varphi}=e^{\imath\psi}$, the function $(\varphi-\psi)/2\pi$ belongs to $(B^s_{p,q}+B^{s_1}_{p_1,q_1})(\Omega ; {\mathbb Z})$. By Lemma \ref{Eunicite}, this implies that $\varphi-\psi$ is constant, and thus $\psi\in B^{s}_{p,q}$, which is a contradiction. \end{proof} \begin{case} \label{xa2} {\it Range.} $0<s<\infty$, $1\le p<\infty$, $1\le q< \infty$, $n\ge 2$, and $1\le sp<2$. Or $0<s<\infty$, $1\le p<\infty$, $ q= \infty$, $n\ge 2$, and $1<sp\le 2$. \smallskip \noindent {\it Conclusion.} $B^s_{p,q}(\Omega ; {\mathbb S}^1)$ does not have the lifting property. \end{case} \begin{proof} The proof is based on the example of a topological obstruction considering the case $n=2$. Consider the map $\displaystyle u(x)=\frac{x}{|x|}$, $\forall\, x\in{\mathbb R}^2$. We first prove that $u \in B^s_{p,q}(\Omega)$ for any smooth bounded domain $\Omega\subset{\mathbb R}^2$. We distinguish two cases: firstly, $q \le \infty$ and $sp <2$ and secondly, $q=\infty$ and $sp=2$. In the first case, let $s_1>s$ such that $s_1$ is not an integer and $1<s_1p<2$, which implies $W^{s_1,p}=B^{s_1}_{p,p} \hookrightarrow B^{s}_{p,q}$. Since $u \in W^{s_1,p}$ \cite[Section 4]{lss}, we find that $u\in B^{s}_{p,q}$. The second case is slightly more involved. By the Gagliardo-Nirenberg inequality (Lemma \ref{gn} below), it suffices to prove that $u \in B^2_{1,\infty}(\Omega)$. Using Proposition \ref{p2.4}, a sufficient condition for this to hold is \begin{equation} \label{qf1} \left\Vert \Delta^3_{h}u\right\Vert_{L^1({\mathbb R}^2)}\lesssim \left\vert h\right\vert^2,\ \forall\, h\in {\mathbb R}^2. \end{equation} Since $u$ is radially symmetric and $0$-homogeneous, this amounts to checking \color{black} that \begin{equation}\label{delta3l1} \|\Delta_{e_1}^3u\|_{L^1(\mathbb R^2)}<\infty. \end{equation} However, by the mean-value theorem, for all $\left\vert x\right\vert\ge 1$ we have \begin{equation}\label{delta3infty} |\Delta^3_{e_1} u(x)|\lesssim 1/|x|^3, \end{equation} while $\Delta^3_{e_1}u$ is bounded in $B(0,1)$ since $u$ is ${\mathbb S}^1$-valued. Using this fact and estimate \eqref{delta3infty}, we obtain \eqref{delta3l1}. We next claim that $u$ has no $B^s_{p,q}$ lifting in $\Omega$ provided $\Omega\subset{\mathbb R}^2$ is a smooth bounded domain containing the origin. Argue by contradiction, and assume that $u=e^{\imath\varphi}$ for some $\varphi\in B^s_{p,q}(\Omega)$. Let, as in \cite[p. 50]{lss}, $ \theta\in C^{\infty}({\mathbb R}^2\setminus ([0,\infty)\times \left\{0\right\}))$ be such that $e^{\imath\theta}=u$. Note that $\theta\in B^s_{p,q}(\omega)$ for every smooth bounded open set $\omega$ such that $\overline\omega\subset{\mathbb R}^2\setminus ([0,\infty)\times \left\{0\right\}))$. Since $(\varphi-\theta)/(2\pi)$ is ${\mathbb Z}$-valued, Lemma \ref{Eunicite} yields that $\varphi-\theta$ is constant a.e. in $\Omega\setminus ([0,\infty)\times \left\{0\right\})$. Thus, $\theta\in B^s_{p,q}(\Omega)$. Similarly, $\widetilde\theta\in B^s_{p,q}(\Omega)$, where $\widetilde\theta \in C^{\infty}({\mathbb R}^2\setminus ((-\infty,0]\times \left\{0\right\}))$ is such that $e^{\imath\widetilde\theta}=u$. We find that $(\theta-\widetilde\theta)/(2\pi)\in B^s_{p,q}(\Omega)$. However, this is a non constant integer-valued function. This contradicts Lemma \ref{Eunicite} and proves non existence of lifting in $B^s_{p, q}$. When $n\ge 3$, the above arguments lead to the following. Let $u(x)=\displaystyle\frac{(x_1, x_2)}{|(x_1, x_2)|}$, and let $\Omega\subset {\mathbb R}^n$ be a smooth bounded domain. Then $u\in B^s_{p, q}(\Omega ; {\mathbb S}^1)$ and, if $0\in\Omega$, then $u$ has no $B^s_{p, q}$ lifting. \end{proof} \section{Open cases} \label{ope} \begin{case} \label{xa1} {\it Range.} $s>1$, $1\le p<\infty$, $p<q<\infty$, $n\ge 3$, and $sp=2$. \smallskip \noindent {\it Discussion.} This case is complementary to Case \ref{Y}. In the above range, we conjecture that the conclusion of Case \ref{Y} still holds, i.e., that the space $B^s_{p,q}(\Omega ; {\mathbb S}^1)$ does not have the lifting property. The non restriction property (Proposition \ref{l7.26}) prevents us from extending the argument used in Case \ref{Y} to Case \ref{xa1}. \end{case} \begin{case} \label{Z} {\it Range.} $s=1$, $1\le p<\infty$, $1\le q<\infty$, $n\ge 3$, and $2\le p<n$. Or: $s=1$, $1\le p<\infty$, $ q=\infty$, $n\ge 3$, and $2< p\le n$. \smallskip \noindent {\it Discussion.} When $p=q=2$, $B^1_{2,2}(\Omega ; {\mathbb S}^1)=H^1(\Omega ; {\mathbb S}^1)$ does have the lifting property \cite[Lemma 1]{bethuelzheng}. The remaining cases are open. The major difficulty arises from the extension of Lemma \ref{at3} to the range considered in Case \ref{Z}. \end{case} \begin{case} \label{T} {\it Range.} $s=0$, $1\le p<\infty$, $1\le q<\infty$ (and arbitrary $n$). \smallskip \noindent {\it Discussion.} As explained in Remark \ref{aa1}, we consider only measurable functions $u:\Omega\to{\mathbb S}^1$. We let $B^0_{p, q}(\Omega ; {\mathbb S}^1):=\{ u:\Omega\to{\mathbb S}^1;\, u \text{ measurable and }u\in B^0_{p,q}\}$, and for $u$ in this space we are looking for a phase $\varphi\in L^1_{loc}\cap B^0_{p,q}$. Note that $B^0_{p,\infty}(\Omega ; {\mathbb S}^1)$ does have the lifting property. Indeed, in this case we have $L^\infty\subset B^0_{p,\infty}$ (Lemma \ref{ia2}) and then it suffices to argue as in the proof of Case \ref{aa1}. More generally, $B^0_{p,q}(\Omega ; {\mathbb S}^1)$ has the lifting property when $L^\infty\hookrightarrow B^0_{p,q}$.\footnote{ A special case of this is $p=q=2$, since $B^0_{2,2}=L^2$. Another special case is $1<p\le 2\le q$. Indeed, in that case we have $L^\infty\hookrightarrow L^p=F^0_{p,2}\hookrightarrow B^0_{p,q}$ \cite[Section 2.3.5, p. 51]{triebel2}, \cite[Section 2.3.2, Proposition 2, p. 47]{triebel2}.} The remaining cases are open. \end{case} \begin{case} \label{xa3} {\it Range.} $0<s\le 1$, $p=1/s$, $q=\infty$ (and arbitrary $n$). \smallskip \noindent {\it Discussion.} We do not know whether $B^s_{p,q}(\Omega ; {\mathbb S}^1)$ does have the lifting property. \end{case} \begin{case} \label{xa4} {\it Range.} $0<s\le 1$, $1< p<\infty$, $q=\infty$, $n\ge 3$, and $sp=n$. \smallskip \noindent {\it Discussion.} We do not know whether $B^s_{p,q}(\Omega ; {\mathbb S}^1)$ does have the lifting property. The difficulty common to Cases \ref{xa3} and \ref{xa4} is that in these ranges $B^s_{p,\infty}\not\subset\text{\rm VMO}$, and thus we are unable to rely on the strategy used in Cases \ref{X} and \ref{kc3}. \end{case} \section{Analysis in Besov spaces} ${}$ The results we state here are valid when $\Omega$ is a smooth bounded domain in ${\mathbb R}^n$, or $(0,1)^n$ or ${\mathbb T}^n$. However, in the proofs we will consider only one of these sets, the most convenient for the proof. \subsection{Embeddings} \label{ape} ${}$ \begin{lemma}\label{Besovemb} Let $0<s_1<s_0<\infty$, $1\le p_0<\infty$, $1\le p_1< \infty$, $1\le q_0\leq \infty$ and $1\le q_1\leq \infty$. Then the following hold. \begin{enumerate} \item If $q_0<q_1$, then $B^s_{p,q_0}\hookrightarrow B^s_{p,q_1}$. \item If $s_0-n/p_0=s_1-n/p_1$, then $ B^{s_0}_{p_0,q_0} \hookrightarrow B^{s_1}_{p_1,q_0}$. \item If $s_0-n/p_0>s_1-n/p_1$, then $ B^{s_0}_{p_0,q_0} \hookrightarrow B^{s_1}_{p_1,q_1}$. \item If $B^{s_0}_{p_0,q_0} \hookrightarrow B^{s_1}_{p_1,q_1}$, then $s_0-n/p_0\geq s_1-n/p_1$. \end{enumerate} Consequently, when $q_0\leq q_1$, \begin{equation} \label{equiv} B^{s_0}_{p_0,q_0} \hookrightarrow B^{s_1}_{p_1,q_1} \Longleftrightarrow s_0-\frac{n}{p_0}\geq s_1-\frac{n}{p_1}. \end{equation} \end{lemma} \begin{proof} For item 1, see \cite[Section 3.2.4]{triebel2}. For items 2 and 3, see \cite[Section 3.3.1]{triebel2} or \cite[Theorem 1, p. 82]{runstsickel}. Item 4 follows from a scaling argument. And \eqref{equiv} is an immediate consequence of items 1--4. \end{proof} For the next result, see e.g. \cite[Section 2.7.1, Remark 2, pp. 130-131]{triebel2}. \begin{lemma} \label{ka1} Let $s>0$, $1\le p<\infty$, $1\le q\le\infty$ be such that $sp>n$. Then $B^s_{p,q}(\Omega)\hookrightarrow C^0(\overline\Omega)$. \end{lemma} \begin{lemma} \label{ia2} Let $s< 0$, $1\le p<\infty$ and $1\le q\le \infty$. Then $L^\infty\hookrightarrow B^s_{p,q}$. Similarly, if $1\le p\le\infty$, then $L^\infty\hookrightarrow B^0_{p,\infty}$. \end{lemma} \begin{proof} We present the argument when $\Omega={\mathbb T}^n$. Let $f\in L^\infty$, with Fourier coefficients $(a_m)_{m\in{\mathbb Z}^n}$. Consider, as in Definition \ref{periodicbesov}, the functions \begin{equation*} f_j(x):=\sum_{m\in{\mathbb Z}^n}a_m\varphi_j(2\pi m)\, e^{2\imath \pi m\cdot x},\ \forall\, j\in{\mathbb N}. \end{equation*} By the (periodic version of) the multiplier theorem \cite[Section 9.2.2, Theorem, p. 267]{triebel2} we have \begin{equation} \label{kb1} \|f_j\|_{L^p}\lesssim \|f\|_{L^p},\ \forall\, 1\le p\le \infty,\ \forall\, j\in{\mathbb N}. \end{equation} We find that $\|f_j\|_{L^p}\lesssim \|f\|_{L^p}\le \|f\|_{L^\infty}$, and thus (by Definition \ref{periodicbesov}, and with the usual modification when $q=\infty$) \begin{equation*} \|f\|_{B^s_{p,q}}\lesssim \left(\sum_j 2^{sjq}\right)^{1/q}<\infty. \end{equation*} The second part of the lemma follows from a similar argument. The proof is left to the reader. \end{proof} An analogous proof leads to the following result. Details are left to the reader. \begin{lemma} \label{kc2} Let $s>0$, $1\le p<\infty$ and $1\le q\le\infty$. Then $B^s_{p,q}\hookrightarrow L^p$. More generally, if $k\in{\mathbb N}$, $s>k$, $1\le p<\infty$, and $1\le q\le \infty$, then $B^s_{p,q}\hookrightarrow W^{k, p}$. \end{lemma} \begin{lemma}\label{B-VMO} Let $0<s<\infty$, $1\le p<\infty$ and $1\le q<\infty$ be such that $sp=n$. Then $\displaystyle B^{s}_{p,q} \hookrightarrow \text{\rm VMO}$.\\ Same conclusion if $0<s<\infty$, $1\le p<\infty$ and $q=\infty$ are such that $sp>n$. \end{lemma} \begin{proof} Assume first that $q<\infty$. Let $p_1>\max\left\{n,p,q\right\}$ and set $s_1:=n/{p_1}$. By Lemma \ref{Besovemb} and the fact that $s_1$ is not an integer, we have \begin{equation*} B^s_{p,q}\hookrightarrow B^{s_1}_{p_1,q}\hookrightarrow B^{s_1}_{p_1,p_1}=W^{s_1,p_1}. \end{equation*} It then suffices to invoke the embedding \begin{equation*} W^{s_1,p_1}\hookrightarrow \text{\rm VMO}\mbox{ when }s_1p_1=n \ \text{\cite[Example 2, p. 210]{brezisnirenberg1}}.\end{equation*} The case where $q=\infty$ is obtained via the first part of the proof. Indeed, it suffices to choose $0<s_1<\infty$, $1\le p_1<\infty$ and $0<q_1<\infty$ such that $s_1p_1=n$ and $B^s_{p,q}\hookrightarrow B^{s_1}_{p_1,q_1}$. Such $s_1$, $p_1$ and $q_1$ do exist, by Lemma \ref{Besovemb}. \end{proof} For the following special case of the Gagliardo-Nirenberg embeddings, see e.g. \cite[Remark 1, pp. 39-40]{runstsickel}. \begin{lemma} \label{gn} Let $0<s<\infty$, $1\le p<\infty$, $1\le q\leq \infty$, and $0<\theta<1$. Then $B^s_{p,q}\cap L^{\infty}\hookrightarrow B^{\theta s}_{p/\theta,q/\theta}$. \end{lemma} \subsection{Restrictions} \label{mo6} ${}$ {\it Captatio benevolenti\ae}. Let $f\in L^1({\mathbb R}^2)$. Then, for a.e., $y\in{\mathbb R}$, the restriction $f(\cdot, y)$ of $f$ to the line ${\mathbb R}\times\{y\}$ belongs to $L^1$. In this section and the next one, we examine some analogues of this property in the framework of Besov spaces. For this purpose, we first introduce some notation for partial functions. Let $\alpha\subset \{1,\ldots, n\}$ and set $\overline\alpha:=\{1,\ldots, n\}\setminus \alpha$. If $x=(x_1,\ldots, x_n)\in{\mathbb R}^n$, then we identify $x$ with the couple $(x_\alpha, x_{\overline\alpha})$, where $x_\alpha:=(x_j)_{j\in\alpha}$ and $x_{\overline\alpha}:=(x_j)_{j\in\overline\alpha}$. Given a function $f=f(x_1,\ldots, x_n)$, we let $f_\alpha=f_\alpha(x_\alpha)$ denote the partial function $x_{\overline\alpha}\mapsto f(x)$. Another useful notation: given an integer $m$ such that $1\le m\le n$, set \begin{equation*} I(n-m,n):=\{\alpha \subset \{1,\ldots, n\};\, \#\alpha=n-m\}. \end{equation*} Thus, when $\alpha\in I(n-m,n)$, $f_\alpha(x_\alpha)$ is a function of $m$ variables. When $q=p$, we have the following result. \begin{lemma} \label{oa1} Let $1\le m<n$. Let $s>0$ and $1\le p<\infty$. Let $f\in B^s_{p,p}({\mathbb R}^n)$. \begin{enumerate} \item Let $\alpha\in I(n-m,n)$. Then, for a.e. $x_\alpha\in{\mathbb R}^{n-m}$, we have $f_\alpha(x_\alpha)\in B^s_{p,p}({\mathbb R}^m)$. \item We have \begin{equation*} \|f\|_{B^s_{p,p}({\mathbb R}^n)}^p\sim \sum_{\alpha\in I(n-m,n)}\int_{{\mathbb R}^{n-m}}\|f_\alpha(x_\alpha)\|_{B^s_{p,p}({\mathbb R}^m)}^p\, dx_\alpha. \end{equation*} \end{enumerate} \end{lemma} \begin{proof} For the case where $m=1$, see \cite[Section 2.5.13, Theorem, (i), p. 115]{triebel2}. The general case is obtained by a straightforward induction on $m$. \end{proof} \begin{lemma} \label{mo7} Let $s>0$, $1\le p<\infty$ and $1\le q\le p$. Let $1\le m< n$ be an integer. Assume that $sp\ge m$ and let $f\in B^s_{p, q}({\mathbb T}^n)$. Then, for every $\alpha\in I(n-m,n)$ and for a.e. $x_\alpha\in{\mathbb T}^{n-m}$, the partial map $f_\alpha(x_\alpha)$ belongs to $\text{\rm VMO}({\mathbb T}^m)$. Same conclusion if $s>0$, $1\le p<\infty$ and $1\le q\le \infty$, and we have $sp>m$. Similar conclusions when $\Omega={\mathbb R}^n$ or $(0,1)^n$. \end{lemma} \begin{proof} In view of the Sobolev embeddings (Lemma \ref{Besovemb}), we may assume that $sp=m$ and $q=p$. By Lemma \ref{oa1} and Lemma \ref{B-VMO}, for a.e. $x_\alpha$ we have $f_\alpha(x_\alpha)\in B^s_{p,p}({\mathbb T}^m)\hookrightarrow \text{\rm VMO}({\mathbb T}^m)$. \end{proof} \begin{lemma} \label{ad1} Let $s>0$, $1\le p<\infty$ and $1\le q<\infty$. Let $M>s$ be an integer. Let $f\in B^s_{p,q}$. For $x'\in {\mathbb T}^{n-1}$, consider the partial map $v(x_n)=v_{x'}(x_n):=f(x',x_n)$, with $x_n\in{\mathbb T}$. Then there exists a sequence $(t_l)\subset (0,\infty)$ such that $t_l\to 0$ and for a.e. $x'\in{\mathbb T}^{n-1}$, we have \begin{equation} \label{ce1} \lim_{l\to\infty}\frac{\left\|\Delta_{t_l}^Mv\right\|_{L^p({\mathbb T})}}{t_l^{s}}=0. \end{equation} More generally, given a finite number of functions $f_j\in B^{s_j}_{p_j,q_j}$, with $s_j>0$, $1\le p_j<\infty$ and $1\le q_j<\infty$, and given an integer $M>\max_j s_j$, we may choose a common set $A$ of full measure in ${\mathbb T}^{n-1}$ and a sequence $(t_l)$ such that the analog of \eqref{ce1}, i.e., \begin{equation} \label{cf1} \lim_{l\to\infty}\frac{\left\|\Delta_{t_l}^Mf_j(x',\cdot)\right\|_{L^{p_j}({\mathbb T})}}{ t_l^{s_j}}=0, \end{equation} holds simultaneously for all $j$ and all $x'\in A$. \end{lemma} \begin{proof} We treat the case of a single function; the general case is similar. Set $g_t:=\left\|\Delta^M_{t e_n}f\right\|_{L^p}$. By \eqref{equivnormhomogrn}, we have $\int_0^1t^{-sq-1}g_t^q\, dt<\infty$, which is equivalent to $ \int_{1/2}^1\sum_{m\ge 0}2^{msq}g_{2^{-m}\sigma}^q\, d\sigma<\infty$. Therefore, there exists some $\sigma\in (1/2,1)$ such that \begin{equation} \label{ce2} \sum_{m\ge 0}2^{msq}g_{2^{-m}\sigma}^q <\infty. \end{equation} By \eqref{ce2} , we find that \begin{equation} \label{ce3} \lim_{m\to \infty}\frac{g_{2^{-m}\sigma}}{(2^{-m}\sigma)^s}=0. \end{equation} Using \eqref{ce3} we find that, along a subsequence $(m_l)$, we have \begin{equation*} \lim_{m\to \infty}\frac{\|\Delta_{2^{-m_l}\sigma}v\|_{L^p}}{(2^{-m_l}\sigma)^s}=0\quad\text{for a.e. }x'\in{\mathbb T}^{n-1}. \end{equation*} This implies \eqref{ce1} with $t_l:=2^{-m_l}\sigma$. \end{proof} \subsection{(Non) restrictions} ${}$ We now address the question whether, given $f\in B^s_{p, q}({\mathbb R}^2)$, we have $f(x, \cdot)\in B^s_{p, q}({\mathbb R})$ for a.e. $x\in{\mathbb R}$. This kind of questions can also be asked in higher dimensions. The answer crucially depends on the sign of $q-p$. We start with a simple result. \begin{proposition} \label{qh1} Let $s>0$ and $1\le q\le p<\infty$. Let $f\in B^s_{p,q}({\mathbb R}^2)$. Then for a.e. $x\in{\mathbb R}$ we have $f(x, \cdot)\in B^s_{p,q}({\mathbb R})$. \end{proposition} \begin{proof} Let $f\in B^s_{p,q}({\mathbb R}^2)$. Using \eqref{equivnormhomogrn} (part 2) and H\"older's inequality, we find that for every finite interval $[a,b]\subset{\mathbb R}$ and $M>s$ we have \begin{equation*} \begin{aligned} \int_a^b |f(x,\cdot)|_{B^s_{p,q}({\mathbb R})}^q\, dx& \sim \int_a^b\int_{\mathbb R} \frac 1{|h|^{sq+1}}\left(\int_{{\mathbb R}}|\Delta^M_{h e_2}f(x, y)|^p\, dy\right)^{q/p}\, dh dx\\ &\le (b-a)^{(p-q)/p}\, \int_{\mathbb R} \frac 1{|h|^{sq+1}}\left(\int_{[a,b]\times{\mathbb R}}|\Delta^M_{h e_2}f(x, y)|^p\, dxdy\right)^{q/p}\, dh\\ &\lesssim |f|_{B^s_{p,q}({\mathbb R}^2)}^q<\infty\end{aligned} \end{equation*} whence the conclusion. \end{proof} When $q>p$, a striking phenomenon occurs. \begin{proposition} \label{l7.26} Let $s>0$ and $1\le p<q\le\infty$. Then there exists some compactly supported $f\in B^s_{p,q}({\mathbb R}^2)$ such that for a.e. $x\in (0,1)$ we have $f(x,\cdot)\not\in B^s_{p,\infty}({\mathbb R})$. In particular, for any $1\le r<\infty$ and a.e. $x\in (0,1)$ we have $f(x,\cdot)\not\in B^s_{p, r}({\mathbb R})$. \end{proposition} Before proceeding to the proof, let us note that if $f\in B^s_{p,q}({\mathbb R}^2)$ then $f\in L^p({\mathbb R}^2)$, and thus the partial function $f(x,\cdot)$ is a well-defined element of $L^p({\mathbb R})$ for a.e. $x$. \begin{proof} Since $B^s_{p, q}({\mathbb R}^2)\subset B^s_{p,\infty}({\mathbb R}^2)$, $\forall\, q$, we may assume that $q<\infty$. We rely on the characterization of Besov spaces in terms of smooth wavelets, as in Section \ref{qa6}. We start by explaining the construction of $f$. Let $\psi_F$ and $\psi_M$ be as in Section \ref{qa6}. With no loss of generality, we may assume that $\operatorname{supp}\psi_M\subset [0,a]$ with $a\in{\mathbb N}$. Consider $(\alpha, \beta)\subset (0,a)$ and $\gamma>0$ such that $\psi_M\ge\gamma$ in $[\alpha, \beta]$. Set $\delta:=\beta-\alpha>0$ and consider some integer $N$ such that $[0,1]\subset [\alpha-N\, \delta, \beta+N\, \delta]$. We look for an $f$ of the form \begin{equation} \label{qb5} f=\sum_{\ell=-N}^N\sum_{j\ge j_0} g^\ell_j, \end{equation} with \begin{equation} \begin{aligned} \label{qb6} g^\ell_j(x,y)=\mu_j\, 2^{-j(s-2/p)}\sum_{m_1\in I_j} &\psi_M(2^j x-m_1-\ell\, \delta)\\ &\times\psi_M(2^j y-m_1-2^{j+1}\, \ell\, a-\ell\, \delta). \end{aligned} \end{equation} Here, the set $I_j$ satisfying \begin{equation} \label{qb7} I_j\subset \{ 0, 1,\ldots, 2^j\}, \end{equation} the integer $j_0$ and the coefficients $\mu_j>0$ will be defined later. We consider the partial sums $f^\ell_J:=\sum_{j=j_0}^J g^\ell_j$. Clearly, we have $f^\ell_J\in C^k$ and, provided $j_0$ is sufficiently large, \begin{equation*} \sup f^\ell_J\subset K_l:=[-N\, \delta , 5/4]\times [2\ell\, a-1/4, (2\ell +1)\, a+1/4]. \end{equation*} We next note that the compacts $K_\ell$ are mutually disjoint. Using Proposition \ref{p2.4} item 2, we easily find that \begin{equation} \label{qb9} \left\|\sum_{\ell=-N}^N f^\ell_J \right\|_{B^s_{p,q}({\mathbb R}^2)}^q\sim \sum_{\ell=-N}^N \left\| f^\ell_J \right\|_{B^s_{p,q}({\mathbb R}^2)}^q. \end{equation} On the other hand, if $\psi_M$ and $\psi_F$ are wavelets such that Proposition \ref{qb1} holds, then so are $\psi_F(\cdot-\lambda)$ and $\psi_M(\cdot-\lambda)$, $\forall\, \lambda\in{\mathbb R}$ \cite[Theorem 1.61 {\it (ii)}, Theorem 1.64]{triebel06}. Combining this fact with \eqref{qb9}, we find that \begin{equation} \label{qc1} \left\|\sum_{\ell=-N}^N f^\ell_J \right\|_{B^s_{p,q}({\mathbb R}^2)}^q\sim \sum_{j=j_0}^J \left( \# I_j\, (\mu_j)^p\right)^{q/p}. \end{equation} We now make the size assumption \begin{equation} \label{qc2} \sum_{j=j_0}^\infty \left( \# I_j\, (\mu_j)^p\right)^{q/p}<\infty. \end{equation} By \eqref{qc1} and \eqref{qc2}, we see that the formal series in \eqref{qb5} defines a compactly supported $f\in B^s_{p,q}({\mathbb R}^2)$, with $\sum_{\ell=-N}^N f^\ell_J\to f$ in $B^s_{p,q}({\mathbb R}^2)$ (and therefore in $L^p({\mathbb R}^2)$) as $J\to\infty$. We next investigate the $B^s_{p,\infty}$ norm of the restrictions $f^\ell_J (x,\cdot)$. As in \eqref{qb9}, we have \begin{equation} \label{qc3} \left\|\sum_{\ell=-N}^N f^\ell_J (x,\cdot)\right\|_{B^s_{p,\infty}({\mathbb R})}\sim \sum_{\ell=-N}^N\|f^\ell_J(x,\cdot)\|_{B^s_{p,\infty}({\mathbb R})}. \end{equation} Rewriting \eqref{qb6} as \begin{equation} \label{qc4} \begin{aligned} g^\ell_j(x,y)=\mu_j\, 2^{-j(s-1/p)}\, 2^{j/p}\sum_{m_1\in I_j} & \psi_M(2^j x-m_1-\ell\, \delta)\\ &\times \psi_M(2^j y-m_1-2^{j+1}\, \ell\, a -\ell\, \delta), \end{aligned} \end{equation} we obtain \begin{equation} \label{qc5} \|f^\ell_J(x,\cdot)\|_{B^s_{p,\infty}({\mathbb R})}^p\sim \sup_{j_0\le j\le J} 2^j\, (\mu_j)^p\, \sum_{m_1\in I_j}|\psi_M (2^j\, x- m_1-\ell\, \delta)|^p. \end{equation} We now make the size assumption \begin{equation} \label{qc6} \sup_{j\ge j_0} 2^j\, (\mu_j)^p\, \sum_{\ell=-N}^N\sum_{m_1\in I_j}|\psi_M (2^j\, x- m_1-\ell\, \delta)|^p=\infty,\ \forall\, x\in [0,1]. \end{equation} Then we claim that for a.e. $x\in (0,1)$ we have \begin{equation} \label{qc7} f(x,\cdot)\not\in B^s_{p,\infty}({\mathbb R}). \end{equation} Indeed, since $\sum_{\ell=-N}^N f^\ell_J\to f$ in $L^p ({\mathbb R}^2)$, for a.e. $x\in{\mathbb R}$ we have \begin{equation} \label{qc8} \sum_{\ell=-N}^\ell f^\ell_J(x,\cdot)\to f(x,\cdot)\ \text{in }L^p({\mathbb R}). \end{equation} We claim that for every $x\in [0,1]$ such that \eqref{qc8} holds, we have $f(x,\cdot)\not\in B^s_{p,\infty}({\mathbb R})$. Indeed, on the one hand \eqref{qc6} implies that for some $\ell$ we have $\lim_{J\to\infty}\|f^\ell_J (x,\cdot)\|_{B^s_{p,\infty}({\mathbb R})}=\infty$. We assume e.g. that this holds when $\ell=0$. Thus \begin{equation} \label{qc80} \sup_{j\ge j_0} 2^j\, (\mu_j)^p\, \sum_{m_1\in I_j}|\psi_M (2^j\, x- m_1)|^p=\infty. \end{equation} On the other hand, assume by contradiction that $f(x,\cdot)\in B^s_{p,\infty}({\mathbb R})$. Then we may write $f(x,\cdot)$ as in \eqref{qb3}, with coefficients as in \eqref{qb40}. In particular, taking into account the explicit formula of $g^\ell_j$ and the fact that $\sum_{\ell=-N}^N f^\ell_J(x,\cdot)\to f(x,\cdot)$ in $L^p({\mathbb R})$, we find that for $k\ge j_0$ and $m_1\in I_j$ we have \begin{equation} \label{qc800} \begin{aligned} \mu^{k, \{ M\}}_{m_1}(f(x,\cdot))&=\mu^{k, \{ M\}}_{m_1}\left(\sum_{j=j_0}^J g^0_j (x, \cdot)\right)=\mu^{k, \{ M\}}_{m_1}(g^0_k (x,\cdot))\\ &=2^{k/p}\, \mu_k\, \psi_M(2^k\, x-m_1),\ \forall\, J\ge k. \end{aligned} \end{equation} We obtain a contradiction combining \eqref{qc80}, \eqref{qc800} and Corollary \ref{qb400}. It remains to construct $I_j$ and $\mu_j$ satisfying \eqref{qb7}, \eqref{qc2} and \eqref{qc6}. We will let $I_j=\llbracket s_j, t_j\rrbracket$, with $0\le s_j\le t_j\le 2^j$ integers to be determined later. Set $t:=q/p \in (1,\infty)$ and \begin{equation*} \mu_j:=\left(\frac 1{(t_j-s_j+1)\, j^{1/t}\, \ln j}\right)^{1/p}. \end{equation*} Clearly, \eqref{qb7} and \eqref{qc2} hold. It remains to define $I_j$ in order to have \eqref{qc6}. Consider the dyadic segment $L_j:=[s_j/2^j, t_j/2^j]$. We claim that \begin{equation} \label{qa11} \sum_{\ell=-N}^N\sum_{m_1\in I_j}|\psi_M (2^j\, x- m_1-\ell\, \delta)|^p\ge \gamma^p,\ \forall\, x\in L_j. \end{equation} Indeed, let $m_1\in [s_j, t_j]$ be the integer part of $2^j\, x$. By the definition of $\delta$ and by choice of $N$, there exists some $\ell\in \llbracket -N, N\rrbracket$ such that $\alpha\le 2^j\, x- m_1-\ell\, \delta\le \beta$, whence the conclusion. By the above, \eqref{qc6} holds provided we have \begin{equation} \label{qc60} \sup_{j\ge j_0}2^j\, (\mu_j)^p\, \ensuremath{\mathonebb{1}}_{L_j(x)}=\infty,\ \forall\, x\in [0,1]. \end{equation} We next note that \begin{equation} \label{qc600} 2^j\, (\mu_j)^{p}\sim \frac 1{|L_j|\, j^{1/t}\, \ln j}=\frac {u_j}{|L_j|}, \end{equation} where $u_j:=1/(j^{1/t}\, \ln j)$ satisfies \begin{equation} \label{qc6000} \sum_{j\ge j_0}u_j=\infty. \end{equation} In view of \eqref{qc600} and \eqref{qc6000}, existence of $I_j$ satisfying \eqref{qc60} is a consequence of Lemma \ref{tempSeq} below. The proof of Proposition \ref{l7.26} is complete.\end{proof} \begin{lemma}\label{tempSeq} Consider a sequence $(u_j)$ of positive numbers such that $\sum_{j\ge j_0}u_j=\infty$. Then there exists a sequence $(L_j)$ of dyadic intervals $L_j=[s_j/2^j, t_j/2^j]$, such that: \begin{enumerate} \item $s_j, t_j\in{\mathbb N}$, $0\le s_j< 2^j$. \item $|L_j|=o(u_j)$ as $j\to\infty$. \item Every $x\in [0,1]$ belongs to infinitely many $L_j$'s. \end{enumerate} \end{lemma} \begin{proof} Consider a sequence $(v_j)$ of positive numbers such that $\sum_{j\ge j_0}v_j\, u_j=\infty$ and $v_j\to 0$. Let $L_{j_0}$ be the largest dyadic interval of the form $[0, t_{j_0}/2^{j_0}]$ of length $\le v_{j_0}\, u_{j_0}$. This defines $s_{j_0}=0$ and $t_{j_0}$. Assuming $L_j=[s_j/2^j, t_j/2^j]=[a_j, b_j]$ constructed for some $j\ge j_0$, one of the following two occurs. Either $b_j<1$ and then we let $L_{j+1}$ be the largest dyadic interval of the form $[2 t_{j}/2^{j+1}, t_{j+1}/2^{j+1}]$ such that $|L_{j+1}|\le v_{j+1}\, u_{j+1}$. Or $b_j\ge 1$, and then we let $L_{j+1}$ be the largest dyadic interval of the form $[0, t_{j+1}/2^{j+1}]$ such that $|L_{j+1}|\le v_{j+1}\, u_{j+1}$. Using the assumption $\sum_{j\ge j_0}v_j\, u_j=\infty$ and the fact that $|L_j|\ge v_j\, u_j-2^{-j}$, we easily find that for every $j\ge j_0$ there exists some $k>j$ such that $L_k=[a_k, b_k]$ satisfies $b_k\ge 1$, and thus the intervals $L_j$ cover each point $x\in [0,1]$ infinitely many times. \end{proof} \begin{remark} \label{r10} Following a suggestion of the first author, Brasseur investigated the non restriction property established in Proposition \ref{l7.26}. In \cite{brasseur} (which is independent of the present work), Brasseur extends Proposition \ref{l7.26} to the full range $0<p<q\le \infty$; the construction is somewhat similar to ours (based on the size of the coefficients $\mu_j$ in the decomposition \eqref{qb6}), but relying on a different decomposition (subatomic instead of wavelets). \cite{brasseur} also contains an interesting positive result: it exhibits function spaces $X$ intermediate between $B^s_{p,q}({\mathbb R})$ and $\displaystyle\bigcup_{\varepsilon>0}B^{s-\varepsilon}_{p, q}({\mathbb R})$ such that, if $f\in B^s_{p,q}({\mathbb R}^2)$, then for a.e. $x\in{\mathbb R}$ we have $f(x,\cdot)\in X$. \end{remark} \subsection{Poincar\'e type inequalities} ${}$ The next Poincar\'e type inequality for Besov spaces is certainly well-known, but we were unable to find a reference in the literature. \begin{lemma} \label{ad2} Let $0<s<1$, $1\leq p<\infty$, and $1\le q\le\infty$. Then we have \begin{equation} \label{PBesov} \left\Vert f-\fint f\right\Vert_{L^p}\lesssim \left\vert f\right\vert_{B^{s}_{p,q}},\quad \forall\, f:\Omega\to{\mathbb R}\text{ measurable function}. \end{equation} \end{lemma} \noindent Recall (Proposition \ref{p2.4}) that the semi-norm in \eqref{PBesov} is given by \begin{equation} \label{aa4} |f|_{B^s_{p,q}}=|f|_{B^s_{p,q}({\mathbb R}^n)}:=\left(\int_{{\mathbb R}^n}|h|^{-sq}\|\Delta_h f\|_{L^p}^q\, \frac{dh}{|h|^n}\right)^{1/q} \end{equation} when $q<\infty$, with the obvious modifications when $q=\infty$ or ${\mathbb R}^n$ is replaced by $\Omega$. \begin{proof} By \eqref{homoglp}, we have $\|f\|_{B^s_{p,q}}\sim \|f\|_{L^p}+|f|_{B^s_{p,q}}$. Recall that the embedding $B^{s}_{p,q}\hookrightarrow L^p$ is compact \cite[Theorem 3.8.3, p. 296]{triebel1}. From this we infer that \eqref{PBesov} holds for every function $f\in B^{s}_{p,q}$. Indeed, assume by contradiction that this is not the case. Then there exists a sequence of functions $(f_j)_{j\geq 1}\subset B^s_{p,q}$ such that, for every $j$, \begin{equation*} 1=\left\Vert f_j-\fint f_j\right\Vert_{L^p}\geq j \left\vert f_j\right\vert_{B^{s}_{p,q}}. \color{black} \end{equation*} Set $g_j:={f_j-\fint f_j}$. Then, up to a subsequence, we have $g_j\to g$ in $L^p$, where $\|g\|_{L^p}=1$ and $\int g=0$. We claim that $g$ is constant in $\Omega$ (and thus $g=0$). Indeed, by the Fatou lemma, for every $h\in {\mathbb R}^n$ we have \begin{equation} \label{aa3} \|\Delta_hg\|_{L^p}\le \liminf \|\Delta_hg_j\|_{L^p}=\liminf \|\Delta_hf_j\|_{L^p}. \end{equation} By \eqref{aa4}, \eqref{aa3} and the Fatou lemma, we have \begin{equation*} |g|_{B^s_{p,q}}\le\liminf |g_j|_{B^s_{p,q}}=\liminf |f_j|_{B^s_{p,q}}=0; \end{equation*} thus $g=0$, as claimed. This contradicts the fact that $\|g\|_{L^p}=1$. Let us now establish \eqref{PBesov} only assuming that $|f|_{B^s_{p,q}}<\infty$. We start by reducing the case where $q=\infty$ to the case where $q<\infty$. This reduction relies on the straightforward estimate \begin{equation*} |f|_{B^\sigma_{p,r}}\lesssim |f|_{B^s_{p,\infty}},\quad \forall\, 0<\sigma<s,\ \forall\, 0<r<\infty. \end{equation*} So let us assume that $q<\infty$. For every integer $k\geq 1$, let $\Phi_k:{\mathbb R}\rightarrow {\mathbb R}$ be given by \begin{equation*} \Phi_k(t):= \begin{cases} t, & \mbox{ if }\left\vert t\right\vert \leq k\\ -k, & \mbox{ if }t\leq -k\\ k, & \mbox{ if }t\geq k \end{cases}. \end{equation*} Clearly, $\Phi_k$ is $1$-Lipschitz, so that \eqref{aa4} easily yields \begin{equation} \label{controlphik} \left\vert \Phi_k(f)\right\vert_{B^s_{p,q}}\leq \left\vert f\right\vert_{B^s_{p,q}} \end{equation} and (by dominated convergence, using $q<\infty$ and \eqref{aa4}) \begin{equation} \label{convphikf} \lim_{k\rightarrow \infty} \left\vert \Phi_k(f)-f\right\vert_{B^{s}_{p,q}}=0. \end{equation} Since $\Phi_k(f)\in L^{\infty}(\Omega)\subset L^p(\Omega)$, one has $\Phi_k(f)\in B^s_{p,q}$ for every $k$. Therefore, \eqref{PBesov} and \eqref{controlphik} imply \begin{equation} \label{phikck} \left\Vert \Phi_k(f)-c_k\right\Vert_{L^p}\lesssim \left\vert \Phi_k(f)\right\vert_{B^{s}_{p,q}}\le \left\vert f\right\vert_{B^s_{p,q}} \end{equation} with $c_k:=\fint \Phi_k(f)$. Thanks to \eqref{convphikf}, we may pick up an increasing sequence of integers $(\lambda_k)_{k\geq 1}$ such that, for every $k$, $\displaystyle \left\vert \Phi_{\lambda_{k+1}}(f)-\Phi_{\lambda_k}(f)\right\vert_{B^s_{p,q}}\leq 2^{-k}$. Applying \eqref{PBesov} to $ \Phi_{\lambda_{k+1}}(f)-\Phi_{\lambda_k}(f)$, one therefore has \begin{equation*} \left\Vert \left(\Phi_{\lambda_{k+1}}(f)-c_{\lambda_{k+1}}\right)-\left(\Phi_{\lambda_k}(f)-c_{\lambda_k}\right)\right\Vert_{L^p}\lesssim \left\vert \Phi_{\lambda_{k+1}}(f)-\Phi_{\lambda_k}(f)\right\vert_{B^s_{p,q}}\leq 2^{-k}, \end{equation*} which entails that $\displaystyle \Phi_{\lambda_k}(f)-c_{\lambda_k}\to g\text{ in }L^p$ as $k\to\infty$. Up to a subsequence, one can also assume that $ \displaystyle \Phi_{\lambda_k}(f)(x)-c_{\lambda_k}\to g(x)$ for a.e. $x\in \Omega$. Take any $x\in \Omega$ such that $\Phi_{\lambda_k}(f)(x)-c_{\lambda_k}\to g(x)$. Since $\displaystyle \Phi_{\lambda_k}(f)(x)\to f(x)$ as $k\to\infty$, one obtains \begin{equation} \label{ckc} \lim_{k\rightarrow \infty} c_{\lambda_k}=c\in {\mathbb C}. \end{equation} Finally, \eqref{phikck}, \eqref{ckc} and the Fatou lemma yield $\displaystyle \left\Vert f-c\right\Vert_{L^p}\lesssim \left\vert f\right\vert_{B^s_{p,q}}$, from which \eqref{PBesov} easily follows. \end{proof} We next state and prove a generalization of Lemma \ref{ad2}. \begin{lemma} \label{ad3} Let $0<s<1$, $1\le p<\infty$, $1\le q\le\infty$, and $\delta\in (0,1]$. Define \begin{equation} \label{ad4} |f|_{B^s_{p,q,\delta}}:=\left(\int_{|h|\le\delta}|h|^{-sq}\|\Delta_h f\|_{L^p}^q\, \frac{dh}{|h|^n}\right)^{1/q} \end{equation} when $q<\infty$, with the obvious modifications when $q=\infty$ or ${\mathbb R}^n$ is replaced by $\Omega$. Then we have \begin{equation} \label{ad5} \left\Vert f-\fint f\right\Vert_{L^p}\lesssim \left\vert f\right\vert_{B^{s}_{p,q, \delta}},\quad \forall\, f:\Omega\to{\mathbb R}\text{ measurable function}. \end{equation} \end{lemma} \begin{proof} Recall that \color{black} $\|f\|_{B^s_{p,q}}\sim \|f\|_{L^p}+|f|_{B^s_{p,q,\delta}}$ (Proposition \ref{p2.4}). We continue as in the proof of Lemma \ref{ad2}. \end{proof} We end with an estimate involving derivatives. \begin{lemma} \label{at4} Let $s>0$, $1< p<\infty$ and $1\le q\le\infty$. Let $f\in {\cal D}'(\Omega)$ be such that $\nabla f\in B^{s-1}_{p,q}(\Omega)$. Then $f\in B^s_{p, q}(\Omega)$ and \begin{equation} \label{at9} \left\| f-\fint f\right\|_{B^s_{p,q}}\lesssim \|\nabla f\|_{B^{s-1}_{p,q}}. \end{equation} \end{lemma} The above result is well-known, but we were unable to find it in the literature; for the convenience of the reader, we present the short argument when $\Omega={\mathbb T}^n$. \begin{proof} We use the notation in Proposition \ref{mm2} and the following result \cite[Lemma 2.1.1, p. 16]{chemin}: we have \begin{equation} \label{mm3} \|f_j\|_{L^p}\sim 2^{-j}\|\nabla f_j\|_{L^p},\quad \forall\, 1\le p\le \infty,\ \forall\, j\ge 1. \end{equation} By combining \eqref{mm3} with Proposition \ref{mm2}, we obtain, e.g. when $q<\infty$: \begin{equation} \label{mm4} \begin{aligned} \left\|f-a_0\right\|_{B^s_{p,q}}^q&=\left\|\sum_{j\ge 1}f_j\right\|_{B^s_{p,q}}^q\sim \sum_{j\ge 1}2^{sjq}\|f_j\|_{L^p}^q \\ &\lesssim \sum_{j\ge 1}2^{sjq}2^{-jq}\|\nabla f_j\|_{L^p}^q \sim \|\nabla f\|_{B^{s-1}_{p,q}}^q. \end{aligned} \end{equation} In particular, $f\in L^1$ (Lemma \ref{kc2}), and thus $a_0=\fint f$. Therefore, \eqref{mm4} is equivalent to \eqref{at9}. \end{proof} \begin{remark} \label{mn41} With more work, Lemma \ref{at4} can be extended to the case where $p=1$. Although this will not be needed here, we sketch below the argument. With the notation in Section \ref{mm6}, consider the Littlewood-Paley decomposition $f=\sum f_j$, with $f_j:=\sum a_m\varphi_j(2\pi m)e^{2\imath\pi m\cdot x}$. Note that the Littlewood-Paley decomposition of $\nabla f$ is simply given by \begin{equation} \label{mn7} \nabla f=\sum \nabla f_j. \end{equation} In the spirit of \cite[Lemma 2.1.1, p. 16]{chemin} (see also \cite[Proof of Lemma 1]{leta}), one may prove that we have the following analog of \eqref{mm3}: \begin{equation} \label{mn6} \|f_j\|_{L^p}\sim 2^{-j}\|\nabla f_j\|_{L^p},\quad \forall\, 1\le p\le \infty,\ \forall\, j\ge 1. \end{equation} Using Definition \ref{periodicbesov}, \eqref{mn7} and \eqref{mn6}, we obtain \eqref{mm4}. We conclude as in the proof of Lemma \ref{at4}. \end{remark} \subsection{Characterization of $B^s_{p,q}$ via extensions} \label{characext} ${}$ The type of results we present in this section are classical for functions defined on the whole ${\mathbb R}^n$ and for the harmonic extension. Such results were obtained by Uspenski\u\i{ } in the early sixties \cite{uspenskii}. For further developments, see \cite[Section 2.12.2, Theorem, p. 184]{triebel2}; see also Section \ref{chha}. When the harmonic extension is replaced by other extensions by regularization, the kind of results we present below were known to experts at least for maps defined on ${\mathbb R}^n$; see \cite[Section 10.1.1, Theorem 1, p. 512]{mazyanew} and also \cite{tracesoldnew} for a systematic treatment of extensions by smoothing. The local variants (involving extensions by averages in domains) we present below could be obtained by adapting the arguments we developed in a more general setting in \cite{tracesoldnew}, and which are quite involved. However, we present here a more elementary approach, inspired by \cite{mazyanew}, sufficient to our purpose. In what follows, we let $|\ |$ denote the $\|\ \|_{\infty}$ norm in ${\mathbb R}^n$. For simplicity, we state our results when $\Omega={\mathbb T}^n$, but they can be easily adapted to arbitrary $\Omega$. \begin{lemma} \label{ab1} Let $0<s<1$, $1\le p<\infty$, $1\le q\le\infty$, and $\delta\in (0,1]$. Set $ V_\delta:={\mathbb T}^n\times (0,\delta)$. \begin{enumerate} \item Let $F\in C^\infty(V_{\delta})$. If \begin{equation} \label{cg6} \left(\int_0^{\delta/2}\varepsilon^{q-sq}\|(\nabla F)(\cdot,\varepsilon)\|_{L^p}^q\, \frac{d\varepsilon}\varepsilon\right)^{1/q}<\infty \end{equation} (with the obvious modification when $q=\infty$), then $F$ has a trace $f\in B^s_{p,q}({\mathbb T}^n)$, satisfying \begin{equation} \label{ab2} |f|_{B^s_{p,q,\delta}}\lesssim \left(\int_0^{\delta/2}\varepsilon^{q-sq}\|(\nabla F)(\cdot,\varepsilon)\|_{L^p}^q\, \frac{d\varepsilon}{\varepsilon}\right)^{1/q}. \end{equation} \item Conversely, let $f\in B^s_{p,q}({\mathbb T}^n)$. Let $\rho\in C^\infty$ be a mollifier supported in $\{ |x|\le 1\}$ and set $F(x,\varepsilon):=f\ast\rho_\varepsilon(x)$, $x\in{\mathbb T}^n$, $0<\varepsilon<\delta$. Then \begin{equation} \label{cg1} \left(\int_0^\delta\varepsilon^{q-sq}\|(\nabla F)(\cdot,\varepsilon)\|_{L^p}^q\, \frac{d\varepsilon}{\varepsilon}\right)^{1/q}\lesssim |f|_{B^s_{p,q,\delta}}. \end{equation} \end{enumerate} \end{lemma} A word about the existence of the trace in item 1 above. We will prove below that for every $0<\lambda<\delta/4$ we have \begin{equation} \label{cg2} \left|F_{|{\mathbb T}^n\times\{\lambda\}}\right|_{B^s_{p,q}}\lesssim \left(\int_0^{\delta/2}\varepsilon^{q-sq}\|(\nabla F)(\cdot,\varepsilon)\|_{L^p}^q\, \frac{d\varepsilon}{\varepsilon}\right)^{1/q}. \end{equation} By Lemma \ref{ad2} and a standard argument, this leads to the existence, in $B^s_{p,q}$, of the limit $\lim_{\varepsilon\to 0}F(\cdot,\varepsilon)$. This limit is the trace of $F$ on ${\mathbb T}^n$ and clearly satisfies \eqref{ab2}. \begin{proof} For simplicity, we treat only the case where $q<\infty$; the case where $q=\infty$ is somewhat simpler and is left to the reader. We claim that in item 1 we may assume that $F\in C^\infty(\overline{V_\delta})$. Indeed, assume that \eqref{ab2} holds (with $\operatorname{tr} F=F(\cdot, 0)$) for such $F$. By Lemma \ref{ad2}, we have the stronger inequality $\left\|\operatorname{tr} F-\fint\operatorname{tr} F\right\|_{B^s_{p,q}}\lesssim I(F)$, where $I(F)$ is the integral in \eqref{cg6}. Then, by a standard approximation argument, we find that \eqref{ab2} holds for every $F$. So let $F\in C^\infty(\overline{V_\delta})$, and set $f(x):=F(x,0)$, $\forall\, x\in{\mathbb T}^n$. Denote by $I(F)$ the quantity in \eqref{cg6}. We have to prove that $f$ satisfies \begin{equation} \label{ab210} |f|_{B^s_{p,q}}\lesssim I(F). \end{equation} If $|h|\le\delta$, then \begin{equation} \label{cg4} |\Delta_hf(x)|\le \left|f(x+h)-F(x+h/2,|h|/2)\right|+\left|f(x)-F(x+h/2,|h|/2)\right|. \end{equation} By symmetry and \eqref{cg4}, the estimate \eqref{ab210} will follow from \begin{equation} \label{cg5} \left(\int_{|h|\le\delta}|h|^{-sq}\|f-F(\cdot+h/2,|h|/2)\|_{L^p}^q\, \frac{dh}{|h|^n}\right)^{1/q}\lesssim I(F). \end{equation} In order to prove \eqref{cg5}, we start from \begin{equation} \label{cg8} \begin{aligned} \left|F(x+h/2,|h|/2)-f(x)\right|&=\left|\int_0^1 (\nabla F)(x+th/2,t|h|/2)\cdot (h/2,|h|/2)\, dt\right|\\ &\le |h|\int_0^1|\nabla F(x+th/2,t|h|/2)|\, dt. \end{aligned} \end{equation} Let $J(F)$ denote the left-hand side of \eqref{cg5}. Using \eqref{cg8} and setting $r:=|h|/2$, we obtain \begin{equation} \label{ch1} \begin{aligned} [J(F)]^q&\le \int_{|h|\le\delta}|h|^{q-sq}\left(\int_0^1\|\nabla F(\cdot+th/2,t|h|/2)\|_{L^p}\, dt\right)^q\, \frac{dh}{|h|^n}\\ &=\int_{|h|\le\delta}|h|^{q-sq}\left(\int_0^1\|\nabla F(\cdot,t|h|/2)\|_{L^p}\, dt\right)^q\, \frac{dh}{|h|^n}\\ &\sim \int_0^{\delta/2}r^{q-sq-1}\left(\int_0^1\|\nabla F(\cdot,tr)\|_{L^p}\, dt\right)^q\, dr\\ &\sim \int_0^{\delta/2}r^{-sq-1}\left(\int_0^r\|\nabla F(\cdot,\sigma)\|_{L^p}\, d\sigma\right)^q\, dr \lesssim [I(F)]^q. \end{aligned} \end{equation} The last inequality is a special case of Hardy's inequality \cite[Chapter 5, Lemma 3.14]{steinweiss}, that we recall here when $\delta =\infty$.\footnote{ But the argument adapts to a finite $\delta$; see e.g. \cite[Proof of Corollary 7.2]{bousquetmironescu}.} Let $1\le q<\infty$ and $1<\rho<\infty$. If $G\in W^{1,1}_{loc}([0,\infty))$, then \begin{equation} \label{e04269} \int_0^\infty\frac{|G(r)-G(0)|^q}{r^\rho}\,dr\leq \left(\frac{q}{\rho-1}\right)^q\int_{0}^\infty\frac{|G'(r)|^q}{r^{\rho-q}}\,dr. \end{equation} We obtain \eqref{ch1} by applying \eqref{e04269} with $G'(r):=\|\nabla F(\cdot, r)\|_{L^p}$ and $\rho:=sq+1$. The proof of item 1 is complete. We next turn to item 2. We have \begin{equation} \label{kh2} \nabla F(x,\varepsilon)=\frac 1\varepsilon f\ast \eta_\varepsilon(x), \end{equation} where $\nabla$ stands for $(\partial_1,\ldots,\partial_n,\partial_{\varepsilon})$. Here, $\eta=(\eta^1,\ldots,\eta^{n+1})\in C^\infty({\mathbb T}^n ; {\mathbb R}^{n+1})$ is supported in $\{ |x|\le 1\}$ and is given in coordinates by \begin{equation} \label{kh3} \eta^j=\partial_j\rho, \ \forall\, j\in \llbracket 1, n\rrbracket,\ \eta^{n+1}=-\operatorname{div} (x\rho). \end{equation} Noting that $\int \eta=0$, we find that \begin{equation} \label{ch2} \begin{aligned} |\nabla F(x,\varepsilon)|&= \frac 1\varepsilon\left|\int_{|y|\le \varepsilon}(f(x-y)-f(x))\eta_\varepsilon(y)\, dy\right|\\ &\lesssim\frac 1{\varepsilon^{n+1}}\int_{|h|\le\varepsilon}|f(x+h)-f(x)|\, dh. \end{aligned} \end{equation} Integrating \eqref{ch2} and using Minkowski's inequality, we obtain \begin{equation} \label{ci1} \|\nabla F(\cdot,\varepsilon)\|_{L^p}\lesssim \frac 1{\varepsilon^{n+1}}\int_{|h|\le\varepsilon}\|\Delta_hf\|_{L^p}\, dh. \end{equation} Let $L(F)$ be the quantity in the left-hand side of \eqref{cg1}. Combining \eqref{ci1} with H\" older's inequality, we find that \begin{equation} \label{mj1} \begin{aligned} [L(F)]^q&\color{black} \lesssim \int_0^{\delta}\frac 1{\varepsilon^{nq+sq+1}}\left(\int_{|h|\le\varepsilon}\|\Delta_hf\|_{L^p}\, dh\right)^q\, d\varepsilon\\ &\lesssim \int_0^{\delta}\frac 1{\varepsilon^{nq+sq+1}}\varepsilon^{n(q-1)}\int_{|h|\le\varepsilon}\|\Delta_hf\|_{L^p}^q\, dh\, d\varepsilon\\ &\lesssim \int_{|h|\le\delta}|h|^{-sq}\|\Delta_hf\|_{L^p}^q\, \frac{dh}{|h|^n}=|f|_{B^s_{p,q,\delta}}^q,\end{aligned} \end{equation} i.e, \eqref{cg1} holds. \end{proof} In the same vein, we have the following result, involving the semi-norm appearing in Proposition \ref{p2.4}, more specifically the quantity \begin{equation} \label{kd4} |f|_{B^1_{p,q,\delta}}:=\left(\int_{|h|\le\delta}|h|^{-q}\|\Delta_h^2 f\|_{L^p}^q\, \frac{dh}{|h|^n}\right)^{1/q} \end{equation} when $q<\infty$, with the obvious modification when $q=\infty$. We first introduce a notation. Given $F\in C^2(V_\delta)$, we let $D^2_\#F$ denote the collection of the second order derivatives of $F$ which are either completely horizontal (that is of the form $\partial_j\partial_k F$, with $j,k\in\llbracket 1,n\rrbracket$), or completely vertical (that is $\partial_{n+1}\partial_{n+1}F$). \begin{lemma} \label{kb2} Let $1\le p<\infty$ and $1\le q\le\infty$. Let $F\in C^\infty(V_{\delta})$ and set \begin{equation*} M(F):=\left(\int_0^{\delta}\varepsilon^{q}\|(\nabla F)(\cdot,\varepsilon)\|_{L^{2p}}^{2q}\, \frac{d\varepsilon}\varepsilon\right)^{1/q} \end{equation*} and \begin{equation*} N(F):=\left(\int_0^{\delta}\varepsilon^{q}\left\|(D^2_\# F)(\cdot,\varepsilon)\right\|_{L^p}^q\frac{d\varepsilon}\varepsilon\right)^{1/q} \end{equation*} (with the obvious modification when $q=\infty$). \begin{enumerate} \item If $M(F)<\infty$ and $N(F)<\infty$, then $F$ has a trace $f\in B^1_{p,q}({\mathbb T}^n)$, satisfying \begin{equation} \label{kf1} \left\|f-\fint f\right\|_{L^p}\lesssim M(F)^{\frac 12}\end{equation} and \begin{equation} \label{kb3} |f|_{B^1_{p,q,\delta}}\lesssim N(F). \end{equation} \item Conversely, let $f\in B^1_{p,q}({\mathbb T}^n ; {\mathbb S}^1)$. Let $\rho\in C^\infty$ be an even mollifier supported in $\{ |x|\le 1\}$ and set $F(x,\varepsilon):=f\ast\rho_\varepsilon(x)$, $x\in{\mathbb T}^n$, $0<\varepsilon<\delta$. Then \begin{equation} \label{kb4} M(F)+N(F)\lesssim |f|_{B^1_{p,q,\delta}}. \end{equation} \end{enumerate} \end{lemma} The above result is inspired by the proof of \cite[ Section 10.1.1, Theorem 1, p. 512]{mazyanew}. The arguments we present also lead to a (slightly different) proof of Lemma \ref{ab1}. We start by establishing some preliminary estimates. We call $H\in{\mathbb R}^n\times{\mathbb R}$ \enquote{pure} if $H$ is either horizontal, or vertical, i.e., either $H\in{\mathbb R}^n\times\{0\}$ or $H\in \{0\}\times{\mathbb R}$. For further use, let us note the following fact, valid for $X\in V_\delta$ and $H\in{\mathbb R}^{n+1}$. \begin{equation} \label{ja2}H\text{ pure}\implies|D^2F(X)\cdot(H,H)|\lesssim |D^2_\#F(X)||H|^2. \end{equation} \begin{lemma} \label{jc1} Let $X, H$ be such that $[X, X+2H]\subset \overline{V_\delta}$. Let $F\in C^2(\overline{V_\delta})$. Then \begin{equation} \label{jc2} |\Delta_H^2F(X)|\le \int_0^2 \tau |D^2F(X+\tau H)\cdot(H,H)|\, d\tau. \end{equation} In particular, if $H$ is pure and we write $H=|H|K$, then \begin{equation} \label{jc3} |\Delta_H^2F(X)|\lesssim \int_0^{2|H|} t |D^2_\# F(X+tK)|\, dt. \end{equation} \end{lemma} \begin{proof} Set \begin{equation*} G(s):=F(X+(1-s)H)+F(X+(1+s)H), \ s\in [0,1], \end{equation*} so that $G\in C^2$ and in addition we have \begin{equation} \label{ja3} G'(0)=0,\ G''(s)=[D^2F(X+(1-s)H)+D^2F(X+(1+s)H)]\cdot(H,H), \end{equation} and \begin{equation} \label{ja4} \int_0^1(1-s)G''(s)\, ds=G(1)-G(0)-G'(0)=\Delta_H^2F(X). \end{equation} Estimate \eqref{jc2} is a consequence of \eqref{ja3} and \eqref{ja4} (using the changes of variable $\tau:=1\pm s$). In the special case where $H$ is pure, we rely on \eqref{ja2} and \eqref{jc2} and obtain \eqref{jc3} via the change of variable $t:=\tau|H|$. \end{proof} If we combine \eqref{jc3} (applied first with $H=(h,0)$, $h\in{\mathbb R}^n$, next with $H=(0,t)$, $t\in [0,\delta/2]$) with Minkowski's inequality, we obtain the two following consequences\footnote{ In \eqref{jb1}, we let $\Delta_h^2F(\cdot,\varepsilon):=F(\cdot+2h,\varepsilon)-2F(\cdot+h,\varepsilon)+F(\cdot,\varepsilon)$.} \begin{equation} \label{jb1} [h\in{\mathbb R}^n,\ 0\le\varepsilon\le\delta]\implies \|\Delta_h^2F(\cdot,\varepsilon)\|_{L^p}\lesssim |h|^2\|D^2_\#F(\cdot,\varepsilon)\|_{L^p}, \end{equation} and\footnote{ With the slight abuse of notation $\Delta_{te_{n+1}}^2F(\cdot,\varepsilon):=F(\cdot,\varepsilon+2t)-2F(\cdot,\varepsilon+t)+F(\cdot,\varepsilon)$.} \begin{equation} \label{jb2} \begin{aligned} [t, \varepsilon\ge 0,\ \varepsilon+2t\le\delta] \implies \|\Delta_{te_{n+1}}^2F(\cdot,\varepsilon)\|_{L^p}&\lesssim \int_0^{2t} r \|D^2_\#F(\cdot,\varepsilon+r)\|_{L^p}\, dr. \end{aligned} \end{equation} \begin{proof}[Proof of Lemma \ref{kb2}] We start by proving \eqref{kf1}. By Lemma \ref{ab1} (applied with $s=1/2$ and with $2p$ (respectively $2q$) instead of $p$ (respectively $q$)), $F$ has, on ${\mathbb T}^n$, a trace $\operatorname{tr} F\in B^{1/2}_{2p,2q}$. By Lemma \ref{ab1}, item $1$, and Lemma \ref{ad3}, we have \begin{equation*} \left\|\operatorname{tr} F-\fint \operatorname{tr} F\right\|_{L^p}\lesssim\left\|\operatorname{tr} F-\fint \operatorname{tr} F\right\|_{L^{2p}}\lesssim M(F)^{1/2} \end{equation*} i.e., \eqref{kf1} holds. We next establish \eqref{kb3}. Arguing as at the beginning of the proof of Lemma \ref{ab1}, one concludes that it suffices to prove \eqref{kb3} when $F\in C^\infty(\overline{V_\delta})$. So let us consider some $F\in C^\infty(\overline{V_\delta})$. We set $f(x)=F(x,0)$, $\forall\, x\in{\mathbb T}^n$. Then \eqref{kb3} is equivalent to \begin{equation} \label{kf2} |f|_{B^1_{p,q,\delta}}\lesssim N(F). \end{equation} We treat only the case where $q<\infty$; the case where $q=\infty$ is slightly simpler and is left to the reader. The starting point is the following identity, valid when $|h|\le\delta$ and with $t:=|h|$ \begin{equation} \label{jd1} \begin{aligned} \Delta_h^2f=&\Delta_{te_{n+1}/2}^2F(\cdot+2h,0)-2\Delta_{te_{n+1}/2}^2F(\cdot+h,0)+\Delta_{te_{n+1}/2}^2F(\cdot,0)\\ &+2\Delta_h^2F(\cdot, t/2)-\Delta_h^2F(\cdot, t). \end{aligned} \end{equation} By \eqref{jb1}, \eqref{jb2} and \eqref{jd1}, we find that \begin{equation} \label{jd2} \begin{aligned} \|\Delta_h^2f\|_{L^p}\lesssim & \int_0^{|h|}r\|D^2_\#F(\cdot, r)\|_{L^p}\, dr+|h|^2\|D^2_\#F(\cdot, |h|/2)\|_{L^p}\\ &+|h|^2\|D^2_\#F(\cdot, |h|)\|_{L^p}. \end{aligned} \end{equation} Finally, \eqref{jd2} combined with Hardy's inequality \eqref{e04269} (applied to the integral $\int_0^\delta$ and with $G'(r):=r\|D^2_\#F(\cdot, r)\|_{L^p}$ and $\rho:=q+1$) yields \begin{equation} \label{kf6} \begin{aligned} |f|_{B^1_{p,q,\delta}}^q&\lesssim \int_{|h|\le\delta}\frac 1{|h|^q}\left(\int_0^{|h|}r \left\| D^2_\#F(\cdot, r)\right\|_{L^p}\, dr\right)^q\, \frac{dh}{|h|^n}+[N(F)]^q\\ &\lesssim [N(F)]^q. \end{aligned} \end{equation} This implies \eqref{kf2} and completes the proof of item 1. We now turn to item 2. We claim that \begin{equation} \label{kg1} |f|_{B^{1/2}_{2p,2q,\delta}}\lesssim |f|_{B^1_{p,q,\delta}}^{1/2}. \end{equation} Indeed, it suffices to note the fact that $ |\Delta_h^2f|^{2p}\lesssim |\Delta_h^2f|^p$ (since $|f|=1$). By combining \eqref{kg1} with Lemma \ref{ab1}, we find that \begin{equation} \label{kg2} M(F)=\left(\int_0^{\delta}\varepsilon^{q}\|(\nabla F)(\cdot,\varepsilon)\|_{L^{2p}}^{2q}\, \frac{d\varepsilon}\varepsilon\right)^{1/q}\lesssim |f|_{B^1_{p,q,\delta}}. \end{equation} Thus, in order to complete the proof of \eqref{kb4}, it suffices to combine \eqref{kg2} with the following estimate \begin{equation} \label{kg3} N(F) \lesssim |f|_{B^1_{p,q,\delta}}, \end{equation} that we now establish. The key argument for proving \eqref{kg3} is the following second order analog of \eqref{ch2}: \begin{equation} \label{ki1} |D^2_\#F(x,\varepsilon)|\lesssim \frac 1{\varepsilon^{n+2}}\int_{|h|\le\varepsilon}|\Delta_h^2f(x-h)|\, dh. \end{equation} The proof of \eqref{ki1} appears in \cite[p. 514]{mazyanew}. For the sake of completeness, we reproduce below the argument. First, differentiating the expression defining $F$, we have \begin{equation} \label{ki2} \partial_j\partial_kF(x,\varepsilon)=\frac 1{\varepsilon^2}f\ast(\partial_j\partial_k\rho)_\varepsilon,\ \forall\, j,\, k\in\llbracket 1,n\rrbracket. \end{equation} Using \eqref{ki2} and the fact that $\partial_j\partial_k\rho$ is even and has zero average, we obtain the identity \begin{equation*} \partial_j\partial_kF(x,\varepsilon)=\frac 1{2\varepsilon^{n+2}}\int_{|h|\le\varepsilon}\partial_j\partial_k\rho(h/\varepsilon)\Delta_h^2f(x-h)\, dh, \end{equation*} and thus \eqref{ki1} holds for the derivatives $\partial_j\partial_kF$, with $j,\, k\in\llbracket 1,n\rrbracket$. We next note the identity \begin{equation} \label{ki4} F(x,\varepsilon)=\frac 1{2\varepsilon^n}\int \rho(h/\varepsilon)\Delta_h^2f(x-h)\, dh+f(x), \end{equation} which follows from the fact that $\rho$ is even. By differentiating twice \eqref{ki4} with respect to $\varepsilon$, we obtain that \eqref{ki1} holds when $j=k=n+1$. The proof of \eqref{ki1} is complete. Using \eqref{ki1} and Minkowski's inequality, we obtain \begin{equation} \label{mi1} \|D^2_\#F(\cdot,\varepsilon)\|_{L^p}\lesssim \frac 1{\varepsilon^{n+2}}\int_{|h|\le\varepsilon}\|\Delta_h^2f\|_{L^p}\, dh, \end{equation} which is a second order analog of \eqref{ci1}. Once \eqref{ci1} is obtained, we repeat the calculation leading to \eqref{mj1} and obtain \eqref{kg3}. The details are left to the reader. The proof of Lemma \ref{kb2} is complete. \end{proof} \begin{remark} \label{av1} One may put Lemmas \ref{ab1} and \ref{kb2} in the perspective of the theory of weighted Sobolev spaces. Let us start by recalling one of the striking achievements of this theory. As it is well-known, we have $\operatorname{tr} W^{1,1}({\mathbb R}^n_+)=L^1({\mathbb R}^{n-1})$, and, when $n\ge 2$, the trace operator has no linear continuous right-inverse $T:L^1({\mathbb R}^{n-1})\to W^{1,1}({\mathbb R}^n)$ \cite{gagliardo}, \cite{peetre}. The expected analogs of these facts for $W^{2,1}({\mathbb R}^n_+)$ are both wrong. More specifically, we have $\operatorname{tr} W^{2,1}({\mathbb R}^n_+)=B^1_{1,1}({\mathbb R}^{n-1})$ (which is a strict subspace of $W^{1,1}({\mathbb R}^{n-1}))$, and the trace operator has a linear continuous right inverse from $B^1_{1,1}({\mathbb R}^{n-1})$ into $W^{2,1}({\mathbb R}^n_+)$. These results are special cases of the trace theory for weighted Sobolev spaces developed by Uspenski\u\i {} \cite{uspenskii}. For a modern treatment of this theory, see e.g. \cite{tracesoldnew}. \end{remark} \subsection{Product estimates} ${}$ Lemma \ref{at3} below is a variant of \cite[Lemma D.2]{lss}. Here, $\Omega$ is either smooth bounded, or $(0, 1)^n$, or ${\mathbb T}^n$. \begin{lemma} \label{at3} Let $s>1$, $1\le p<\infty$ and $1\le q\le \infty$. If $u, v\in B^s_{p,q}\cap L^\infty(\Omega)$, then $u\nabla v\in B^{s-1}_{p,q}$. \end{lemma} \begin{proof} After extension to ${\mathbb R}^n$ and cutoff, we may assume that $u, v\in B^s_{p,q}\cap L^\infty$. It thus suffices to prove that $u, v\in B^s_{p,q}\cap L^\infty({\mathbb R}^n)$$\implies$$u\nabla v\in B^{s-1}_{p,q}({\mathbb R}^n)$. In order to prove the above, we argue as follows. Let $u=\sum u_j$ and $v=\sum v_j$ be the Littlewood-Paley decompositions of $u$ and $v$. Set \begin{equation*} f^j:=\sum_{k\le j}u_k\nabla v_j+\sum_{k<j}u_j\nabla v_k. \end{equation*} Since $\operatorname{supp}{\mathcal F} (u_k\nabla v_j)\subset B(0, 2^{\max\{ k, j\}+2})$, we find that $u\nabla v=\sum f^j$ is a Nikolski\u\i {} decomposition of $u\nabla v$; see Section \ref{mm8}. Assume e.g. that $q<\infty$. In view of Proposition \ref{mm9}, the conclusion of Lemma \ref{at3} follows if we prove that \begin{equation} \label{mn1} \sum 2^{(s-1)jq}\|f^j\|_{L^p}^q<\infty. \end{equation} In order to prove \eqref{mn1}, we rely on the elementary estimates \cite[Lemma 2.1.1, p. 16]{chemin}, \cite[formulas (D.8), (D.9), p. 71]{lss} \begin{equation} \label{mn2} \left\|\sum_{k\le j}u_k\right\|_{L^\infty}\lesssim \|u\|_{L^\infty}, \quad \forall\, j\ge 0, \end{equation} \begin{equation} \label{mn3} \left\| \sum_{k< j}\nabla v_k\right\|_{L^\infty}\lesssim 2^j\|v\|_{L^\infty}, \quad \forall\, j\ge 0, \end{equation} and \begin{equation} \label{mn4} \|\nabla v_j\|_{L^p}\lesssim 2^j\|v_j\|_{L^p},\quad \forall\, j\ge 0. \end{equation} By combining \eqref{mn2}-\eqref{mn4}, we obtain \begin{equation*} \begin{aligned} \sum 2^{(s-1)jq}\|f^j\|_{L^p}^q&\lesssim \sum 2^{(s-1)jq}\left(\left\|\sum_{k\le j}u_k\right\|_{L^\infty}^q\|\nabla v_j\|_{L^p}^q+\left\|\sum_{k<j}\nabla v_k\right\|_{L^\infty}^q\| u_j\|_{L^p}^q\right)\\ &\lesssim \|u\|_{L^\infty}^q\sum 2^{sjq}\|v_j\|_{L^p}^q+\|v\|_{L^\infty}^q\sum 2^{sjq}\|u_j\|_{L^p}^q \\ &\lesssim \|u\|_{L^\infty}^q\|v\|_{B^s_{p,q}}^q+\|v\|_{L^\infty}^q\|u\|_{B^s_{p,q}}^q, \end{aligned} \end{equation*} and thus \eqref{mn1} holds. \end{proof} \subsection{Superposition operators} ${}$ In this section, we examine the mapping properties of the operator \begin{equation*} T_\Phi,\ \psi\xmapsto{T_\Phi} \Phi\circ\psi. \end{equation*} We work in $\Omega$ smooth bounded, or $(0,1)^n$, or ${\mathbb T}^n$. The next result is classical and straightforward; see e.g. \cite[Section 5.3.6, Theorem 1]{runstsickel}. \begin{lemma} \label{eipsi} Let $0<s<1$, $1\le p<\infty$, and $1\le q<\infty$. Let $\Phi:{\mathbb R}^k\to{\mathbb R}^l$ be a Lipschitz function . Then $ T_\Phi$ maps $B^{s}_{p,q}(\Omega ; {\mathbb R}^k)$ into $B^{s}_{p,q}(\Omega ; {\mathbb R}^l)$. Special case: $\psi\mapsto e^{{\imath} \psi}$ maps $B^s_{p,q}(\Omega ; {\mathbb R})$ into $B^s_{p,q}(\Omega ; {\mathbb S}^1)$. In addition, when $q<\infty$, $T_\Phi$ is continuous. \end{lemma} For the next result, see \cite[Section 5.3.4, Theorem 2, p. 325]{runstsickel}. \begin{lemma} \label{ka2} Let $s>0$, $1\le p<\infty$ and $1\le q\le\infty$. Let $\Phi\in C^\infty({\mathbb R}^k ; {\mathbb R}^l)$. Then $T_\Phi$ maps $(B^{s}_{p,q}\cap L^\infty)(\Omega ; {\mathbb R}^k)$ into $(B^{s}_{p,q}\cap L^\infty)(\Omega ; {\mathbb R}^l)$. Special case: $\psi\mapsto e^{{\imath} \psi}$ maps $(B^{s}_{p,q}\cap L^\infty)(\Omega ; {\mathbb R})$ into $(B^{s}_{p,q}\cap L^\infty)(\Omega ; {\mathbb S}^1)$. \end{lemma} \subsection{Integer valued functions} ${}$ The next result is a cousin of \cite[Appendix B]{lss},\footnote{ The context there is the one of the Sobolev spaces.} but the argument in \cite{lss} does not seem to apply in our situation. Lemma \ref{Eunicite} can be obtained from the results in \cite{bbmuni}, but we present below a simpler direct argument. \begin{lemma}\label{Eunicite}Let $s>0$, $1\le p<\infty$ and $1\le q<\infty$ be such that $sp\ge 1$. Then the functions in $B^{s}_{p,q}(\Omega ;\mathbb{Z})$ are constant. Same result when $s>0$, $1\le p<\infty$, $q=\infty$ and $sp>1$. The same conclusion holds for functions in $\sum_{j=1}^k B^{s_j}_{p_j,q_j}(\Omega ; \mathbb{Z})$, provided we have for all $j\in\llbracket 1,k\rrbracket$: either $s_jp_j=1$ and $1\le q_j<\infty$, or $s_jp_j>1$ and $1\le q_j\le\infty$. \end{lemma} \begin{proof} The case where $n=1$ is simple. Indeed, by Lemma \ref{B-VMO} we have $B^{s}_{p,q}\hookrightarrow \text{\rm VMO}$ (and similarly $\sum_{j=1}^k B^{s_j}_{p_j,q_j}\hookrightarrow \text{\rm VMO}$). The conclusion follows from the fact that $\text{\rm VMO}((0,1) ;{\mathbb Z})$ functions are constant \cite[Step 5, p. 229]{brezisnirenberg1}. We next turn to the general case. Let $f=\sum_{j=1}^kf_j$, with $f_j\in B^{s_j}_{p_j,q_j}(\Omega ; {\mathbb Z})$, $\forall\, j\in\llbracket 1, k\rrbracket$. In view of the conclusion, we may assume that $\Omega =(0,1)^n$. By the Sobolev embeddings, we may assume that for all $j$ we have $s_jp_j=1$ (and thus either $1<p_j<\infty$ and $s_j=1/p_j$, or $p_j=1$ and $s_j=1$) and $1\le q_j<\infty$. Let, as in Lemma \ref{ad1}, $A\subset (0,1)^{n-1}$ be a set of full measure such that \eqref{cf1} holds with $M=2$. The proof of the lemma relies on the following key implication: \begin{equation} \label{cf2} [x_1+\cdots+x_k\in{\mathbb Z},\ 1\le p_1,\ldots, p_k<\infty]\implies |x_1+\cdots+x_k|\lesssim |x_1|^{p_1}+\cdots+|x_k|^{p_k}. \end{equation} This leads to the following consequence: if $g:=g_1+\cdots+g_k$ is integer-valued, then \begin{equation} \label{aaa1} \|\Delta_h^2g\|_{L^1}\lesssim \|\Delta_h^2g_1\|_{L^{p_1}}^{p_1}+\cdots+\|\Delta_h^2g_k\|_{L^{p_k}}^{p_k}. \end{equation} By combining \eqref{cf1} with \eqref{aaa1}, we find that \begin{equation} \label{cf3} \lim_{l\to\infty}\frac{\left\|\Delta_{t_le_n}^2f(x',\cdot)\right\|_{L^{1}((0,1))}}{t_l}=0,\quad\forall\, x'\in A,\text{ for some sequence }t_l\to 0. \end{equation} By Lemma \ref{cf4} below, we find that $f(x',\cdot)$ is constant, for every $x'\in A$. By a permutation of the coordinates, we find that for every $i\in \llbracket 1, n\rrbracket$, the function \begin{equation} \label{aa2} \text{$t\mapsto f(x_1,...,x_{i-1},t, x_{i+1},...,x_n)$ is constant, $\forall\, i\in \llbracket 1, n\rrbracket$, a.e. $\widehat x_i \in (0,1)^{n-1}$;} \end{equation} here, $\widehat x_i:=(x_1,...,x_{i-1},x_{i+1},...,x_n) \in (0,1)^{n-1}$. We next invoke the fact that every measurable function satisfying \eqref{aa2} is constant \cite[Lemma 2]{blmn}. \end{proof} \begin{lemma} \label{cf4} Let $g\in L^1((0,1) ; {\mathbb Z})$ be such that, for some sequence $t_l\to 0$, we have \begin{equation} \label{cf5} \lim_{l\to\infty}\frac{\left\|\Delta_{t_l}^2g\right\|_{L^{1}((0,1))}}{t_l}=0. \end{equation} Then $g$ is constant. \end{lemma} \begin{proof} In order to explain the main idea, let us first assume that $g=\ensuremath{\mathonebb{1}}_B$ for some $B\subset (0,1)$. Let $h\in (0,1)$. If $x\in B$ and $x+2h\not\in B$, then $\Delta_h^2g(x)$ is odd, and thus $|\Delta_h^2g(x)|\ge 1$. The same holds if $x\not\in B$ and $x+2h\in B$. On the other hand, we have $|\Delta_{2h}g(x)|\le 1$, with equality only when either $x\in B$ and $x+2h\not\in B$, or $x\not\in B$ and $x+2h\in B$. By the preceding, we obtain the inequality \begin{equation} \label{cf6}|\Delta^2_hg(x)|\ge |\Delta_{2h}g(x)|,\quad\forall\, x,\, \forall\, h. \end{equation} Using \eqref{cf5} and \eqref{cf6}, we obtain \begin{equation} \label{cf7} g'=\lim_{l\to\infty}\frac{\Delta_{2t_l}g}{2t_l}=0.\footnotemark \end{equation} \footnotetext{ In \eqref{cf7}, the first limit is in ${\cal D}'$, the second one in $L^1$.} Thus either $g=0$, or $g=1$. We next turn to the general case. Consider some $k\in{\mathbb Z}$ such that the measure of the set $g^{-1}(\{k\})$ is positive. We may assume that $k=0$, and we will prove that $g=0$. For this purpose, we set $B:=g^{-1}(2{\mathbb Z})$, and we let $\overline g:=\ensuremath{\mathonebb{1}}_B$. Arguing as above, we have $ |\Delta^2_h g(x)|\ge |\Delta_{2h}\overline g(x)|$, $\forall\, x$, $\forall\, h$, and thus $\overline g=0$. We find that $g$ takes only even values. We next consider the integer-valued map $g/2$. By the above, $g/2$ takes only even values, and so on. We find that $g=0$. \end{proof} \subsection{Disintegration of the Jacobians} \label{au1} ${}$ The purpose of this section is to prove and generalize the following result, used in the analysis of Case \ref{Y}. \begin{lemma} \label{at6} Let $s>1$, $1\le p<\infty$, $1\le q\le p$ and $n\ge 3$, and assume that $sp\ge 2$. Let $u\in B^s_{p,q}(\Omega ; {\mathbb S}^1)$ and set $F:=u\wedge\nabla u$. Then $\operatorname{curl} F=0$. Same conclusion if $s>1$, $1\le p<\infty$, $1\le q\le \infty$ and $n\ge 2$, and we have $sp>2$. Same conclusion if $s>1$, $1\le p<\infty$, $1\le q< \infty$ and $n=2$, and we have $sp=2$. \end{lemma} In view of the conclusion, we may assume that $\Omega=(0,1)^n$. Note that in the above we have $n\ge 2$; for $n=1$ there is nothing to prove. Since the results we present in this section are of independent interest, we go beyond what is actually needed in Case \ref{Y}. The conclusion of (the generalization of) Lemma \ref{at6} relies on three ingredients. The first one is that it is possible to define, as a distribution, the product $F: =u\wedge\nabla u$ for $u$ in a low regularity Besov space; this goes back to \cite{lddjr} when $n=2$, and the case where $n\ge 3$ is treated in \cite{bousquetmironescu}. The second one is a Fubini (disintegration) type result for the distribution $\operatorname{curl} F$. Again, this result holds even in Besov spaces with lower regularity than the ones in Lemma \ref{at6}; see Lemma \ref{mo2} below. The final ingredient is the fact that when $u\in\text{\rm VMO}((0,1)^2 ; {\mathbb S}^1)$ we have $\operatorname{curl} F=0$; see Lemma \ref{mo3}. Lemma \ref{at6} is obtained by combining Lemmas \ref{mo2} and \ref{mo3} via a dimensional reduction (slicing) based on Lemma \ref{mo7}; a more general result is presented in Lemma \ref{mo4}. Now let us proceed. First, following \cite{lddjr} and \cite{bousquetmironescu}, we explain how to define the Jacobian $Ju:=1/2\operatorname{curl} F$ of low regularity unimodular maps $u\in W^{1/p,p}((0,1)^n ; {\mathbb S}^1)$, with $1\le p<\infty$.\footnote{ In \cite{lddjr} and \cite{bousquetmironescu}, maps are from ${\mathbb S}^n$ (instead of $(0,1)^n$) into ${\mathbb S}^1$, but this is not relevant for the validity of the results we present here.} Assume first that $n=2$ and that $u$ is smooth. Then, in the distributions sense, we have \begin{equation} \label{oa2} \begin{aligned} \langle Ju,\zeta\rangle&=\frac 12\int_{(0,1)^2}\operatorname{curl} F\, \zeta=-\frac 12\int_{(0,1)^2}\nabla\zeta\wedge(u\wedge\nabla u)\\&=\frac 12\int_{(0,1)^2}[(u\wedge\partial_1u)\partial_2\zeta-(u\wedge\partial_2u)\partial_1\zeta]\\ &=\frac 12\int_{(0,1)^2}(u_1\nabla u_2\wedge\nabla\zeta-u_2\nabla u_1\wedge\nabla\zeta),\quad\forall\, \zeta\in C^\infty_c((0,1)^2). \end{aligned} \end{equation} In higher dimensions, it is better to identify $Ju$ with the $2$-form (or rather a $2$-current) $Ju\equiv 1/2\, d(u\wedge du)$.\footnote{ We recover the two-dimensional formula \eqref{oa2} via the usual identification of $2$-forms on $(0,1)^2$ with scalar functions (with the help of the Hodge $\ast$-operator).} With this identification and modulo the action of the Hodge $\ast$-operator, $Ju$ acts either or $(n-2)$-forms, or on $2$-forms. The former point of view is usually adopted, and is expressed by the formula \begin{equation} \label{oa3} \begin{aligned} \langle Ju,\zeta\rangle&=\frac {(-1)^{n-1}}2\int_{(0,1)^n}d\zeta\wedge(u\wedge\nabla u)\\ &=\frac {(-1)^{n-1}}2\int_{(0,1)^n}d\zeta\wedge(u_1\, du_2-u_2\, du_1),\quad\forall\, \zeta\in C^\infty_c(\Lambda^{n-2}(0,1)^n). \footnotemark \end{aligned} \end{equation} \footnotetext{ Here, $C^\infty_c(\Lambda^{n-2}(0,1)^n)$ denotes the space of smooth compactly supported $(n-2)$-forms on $(0,1)^n$.} The starting point in extending the above formula to lower regularity maps $u$ is provided by the identity \eqref{oa4} below; when $u$ is smooth, \eqref{oa4} is obtained by a simple integration by parts. More specifically, consider any smooth extension $U:(0,1)^n\times [0,\infty)\to{\mathbb C}$, respectively $\varsigma\in C^\infty_c(\Lambda^{n-2}((0,1)^n\times [0,\infty)))$ of $u$, respectively of $\zeta$.\footnote{ We do not claim that $U$ is ${\mathbb S}^1$-valued. When $u$ is not smooth, existence of ${\mathbb S}^1$-valued extensions is a delicate matter \cite{soreview}.} Then we have the identity \cite[Lemma 5.5]{bousquetmironescu} \begin{equation} \label{oa4} \langle Ju,\zeta\rangle=(-1)^{n-1}\int_{(0,1)^n\times (0,\infty)}d\varsigma\wedge\ dU_1\wedge dU_2. \end{equation} For a low regularity $u$ and for a well-chosen $U$, we take the right-hand side of \eqref{oa4} as the definition of $Ju$. More specifically, let $\Phi\in C^\infty({\mathbb R}^2; {\mathbb R}^2)$ be such that $\Phi(z)=z/|z|$ when $|z|\ge 1/2$, and let $v$ be a standard extension of $u$ by averages, i.e., $v(x,\varepsilon)=u\ast\rho_\varepsilon(x)$, $x\in (0,1)^n$, $\varepsilon>0$, with $\rho$ a standard mollifier. Set $U:=\Phi(v)$. With this choice of $U$, the right-hand side of \eqref{oa4} does not depend on $\varsigma$ (once $\zeta$ is fixed) \cite[Lemma 5.4]{bousquetmironescu} and the map $u\mapsto Ju$ is continuous from $W^{1/p,p}((0,1)^n ; {\mathbb S}^1)$ into the set of $2$- (or $(n-2)$-)currents. When $p=1$, continuity is straightforward. For the continuity when $p>1$, see \cite[Theorem 1.1 item 2]{bousquetmironescu}. In addition, when $u$ is sufficiently smooth (for example when $u\in W^{1,1}((0,1)^n ; {\mathbb S}^1)$), $Ju$ coincides\footnote{ Up to the action of the $\ast$ operator.} with $\operatorname{curl} F$ \cite[Theorem 1.1 item 1]{bousquetmironescu}. Finally, we have the estimate \cite[Theorem 1.1 item 3]{bousquetmironescu} \begin{equation} \label{oa6} |\langle Ju,\zeta\rangle|\lesssim |u|_{W^{1/p,p}}^p\|d\zeta\|_{L^\infty},\quad\forall\, \zeta\in C^\infty_c(\Lambda^{n-2}(0,1)^n). \end{equation} We are now in position to explain disintegration along two-planes. We use the notation in Section \ref{mo6}. Let $u\in W^{1/p,p}((0,1)^n ; {\mathbb S}^1)$, with $n\ge 3$. Let $\alpha\in I(n-2, n)$. Then for a.e. $x_\alpha\in (0,1)^{n-2}$, the partial map $u_\alpha(x_\alpha)$ belongs to $W^{1/p,p}((0,1)^2 ; {\mathbb S}^1)$ (Lemma \ref{oa1}), and therefore $Ju_\alpha(x_\alpha)$ makes sense and acts on functions.\footnote{ Or rather on $2$-forms, in order to be consistent with our construction in dimension $\ge 3$.} Let now $\zeta\in C^\infty_c(\Lambda^{n-2}(0,1)^n)$. Then we may write \begin{equation*} \zeta=\sum_{\alpha\in I(n-2,n)}\zeta^\alpha\, dx^{\alpha}=\sum_{\alpha\in I(n-2,n)}\left(\zeta^\alpha\right)_\alpha(x_{\overline\alpha})\, dx^{\alpha}. \end{equation*} Here, $dx^{\alpha}$ is the canonical $(n-2)$-form induced by the coordinates $x_j$, $j\in\alpha$, and $(\zeta^\alpha)_\alpha(x_{\overline\alpha})=\zeta^\alpha(x_\alpha, x_{\overline\alpha})$ belongs to $C^\infty_c((0,1)^2)$ (for fixed $x_\alpha$). We next note the following formal calculation. Fix $\alpha\in I(n-2,n)$, and let $\overline\alpha=\{ j, k\}$, with $j<k$. Then \begin{equation*} \begin{aligned} 2(-1)^{n-1}\langle Ju,\zeta^\alpha\, dx^\alpha\rangle&=\int_{(0,1)^n}d(\zeta^\alpha\, dx^\alpha)\wedge(u\wedge \nabla u)\\ &=\int_{(0,1)^n}(\partial_j\zeta^\alpha\, dx_j+\partial_k\zeta^\alpha\, dx_k)\wedge dx^\alpha\wedge u\wedge (\partial_j u\, dx_j+\partial_k u\, dx_k)\\ &=\int_{(0,1)^n}(\partial_j\zeta^\alpha\, u\wedge \partial_k u-\partial_k\zeta^\alpha\, u\wedge \partial_j u)\, dx_j\wedge dx^\alpha\wedge dx_k, \end{aligned} \end{equation*} that is, \begin{equation} \label{oc2} \langle Ju,\zeta\rangle=\frac 12\ \sum_{\alpha\in I(n-2,n)}\varepsilon(\alpha)\int_{(0,1)^{n-2}}\langle Ju_\alpha,\left(\zeta^\alpha\right)_\alpha(x_\alpha)\rangle\, dx_\alpha, \end{equation} where $\varepsilon(\alpha)\in \{-1, 1\}$ depends on $\alpha$. When $u\in W^{1,1}((0,1)^n ; {\mathbb S}^1)$, it is easy to see that \eqref{oc2} is true (by Fubini's theorem). The validity of \eqref{oc2} under weaker regularity assumptions is the content of our next result. \begin{lemma} \label{mo2} Let $1\le p<\infty$ and $n\ge 3$. Let $u\in W^{1/p,p}((0,1)^n ; {\mathbb S}^1)$. Then \eqref{oc2} holds. \end{lemma} \begin{proof} The case $p=1$ being clear, we may assume that $1<p<\infty$. We may also assume that $\zeta=\zeta^\alpha\, dx^\alpha$ for some fixed $\alpha\in I(n-2,n)$. A first ingredient of the proof of \eqref{oc2} is the density of $W^{1,1}((0,1)^n ; {\mathbb S}^1)\cap W^{1/p,p}((0,1)^n ; {\mathbb S}^1)$ into $W^{1/p,p}((0,1)^n ; {\mathbb S}^1)$ \cite[Lemma 23]{bbmihes}, \cite[Lemma A.1]{lddjr}. Next, we note that the left-hand side of \eqref{oc2} is continuous with respect to the $W^{1/p,p}$ convergence of unimodular maps \cite[Theorem 1.1 item 2]{bousquetmironescu}. In addition, as we noted, \eqref{oc2} holds when $u\in W^{1,1}((0,1)^n ; {\mathbb S}^1)$. Therefore, it suffices to prove that the right-hand side of \eqref{oc2} is continuous with respect to $W^{1/p,p}$ convergence of ${\mathbb S}^1$-valued maps. This is proved as follows. Let $u_j, u\in W^{1/p,p}((0,1)^n ; {\mathbb S}^1)$ be such $u_j\to u$ in $W^{1/p,p}$. By a standard argument, since the right-hand side of \eqref{oc2} is uniformly bounded with respect to $j$ by \eqref{oa6}, it suffices to prove that the right-hand side of \eqref{oc2} corresponding to $u_j$ tends to the one corresponding to $u$ possibly along a subsequence. In turn, convergence up to a subsequence is proved as follows. Recall the following vector-valued version of the \enquote{converse} to the dominated convergence theorem \cite[Theorem 4.9, p. 94]{brezisfa}. If $X$ is a Banach space, $\omega$ a measured space and $f_j\to f$ in $L^p(\omega, X)$, then (possibly along a subsequence) for a.e. $\varpi\in\omega$ we have $f_j(\varpi,\cdot )\to f(\varpi,\cdot)$ in $X$, and in addition there exists some $g\in L^p(\omega)$ such that $\|f_j(\varpi,\cdot)\|_{X}\le g(\varpi)$ for a.e. $\varpi\in \omega$. Using the above and Lemma \ref{oa1} item 2 (applied with $s=1/p$), we find that, up to a subsequence, we have \begin{equation} \label{oc3} (u_j)_\alpha(x_\alpha)\to u_\alpha(x_\alpha)\text{ in }W^{1/p,p}((0,1)^2 ; {\mathbb S}^1)\text{ for a.e. }x_\alpha\in (0,1)^{n-2}, \end{equation} and in addition we have, for some $g\in L^p((0,1)^{n-2})$, \begin{equation} \label{oc4} |(u_j)_\alpha(x_\alpha)|_{W^{1/p,p}((0,1)^2)}\le g(x_\alpha)\text{ for a.e. }x_\alpha\in(0,1)^{n-2}. \end{equation} The continuity of the right-hand side of \eqref{oc2} (along some subsequence) is obtained by combining \eqref{oc3} and \eqref{oc4} with \eqref{oa6} (applied with $n=2$).\footnote{In order to be complete, we should also check that the right-hand side of \eqref{oc2} is measurable with respect to $x_\alpha$. This is clear when $u\in W^{1,1}((0,1)^n ; {\mathbb S}^1)$. The general case follows by density and \eqref{oc3}.} \end{proof} \begin{lemma} \label{mo3} Let $1\le p<\infty$. Let $u\in W^{1/p,p}\cap\text{\rm VMO}((0,1)^2 ; {\mathbb S}^1)$. Then $Ju=0$. \end{lemma} \begin{proof} Assume first that in addition we have $u\in C^\infty$. Then $u=e^{\imath\varphi}$ for some $\varphi\in C^\infty$, and thus $Ju=1/2\operatorname{curl} (u\wedge\nabla u)=1/2\operatorname{curl}\nabla\varphi=0$. We now turn to the general case. Let $F(x,\varepsilon):=u\ast\rho_\varepsilon(x)$, with $\rho$ a standard mollifier. Since $u\in\text{\rm VMO}((0,1)^2 ; {\mathbb S}^1)$, there exists some $\delta>0$ such that $1/2<|F(x,\varepsilon)|\le 1$ when $0<\varepsilon<\delta$ (see \eqref{boundsv} and the discussion in Case \ref{X}). Let $\Phi\in C^\infty({\mathbb R}^2 ; {\mathbb R}^2)$ be such that $\Phi(z):=z/|z|$ when $|z|\ge 1/2$, and define $F_\varepsilon(x):=F(x,\varepsilon)$ and $u_\varepsilon:=\Phi\circ F_\varepsilon$, $\forall\, 0<\varepsilon<\delta$. Then $F_\varepsilon\to u$ in $W^{1/p,p}$ and (by Lemma \ref{eipsi} when $p>1$, respectively by a straightforward argument when $p=1$) we have $u_\varepsilon=\Phi(F_\varepsilon)\to \Phi(u)=u$ in $W^{1/p,p}((0,1)^2 ; {\mathbb S}^1)$ as $\varepsilon\to 0$. Since (by the beginning of the proof) we have $Ju_\varepsilon=0$, we conclude via the continuity of $J$ in $W^{1/p,p}((0,1)^2 ; {\mathbb S}^1)$ \cite[Theorem 1.1 item 2]{bousquetmironescu}. \end{proof} We may now state and prove the following generalization of Lemma \ref{at6}. \begin{lemma} \label{mo4} Let $s>0$, $1\le p<\infty$, $1\le q\le p$, $n\ge 3$, and assume that $sp\ge 2$. Let $u\in B^s_{p,q}(\Omega ; {\mathbb S}^1)$. Then $Ju=0$. Same conclusion if $s>0$, $1\le p<\infty$, $1\le q\le\infty$, $n\ge 2$, and we have $sp>2$. Same conclusion if $s>0$, $1\le p<\infty$, $1\le q<\infty$, $n= 2$, and we have $sp=2$. \end{lemma} \begin{proof} We may assume that $\Omega=(0,1)^n$. By the Sobolev embeddings (Lemma \ref{Besovemb}), it suffices to consider the limiting case where: \begin{enumerate} \item $s>0$, $1\le p<\infty$, $1\le q<\infty$, $n=2$, and $sp=2$. Or \item $s>0$, $1\le p<\infty$, $q= p$, $n\ge 3$, and $sp=2$. \end{enumerate} In view of Lemmas \ref{Besovemb} and \ref{B-VMO}, the case where $n=2$ is covered by Lemma \ref{mo3}. Assume that $n\ge 3$. Then the desired conclusion is obtained by combining Lemmas \ref{oa1}, \ref{mo7}, \ref{mo2} and \ref{mo3}. \end{proof} \begin{remark} \label{oc1} Arguments similar to the one developed in this section lead to the conclusion that the Jacobians of maps $u\in W^{s,p}((0,1)^n ; {\mathbb S}^k)$, defined when $sp\ge k$ \cite{lddjr}, \cite{bousquetmironescu}, disintegrate over $(k+1)$-planes. When $s=1$ and $p\ge k$, this assertion is implicit in \cite[Proof of Proposition 2.2, pp. 701-704]{isobe2}. \end{remark}
{ "timestamp": "2017-06-20T02:10:47", "yymm": "1705", "arxiv_id": "1705.04271", "language": "en", "url": "https://arxiv.org/abs/1705.04271", "abstract": "Let $\\Omega$ be a smooth bounded domain in $\\mathbb R^n$ and u be a measurable function on $\\Omega$ such that $|u(x)|=1$ almost everywhere in $\\Omega$. Assume that u belongs to the $B^s_{p,q}(\\Omega)$ Besov space. We investigate whether there exists a real-valued function $\\varphi \\in B^s_{p,q}$ such that $u=e^{i\\varphi}$. This extends the corresponding study in Sobolev spaces due to Bourgain, Brezis and the first author. The analysis of this lifting problem leads us to prove some interesting new properties of Besov spaces, in particular a non restriction property when $q>p$.", "subjects": "Classical Analysis and ODEs (math.CA); Analysis of PDEs (math.AP)", "title": "Lifting in Besov Spaces", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.983596967483837, "lm_q2_score": 0.8152324848629214, "lm_q1q2_score": 0.8018601999054825 }
https://arxiv.org/abs/math/0512131
Unimodality and convexity of f-vectors of polytopes
We consider unimodality and related properties of f-vectors of polytopes in various dimensions. By a result of Kalai (1988), f-vectors of 5-polytopes are unimodal. In higher dimensions much less can be said; we give an overview on current results and present a potentially interesting construction as well as a conjecture arising from this.
\section{Introduction} \label{sec:introduction} Let $f = (f_0,\ldots,f_{d-1})$ be the $f$-vector of a $d$-polytope. It is natural to ask whether the $f$-vector necessarily has one (or more) of the following properties: \begin{itemize} \item[{\rm (C)}] convexity: $f_k \geq (f_{k-1}+f_{k+1})/2$ for all $k \in \{ 1,\ldots,d-2 \}$ \item[{\rm (L)}] logarithmic convexity: $f_k^2 \geq f_{k-1} f_{k+1}$ for all $k \in \{ 1,\ldots,d-2 \}$ \item[{\rm (U)}] unimodality: $f_0 \leq \ldots \leq f_k \geq \ldots \geq f_{d-1}$ for some $k \in \{ 0,\ldots,d-1 \}$ \item[{\rm (B)}] B\'ar\'any's property: $f_k \geq \min \{ f_0,f_{d-1} \}$ for all $k \in \{ 1,\ldots,d-2 \}$ \end{itemize} Clearly each property implies the next one: {\rm (C)}\ $\Rightarrow$ {\rm (L)}\ $\Rightarrow$ {\rm (U)}\ $\Rightarrow$ {\rm (B)}. Unimodality is known to be false in general for $d \geq 8$ and (rather trivially) true for $d \leq 4$. For simplicial (and therefore also for simple) polytopes of arbitrary dimension a weaker version of unimodality was proved by Bj\"orner \cite[Section 8.6]{MR96a:52011}. Similarly, convexity is trivially true up to $d \leq 3$ and for $d=4$ follows easily from $f_0 \geq 5$ and $f_2 \geq 2 f_3$ together with Euler's equation and duality. \subsection*{Toric $g$-vectors} \label{sec:toric-g-vector} To every $d$-polytope $P$ we can assign a $(\lfloor d/2 \rfloor+1)$-dimensional vector $g(P) = (g_0^d(P), \ldots, g_{\lfloor d/2 \rfloor}^d(P))$, the \emph{toric $g$-vector} of $P$. Its entries can be calculated via recursion \cite[Section 3.14]{MR98a:05001}, and interpreted geometrically for simplicial polytopes. It is well known \cite{MR89f:52016} that $g_i^d(P) \geq 0$ for rational polytopes $P$ and only recently Karu \cite{MR2076929} showed that nonnegativity also holds for nonrational polytopes. The entries of the toric $g$-vector of a polytope $P$ can be rewritten as a linear combination of entries of the flag vector of $P$. Some special cases which we will need are $g_0^d (P) = 1$ and $g_1^d (P) = f_0 - (d+1)$ for $d$-polytopes $P$ (note that $1 = f_\emptyset(P)$). See \cite{MR2132764} for a general description. \subsection*{Convolutions} \label{sec:convolution} Let $m_1$ and $m_2$ be linear forms on flag vectors of $d_1$-, resp. $d_2$-polytopes. Then we obtain a linear form $m = m_1 * m_2$ by defining \begin{displaymath} m (P) \; := \; \sum_{F \; d_1\text{-face of } P} m_1 (F) \, m_2 (P/F) \end{displaymath} for every $(d_1+d_2+1)$-polytope $P$. Alternatively, the convolution can be described by defining \begin{displaymath} f_S * f_T \; := \; f_{S \cup \{d_1\} \cup (T+d_1+1)} \end{displaymath} for $S \subseteq \{ 0,\ldots,d_1 \}$, $T \subseteq \{ 0,\ldots,d_2 \}$ (where $M+x := \{ m+x \mid m \in M \})$ and extending linearly \cite[Section 3]{MR90a:52012} \cite[Section 7]{MR2001c:52009}. We will use this notation, occasionally writing $f_S^d$ to indicate the dimension $d$ of the polytopes the respective flag vector refers to. \subsection*{{{\bf cd}}-index} \label{sec:cd-index} Connected with every polytope (in fact with every Eulerian poset) is its {{\bf cd}}-index, which is a polynomial in the non-commuting variables $\c$ and $\d$. The coefficients of the {{\bf cd}}-index can again be viewed as linear combinations of flag vector entries \cite[Section 7]{MR2001c:52009}. Stanley \cite{MR96b:06006} showed that the coefficients of the {{\bf cd}}-index of a polytope are nonnegative, which again yields inequalities for the flag vector. Further useful results were obtained by Ehrenborg \cite{MR2132764}. From there we will adopt the following notation: write $\langle u \mid \Psi(P) \rangle$ for the coefficient of the {{\bf cd}}-monomial $u$ in the {{\bf cd}}-index of the polytope $P$. Using linearity we can then define the number $\langle p \mid \Psi(P) \rangle$ for any {{\bf cd}}-polynomial $p$. \bigskip In some of the following proofs we omit the longer calculations. For more details see the appendix. \section{Dimension 5} \label{sec:dimension-5} \begin{theorem} \label{thm:unimodal5} Unimodality {\rm (U)}\ holds for $f$-vectors of polytopes of dimension $d \leq 5$. \end{theorem} \begin{proof} Let $P$ be a $5$-polytope and $f(P) = (f_0,f_1,f_2,f_3,f_4)$ its $f$-vector. Trivially, $5f_0 \leq 2f_1$ and $5f_4 \leq 2f_3$, therefore $f_0 < f_1$ and $f_3 > f_4$. Kalai \cite{MR90a:52012} showed that $3f_2 \geq 2f_1 + 2f_3$, hence \begin{displaymath} f_2 \; \geq \; \frac{2}{3} \, (f_1+f_3) \; > \; \frac{f_1+f_3}{2} \end{displaymath} which implies that ``there cannot be a dip'' at $f_2$. Therefore $f(P)$ is unimodal. \end{proof} \begin{theorem} Convexity {\rm (C)}\ fails to hold for $d \geq 5$, that is, the $f$-vectors of $d$-polytopes are not convex in general. \end{theorem} \begin{proof} For dimension $5$ the $f$-vector of the cyclic polytope with $n$ vertices is \begin{displaymath} f (\cyclic{5}{n}) \; = \; \left( n, \, \tfrac{n(n-1)}{2}, \, 2(n^2-6n+10), \, \tfrac{5(n-3)(n-4)}{2}, \, (n-3)(n-4) \right) \end{displaymath} (cf.~\cite[Chapter~8]{MR96a:52011}), which implies \begin{displaymath} f_1 \; = \; \frac{n^2-n}{2} \; < \; \frac{2n^2-11n+20}{2} \; = \; \frac{f_0+f_2}{2} \end{displaymath} for $n \geq 8$; see Figure \ref{fig:f-c-5-8}. \begin{figure}[tb] \centering \includegraphics{cyclic5-8.f-vector.eps} \caption{(non-convex) $f$-vector of $\cyclic{5}{8}$} \label{fig:f-c-5-8} \end{figure} \par For $d \geq 6$, cyclic $d$-polytopes are $2$-neighbourly, therefore $f_1 = {f_0 \choose 2}$ and $f_2 = {f_0 \choose 3}$ for $f_0 \geq d+1$. We conclude that \begin{displaymath} f_0 + f_2 - 2f_1 \; = \; \frac{1}{6} \, f_0 (f_0-2)(f_0-7) > 0 \end{displaymath} for cyclic $d$-polytopes with $f_0 \geq \max\{d+2,8\}$ vertices. Thus for $d \geq 7$ already the simplex is a counterexample for {\rm (C)}. \end{proof} \section{Dimension 6} \label{sec:dimension-6} Concerning unimodality for $f$-vectors of $6$-polytopes, we have a couple of trivial facts, such as $f_0 < f_1$ and $f_4 > f_5$. Unimodality would therefore simply follow from the statement $(*) \; f_1 \leq f_2$ or equivalently from $f_3 \geq f_4$ by duality. Bj\"orner showed that the latter is true for simplicial polytopes (cf. \cite{MR96a:52011}, Theorem 8.39), therefore in particular for cyclic polytopes, which seems to indicate that it is true in general. However, it does not follow from the yet known inequalities -- we only have a weaker statement. \begin{proposition} \label{prop:barany6} Let $f=(f_0,\ldots,f_5)$ be the $f$-vector of a $6$-polytope. Then \[ f_2 \; \geq \; \frac{2}{3} \, f_1 + 63 . \] \end{proposition} \begin{proof} We claim that the following inequalities hold for $f$: \begin{eqnarray} \label{in6-1} f_1 - 3f_0 & \geq & 0 \\ \label{in6-2} f_0 - f_1 + f_2 - 21 & \geq & 0 \end{eqnarray} The assertion then follows by multiplying \eqref{in6-2} by $3$ and adding \eqref{in6-1}. \par Inequality \eqref{in6-1} is trivial, simply stating that every vertex is in at least 6 edges. For the proof of \eqref{in6-2} we use \cite[Theorem 3.7]{MR2132764}, which implies that $\langle \c^2\d\c^2 - 19 \c^6 \mid \Psi(P) \rangle \geq 0$. Expressing the {{\bf cd}}-polynomial $\c^2\d\c^2 - 19 \c^6$ as linear combination of flag vector entries gives $f_0-f_1+f_2-21$ and therefore yields inequality \eqref{in6-2}. See the last section for detailed calculations. \end{proof} \begin{corollary} The $f$-vectors of $6$-polytopes satisfy B\'ar\'any's property {\rm (B)}. \end{corollary} \begin{proof} Let $f=(f_0,\ldots,f_5)$ be the $f$-vector of a $6$-polytope. Clearly, $f_1 \geq 3 f_0 > f_0$, thus by Proposition~\ref{prop:barany6} \begin{displaymath} f_2 \; \geq \; \frac{2}{3} \, f_1 + 21 \; \geq \; 2 f_0 + 21 \; > \; f_0 . \end{displaymath} Dually, we have $f_3 > f_5$ and $f_4 > f_5$. \end{proof} As the desired inequality $(*)$ for unimodality does not follow from the known linear inequalities, one can find vectors that satisfies all these, but not $(*)$. An example for a family of vectors is \begin{align*} f^{(\ell)} = ( & f_0 , f_1 , f_2 , f_3 , f_4 ; \\ & f_{02} , f_{03} , f_{04} , f_{13} , f_{14} , f_{24} ; \\ & f_{024} ) \\ \phantom{f^{(\ell)} }= ( & 22 + \ell , 111 + 3\ell , 110 + 2\ell , 35 + 4\ell , 21 + 6\ell ; \\ & 780 + 15\ell , 1340 + 50\ell , 1080 + 51\ell , 2010 + 90\ell , 2160 + 132\ell , 1260 + 114\ell ; \\ & 6480 + 396\ell ) \end{align*} for $\ell \geq 0$. The other components of these (potential) flag vectors can be calculated from the Generalized Dehn--Sommer\-ville equations. In particular, the number of facets is $f_5 = 7+2\ell$. However it is not at all clear that there exist polytopes having these as flag vectors. \section{Dimension 7} \label{sec:dimension-7} A similar statement to the one in Proposition~\ref{prop:barany6} holds for $7$-polytopes. Nevertheless, this is not enough to prove even B\'ar\'any's property {\rm (B)}, since we yet have no condition for $f_3$. \begin{proposition} \label{prop:barany7} Let $f=(f_0,\ldots,f_6)$ be the $f$-vector of a $7$-polytope. Then \[ f_2 \; \geq \; \frac{5}{7} \, f_1 + 36 \] \end{proposition} \begin{proof} As before, we consider two valid inequalities for $f$ which together imply the assertion: \begin{eqnarray} \label{in7-1} 2f_1 - 7f_0 & \geq & 0 \\ \label{in7-2} f_0 - f_1 + f_2 - 36 & \geq & 0 \end{eqnarray} Again, \eqref{in7-1} is trivial. The nonnegativity of $\langle \c^2\d\c^3-34\c^7 \mid \Psi(P) \rangle$ gives inequality \eqref{in7-2}; see the last section. \end{proof} Again, one can find vectors satisfying all known linear inequalities, but violating both $f_3 \geq f_0$ and $f_3 \geq f_6$; take, for instance, the potential flag vector \begin{align*} f \; & = \; ( f_0 , f_1 , f_2 , f_3 , f_4 , f_5 ; \\ & \qquad \quad f_{02} , f_{03} , f_{04} , f_{05} , f_{13} , f_{14} , f_{15} , f_{24} , f_{25} , f_{35} ; \\ & \qquad \quad f_{024} , f_{025} , f_{035} , f_{135} ) \\ & = \; ( 134 , 469 , 371 , 70 , 371 , 469 ; \\ & \qquad \quad 2814 , 6580 , 10360 , 8484 , 9870 , 20720 , 21210 , 13790 , 20720 , 9870 ; \\ & \qquad \quad 62160 , 84840 , 84840 , 127260 ) . \end{align*} From Euler's equation, we get $f_6 = 134$; nevertheless, it is again open whether this really is the flag vector of some $7$-polytope. As it is an open question whether logarithmic convexity holds for $f$-vectors of $7$-polytopes, one could try to find counterexamples. Most promising may be connected sums of cyclic polytopes, since this construction yields counterexamples for unimodality in dimension 8 (see \cite[pp.~274f]{MR96a:52011}). \begin{definition} Let $P$ and $Q$ be polytopes of the same dimension. If $P$ is simplicial and $Q$ simple, then a \emph{connected sum} $P \# Q$ of $P$ and $Q$ is obtained by cutting one vertex off $Q$ and stacking the result --- with the newly created facet --- onto $P$ (cf.~\cite[p.~274]{MR96a:52011}). \end{definition} The effect of these construction on the $f$-vector of the involved polytopes can be described as follows. \begin{proposition} \label{prop:f-consum} Let $d \geq 3$ and $P$ a simplicial and $Q$ a simple $d$-polytope. Then the $f$-vector of $P\#Q$ is given by \begin{displaymath} f_i(P\#Q) \; = \; \left\{ \begin{array}{l@{\quad\text{ if }}l} f_i(P) + f_i(Q) & 1 \leq i \leq d-2 \\ f_i(P) + f_i(Q) - 1 & i = 0 \text{ or } i = d-1 \end{array} \right. \end{displaymath} Additionally, the $f$-vector of the connected sum $P\#\polar{P}$ of a polytope $P$ with its dual is symmetric. \end{proposition} \begin{proof} Cutting one vertex $v$ off $Q$ decreases $f_0(Q)$ by $1$ and creates a new facet $F$, isomorphic to a $(d-1)$-simplex. Therefore, $f_i(Q)$ increases by ${d \choose i+1}$ if $i>0$ and by $d-1$ if $i=0$. Afterwards all faces of both polytopes are again faces of $P\#Q$, except the facet $F$ in both polytopes (which completely disappears) and the new faces of $F$ in $Q$ (which are identified with their counterparts in $P$). \par The $f$-vector of $P\#\polar{P}$ is obviously symmetric, since \begin{displaymath} f_i(P\#\polar{P}) \; = \; \left\{ \begin{array}{l@{\quad\text{ if }}l} f_i(P) + f_{d-1-i}(P) & 1 \leq i \leq d-2 \\ f_i(P) + f_{d-1-i}(P) - 1 & i = 0 \text{ or } i = d-1 \end{array} \right. \end{displaymath} \end{proof} \begin{proposition} \label{prop:logconv7} For all $n \geq 8$, the $f$-vector of $P_7^n := \cyclic{7}{n} \# \cyclic{7}{n}^\Delta$ is logarithmically convex and \[ \frac{f_3(P_7^n)^2}{f_2(P_7^n) f_4(P_7^n)} \; \stackrel{n \rightarrow \infty}{\longrightarrow} \; 1 \] \end{proposition} \begin{proof} The proof is done by straightforward calculation; see the last section for the main steps. \end{proof} So in a sense, the connected sums of cyclic $7$-polytopes are as close as polytopes can get to logarithmic non-convexity. \section{Summary} \label{sec:summary} The results can be summarized as in Table~\ref{tab:summary}. \begin{table}[bt] \centering \begin{tabular}{|c|ccccc|} \hline Dimension & $\leq 4$ & $5$ & $6$ & $7$ & $\geq 8$ \\ \hline {\rm (C)} & \ding{52} & \ding{56} & \ding{56} & \ding{56} & \ding{56} \\ {\rm (L)} & \ding{52} & ? & ? & ? & \ding{56} \\ {\rm (U)} & \ding{52} & \ding{52} & ? & ? & \ding{56} \\ {\rm (B)} & \ding{52} & \ding{52} & \ding{52} & ? & ? \\ \hline \end{tabular} \caption{Summary of known properties for polytopes --- a \ding{52}, resp.~\ding{56}\/ indicates that the given property holds, resp.~does not hold for all polytopes of the given dimension.} \label{tab:summary} \end{table} In the light of Proposition~\ref{prop:logconv7}, the following conjecture seems natural. \begin{conjecture} {\rm (L)}\ holds for $d$-polytopes of dimension $d \leq 7$. \end{conjecture} \begin{appendix} \section*{Detailed calculations} \label{sec:deta-calc} \subsection*{Proof of Theorem \ref{thm:unimodal5} \\ {\normalsize \rm (cf.~\cite[Section 7]{MR90a:52012})}} \label{sec:proof-theor-refthm} Let $P$ be a $5$-polytope. \begin{eqnarray*} (g_0^1 * g_1^2 * g_0^0)(P) & = & f_\emptyset^1 * (f_0^2-3) * f_\emptyset^0 \; = \; (f_{12}^4 - 3f_1^4) * f_\emptyset^0 \; = \; f_{124} - 3f_{14} \\ & = & f_{123} - 3 \, (2f_1 - f_{12} + f_{13}) \; = \; -6f_1 + 3f_{02} - f_{13} \\ (g_0^0 * g_1^2 * g_0^1)(P) & = & f_\emptyset^0 * (f_0^2-3) * f_\emptyset^1 \; = \; (f_{01}^3 - 3f_0^3) * f_\emptyset^1 \; = \; f_{013} - 3f_{03} \\ & = & 2f_{13} - 3f_{03} \\ (g_1^2 * g_1^2)(P) & = & (f_0^2 - 3) * (f_0^2 - 3) \; = \; f_{023} - 3f_{02} - 3f_{23} + 9f_2 \\ & = & f_{013} - 3f_{02} - 3 \, (2f_3 - f_{03} + f_{13}) + 9f_2 \\ & = & -f_{13} - 3f_{02} - 6f_3 + 3f_{03} + 9f_2 \end{eqnarray*} by the rules of convolution and the Generalized Dehn--Sommerville equations \cite{MR86f:52010b}. Hence we have \begin{displaymath} \begin{array}{rrcrcrcrcrcl} -6 f_1 & & & & + & 3 f_{02} & & & - & f_{13} & \geq & 0 \\ & & & & & & - & 3 f_{03} & + & 2 f_{13} & \geq & 0 \\ & 9 f_2 & - & 6 f_3 & - & 3 f_{02} & + & 3 f_{03} & - & f_{13} & \geq & 0 \end{array} \end{displaymath} Adding up all three inequalities yields $ -6 f_1 + 9 f_2 - 6 f_3 \geq 0 $, that is the assertion $3f_2 \geq 2f_1 + 2f_3$. \qed \subsection*{Proof of Inequality~\eqref{in6-2}} \label{sec:proof:prop:barany6} We express the {{\bf cd}}-word $\c^2\d\c^2$ in terms of the flag vector of the $6$-polytope $P$ by applying \cite[Proposition~7.1]{MR2001c:52009}: \begin{displaymath} \langle \c^2\d\c^2 \mid \Psi(P) \rangle \; = \; \sum_{i=0}^2 (-1)^{4-i} k_i \; = \; k_0 - k_1 + k_2 . \end{displaymath} For the sparse flag $k$-vector we have \begin{displaymath} k_i \; = \; \sum_{T \subseteq \{i\}} (-2)^{1-|T|} f_T \; = \; -2 f_\emptyset + f_i \; = \; f_i - 2 \end{displaymath} and therefore $\langle \c^2\d\c^2 \mid \Psi(P) \rangle = f_0-f_1+f_2-2$. The trivial {{\bf cd}}-word $\c^6$ translates into $f_\emptyset = 1$, hence \begin{displaymath} \langle \c^2\d\c^2 - 19\c^6 \mid \Psi(P) \rangle \; = \; f_0-f_1+f_2-21 . \end{displaymath} \qed \subsection*{Proof of Inequality~\eqref{in7-2}} \label{sec:proof:prop:barany7} We calculate for the $7$-polytope $P$ exactly as above (the additional $\c$ at the end has no influence whatsoever on the calculation): \begin{displaymath} \langle \c^2\d\c^3 \mid \Psi(P) \rangle \; = \; f_0-f_1+f_2-2 . \end{displaymath} Together with $\c^7$, which again represents $f_\emptyset$, we get \begin{displaymath} \langle \c^2\d\c^3 - 34\c^7 \mid \Psi(P) \rangle \; = \; f_0-f_1+f_2-36 . \end{displaymath} \qed \subsection*{Proof of Proposition~\ref{prop:logconv7}} The $f$-vector of the cyclic $7$-polytope on $n$ vertices is given by \begin{align*} f(\cyclic{7}{n}) \; = \; \big( n & , \, \tfrac{n(n-1)}{2} , \, \tfrac{n(n-1)(n-2)}{6} , \, \tfrac{5(n-4)(n^2-8n+21)}{6} , \\ & \tfrac{(n-4)(3n^2-31n+84)}{2} , \, \tfrac{7(n-4)(n-5)(n-6)}{6} , \, \tfrac{(n-4)(n-5)(n-6)}{3} \big) \end{align*} (cf.~\cite[Chapter~8]{MR96a:52011}). From this we obtain $f(P_7^n)=(f_0(n),\ldots,f_6(n))$ by Proposition~\ref{prop:f-consum}: \begin{align*} f_0(n) \; = \; \tfrac{(n-3)(n^2-12n+41)}{3} \quad , & \quad f_1(n) \; = \; \tfrac{7n^3-102n^2+515n-840}{6} , \\ f_2(n) \; = \; \tfrac{5n^3-66n^2+313n-504}{3} \quad , & \quad f_3(n) \; = \; \tfrac{5(n-4)(n^2-8n+21)}{3} \end{align*} By symmetry of $f(P_7^n)$, these entries suffice to verify logarithmic convexity. We get \begin{align*} \frac{f_1(n)^2}{f_0(n)f_2(n)} \; & = \; \frac{(7n^3-102n^2+515n-840)^2}{4(n-3)(n^2-12n+41)(5n^3-66n^2+313n-504)} \; > \; 1 \, , \\ \frac{f_2(n)^2}{f_1(n)f_3(n)} \; & = \; \frac{2(5n^3-66n^2+313n-504)^2}{5(n-4)(n^2-8n+21)(7n^3-102n^2+515n-840)} \; > \; 1 \, , \\ \frac{f_3(n)^2}{f_2(n)f_4(n)} \; & = \; \frac{25(n-4)^2(n^2-8n+21)^2}{(5n^3-66n^2+313n-504)^2} \; > \; 1 \end{align*} for $n \geq 8$. Since the leading coefficients of the polynomials in the numerator and the denominator of the last fraction are equal, \begin{displaymath} \frac{f_3(P_7^n)^2}{f_2(P_7^n) f_4(P_7^n)} \; \stackrel{n \rightarrow \infty}{\longrightarrow} \; 1 . \end{displaymath} \qed \end{appendix} \vfill \bibliographystyle{siam}
{ "timestamp": "2005-12-06T18:54:42", "yymm": "0512", "arxiv_id": "math/0512131", "language": "en", "url": "https://arxiv.org/abs/math/0512131", "abstract": "We consider unimodality and related properties of f-vectors of polytopes in various dimensions. By a result of Kalai (1988), f-vectors of 5-polytopes are unimodal. In higher dimensions much less can be said; we give an overview on current results and present a potentially interesting construction as well as a conjecture arising from this.", "subjects": "Combinatorics (math.CO)", "title": "Unimodality and convexity of f-vectors of polytopes", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9835969665221771, "lm_q2_score": 0.8152324848629214, "lm_q1q2_score": 0.8018601991215062 }
https://arxiv.org/abs/2110.03755
Fast and stable approximation of analytic functions from equispaced samples via polynomial frames
We consider approximating analytic functions on the interval $[-1,1]$ from their values at a set of $m+1$ equispaced nodes. A result of Platte, Trefethen \& Kuijlaars states that fast and stable approximation from equispaced samples is generally impossible. In particular, any method that converges exponentially fast must also be exponentially ill-conditioned. We prove a positive counterpart to this `impossibility' theorem. Our `possibility' theorem shows that there is a well-conditioned method that provides exponential decay of the error down to a finite, but user-controlled tolerance $\epsilon > 0$, which in practice can be chosen close to machine epsilon. The method is known as \textit{polynomial frame} approximation or \textit{polynomial extensions}. It uses algebraic polynomials of degree $n$ on an extended interval $[-\gamma,\gamma]$, $\gamma > 1$, to construct an approximation on $[-1,1]$ via a SVD-regularized least-squares fit. A key step in the proof of our main theorem is a new result on the maximal behaviour of a polynomial of degree $n$ on $[-1,1]$ that is simultaneously bounded by one at a set of $m+1$ equispaced nodes in $[-1,1]$ and $1/\epsilon$ on the extended interval $[-\gamma,\gamma]$. We show that linear oversampling, i.e., $m = c n \log(1/\epsilon) / \sqrt{\gamma^2-1}$, is sufficient for uniform boundedness of any such polynomial on $[-1,1]$. This result aside, we also prove an extended impossibility theorem, which shows that such a possibility theorem (and consequently the method of polynomial frame approximation) is essentially optimal.
\section{Introduction} In this paper, we consider the problem of approximating an analytic function $f : [-1,1] \rightarrow \bbC$ from its values at $m+1$ equispaced points in $[-1,1]$. Several years ago, Platte, Trefethen \& Kuijlaars \cite{platte2011impossibility} demonstrated that this problem is intrinsically difficult. They proved an `impossibility' theorem which states that any method that offers exponential rates of convergence in $m$ for all functions analytic in a fixed, but arbitrary region of the complex plane must necessarily be exponentially ill-conditioned. Furthermore, the best rate of convergence achievable by a stable method is necessarily subexponential -- specifically, root-exponential in $m$. This result generalizes what has long been known for polynomial interpolation at equispaced nodes: namely, Runge's phenomenon. Polynomial interpolation is divergent for functions that not analytic in a sufficiently large complex region (the Runge region). And while it converges exponentially fast for functions that are analytic in this region, its condition number is also exponentially large. Such ill-conditioning means that, when computed in floating point arithmetic, the error of polynomial interpolation eventually increases, even for entire functions, due to the accumulation of round-off error. Many methods have been proposed to overcome Runge's phenomenon by stably and accurately approximating analytic functions from equispaced nodes (see, e.g., \cite{adcock2016mapped,boyd2009exponentially,platte2011impossibility} and references therein). Several such methods appear to offer fast convergence in practice{, which, for some functions at least, appears faster than root-exponential.} Yet full mathematical explanations for this phenomenon have hereto been lacking. The purpose of this paper is to fully analyze one such method in view of the impossibility theorem. This method is termed \textit{polynomial frame approximation} or \textit{polynomial extensions} \cite{adcock2020approximating}, and is closely related to so-called \textit{Fourier extensions} \cite{matthysen2017function,lyon2012fast,huybrechs2010fourier,bruno2007accurate,boyd2002fourier,pasquetti1996spectral}. It approximates a function $f$ on $[-1,1]$ using orthogonal polynomials on an \textit{extended} interval $[-\gamma,\gamma]$ for some fixed $\gamma > 1$. The approximation is then computed by solving a regularized least-squares problem, with user-controlled regularization parameter $\epsilon > 0$. Our main contribution is to show that this method offers a positive counterpart to the impossibility theorem of \cite{platte2011impossibility}. We prove a `possibility' theorem, which asserts that the polynomial frame approximation method is well-conditioned and, for all functions that are analytic in a sufficiently large region (related to the parameter $\gamma$), its error decreases exponentially fast down to roughly $\ordu{m^{3/2} \epsilon}$, where $\epsilon$ is the user-determined tolerance. This tolerance may be taken to be on the order of machine epsilon without impacting the conditioning of the method, thus rendering the approach suitable for practical purposes. But it may also be taken larger (with benefits in terms of stability) if fewer digits of accuracy are required. While the impossibility theorem dictates that exponential convergence to zero cannot be achieved by a well-conditioned method, we show that exponential decrease of the error down to an arbitrary tolerance, multiplied by the slowly growing factor proportional to $m^{3/2}$, is indeed possible. Additionally, we establish an `extended' impossibility theorem, which relates conditioning and error decay for approximation methods that achieve only a finite accuracy $\epsilon$. This theorem explains how the method we consider is essentially optimal, in a suitable sense. Our main result hinges on a new bound for the maximal behaviour of polynomials that are bounded simultaneously at a set of $m+1$ equispaced nodes on $[-1,1]$ and on the extended interval $[-\gamma,\gamma]$. The impossibility theorem of \cite{platte2011impossibility} uses a classical result of Coppersmith and Rivlin \cite{coppersmith1992growth}, which states that a polynomial of degree $n$ that is bounded by one at $m \geq n$ equispaced nodes can grow as large as $\alpha^{n^2/m}$ on $[-1,1]$ outside of these nodes, where $\alpha > 1$ is a constant. In particular, $m = c n^2$ equispaced points are both sufficient and necessary to ensure boundedness of such a polynomial on $[-1,1]$. We consider a nontrivial variation of this setting, where the polynomial is also assumed to be no larger than $1/\epsilon$ on $[-\gamma,\gamma]$. Our key result shows that $m = c n \log(1/\epsilon) / \sqrt{\gamma^2-1}$ equispaced points suffice for ensuring boundedness of any such polynomial on $[-1,1]$. While we use this bound to analyze polynomial frames, we expect it to be of independent interest from a pure approximation-theoretic perspective. { More broadly, our work has some connections with optimal recovery. We note that optimal recovery of functions in general, and from equispaced samples in particular has been actively studied in approximation theory, with emphasis on optimal approximation methods without taking stability issues into considerations. See, for instance, \cite{temlyakov1993approximate,michelli1985lecture,temlyakov2018multivariate}. } The outline of the remainder of this paper is as follows. In \S \ref{s:overview} we present an overview of the main components of the paper. We introduce notation, describe the impossibility theorem of \cite{platte2011impossibility} and then present polynomial frame approximation. We next state our main results, then conclude with a discussion of related work. In \S \ref{s:acc-stab-poly-frame} we analyze the accuracy and conditioning of polynomial frame approximation. Then in \S \ref{s:maximal-behaviour-proof} we give the proof of the aforementioned result on the maximal behaviour of polynomials bounded at equispaced nodes and on the extended interval $[-\gamma,\gamma]$. With this in hand, in \S \ref{s:main-thm-proof} we give the proof of the main results: the possibility theorem for polynomial frame approximation and the extended impossibility theorem. We then present several numerical examples in \S \ref{s:numerical}, before offering some concluding remarks in \S \ref{s:conclusions}. \section{Overview of the paper}\label{s:overview} \subsection{Notation} Throughout, $\bbP_n$ denotes the space of polynomials of degree at most $n$. For $m \geq 1$, we let $\{ x_i \}^{m}_{i=0}$ be a set of $m+1$ equispaced points in $[-1,1]$ including endpoints, i.e.\ $x_i = -1 + 2 i /m$. Given an interval $I$, we let $C(I)$ be the space of continuous functions on $I$ and \bes{ \nm{g}_{I,\infty} = \sup_{x \in I} |g(x)|,\quad g \in C(I), } be the uniform norm over $I$. We also let \bes{ \ip{f}{g}_{I,2} = \int_{I} f(x) \overline{g(x)} \D x,\quad f , g \in C(I), } be the usual $L^2$-inner product over $I$ and $\nm{\cdot}_{I,2} = \sqrt{\ip{\cdot}{\cdot}_{I,2}}$ be the corresponding $L^2$-norm. Next, we define several discrete semi-norms and semi-inner products. We define \bes{ \nm{g}_{m,\infty} = \max_{i = 0,\ldots,m} |g(x_i) |,\quad g \in C(I), } where $\{ x_i \}^{m}_{i=0}$ is the equispaced grid on $[-1,1]$, and \bes{ \ip{f}{g}_{m,2} = \frac{2}{m+1} \sum^{m}_{i=0} f(x_i) \overline{g(x_i)},\quad f , g \in C(I). } We also let $\nm{\cdot}_{m,2} = \sqrt{\ip{\cdot}{\cdot}_{m,2}}$ be the corresponding discrete semi-norm. Note that $\ip{\cdot}{\cdot}_{m,2}$ is an inner product on $\bbP_n$ for any $m \geq n$, since a polynomial of degree $n$ cannot vanish at $m+1 \geq n+1$ distinct points unless it is the zero polynomial. Observe also that \be{ \label{discnormbds} \sqrt{2/(m+1)} \nm{f}_{m,\infty} \leq \nm{f}_{m,2} \leq \sqrt{2} \nm{f}_{m,\infty} \leq \sqrt{2} \nm{f}_{[-1,1],\infty}, } for any $f \in C(I)$. Finally, given a compact set $E \subset \bbC$, we write $B(E)$ for the set of functions that are continuous on $E$ and analytic in its interior. We also define $\nm{f}_{E,\infty} = \sup_{z \in E} |f(z)|$. \subsection{The impossibility theorem} Throughout this paper, we consider families of mappings \bes{ \cR_m : C([-1,1]) \rightarrow C([-1,1]), } where, for each $m \geq 1$ and $f \in C([-1,1])$, $\cR_m(f)$ depends only on the values $\{f(x_i)\}^{m}_{i=0}$ of $f$ on the equispaced grid $\{x_i\}^{m}_{i=0}$. We define the (absolute) condition number of $\cR_m$ (in terms of the continuous and discrete uniform norms) as \be{ \label{kappa-def} \kappa(\cR_m) = \sup_{f \in C([-1,1])} \lim_{\delta \rightarrow 0^+} \sup_{\substack{h \in C([-1,1]) \\ 0 < \nm{h}_{m,\infty} \leq \delta }} \frac{\nmu{\cR_{m}(f+h) - \cR_{m}(f) }_{[-1,1],\infty}}{\nm{h}_{m,\infty}}. } We are now ready to state the impossibility theorem: \thm{ [The impossibility theorem, \cite{platte2011impossibility}] \label{t:impossibility} Let $E \subset \bbC$ be a compact set containing $[-1,1]$ in its interior and $\{ \cR_m \}^{\infty}_{m=1}$ be an approximation procedure based on equispaced grids of $m+1$ points such that, for some $C,\rho > 1$ and $1/2 < \tau \leq 1$, we have \bes{ \nmu{f - \cR_m(f) }_{[-1,1],\infty} \leq C \rho^{-m^{\tau}} \nm{f}_{E,\infty},\quad \forall m \in \bbN,\ f \in B(E). } Then the condition numbers \R{kappa-def} satisfy \bes{ \kappa(\cR_m) \geq \sigma^{m^{2 \tau - 1}} } for some $\sigma > 1$ and all sufficiently large $m$. } When $\tau = 1$, this implies that any approximation procedure that achieves exponential convergence at a geometric rate must also be exponentially ill-conditioned at a geometric rate. Furthermore, the best possible (and achievable \cite{adcock2019optimal}) rate of convergence of a stable method is root-exponential in $m$, i.e.\ the error decays like $\rho^{-\sqrt{m}}$ for some $\rho > 1$. \subsection{Polynomial frame approximation} We now describe polynomial frame approximation. As observed, this method was formalized in \cite{adcock2020approximating}, and is related to earlier works on Fourier extensions \cite{adcock2014parameter,adcock2014numerical,matthysen2017function,lyon2012fast,huybrechs2010fourier,bruno2007accurate,boyd2002fourier,pasquetti1996spectral}, and more generally, numerical approximations with frames \cite{adcock2019frames,adcock2020frames}. Let $\gamma > 1$. Polynomial frame approximation uses orthogonal polynomials on an extended interval $[-\gamma,\gamma]$ to construct an approximation to a function over $[-1,1]$. In this paper, we use orthonormal Legendre polynomials, although we remark in passing that other orthogonal polynomials such as Chebyshev polynomials could also be employed. Note that an orthonormal basis on $[-\gamma,\gamma]$ fails to constitute a basis when restricted to the smaller interval $[-1,1]$. It forms a so-called \textit{frame} \cite{adcock2019frames,christensen2016introduction}, hence the name polynomial `frame' approximation. Let $P_i(x)$ be the classical Legendre polynomial on $[-1,1]$, normalized so that $P_i(1) = 1$. Since $\int^{1}_{-1} |P_i(x)|^2 \D x = (i+1/2)^{-1}$, we define Legendre polynomial frame on $[-1,1]$ as $\{ \psi_i \}^{\infty}_{i=0}$, where the $i$th such function is given by \be{ \label{psii-def} \psi_i(x) = \sqrt{i+1/2} P_i (x/\gamma ) /\sqrt{\gamma},\quad x \in [-1,1]. } Let $m,n \geq 0$ and consider a function $f \in C([-1,1])$. Our aim is to compute a polynomial approximation to $f$ of the form \bes{ f \approx \hat{f} = \sum^{n}_{i=0} \hat{c}_i \psi_i \in \bbP_n } for suitable coefficients $\hat{c}_i$. It is natural to strive to do this via a least-squares fit, i.e. \be{ \label{LScoeff} \hat{c} = (\hat{c}_i)^{n}_{i=0} \in \argmin{c \in \bbC^{n+1}} \nm{A c - b}_{2}, } where \be{ \label{Adef} A = \sqrt{2/(m+1)} (\psi_j(x_i))^{m,n}_{i,j=0} \in \bbC^{m \times n},\qquad b = \sqrt{2/(m+1)} (f(x_i))^{m}_{i=0} \in \bbC^m. } Note that $\sqrt{2/(m+1)}$ is simply a normalization factor that is included for convenience. Unfortunately, as described in \cite{adcock2020approximating}, this least-squares problem is ill-conditioned for large $n$ (even when $m \gg n$) due to the use of a frame rather than a basis \cite{adcock2019frames}. Therefore, we instead solve a suitably regularized least-squares problem. There are a number of different ways to do this, but, following previous works we consider an $\epsilon$-truncated Singular Value Decomposition (SVD). Suppose that the least-squares matrix \R{Adef} has SVD $A = U \Sigma V^*$, where $\Sigma = \diag(\sigma_0,\ldots,\sigma_n) \in \bbR^{m \times n}$ is the diagonal matrix of singular values. Recall that the minimal $2$-norm solution $\hat{c}$ of \R{LScoeff} is given by \bes{ \hat{c} = A^{\dag} b = V \Sigma^{\dag} U^* b, } where $\dag$ denotes the pseudoinverse. Given $\epsilon > 0$, we define $\Sigma^{\epsilon}$ as $\epsilon$-regularized version of $\Sigma$ as \bes{ (\Sigma^{\epsilon} )_{ii} = \begin{cases} \sigma_i & \sigma_i > \epsilon \\ 0 & \mbox{otherwise} \end{cases} , } and let $\Sigma^{\epsilon,\dag}$ be its pseudoinverse, i.e. \bes{ (\Sigma^{\epsilon,\dag} )_{ii} = \begin{cases} 1/\sigma_i & \sigma_i > \epsilon \\ 0 & \mbox{otherwise} \end{cases} . } Then we define the $\epsilon$-regularized approximation of \R{LScoeff} as $\hat{c}^{\epsilon} = V \Sigma^{\epsilon,\dag} U^* b$ and the corresponding approximation to $f$ as \bes{ \hat{f}^{\epsilon} = \sum^{n}_{i=0} \hat{c}^{\epsilon}_i \psi_i. } With this in hand, we define the overall approximation procedure as the mapping \be{ \label{poly-frame-approx-1} \cP^{\epsilon,\gamma}_{m,n} : C([-1,1]) \rightarrow C([-1,1]),\ f \mapsto \hat{f}^{\epsilon} = \sum^{n}_{i=0} \hat{c}^{\epsilon}_{i} \psi_i, } where \be{ \label{poly-frame-approx-2} \hat{c}^{\epsilon} = (\hat{c}^{\epsilon}_i)^{n}_{i=0} = V \Sigma^{\epsilon,\dag} U^* b,\qquad b = \sqrt{2/(m+1)} (f(x_i))^{m}_{i=0} . } \rem{ [Why not use orthogonal polynomials on the original interval] \label{r:why-not-unit-interval} The use of orthogonal polynomials on an extended interval may at first seem bizarre, since, as noted, the infinite collection of such functions no longer forms a basis when restricted to $[-1,1]$, but a frame. And even though the first $n+1$ such functions $\psi_0,\ldots,\psi_n$ constitute a basis for $\bbP_n$, they are extremely ill-conditioned as $n \rightarrow \infty$. To be precise, the condition number of their Gram matrix $G = (\ip{\psi_j}{\psi_i}_{[-1,1],2} )^{n}_{i,j=0}$ is exponentially-large in $n$. In turn, the matrix $A$ is also exponentially ill-conditioned in $n$, even when $m \gg n$. To understand why, observe that $A^* A = (\ip{\psi_j}{\psi_i}_{m,2} )^{n}_{i,j=0}$ is simply a discrete approximation to $G$. Why then, do we not consider Legendre polynomials in $[-1,1]$? These constitute a perfectly well-conditioned basis, which means that the corresponding matrix $A$ would also be well-conditioned for $m \gg n$. Thus there is no need for regularization, and we may simply compute the least-squares fit by solving \R{LScoeff}. Note that this simply corresponds to $\cP^{\epsilon,\gamma}_{m,n}$ with $\epsilon = 0$ and $\gamma = 1$. The problem is that such an approximation, which we term \textit{polynomial least-squares approximation}, {behaves exactly as the} impossibility theorem (Theorem \ref{t:impossibility}) {predicts in the best case}. It is ill-conditioned if $m = o(n^2)$ as $n \rightarrow \infty$, and in particular, if $m \sim c n$ as $n \rightarrow \infty$, then the condition number of the mapping grows exponentially fast. On the other hand, with \textit{quadratic oversampling}, i.e.\ $m \sim c n^2$ as $n \rightarrow \infty$, the approximation is well-conditioned and its convergence rate is root-exponential in $m$ for all analytic functions. See \cite{adcock2019optimal} for a discussion on polynomial least-squares and the impossibility theorem. See also Remark \ref{r:least-squares-cond} below. } \subsection{Maximal behaviour of polynomials bounded at equispaced nodes} As mentioned, both the impossibility theorem and the subsequent possibility theorem rely on estimates for the maximal behaviour of polynomials that are bounded at equispaced nodes. The former is based on a classical result due to Coppersmith and Rivlin concerning the maximal growth of a polynomial $p \in \bbP_n$ that is at most one at the $m+1$ equispaced nodes $\{ x_i \}^{m}_{i=0}$. Specifically, in \cite{coppersmith1992growth} they showed that there exist constants $\beta \geq \alpha > 1$ such that, if \bes{ B(m,n) = \sup \{ \nm{p}_{[-1,1],\infty} : p \in \bbP_n,\ \nm{p}_{m,\infty} \leq 1 \}, } then \be{ \label{coppersmith-rivlin} \alpha^{n^2/m} \leq B(m,n) \leq \beta^{n^2/m},\quad \forall 1 \leq n \leq m. } \rem{ [The condition number of polynomial least-squares approximation] \label{r:least-squares-cond} It is not difficult to show that the condition number of polynomial least-squares approximation $\cP_{m,n} = \cP^{0,1}_{m,n}$ satisfies \bes{ B(m,n) \leq \kappa(\cP_{m,n}) \leq \sqrt{m+1} B(m,n), } where $B(m,n)$ is as in \R{coppersmith-rivlin} (this follows by setting $\gamma = 1$ and $\epsilon = 0$ in a result we show later, Lemma \ref{l:poly-frame-accuracy-stability}). Thus, \R{coppersmith-rivlin} immediately explains why this approximation is ill-conditioned when $m = o(n^2)$ as $n \rightarrow \infty$, and only well-conditioned when $m \sim c n^2$ (or faster) as $n \rightarrow \infty$. } As described above, the polynomial frame approximation is constructed via an $\epsilon$-truncated SVD. We will see later in \S \ref{s:acc-stab-poly-frame} that such truncation means that the approximation $\cP^{\epsilon,\gamma}_{m,n}(f)$ of a function $f$ belongs to (a subspace of) the set of polynomials \bes{ P^{\epsilon,\gamma}_{m,n} = \left \{ p \in \bbP_n : \nm{p}_{[-\gamma,\gamma],2} \leq \nm{p}_{m,2} / \epsilon \right \} \subseteq \bbP_n } whose $L^2$-norm over the extended interval $[-\gamma,\gamma]$ is at most $1/\epsilon$ times larger than their discrete $2$-norm over the equispaced grid. In other words, the effect of regularization via the truncated SVD is to restrict the type of polynomial the approximation can take to one that does not grow too large on the extended interval. After interchanging the $2$-norms for uniform norms, this observation motivates the study of the quantity \be{ \label{C-def-general} C(m,n,\gamma,\epsilon) = \sup \{ \nm{p}_{[-1,1],\infty} : p \in \bbP_n,\ \| p \|_{m,\infty} \leq 1,\ \| p \|_{[-\gamma,\gamma],\infty} \leq 1/\epsilon \}. } In \S \ref{s:acc-stab-poly-frame}, we show that the condition number of the polynomial frame approximation $\cP^{\epsilon,\gamma}_{m,n}$ satisfies \bes{ \kappa(\cP^{\epsilon,\gamma}_{m,n}) \leq \sqrt{m+1} C(m,n,\gamma,\epsilon). } Notice that $C(m,n,\gamma,0) = B(m,n)$ and, in general, \bes{ C(m,n,\gamma,\epsilon) \leq B(m,n), } where $B(m,n)$ is the as in \R{coppersmith-rivlin}. However, whereas $B(m,n)$ is only bounded as $m,n\rightarrow\infty$ in the quadratic oversampling regime (i.e.\ $m \sim c n^2$ for some $c>0$), for $C(m,n,\gamma,\epsilon)$ we show that linear oversampling is sufficient. Specifically: \thm{ [Maximal behaviour of polynomials bounded at equispaced nodes and extended intervals] \label{t:polynomial-inequality-main} Let $0 < \epsilon \leq 1/\E$, $\gamma > 1$ and $n \geq \sqrt{\gamma^2-1} \log(1/\epsilon)$, and consider the quantity $C(m,n,\gamma,\epsilon)$ defined in \R{C-def-general}. Suppose that \be{ \label{m-n-eps-poly-growth} m \geq 36 n \log(1/\epsilon) / \sqrt{\gamma^2-1}. } Then \bes{ C(m,n,\gamma,\epsilon) \leq c, } for some numerical constant $c > 0$. Specifically, $c$ can be taken as $c = 4 \beta + 3$ for $\beta$ as in \R{coppersmith-rivlin}. } This theorem is a direct consequence of a more general result (Theorem \ref{t:poly-inequality-general}) that we state and prove in \S \ref{s:maximal-behaviour-proof}. Note that the factor $36$ appearing in \R{m-n-eps-poly-growth} was chosen for convenience. Theorem \ref{t:poly-inequality-general} describes in general how a condition roughly of the form $m \geq c_1 n \log(1/\epsilon) / \sqrt{\gamma^2-1}$ leads to a bound $C(m,n,\gamma,\epsilon) \leq h(c_1)$ for some function $h(c_1)$ depending on $c_1$. \subsection{The motivations for considering polynomials on an extended interval} Above we asserted that, by considering orthogonal polynomials on an extended interval and using regularization, the polynomial frame approximation constructs an approximation in a space within which the polynomials cannot grow too large on $[-\gamma,\gamma]$. As Theorem \ref{t:polynomial-inequality-main} makes clear, this prohibits such polynomials from behaving too wildly on $[-1,1]$ away from the equispaced grid, whenever $m$ scales linearly with $n$. Thus, using orthogonal polynomials on an extended interval allows for a reduction in the oversampling rate from quadratic in $n$ (as is the case for standard polynomial-least squares approximation -- see Remark \ref{r:least-squares-cond}) to linear in $n$. On the other hand, by restricting the approximation space in this way, we potentially limit the ability of the scheme to accurately approximate analytic functions. In Theorem \ref{t:acc-cond-poly-frame}, we establish an error bound for the polynomial frame approximation that takes the form \bes{ \nmu{f - \cP^{\epsilon',\gamma}_{m,n}(f)}_{[-1,1],\infty} \leq 2 c \sqrt{m+1} \inf_{p \in \bbP_n} \left \{ \nm{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty} \right \}, } where $\epsilon' = \epsilon (n+1) / \sqrt{\gamma}$ (the reasons for this choice of $\epsilon'$ are discussed further below). This holds for any $c > 1$ and $f \in C([-1,1])$, provided $m$ and $n$ are chosen so that \bes{ C(m,n,\gamma,\epsilon) \leq c. } Such a bound is very similar to those shown previously for both Fourier extensions \cite{adcock2014numerical} and polynomial frame approximations \cite{adcock2020approximating}, the main difference being the use of the $L^{\infty}$-norm instead of the $L^2$-norm. The key component of it is the best approximation term \bes{ \inf_{p \in \bbP_n} \left \{ \nm{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty} \right \}. } In other words, the effect of the truncated SVD regularization is to replace the classical best approximation error term \bes{ \inf_{p \in \bbP_n} \nm{f - p}_{[-1,1],\infty}, } (which arises in the case $\epsilon = 0$, i.e.\ standard polynomial least-squares approximation) by one that also involves a term depending on $\epsilon$ multiplied by the norm of $p$ over the extended interval. As a result, the overall approximation error depends on how well $f$ can be approximated by a polynomial $p \in \bbP_n$ uniformly on $[-1,1]$ (the term $\nm{f - p}_{[-1,1],\infty}$) that does not grow too large on the extended interval $[-\gamma,\gamma]$ (the term $ \nm{p}_{[-\gamma,\gamma],\infty} $). Our main result, stated next, arises by bounding this best approximation term for functions that are analytic in sufficiently large complex regions. \subsection{Main result: the possibility theorem} We now state our main result (see \S \ref{s:main-thm-proof} for the proof). For this, we first recall the definition of the Bernstein ellipse with parameter $\theta > 1$: \be{ \label{Bernstein-ellipse} E_{\theta} = \left \{ \frac{z+z^{-1}}{2} : z \in \bbC,\ 1 \leq | z | \leq \theta \right \}. } A classical theorem in approximation theory states that any function $f \in B(E_{\theta})$ is approximated to exponential accuracy by polynomials. Specifically, \be{ \label{poly-BA} \inf_{p \in \bbP_n} \nm{f - p}_{[-1,1],\infty} \leq \frac{2}{\theta-1} \nm{f}_{E_{\theta},\infty} \theta^{-n}, } (see also Lemma \ref{poly-approx-bounds} later). Our main results asserts that polynomial frame approximation can achieve a similar rate of decay in $n$, subject to linear oversampling in $m$. Specifically: \thm{ [The possibility theorem] \label{t:possibility-thm} Let $0 < \epsilon \leq 1/\E$, $\gamma > 1$ and $n \geq \sqrt{\gamma^2-1} \log(1/\epsilon)$, and consider the polynomial frame approximation $\cP^{\epsilon',\gamma}_{m,n}$ defined in \R{poly-frame-approx-1}--\R{poly-frame-approx-2}, where \be{ \label{m-n-relation} m = \left \lceil 36 n \log(1/\epsilon) \big / \sqrt{\gamma^2-1} \right \rceil ,\qquad \epsilon' = \frac{\epsilon(n+1)}{\sqrt{\gamma}}. } Then the condition number of the mapping $\cP^{\epsilon',\gamma}_{m,n}$ satisfies \bes{ \kappa(\cP^{\epsilon',\gamma}_{m,n}) \leq c \sqrt{m+1} , } where $c$ is as in Theorem \ref{t:polynomial-inequality-main}. Moreover, if $E_{\theta}$ is a Bernstein ellipse with parameter \be{ \label{theta-cond} \theta > \gamma + \sqrt{\gamma^2-1} } then, for all $f \in B(E_{\theta})$, \be{ \label{main-err-bd} \begin{split} \nmu{f - \cP^{\epsilon',\gamma}_{m,n}(f)}_{[-1,1],\infty} & \leq c g(\theta,\gamma) \sqrt{m} \left ( \theta^{-n} + n \epsilon \right ) \nm{f}_{E_{\theta},\infty} \\ & \leq c g(\theta,\gamma) \sqrt{m} \left ( \rho^{1-m} + m \epsilon \right ) \nm{f}_{E_{\theta},\infty}, \end{split} } where $g(\theta,\gamma)$ depends on $\theta$ and $\gamma$ only and \be{ \label{rho-main-thm} \rho = \theta^{c_*},\qquad c_* = \frac{\sqrt{\gamma^2-1} }{36 \log(1/\epsilon)} . } } This result shows that polynomial frame approximation is well-conditioned when $n$ scales linearly with $m$ (specifically, \R{m-n-relation} holds), with its condition number being at worst $\ordu{\sqrt{m}}$ as $m \rightarrow \infty$. Moreover, for functions that are analytic in $E_{\theta}$ (note that this region contains the extended interval $[-\gamma,\gamma]$ in its interior, due to the condition \R{theta-cond}) its error decreases exponentially fast in $m$ down to the level $\ordu{m^{3/2} \epsilon}$. Recall that the rate $\theta^{-n}$ in \R{main-err-bd} is the same as in \R{poly-BA} for the best polynomial approximation of a function in $B(E_{\theta})$. Thus, one can achieve a near-optimal error decay rate in $n$ with only linear oversampling in $m$. Overall, Theorem \ref{t:possibility-thm} provides a positive counterpart to the impossibility theorem (Theorem \ref{t:impossibility}). It is important emphasize that it does not violate Theorem \ref{t:impossibility}: indeed, exponential decay of the error is only guaranteed down to a finite accuracy. \rem{ [Varying $\epsilon$ with $n$] \label{rem:varying-eps} It is worth observing how the theorem changes if one strives to scale $\epsilon$ with $n$ so as to achieve exponential convergence of the error down to zero{, rather than exponential decay down to a finite level of accuracy}. If $\epsilon$ is chosen as $\epsilon = \theta^{-n}$ then the scaling between $m$ and $n$ becomes quadratic, since $\log(1/\epsilon) = n \log(\theta)$ in this case. When substituted into Theorem \ref{t:polynomial-inequality-main}, this implies quadratic scaling of $m$ with $n$, and therefore root-exponential convergence {(to zero)} of the approximation with respect to $m$. {This is in precise agreement with the impossibility theorem.} } Another key aspect of Theorem \ref{t:possibility-thm} is the dependence on $\epsilon$ in \R{m-n-relation} and, in turn, the exponential rate \R{rho-main-thm}. Since $\epsilon > 0$ dictates the limiting accuracy of the approximation scheme, it is often desirable to choose $\epsilon$ close to machine epsilon $\epsilon_{\mathrm{mach}}$, which in IEEE double precision is roughly $\epsilon_{\mathrm{mach}} \approx 1.1 \times 10^{-16}$. Thus, the scaling $\log(1/\epsilon)$ -- which is proportional to the number of digits of accuracy desired -- is highly appealing. A scaling of, for example, $1/\epsilon$, would be meaningless for practical purposes. Note that Theorem \ref{t:possibility-thm} does not say anything about the rate of the decay of the error for functions that are not analytic in a Bernstein ellipse $E = E_{\theta}$ that is large enough to contain the extended interval $[-\gamma,\gamma]$. We discuss the behaviour of the error for such functions in \S \ref{ss:lower-regularity}. On the other hand, this theorem also offers some insight into the effect of the choice of $\gamma$ on the approximation. Specifically, choosing a smaller $\gamma$ means that \R{theta-cond} holds for smaller values of $\theta$, thus the analyticity requirement $f \in B(E_{\theta})$ becomes less stringent. However, this also leads to a slower rate of exponential convergence in $m$, since $\rho$ is an increasing function of both $\gamma$ and $\theta$. We discuss this matter further in \S \ref{s:numerical}. Finally, we remark that this theorem actually considers a polynomial frame approximation with parameter $\epsilon' = \epsilon(n+1)/\sqrt{\gamma}$ that grows linearly in $n$. The reason for this can be traced to the need to switch between the $L^2$-norm (or corresponding discrete seminorm) and the $L^{\infty}$-norm (or corresponding discrete seminorm) at various stages in the proof. See the proofs of Lemma \ref{C1-C2-as-C} and Theorem \ref{t:acc-cond-poly-frame} for the precise details. This choice of scaling is made to ensure the first term in the error bound decreases exponentially fast in $m$, which in turn follows from the linear relationship $m \approx 24 n \log(1/\epsilon)/\sqrt{\gamma^2-1}$. It is also possible to use $\epsilon$ as the truncation parameter rather than $\epsilon'$. Following much the same arguments, one can show that this choice results in a log-linear relationship between $m$ and $n$, which in turn leads to subexponential convergence of the form $\sigma^{-m / \log(m)}$ for large $m$, where, like $\rho$, $\sigma > 1$ depends on $\theta$, $\gamma$ and $\epsilon$. \subsection{Error decay rates for functions of lower regularity}\label{ss:lower-regularity} Theorem \ref{t:possibility-thm} only asserts exponential decay of the error for functions that are analytic in complex regions containing the extended interval $[-\gamma,\gamma]$. We now consider arbitrary analytic functions. The following result shows that the error for such functions decays exponentially fast with the same rate $\theta^{-n}$, but only down to a larger tolerance. \thm{ [Error decay for arbitrary analytic functions] \label{t:possibility-slower-exp} Consider the setup of Theorem \ref{t:possibility-thm}. Let $E_{\theta}$ be the Bernstein ellipse with parameter \bes{ 1 < \theta < \tau : =\gamma + \sqrt{\gamma^2-1}. } Then, for all $f \in B(E_{\theta})$, \be{ \label{main-err-bd2} \begin{split} \nmu{f - \cP^{\epsilon',\gamma}_{m,n}(f)}_{[-1,1],\infty} & \leq c g(\theta,\gamma) \sqrt{m} \left ( \theta^{-n} + n \epsilon^{\frac{\log(\theta)}{\log(\tau)}} \right ) \nm{f}_{E_{\theta},\infty} \\ & \leq c g(\theta,\gamma) \sqrt{m} \left ( \rho^{1-m} + m \epsilon^{\frac{\log(\theta)}{\log(\tau)}} \right ) \nm{f}_{E_{\theta},\infty}, \end{split} } where $g(\theta,\gamma)$ and $\rho$ are also as in Theorem \ref{t:possibility-thm}. } This result shows decrease of the error exponentially fast down to roughly $\epsilon^{\frac{\log(\theta)}{\log(\tau)}}$, i.e.\ some fractional power of $\epsilon$ depending on the relative sizes of $\theta$ and $\tau = \gamma + \sqrt{\gamma^2-1}$. It raises an immediate question: what happens after such an accuracy level is reached? As we show later through an extended impossibility theorem, we cannot expect exponential decay down to $\epsilon$ in general. However, we now show that superalgebraic decay -- i.e.\ faster than any power of $m^{-k}$ -- is indeed possible down to this level. We do this by first noting that an analytic function is infinitely smooth, and then by establishing an error bound for functions that are $k$-times continuously differentiable. Specifically, in the following result we consider the space $C^{k}([-1,1])$ of functions that are $k$-times continuously differentiable on $[-1,1]$. We define the norm on this space as \bes{ \nm{f}_{C^k([-1,1])} = \max_{j = 0,\ldots,k} \nmu{f^{(j)}}_{[-1,1],\infty}. } \thm{ [Error decay for $C^k$ functions] \label{t:possibility-algebraic} Consider the setup of Theorem \ref{t:possibility-thm}. Then, for all $k \in \bbN$ and $f \in C^k([-1,1])$, \bes{ \nmu{f - \cP^{\epsilon',\gamma}_{m,n}(f)}_{[-1,1],\infty} \leq c g(k,\gamma) \sqrt{m} \left ( n^{-k} + n \epsilon \right ) \nm{f}_{C^k([-1,1])}, } where $g(k,\gamma)$ depends on $k$ and $\gamma$ only. } Proofs of Theorems \ref{t:possibility-slower-exp} and \ref{t:possibility-algebraic} can be found in \S \ref{s:main-thm-proof}. Note that all these observations about the rate of error decay are seen in practice in numerical examples. We present a series of experiments confirming these results in \S \ref{s:numerical}. We remark in passing that superalgebraic decay is slower than root-exponential decay, which is the best possible stipulated by the impossibility theorem for analytic function approximation. Whether or not polynomial frame approximation exhibits root-exponential decay in $m$ after the breakpoint $\epsilon^{\frac{\log(\theta)}{\log(\tau)}}$ is an open problem. \subsection{An extended impossibility theorem} Theorems \ref{t:possibility-thm} and \ref{t:possibility-slower-exp} show that polynomial frame approximation can achieve roughly the same exponential rate $\theta^{-n}$ as the best polynomial approximation for functions in $B(E_{\theta})$ when subject to linear oversampling. However, it only maintains this rate down to roughly $\epsilon$ for sufficiently large $\theta$. We now ask whether or not there exists an approximation scheme that can perform better than this: namely, whether an exponential rate $\theta^{-n}$ down to $\epsilon$ can be attained for any $\theta > 1$ in the linear oversampling regime. The following extended impossibility theorem shows that the answer to this question is no. \thm{[Extended impossibility theorem] \label{t:impossibility-extended} Let $\{ \cR_m \}^{\infty}_{m=1}$ be an approximation procedure based on equispaced grids of $m+1$ points such that, for some $c > 0$, $C,\theta^* > 1$, $0 < \epsilon \leq (4C)^{-2}$ and $1/2 < \tau \leq 1$, we have \be{ \label{extended-error-bound} \nmu{f - \cR_m(f) }_{[-1,1],\infty} \leq C \left ( \theta^{-c m^{\tau}} + \epsilon \right ) \nm{f}_{E_{\theta},\infty},\quad } for all $m \in \bbN$, $f \in B(E_{\theta})$ and $1 < \theta \leq \theta^*$. Then the condition numbers \R{kappa-def} satisfy \bes{ \kappa(\cR_m) \geq \sigma^{m^{2 \tau - 1}} } for some $\sigma > 1$ and all sufficiently large $m$. } The proof of this theorem can be found in \S \ref{s:numerical}. It follows essentially the same steps as that of the impossibility theorem. This theorem has several consequences. First, it extends the impossibility theorem by showing that exactly the same relationship between fast error decay and conditioning holds even when the overall error decreases only down to a constant tolerance $\epsilon$. To do this, it makes the stronger assumption that the scheme yields exponential decay for all analytic functions -- including those with singularities arbitrarily close to $[-1,1]$ -- with the rate of exponential decay being dependent on the size of the region of analyticity. Specifically, $f \in B(E_{\theta})$ implies a term of the form $\theta^{-c m^{\tau}}$ for \textit{all} $1 < \theta \leq \theta^*$. Second, it implies the following. Any well-conditioned method must either (i) yield root-exponential decrease of the error down to $\epsilon$, i.e. $\theta^{-c \sqrt{m}} + \epsilon$; or (ii) fail to yield exponential convergence for all analytic functions, i.e.\ \R{extended-error-bound} holds with $\tau = 1$ only for $\theta^{**} \leq \theta \leq \theta^*$ for some $1 < \theta^{**} \leq \theta^*$. As discussed previously, (ii) is exactly how polynomial frame approximation behaves, up to small algebraic factors in $m$. In other words, it is nearly optimal. \rem{ [Varying $\gamma$ with $n$] Recall that Theorem \ref{t:possibility-thm} only asserts exponential decay down to $\epsilon$ for functions that are analytic Bernstein ellipses containing the interval $[-\gamma,\gamma]$. A natural idea is therefore to decrease $\gamma$ with $n$ so that, for any fixed, analytic function, the interval $[-\gamma,\gamma]$ is included in its region of analyticity for all large $n$. To determine a suitable scaling for $\gamma$, one can use similar ideas to those employed in so-called \textit{mapped polynomial} spectral methods \cite{adcock2016mapped,don1997accuracy,boyd1989chebyshev,hale2008new,kosloff1993modified}: namely, choose $\gamma$ to match the terms in the error bound \R{main-err-bd}. Ignoring the term $n$ (for simplicity), we therefore solve the equation $(\gamma + \sqrt{\gamma^2-1})^{-n} = \epsilon$ with respect to $\gamma$, which yields \be{ \label{gamma-scale} \gamma = \gamma(n,\epsilon) = \frac{\epsilon^{1/n} + \epsilon^{-1/n}}{2}. } Observe that $\gamma \rightarrow 1^+$ as $n \rightarrow \infty$ for fixed $n$. By combining Theorems \ref{t:possibility-thm} (for large $n$, since $\gamma < \theta$ for all large $n$) and \ref{t:possibility-slower-exp} (for small $n$, noting that $\epsilon^{\frac{\log(\tau)}{\log(\theta)}} = \theta^{-n}$ with this choice of $\gamma$), one can show that \be{ \label{gamma-scale-err-bound} \nmu{f - \cP^{\epsilon',\gamma}_{m,n}(f)}_{[-1,1],\infty} \leq c g(\theta,\gamma) \sqrt{m} \left ( \theta^{-n} + n \epsilon \right ) \nm{f}_{E_{\theta},\infty}, } for $f \in B(E_{\theta})$ and \textit{any} $\theta > 1$. Yet, scaling $\gamma$ as in \R{gamma-scale} causes the relation between $m$ and $n$ to become quadratic. Indeed, $\sqrt{\gamma^2 - 1} = \log(1/\epsilon)/n + \ord{(\log(1/\epsilon)/n)^3}$ as $n \rightarrow \infty$, and therefore the condition \R{m-n-eps-poly-growth} becomes $m > 36 n^2$ for all large $n$. Of course, this is to be expected in view of Theorem \ref{t:impossibility-extended}, since the error bound \R{gamma-scale-err-bound} holds for any $\theta > 1$. } \section{Accuracy and conditioning of $\cP^{\epsilon,\gamma}_{m,n}$}\label{s:acc-stab-poly-frame} The next three sections develop the proofs of Theorems \ref{t:polynomial-inequality-main}--\ref{t:impossibility-extended}. We commence in this section by analyzing the error and condition number of the polynomial frame approximation $\cP^{\epsilon,\gamma}_{m,n}$. Our analysis follows similar lines to that of \cite{adcock2020frames} and, in particular, \cite{adcock2020approximating}, the main difference being that we work in the $L^{\infty}$-norm, rather than the $L^2$-norm. \subsection{Reformulation in terms of singular vectors} Let $v_0,\ldots,v_{n} \in \bbC^n$ be the right singular vectors of the matrix $A$ defined in \R{Adef}. Then, to each such vector, we can associate a polynomial in $\bbP_n$: \bes{ \xi_i = \sum^{n}_{j=0} (v_i)_j \psi_j \in \bbP_n,\quad i =0,\ldots,n. } Here the $\psi_j$ are as in \R{psii-def}. Since these functions are orthonormal on $[-\gamma,\gamma]$ and the $v_i$ are orthonormal vectors, the functions $\xi_i$ are themselves orthonormal on $[-\gamma,\gamma]$: \bes{ \ip{\xi_i}{\xi_j}_{[-\gamma,\gamma],2} = v^*_j v_i = \delta_{ij},\quad i,j = 0,\ldots,n. } However, the functions $\xi_i$ are also orthogonal with respect to the discrete inner product $\ip{\cdot}{\cdot}_{m,2}$. Indeed, since $v_j$ and $v_k$ are singular vectors, we have \eas{ \ip{\xi_j}{\xi_k}_{m,2} = \frac{2}{m+1} \sum^{m}_{i=0} \xi_j(x_i) \overline{\xi_k(x_i)} & = \frac{2}{m+1} \sum^{m}_{i=0} \sum^{n}_{s=0} \sum^{n}_{t=0} (v_j)_s \psi_s(x_i) \overline{(v_k)_t} \overline{\psi_t(x_i)} \\ & = v^*_k A^* A v_j \\ & = \sigma^2_j \delta_{jk}. } With this in hand, we now define the subspace \bes{ \bbP^{\epsilon,\gamma}_{m,n} = \spn \{ \xi_i : \sigma_i > \epsilon \} \subseteq \bbP_n. } We note in passing that this space coincides with $\bbP_n$ whenever $m \geq n$ and $\sigma_{\min} = \sigma_n > \epsilon$. In particular, $\bbP^{0,\gamma}_{m,n} = \bbP_n$ for $m \geq n$. On the other hand, when $\epsilon > 0$ we have $\bbP^{\epsilon,\gamma}_{m,n} \subseteq P^{\epsilon,\gamma}_{m,n}$, where \bes{ P^{\epsilon,\gamma}_{m,n} = \left \{ p \in \bbP_n : \nm{p}_{[-\gamma,\gamma],2} \leq \nm{p}_{m,2} / \epsilon \right \} \subseteq \bbP_n, } is, as before, the set of polynomials of degree at most $n$ whose $L^2$-norm over the extended interval $[-\gamma,\gamma]$ is at most $1/\epsilon$ times larger than their discrete $2$-norm over the equispaced grid. To see why this holds, let $p = \sum_{\sigma_i > \epsilon} c_i \xi_i \in \bbP^{\epsilon,\gamma}_{m,n}$. Then, by the double orthogonality of the $\xi_i$, \bes{ \nm{p}^2_{[-\gamma,\gamma,2]} = \sum_{ \sigma_i > \epsilon} |c_i|^2 \leq \frac{1}{\epsilon^2} \sum_{ \sigma_i > \epsilon} \sigma^2_i |c_i|^2 = \frac{\nm{p}^2_{m,2}}{\epsilon^2}. } Hence $p \in P^{\epsilon,\gamma}_{m,n}$, as required. Finally, observe that the polynomial frame approximation $\cP^{\epsilon,\gamma}_{m,n}(f)$ of $f \in C([-1,1])$ belongs to the space $\bbP^{\epsilon,\gamma}_{m,n}$. In fact, it is the orthogonal projection onto this space with respect to the discrete inner product $\ip{\cdot}{\cdot}_{m,2}$. Therefore, by orthogonality, we may write \be{ \label{Feps-in-terms-of-SVs} \cP^{\epsilon,\gamma}_{m,n}(f) = \sum_{\sigma_i > \epsilon} \frac{\ip{f}{\xi_i}_{m,2}}{\sigma^2_i} \xi_i,\qquad f \in C([-1,1]). } \subsection{Accuracy and conditioning up to constants} We can now examine the accuracy and conditioning of $\cP^{\epsilon,\gamma}_{m,n}$. We first require the following: \lem{ \label{l:poly-frame-accuracy-stability} Let $\epsilon \geq 0$, $\gamma \geq 1$ and $\cP^{\epsilon,\gamma}_{m,n}$ the corresponding polynomial frame approximation. Then, for any $f \in C([-1,1])$, \be{ \label{acc-bound} \nmu{f - \cP^{\epsilon,\gamma}_{m,n}(f)}_{[-1,1],\infty} \leq \left ( 1 + \sqrt{m+1} C_1 \right )\nmu{f - p}_{[-1,1],\infty} + C_2 \epsilon \nmu{p}_{[-\gamma,\gamma],\infty},\quad \forall p \in \bbP_n, } where \be{ \label{C1-C2-def} \begin{split} C_1 & = C_1(m,n,\gamma,\epsilon) = \sup \{ \nmu{p}_{[-1,1],\infty} : p \in \bbP^{\epsilon,\gamma}_{m,n},\ \nm{p}_{m,\infty} \leq 1 \}, \\ C_2 & = C_2(m,n,\gamma,\epsilon) = \{ \nmu{p -\cP^{\epsilon,\gamma}_{m,n}(p)}_{[-1,1],\infty} : p \in \bbP_{n},\ \nm{p}_{[-\gamma,\gamma],\infty} \leq \epsilon^{-1} \}. \end{split} } Moreover, the condition number $\kappa(\cP^{\epsilon,\gamma}_{m,n})$ satisfies \be{ \label{stab-bound} C_1 \leq \kappa(\cP^{\epsilon,\gamma}_{m,n}) \leq \sqrt{m+1} C_1. } } These constants are interpreted as follows. The first, $C_1$, measures how large an element of the space $\bbP^{\epsilon,\gamma}_{m,n}$ can be uniformly on $[-1,1]$ relative to its discrete uniform norm over the grid. The second, $C_2$, examines the effect of the $\epsilon$-truncation. Specifically, it measures the error in reconstructing a polynomial $p \in \bbP_n$ relative to its size on the extended interval. Note, in particular, that $C_2 = 0$ when $\epsilon = 0$ and $m \geq n$. We also have $C_1(m,n,\gamma,0) = B(m,n)$ in this case, where $B(m,n)$ is as in \R{coppersmith-rivlin}. \prf{ Let $f \in C([-1,1])$ and $p \in \bbP_n$. Then \eas{ \| f - \cP^{\epsilon,\gamma}_{m,n}(f) \|_{[-1,1],\infty} & \leq \nmu{f - p}_{[-1,1],\infty} + \nmu{\cP^{\epsilon,\gamma}_{m,n}(f) - \cP^{\epsilon,\gamma}_{m,n}(p)}_{[-1,1],\infty} + \nmu{p - \cP^{\epsilon,\gamma}_{m,n}(p)}_{[-1,1],\infty} \\ & \leq \nmu{f - p}_{[-1,1],\infty} + C_1 \nmu{\cP^{\epsilon,\gamma}_{m,n}(f) - \cP^{\epsilon,\gamma}_{m,n}(p)}_{m,\infty} + C_2 \epsilon \nmu{p}_{[-\gamma,\gamma],\infty}. } Here, in the second step, we used the fact that $\cP^{\epsilon,\gamma}_{m,n}$ is linear and its range is the space $\bbP^{\epsilon,\gamma}_{m,n}$. Now recall that $\cP^{\epsilon,\gamma}_{m,n}$ is an orthogonal projection with respect to $\ip{\cdot}{\cdot}_{m,2}$. Thus, using \R{discnormbds}, \eas{ \nmu{\cP^{\epsilon,\gamma}_{m,n}(f) - \cP^{\epsilon,\gamma}_{m,n}(p)}_{m,\infty} & \leq \sqrt{(m+1)/2} \nmu{\cP^{\epsilon,\gamma}_{m,n}(f) - \cP^{\epsilon,\gamma}_{m,n}(p)}_{m,2} \\ & \leq \sqrt{(m+1)/2} \nmu{f - p}_{m,2} \\ & \leq \sqrt{m+1} \nmu{f-p}_{m,\infty} \leq \sqrt{m+1} \nmu{f-p}_{[-1,1],\infty}. } Substituting this into the previous expression now gives \R{acc-bound}. For the second result, we use the fact that $\cP^{\epsilon,\gamma}_{m,n}$ is linear once more to write \bes{ \kappa(\cP^{\epsilon,\gamma}_{m,n}) = \sup_{\substack{f \in C([-1,1]) \\ \nm{f}_{m,\infty} \neq 0}} \frac{\nmu{\cP^{\epsilon,\gamma}_{m,n}(f) }_{[-1,1],\infty} }{\nm{f}_{m,\infty}}. } Notice that $\cP^{\epsilon,\gamma}_{m,n}(p) = p$ for all $p \in \bbP^{\epsilon,\gamma}_{m,n}$. Hence \bes{ \kappa(\cP^{\epsilon,\gamma}_{m,n}) \geq \sup_{\substack{p \in \bbP^{\epsilon,\gamma}_{m,n} \\ \nm{p}_{m,\infty} \neq 0}} \frac{\nm{p}_{[-1,1],\infty} }{\nm{p}_{m,\infty}} = C_1, } which gives the lower bound in \R{stab-bound}. On the other hand, using \R{discnormbds} and the fact that $\cP^{\epsilon,\gamma}_{m,n}$ is an orthogonal projection once more, we have \eas{ \nmu{\cP^{\epsilon,\gamma}_{m,n}(f) }_{[-1,1],\infty} &\leq C_1 \nmu{\cP^{\epsilon,\gamma}_{m,n}(f) }_{m,\infty} \leq C_1 \sqrt{(m+1)/2} \nmu{\cP^{\epsilon,\gamma}_{m,n}(f) }_{m,2} \\ & \leq C_1 \sqrt{(m+1)/2} \nmu{f}_{m,2} \\ & \leq C_1 \sqrt{m+1} \nm{f}_{m,\infty}. } This gives the upper bound in \R{stab-bound}. } \subsection{Bounding the constants} The next step is to estimate the constants $C_1$ and $C_2$ appearing in this lemma. \lem{ \label{C1-C2-as-C} Consider the setup of the previous lemma. Then the constants $C_1$ and $C_2$ defined in \R{C1-C2-def} satisfy \bes{ C_1 \leq C \left(m,n,\gamma,\epsilon / (\sqrt{2} c_{n,\gamma}) \right ), } and \bes{ C_2 \leq \sqrt{\gamma(m+1)} \cdot C \left ( m , n , \gamma , \sqrt{m+1} \epsilon / (\sqrt{2} c_{n,\gamma}) \right ), } where $C$ is as in \R{C-def-general} and $c_{n,\gamma} = \frac{n+1}{\sqrt{2\gamma}}$. } \prf{ Consider $C_1$ first. Let $p \in \bbP^{\epsilon,\gamma}_{m,n}$ with $\nm{p}_{m,\infty} \leq 1$. Then we can write $p = \sum_{\sigma_i > \epsilon} c_i \xi_i$ and, using the orthogonality of the $\xi_i$, we see that \bes{ \nm{p}^2_{m,2} = \sum_{\sigma_i > \epsilon} \sigma^2_i |c_i|^2 ,\qquad \nm{p}^2_{[-\gamma,\gamma],2} = \sum_{\sigma_i > \epsilon} |c_i|^2. } Therefore \bes{ \nm{p}_{[-\gamma,\gamma],2} \leq \epsilon^{-1} \nm{p}_{m,2} \leq \sqrt{2} \epsilon^{-1} \nm{p}_{m,\infty} \leq \sqrt{2} \epsilon^{-1}. } We also recall the following inequality over $\bbP_n$: \be{ \label{poly-inf-2} \nm{q}_{[-\gamma,\gamma],\infty} \leq c_{n,\gamma} \nm{q}_{[-\gamma,\gamma],2},\quad \forall q \in \bbP_n. } This can be show directly by recalling that the classical Legendre polynomial $P_i(x)$ attains its maximum at $x = 1$ and takes value $P_i(1) = 1$. Hence, writing $q \in \bbP_n$ as $q = \sum^{n}_{i=0} c_i \psi_i$ and recalling \R{psii-def}, we get \eas{ \nm{q}_{[-\gamma,\gamma],\infty} & \leq \sum^{n}_{i=0} |c_i| \sqrt{\frac{i+1/2}{\gamma}} \leq \left ( \sum^{n}_{i=0} |c_i|^2 \right )^{1/2} \left ( \sum^{n}_{i=0} \frac{i+1/2}{\gamma} \right )^{1/2} = c_{n,\gamma} \nm{q}_{[-\gamma,\gamma],2}, } which establishes \R{poly-inf-2}. Therefore, since $p \in \bbP^{\epsilon,\gamma}_{m,n} \subseteq \bbP_n$, we get \bes{ \nm{p}_{[-\gamma,\gamma],\infty} \leq c_{n,\gamma} \nm{p}_{[-\gamma,\gamma],2} \leq \sqrt{2} c_{n,\gamma} \epsilon^{-1}. } We deduce that \eas{ C_1 &\leq \sup \{ \| p \|_{[-1,1],\infty} : p \in \bbP^{\epsilon,\gamma}_{m,n},\ \nm{p}_{m,\infty} \leq 1,\ \nm{p}_{[-\gamma,\gamma],\infty} \leq \sqrt{2} c_{n,\gamma} \epsilon^{-1} \} \\ & = C(m,n,\gamma,\epsilon / (\sqrt{2} c_{n,\gamma} ) ), } which gives the first result. We now consider $C_2$. Let $p \in \bbP_n$ with $\nm{p}_{[-\gamma,\gamma],\infty} \leq \epsilon^{-1}$. Since $p \in \bbP_n$, we may write \bes{ p = \sum^{n}_{i = 0} \frac{\ip{p}{\xi_i}_{m,2}}{\sigma^2_i} \xi_i, } and using \R{Feps-in-terms-of-SVs}, we may also write \bes{ \cP^{\epsilon,\gamma}_{m,n}(p) = \sum_{\sigma_i > \epsilon} \frac{\ip{p}{\xi_i}_{m,2}}{\sigma^2_i} \xi_i. } Therefore \bes{ p - \cP^{\epsilon,\gamma}_{m,n}(p) = \sum_{\sigma_i \leq \epsilon} \frac{\ip{p}{\xi_i}_{m,2}}{\sigma^2_i} \xi_i. } Using the fact that the $\xi_i$ are orthonormal over $[-\gamma,\gamma]$, we deduce that \be{ \label{p-minus-Feps-p-1} \nmu{p - \cP^{\epsilon,\gamma}_{m,n}(p)}^2_{[-\gamma,\gamma],2} = \sum_{\sigma_i \leq \epsilon} \frac{|\ip{p}{\xi_i}_{m,2} |^2}{\sigma^4_i} \leq \sum^{n}_{i = 0} \frac{|\ip{p}{\xi_i}_{m,2} |^2}{\sigma^4_i} = \nm{p}^2_{[-\gamma,\gamma],2}, } and, using the fact that the $\xi_i$ are orthogonal with respect to the discrete semi-inner product $\ip{\cdot}{\cdot}_{m,2}$, we see that \be{ \label{p-minus-Feps-p-2} \nmu{p - \cP^{\epsilon,\gamma}_{m,n}(p)}^2_{m,2} = \sum_{\sigma_i \leq \epsilon} \frac{|\ip{p}{\xi_i}_{m,2} |^2}{\sigma^2_i} \leq \epsilon^2 \sum_{\sigma_i \leq \epsilon} \frac{|\ip{p}{\xi_i}_m |^2}{\sigma^4_i} \leq \epsilon^2 \nm{p}^2_{[-\gamma,\gamma],2}. } Now observe that we can write \bes{ C_2 = \max \{ \nmu{q}_{[-1,1],\infty} : q \in \cA \},\qquad \cA = \left \{ q : q = p - \cP^{\epsilon,\gamma}_{m,n}(p),\ p \in \bbP_n,\ \nmu{p}_{[-\gamma,\gamma],\infty} \leq \epsilon^{-1} \right \}. } Let $q = p - \cP^{\epsilon,\gamma}_{m,n}(p) \in \cA$. Then $q \in \bbP_n$ and, due to \R{poly-inf-2} and \R{p-minus-Feps-p-1}, \bes{ \nmu{q}_{[-\gamma,\gamma],\infty} \leq c_{n,\gamma} \nmu{q}_{[-\gamma,\gamma],2} \leq c_{n,\gamma} \nmu{p}_{[-\gamma,\gamma],2} \leq \sqrt{2 \gamma} c_{n,\gamma} \nmu{p}_{[-\gamma,\gamma],\infty} \leq \sqrt{2 \gamma} c_{n,\gamma} /\epsilon. } Also, by \R{discnormbds}, \R{p-minus-Feps-p-2} and the fact that $\nm{p}_{[-\gamma,\gamma],\infty} \leq \epsilon^{-1}$, \eas{ \nmu{q}_{m,\infty}& \leq \sqrt{(m+1)/2} \nmu{q}_{m,2} \leq \sqrt{(m+1)/2} \epsilon \nm{p}_{[-\gamma,\gamma],2} \leq \sqrt{(m+1)/2} \sqrt{2 \gamma} \epsilon \nm{p}_{[-\gamma,\gamma],\infty}, } and therefore $ \nmu{q}_{m,\infty} \leq \sqrt{(m+1)/2} \sqrt{2 \gamma}$. Hence \bes{ q \in \cB : = \left \{ q \in \bbP_n : \nmu{q}_{[-\gamma,\gamma],\infty} \leq \sqrt{2 \gamma} c_{n,\gamma} / \epsilon,\ \nmu{q}_{m,\infty} \leq \sqrt{(m+1)/2} \sqrt{2 \gamma} \right \}, } which implies that $C_2 \leq \max \{ \nmu{q}_{[-1,1],\infty} : q \in \cB \}$ and, after renormalizing, \bes{ C_2 \leq \sqrt{(m+1)/2} \sqrt{2 \gamma} \max \left \{ \nmu{p}_{[-\gamma,\gamma],\infty} : p \in \bbP_{n},\ \nm{p}_{m,\infty} \leq 1,\ \nm{p}_{[-\gamma,\gamma],\infty} \leq \frac{c_{n,\gamma} }{\sqrt{(m+1)/2} \epsilon} \right \}. } This gives the second result. } \subsection{Main result on accuracy and conditioning} We now summarize these two lemmas with the following theorem: \thm{ [Accuracy and conditioning of polynomial frame approximation] \label{t:acc-cond-poly-frame} Let $\epsilon > 0$, $\gamma \geq 1$, $c > 1$ and $m , n \geq 1$ be such that \be{ \label{C-cond-main-acc-stab} C(m,n,\gamma,\epsilon ) \leq c. } Then the polynomial frame approximation $\cP^{\epsilon',\gamma}_{m,n}$ with $\epsilon' = \epsilon (n+1) / \sqrt{\gamma}$ satisfies \bes{ \kappa(\cP^{\epsilon',\gamma}_{m,n}) \leq c\sqrt{m+1}, } and, for any $f \in C([-1,1])$, \bes{ \nmu{f - \cP^{\epsilon',\gamma}_{m,n}(f)}_{[-1,1],\infty} \leq 2 c \sqrt{m+1} \inf_{p \in \bbP_n} \left \{ \nm{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty} \right \}. } } \prf{ Observe that $C(m,n,\gamma,\epsilon)$ is a decreasing function of $\epsilon$. Hence, by this, Lemma \ref{C1-C2-as-C} and the fact that $\epsilon' = \sqrt{2} c_{n,\gamma} \epsilon$, \bes{ C_1(m,n,\gamma,\epsilon') \leq C(m,n,\gamma,\epsilon' / (\sqrt{2} c_{n,\gamma}) ) = C(m,n,\gamma,\epsilon), } and \bes{ C_{2}(m,n,\gamma,\epsilon') \leq \sqrt{\gamma(m+1)}C(m,n,\gamma,\sqrt{m+1} \epsilon' / (\sqrt{2} c_{n,\gamma}) ) \leq \sqrt{\gamma(m+1)} C(m,n,\gamma,\epsilon). } Therefore, the condition \R{C-cond-main-acc-stab} implies that \bes{ C_1(m,n,\gamma,\epsilon' ) \leq c,\quad C_2(m,n,\gamma,\epsilon' ) \leq c \sqrt{\gamma(m+1)}. } We now apply Lemma \ref{l:poly-frame-accuracy-stability} to get that $\kappa(\cP^{\epsilon',\gamma}_{m,n}) \leq c \sqrt{m+1}$. For the error bound, we have \eas{ \nmu{f - \cP^{\epsilon',\gamma}_{m,n}(f)}_{[-1,1],\infty} \leq & \left ( 1 + c\sqrt{m+1} \right ) \nm{f-p}_{[-1,1],\infty} + c \sqrt{\gamma(m+1)} \epsilon' \nm{p}_{[-\gamma,\gamma],\infty} \\ \leq & \left ( 1 + c\sqrt{m+1} \right ) \left ( \nm{f-p}_{[-1,1],\infty} + (n+1)\nm{p}_{[-\gamma,\gamma],\infty} \right ), } where in the second step we used the definition of $\epsilon'$. The result now follows, since $1 + c\sqrt{m+1} \leq 2 c \sqrt{m+1}$. } \section{Maximal behaviour of polynomials bounded at equispaced nodes}\label{s:maximal-behaviour-proof} In this section, we prove Theorem \ref{t:polynomial-inequality-main}. \subsection{Pointwise Markov inequality } We first require a pointwise Markov inequality. The following lemma may be viewed as a generalization of the pointwise Bernstein inequality for algebraic polynomials $p \in \bbP_n$: \be{\lb{B} |p'(x)| \le \frac{n}{\sqrt{1-x^2}} \|p\|_{[-1,1],\infty}, \qquad |x| < 1, } to higher derivatives. It appeared in \cite{konyagin2021stable} in a slightly less general form. We provide essentially the same proof for completeness. \begin{lemma} \lb{M} For any $k,n \in \bbN$ and $\delta \in (0,1)$ such that $k < n \sqrt{(1-\delta^2)/2}$, and for any polynomial $p \in \bbP_n$, we have \be{\lb{k} |p^{(k)}(x)| \le \frac{1.251 n^k}{(1-x^2)^{k/2}} \|p\|_{[-1,1],\infty}, \qquad |x| \le \delta. } \end{lemma} Note that we cannot get \rf[k] just by iterating the Bernstein inequality \rf[B]. Such iterated use produces a much weaker result $$ |p^{(k)}(x)| \le \frac{n^k k^{k/2}}{(1-x^2)^{k/2}} \|p\|_{[-1,1],\infty}, \qquad |x| < 1. $$ \prf{ We will use the following known results. First, Shaeffer and Duffin \cite{schaeffer1938some} proved that, for any $k,n \in \bbN$ and $p \in \bbP_n$, \be{\label{sd1} |p^{(k)}(x)| \le D_{n,k}(x) \|p\|_{[-1,1],\infty}, \quad |x| < 1, } where \be{\lb{D} D_{n,k}(x) = \left|(\cos n \arccos x)^{(k)} + i (\sin n \arccos x)^{(k)}\right|. } Second, Shadrin \cite{shadrin2004twelve} derived the explicit expression for $D_{n,k}(\cdot)$. Specifically, \be{\lb{Df} \frac{1}{n^2}\big(D_{n,k}(x)\big)^2 = \sum_{m=0}^{k-1} \frac{b_{m,n}}{(1-x^2)^{k+m}}, } where \ea{ b_{m,n} &= c_{m,k} (n^2-(m+1)^2)\cdots (n^2-(k-1)^2), \lb{b} \\ c_{m,k} &:= \begin{cases} 1 & m=0, \\ {k-1+m \choose 2m} (2m-1)!!^2 & m\geq 1 \end{cases}. \lb{c} } In particular, \eas{ \frac{1}{n^2}(D_{n,1}(x))^2 & = \frac{1}{1-x^2} , \\ \frac{1}{n^2}(D_{n,2}(x))^2 & = \frac{(n^2-1)}{(1-x^2)^2} + \frac{1}{(1-x^2)^3},\\ \frac{1}{n^2} (D_{n,3}(x))^2 & = \frac{(n^2-1)(n^2-4)}{(1-x^2)^3} + \frac{3(n^2-4)}{(1-x^2)^4} + \frac{9}{(1-x^2)^5}. } Now, using \rf[Df] and \rf[b], we see that \eas{ (D_{n,k}(x))^2 &= n^2 \sum_{m=0}^{k-1} \frac{c_{m,k}}{(1-x^2)^{k+m}} (n^2-(m+1)^2)\cdots(n^2-(k-1)^2) \\ &= \frac{n^2(n^2\!-\!1^2)\cdots(n^2\!-\!(k\!-\!1)^2)} {(1-x^2)^k} \left(1 + \sum_{m=1}^{k-1} \frac{c_{m,k}}{(1-x^2)^m} \frac{1}{(n^2\!-\!1^2)\cdots(n^2\!-\!m^2)} \right) \\ &=: (A_{n,k}(x))^2 \left( 1 + B_{n,k}(x) \right). } Clearly $$ (A_{n,k}(x))^2 \le \frac{n^{2k}}{(1-x^2)^k}. $$ We now find an upper bound for the sum $B_{n,k}(x)$. Using \rf[c] we expand $c_{m,k}$ as $$ c_{m,k} = {k-1+m \choose 2m}(2m-1)!!^2 = \frac{(2m-1)!!^2}{(2m)!} \frac{(k-1+m)!}{(k-1-m)!}. $$ In the latter expression, we estimate the first factor using Wallis' inequality (see, e.g., \cite[Chpt.\ 22]{bullen2015dictionary}): $$ \frac{(2m-1)!!^2}{(2m)!} = \frac{(2m-1)!!}{(2m)!!} < \frac{1}{\sqrt{\pi m}} \le \frac{1}{\sqrt{\pi}}. $$ For the second factor, we observe that $$ \frac{(k-1+m)!}{(k-1-m)!} = \frac{k}{k+m} (k^2-1^2)\cdots(k^2-m^2) \\ < (k^2-1^2)\cdots(k^2-m^2). $$ Therefore $$ B_{n,k}(x) \le \frac{1}{\sqrt{\pi}} \sum_{m=1}^{k-1} \frac{1}{(1-x^2)^m} \frac{(k^2-1^2)\cdots(k^2-m^2)}{(n^2\!-\!1^2)\cdots(n^2\!-\!m^2)} \le \frac{1}{\sqrt{\pi}} \sum_{m=1}^{k-1} \left( \frac{1}{1-x^2} \frac{k^2}{n^2}\right)^m, $$ where in the second step we used the inequality $\frac{k^2-s^2}{n^2-s^2} < \frac{k^2}{n^2}$. Finally, if $k \le n \sqrt{(1-\delta^2)/2}$ and $|x| \le \delta$, then $\frac{1}{1-x^2} \frac{k^2}{n^2} \le \frac{1}{2}$, and $$ B_{n,k}(x) \le \frac{1}{\sqrt{\pi}} \sum_{m=1}^{k-1} \frac{1}{2^m} < \frac{1}{\sqrt{\pi}}. $$ Altogether, this gives $$ D_{n,k}(x) < \frac{c_1 n^k}{(1-x^2)^{k/2}}, \quad |x| \le \delta, \qquad c_1 = (1+ 1/\sqrt{\pi})^{1/2} < 1.251, $$ as required. } \begin{corollary} \lb{1/2} For any $k,n \in \bbN$ and $\delta \in (0,1)$ such that $k < n \sqrt{(1-\delta^2)/2}$, and for any polynomial $p \in \bbP_n$, we have \be{\lb{Markov} \|p^{(k)}\|_{[-\delta,\delta],\infty} \le \frac{1.251 n^k}{(1-\delta^2)^{k/2}} \|p\|_{[-1,1],\infty}. } \end{corollary} \subsection{A lemma on best approximation} Next, we require the following lemma on best approximation by polynomials. Although well known, we give a proof for completeness. \begin{lemma} \lb{T} Let $f \in C^r([a,b])$. Then $$ E_{r-1}(f) := \inf_{p \in \bbP_{r-1}} \|f - p\|_{[a,b],\infty} \le \frac{2}{r!} \left( \frac{b-a}{4} \right)^r \|f^{(r)}\|_{[a,b],\infty}. $$ \end{lemma} \prf{ Let $p_\Delta \in \bbP_{r-1}$ be the polynomial that interpolates $f$ at the $r$ points of the set $\Delta = (x_i)^{r}_{i=1}$. By the Lagrange interpolation formula, for any $x \in [a,b]$ we have $$ f(x) - p_\Delta(x) = \frac{1}{r!} \omega_\Delta(x) f^{(r)}(\xi), \qquad \omega_\Delta(x) := \prod_{i=1}^r (x-x_i), $$ for some $\xi = \xi_x \in [a,b]$. It follows that $$ E_{r-1}(f) \le \inf_{\substack{\Delta\subset[a,b] \\ |\Delta| = r}} \|f - p_\Delta\|_{[a,b],\infty} \le \frac{1}{r!} \inf_{\substack{\Delta\subset[a,b] \\ |\Delta| = r}} \|\omega_\Delta\|_{[a,b],\infty} \cdot \|f^{(r)}\|_{[a,b],\infty}. $$ It is well-known that the infimum of $\|\omega_\Delta\|_{[a,b],\infty}$ is attained for the set $\Delta_*$ of zeros of the Chebyshev polynomial $T_r^*$ on $[a,b]$, and that for this set we have $$ \|\omega_{\Delta_*}\|_{[a,b],\infty} = \frac{1}{2^{r-1}} \left( \frac{b-a}{2} \right)^r. $$ This completes the proof. } \subsection{Bounding the norm on a subinterval} We are now in a position to state and prove the main result of this section. Note that, for convenience, we formulate this result in terms of intervals $[-\delta,\delta]$ and $[-1,1]$ as opposed to intervals $[-1,1]$ and $[-\gamma,\gamma]$. The analogous result for the latter is obtained by setting $\delta = 1/\gamma$. \begin{theorem}\label{t:poly-inequality-general} Given $\epsilon> 0$, let $p \in \bbP_n$ satisfy $$ \|p\|_{[-1,1],\infty} \le 1/\epsilon, $$ and assume that, for some $\delta \in (0,1)$ and $m \in \bbN$, we also have $$ |p(x_i)| \le 1, \qquad x_i = - \delta + 2\delta i/m,\quad i = 0,\ldots, m. $$ If \be{\lb{m} m \ge \max \left\{ 12 c_1 n \log (1/\epsilon)\frac{\delta}{\sqrt{1-\delta^2}}, c_1 \log^2(1/\epsilon) \right\}, } for some $0 < \epsilon \leq 1/\E$ and $c_1 \ge \max \{1/\lfloor \log(1/\epsilon) \rfloor, 3/n\}$, then $$ \|p\|_{[-\delta,\delta]} \le C_0 := \beta^{3/c_1}(1+8\epsilon) +8 \epsilon, $$ where $\beta > 1$ is the upper constant in the Coppersmith and Rivlin bound \R{coppersmith-rivlin}. \end{theorem} This theorem is essentially a more general version of Theorem \ref{t:polynomial-inequality-main} in which the dependence of the bound on $C_0$ in the constant factor $c_1$ in \R{m} is made explicit. We now show how it implies Theorem \ref{t:polynomial-inequality-main}. \prf{ [Proof of Theorem \ref{t:polynomial-inequality-main}] We use Theorem \ref{t:poly-inequality-general} with $\delta = 1/\gamma$ and $c_1 = 3$. Recall that $0 < \epsilon \leq 1/\E$ and $n \geq \sqrt{\gamma^2-1} \log(1/\epsilon) > 0$ by assumption. In particular, $n \geq 1$ since it is an integer. This implies that \bes{ \max \{1/\lfloor\log(1/\epsilon) \rfloor, 3/n\} \leq \max \{ 1/\lfloor\log(\E) \rfloor, 3 \} = 3. } Hence the value $c_1 = 3$ is permitted. Observe now that, since $\delta = 1/\gamma$, \eas{ \max \left\{ 12 c_1 n\log (1/\epsilon) \frac{\delta}{\sqrt{1-\delta^2}}, c_1 \log^2(1/\epsilon) \right\} &\leq 12 c_1 n \log (1/\epsilon) \frac{\delta}{\sqrt{1-\delta^2}} \\ & = 36 n \log(1/\epsilon) \frac{1}{\sqrt{\gamma^2-1}}. } Therefore, \R{m-n-eps-poly-growth} implies that \R{m} holds. We deduce that \bes{ C(m,n,\gamma,\epsilon) \leq C_0 = \beta (1+\epsilon) +\epsilon \leq \beta \left ( 1 + 8/\E \right ) +8/\E < 4 \beta + 3, } as required. } \prf{ [Proof of Theorem \ref{t:poly-inequality-general}] For a given $p \in \bbP_n$ that satisfies assumptions of the theorem, set $$ k := \lfloor \log (1/\epsilon) \rfloor. $$ If $k > n \sqrt{(1-\delta^2)/2}$, then the condition \rf[m] implies that $$ m > c_1 n^2 \max \left \{3\sqrt{2} \delta, (1-\delta^2)/2 \right\} > c_1 n^2 /3. $$ Hence the polynomial $p$ is bounded on the interval $[-\delta, \delta]$ at $m+1 > c_1 n^2/3 + 1$ equidistant points. We deduce that \be{\lb{p_i} \|p\|_{[-\delta,\delta],\infty} \le \beta^{3/c_1}, } by the Coppersmith and Rivlin bound \R{coppersmith-rivlin}. We now suppose that $k \le n \sqrt{(1-\delta^2)/2}$, so that the Markov-type inequality \rf[Markov] is applicable. First, consider partition of the interval $[-\delta,\delta]$ with the $m+1$ equispaced points $$ x_i = - \delta + 2 \delta i/m, \qquad i = 0,\ldots,m, \qquad m \ge \max \left \{12 c_1 n k \frac{\delta}{\sqrt{1-\delta^2}}, c_1 k^2 \right\}. $$ Take any subinterval $I = [x_r, x_s] \subset [-\delta,\delta]$ containing $\lfloor c_1 k^2 \rfloor + 1$ points, so that \be{\lb{I} |I| = \frac{\lfloor c_1 k^2 \rfloor}{m} 2 \delta \le \frac{c_1 k^2 }{m} 2\delta \le \frac{k \sqrt{1-\delta^2}}{6n}, } where we used the inequality $m \ge 12 c_1 n k \frac{\delta}{\sqrt{1-\delta^2}}$. Now, with the same $k = \lfloor \log(1/\epsilon) \rfloor$, let $Q \in \bbP_k$ be the polynomial of best approximation to $p\in\bbP_n$ from $\bbP_k$ on $I$: $$ \|p - Q\|_{I,\infty} = \inf_{q \in \bbP_k} \|p -q\|_{I,\infty} =: E_k(p) \le E_{k-1}(p). $$ By Lemma \ref{T}, $$ \|p-Q\|_{I,\infty} \le \frac{2}{k!} \left(\frac{|I|}{4}\right)^k \|p^{(k)}\|_{I,\infty}, $$ and by Corollary \ref{1/2} $$ \|p^{(k)}\|_{I,\infty} \le \|p^{(k)}\|_{[-\delta,\delta],\infty} \le 1.251 n^k (1/\sqrt{1-\delta^2})^k \|p\|_{[-1,1],\infty}. $$ Hence, using the well-known estimate $k! \ge \sqrt{2\pi}\sqrt{k} (k/\E)^k$, and the bound \rf[I] for $|I|$, we obtain \eas{ \|p-Q\|_{I,\infty} &\le \frac{2.502}{\sqrt{2\pi} \sqrt{k}} \frac{\E^k}{k^k} \left( \frac{k \sqrt{1-\delta^2}}{4 \cdot 6n} \right)^k n^k (1/\sqrt{1-\delta^2})^k \|p\|_{[-1,1],\infty} \\ &\le \rho^k \|p\|_{[-1,1],\infty} , \qquad \rho = \frac{\E}{24} < \frac{1}{\E^2}. } Now, recalling that $k = \lfloor \log (1/\epsilon) \rfloor$, whereas $\|p\|_{[-1,1],\infty} \le 1/\epsilon$, we conclude that \be{\lb{pQ} \|p-Q\|_{I,\infty} \le \E^{-2 \log(1/\epsilon) + 2} 1/\epsilon \le \E^2 \epsilon < 8 \epsilon. } On the other hand, by construction the interval $I = [x_r, x_s]$ contains $\lfloor c_1 k^2 \rfloor +1$ equispaced points from the $(x_i)$. By \R{coppersmith-rivlin}, for $Q \in \bbP_k$, we have \be{\lb{Qi} \|Q\|_{I,\infty} \le C_2 \max_{x_i \in I} |Q(x_i)|, \qquad C_2 = \beta^{c_3}, \qquad c_3 = \frac{k^2}{\lfloor c_1 k^2 \rfloor} \le \frac{k^2}{c_1 k^2/2} \le \frac{2}{c_1}. } Here, for the Copppersmith-Rivlin bound, the assumption $c_1k^2 \ge k$, i.e., boundedness of $Q \in \bbP_k$ on at least $k+1$ points, is required, and similarly we required $c_1 n^2/3 \ge n$ in obtaining \rf[p_i]. This is where the theorem's condition $$ c_1 \ge \max\{1/k,3/n\} = \max\{1/ \lfloor\log(1/\epsilon) \rfloor, 3/n\} $$ came from. Also, since $c_1 k^2 \ge k \ge 1$, we have the inequality $\lfloor c_1 k^2 \rfloor \ge c_1 k^2/2$ which we used in \rf[Qi] We now conclude the proof. Using \rf[pQ] and \rf[Qi], we obtain \eas{ \|p\|_{I,\infty} \le \|Q\|_{I,\infty} + 8 \epsilon \le C_2 \max_{x_i \in I} |Q(x_i)| + 8 \epsilon \le C_2 (\max_{x_i \in I} |p(x_i)| + 8 \epsilon) + 8 \epsilon &\le C_2 (1+8\epsilon) + 8\epsilon, } as required. } \section{Proofs of the possibility and impossibility theorems}\label{s:main-thm-proof} We now prove Theorems \ref{t:possibility-thm}--\ref{t:impossibility-extended}. For these, we first require the following result. \lem{ \label{poly-approx-bounds} Suppose that $E_{\theta} \subset \bbC$ is the Bernstein ellipse \R{Bernstein-ellipse} with parameter $\theta > 1$ and $f \in B(E_{\theta})$. Then there exists a polynomial $p \in \bbP_n$ for which \bes{ \nmu{f - p}_{[-1,1],\infty} \leq \frac{2}{\theta - 1} \nm{f}_{E_{\theta},\infty} \theta^{-n}. } Moreover, this polynomial satisfies \bes{ \nm{p}_{E_{\tau},\infty} \leq \frac{2}{1-\tau/\theta} \nm{f}_{E_{\theta},\infty},\qquad 1 < \tau < \theta } and \bes{ \nm{p}_{E_{\tau},\infty} \leq \left ( \frac{\tau}{\theta} \right )^n \frac{2\tau/\theta}{\tau/\theta-1} \nm{f}_{E_{\theta},\infty},\qquad \tau > \theta. } } \prf{ This result is essentially standard. We repeat it to obtain the explicit bounds for $p$. Since $f \in B(E_{\theta})$ its Chebyshev expansion converges uniformly on $[-1,1]$, i.e. \bes{ f(x) = \sum^{\infty}_{n=0} c_k \phi_k(x),\quad \phi_k(x) = \cos(n \arccos(x)), } and its coefficients satisfy $|c_k| \leq 2 \nm{f}_{E_{\theta}} \theta^{-k}$ \cite[Thm.\ 8.1]{trefethen2013approximation}. Let $p = \sum^{n}_{k=0} c_k \phi_k$. Then \bes{ \nmu{f-p}_{[-1,1],\infty} \leq 2 \nm{f}_{E_{\theta}} \sum_{k > n} \theta^{-k} . } Evaluating the sum gives the first result. For the other results, recall that the Bernstein ellipse is given by $E_{\tau} = \left \{ J(z) : z \in \bbC,\ 1 \leq |z | \leq \tau \right \}$, where $J(z) = \frac12(z+z^{-1})$ is the Joukowsky map, and the Chebyshev polynomials satisfy $\psi_k(J(z)) = \frac12 \left ( z^n + z^{-n} \right )$. Hence $\nm{\psi_k }_{E_{\tau},\infty} \leq \tau^{n}$. Therefore \bes{ \nm{p}_{E_{\tau},\infty} \leq 2\nm{f}_{E_{\theta}} \sum^{n}_{k=0} (\tau / \theta)^k, } which yields the result. } \prf{ [Proof of Theorem \ref{t:possibility-thm}] We use Theorem \ref{t:acc-cond-poly-frame}. Note that the condition \R{m-n-relation} implies that \R{m-n-eps-poly-growth} holds. Therefore Theorem \ref{t:polynomial-inequality-main} implies that \R{C-cond-main-acc-stab} holds with $c$ as defined therein. The desired bound for the condition number follows immediately. For the error bound, let $f \in B(E_{\theta})$, where $\theta > \gamma + \sqrt{\gamma^2-1}$, and let $p \in \bbP_n$ be as in the previous lemma. Since $[-\gamma,\gamma] \subset E_{\tau}$ where $\tau = \gamma + \sqrt{\gamma^2-1} < \theta$ by assumption, this lemma gives \eas{ \nmu{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty} & \leq \left ( \frac{2}{\theta-1} \theta^{-n} + \frac{2(n+1)\epsilon}{1-(\gamma + \sqrt{\gamma^2-1}) / \theta} \right ) \nm{f}_{E_{\theta},\infty} \\ & \leq g(\theta,\gamma) \left ( \theta^{-n} + n \epsilon \right ) \nm{f}_{E_{\theta},\infty}, } for some function $g(\theta,\gamma)$ depending on $\theta$ and $\gamma$ only. Here we also use the fact that $n \geq 1$, which follows from the fact that $n$ is an integer and $n \geq \log(1/\epsilon) > 0$ by assumption. Supposing that \R{C-cond-main-acc-stab} holds, Theorem \ref{t:acc-cond-poly-frame} now gives, up to a constant change in $g(\theta,\gamma)$, \bes{ \nmu{f - \cP^{\epsilon',\gamma}_{m,n}(f) }_{[-1,1],\infty} \leq c \sqrt{m} g(\theta,\gamma) \left ( \theta^{-n} + n \epsilon \right ) \nm{f}_{E_{\theta},\infty}, } which is the desired error bound with respect to $n$. To obtain the error bound with respect to $m$, we simply notice that \R{m-n-relation} implies that $m \leq 36 n \log(1/\epsilon) / \sqrt{\gamma^2-1} + 1 = n/c^*+1$. Hence $\theta^{-n} \leq \theta^{c^*(1-m)} = \rho^{1-m}$, as required. } \prf{ [Proof of Theorem \ref{t:possibility-slower-exp}] The overall argument is similar to the previous proof. The only difference is the estimation of the term \be{ \label{best-approx-term} \nmu{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty}. } Suppose that $f$ is analytic in $E = E_{\theta}$ for some $\theta$ with $1 < \theta < \tau : = \gamma + \sqrt{\gamma^2-1}$. In other words, the Bernstein ellipse $E_{\theta}$ does not contain the extended interval $[-\gamma,\gamma]$. Let $1 \leq k \leq n$ and $p \in \bbP_k$ be the polynomial guaranteed by Lemma \ref{poly-approx-bounds}. Then \eas{ \nmu{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty} & \leq \left ( \frac{2}{\theta-1} \theta^{-k} + \left ( \frac{\tau}{\theta} \right )^k \frac{2 \tau / \theta}{\tau/\theta-1} (n+1)\epsilon \right )\nm{f}_{E_{\theta},\infty} \\ & \leq g(\theta,\gamma) \left ( \theta^{-k} + \left ( \frac{\tau}{\theta} \right )^k n \epsilon \right ) \nm{f}_{E_{\theta},\infty}. } We now choose \bes{ k = \min \left \{ n , \left \lfloor \frac{\log(1/\epsilon)}{\log(\tau)} \right \rfloor \right \}, } so that \bes{ \left ( \frac{\tau}{\theta} \right )^k \epsilon \leq \epsilon^{1-\frac{\log(\tau/\theta) }{ \log(\tau) }} = \epsilon^{\frac{\log(\theta)}{\log(\tau)} }, } and \bes{ \theta^{-k} \leq \theta^{-n} + \theta^{1-\frac{\log(1/\epsilon)}{\log(\tau)}} = \theta^{-n} + \theta \epsilon^{\frac{\log(\theta)}{\log(\tau)}}. } Since $n \geq 1$, we deduce that \bes{ \nmu{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty} \leq 2 g(\theta,\gamma) \left ( \theta^{-n} + n \epsilon^{\frac{\log(\theta)}{\log(\tau)}} \right ) \nm{f}_{E_{\theta},\infty}, } as required. } \prf{ [Proof of Theorem \ref{t:possibility-algebraic}] As in the previous proof, we only need to obtain the desired estimate for the term \R{best-approx-term}. Let $f \in C^{k}([-1,1])$. Then $f$ has a $C^k$-extension to the interval $[-\gamma,\gamma]$. Specifically, there is a function $\tilde{f} \in C^{k}([-\gamma,\gamma])$ satisfying $\tilde{f}(x) = f(x)$ for all $x \in [-1,1]$ and \be{ \label{bounded-extension} \nmu{\tilde{f}}_{C^{k}([-\gamma,\gamma])} \leq c_{k,\gamma} \nm{f}_{C^{k}([-1,1])}, } for some constant $c_{k,\gamma} \geq 1$ depending on $k$ and $\gamma$ only. Since $\tilde{f} \in C^{k}([-\gamma,\gamma])$ a classical result (see, e.g., \cite[\S 4.6]{cheney1982introduction}) gives that \bes{ \inf_{p \in \bbP_n} \nmu{\tilde{f} - p}_{[-\gamma,\gamma],\infty} \leq c'_{k,\gamma} n^{-k} \nmu{\tilde{f}^{(k)}}_{[-\gamma,\gamma],\infty}, } for some $c'_{k,\gamma} > 0$. Observe that \bes{ \nmu{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty} \leq 2 \nmu{\tilde{f} - p}_{[-\gamma,\gamma],\infty} + (n+1) \epsilon \nmu{\tilde{f}}_{[-\gamma,\gamma],\infty}. } Therefore \eas{ \inf_{p \in \bbP_n} \left \{ \nmu{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty} \right \} & \leq 2 c'_{k,\gamma} n^{-k} \nmu{\tilde{f}^{(k)}}_{[-\gamma,\gamma],\infty} + (n+1) \epsilon \nmu{\tilde{f}}_{[-\gamma,\gamma],\infty} \\ & \leq \left ( 2 c'_{k,\gamma} n^{-k} + (n+1) \epsilon \right ) \nmu{\tilde{f}}_{C^k([-\gamma,\gamma])} \\ & \leq 2 \max \{c'_{k,\gamma} ,1 \} c_{k,\gamma} \left ( n^{-k} + n \epsilon \right ) \nmu{f}_{C^k([-1,1])}, } where in the last step we used \R{bounded-extension} and the fact that $n \geq 1$. This gives the desired result. } We conclude this section with the proof of Theorem \ref{t:impossibility-extended}. \prf{ [Proof of Theorem \ref{t:impossibility-extended}] The proof is similar to that of the impossibility theorem shown in \cite{platte2011impossibility}. First observe that $\cR_m(0) = 0$ for any approximation procedure satisfying \R{extended-error-bound}. Now let $k \in \bbN_0$. Then the condition number \R{kappa-def} satisfies \bes{ \kappa(\cR_m) \geq \lim_{\delta \rightarrow 0^+} \sup_{\substack{q \in \bbP_k \\ 0 < \nm{q}_{m,\infty} \leq \delta}} \frac{\nm{\cR_m(q)}_{[-1,1],\infty}}{\nm{q}_{m,\infty}}. } Let $p \in \bbP_k$ with $\nm{p}_{m,\infty} \neq 0$ and set $q = \delta p / \nm{p}_{m,\infty}$ so that $q \in \bbP_k$ with $\nm{q}_{m,\infty} = \delta$. Since polynomials are entire functions, \R{extended-error-bound} gives that \eas{ \nmu{\cR_m(q)}_{[-1,1],\infty} & \geq \nm{q}_{[-1,1],\infty} - C \left ( \theta^{-c m^{\tau}} + \epsilon \right ) \nm{q}_{E_{\theta},\infty} \geq \nm{q}_{[-1,1],\infty} \left ( 1 - C \left ( \theta^{-c m^{\tau}} + \epsilon \right ) \theta^k \right ). } Here, in the second step, we used the classical inequality $\nm{q}_{E_{\theta},\infty} \leq \theta^k \nm{q}_{[-1,1],\infty}$, $\forall q \in \bbP_k$. Recalling the definition of $q$, we deduce that \bes{ \kappa(\cR_m) \geq \left ( 1 - C \left ( \theta^{-c m^{\tau}} + \epsilon \right ) \theta^k \right ) \sup_{\substack{p \in \bbP_k \\ \nm{p}_{m,\infty} \neq 0 }} \frac{\nm{p}_{[-1,1],\infty}}{\nm{p}_{m,\infty}}. } Now let \bes{ k = \left \lfloor \min \left \{ \frac{\log(\frac{1}{4C \epsilon})}{\log(\theta)} , c m^{\tau} + \frac{\log(\frac{1}{4C})}{\log(\theta)} \right \} \right \rfloor. } This choice of $k$ gives $C \left ( \theta^{-c m^{\tau}} + \epsilon \right ) \theta^k \leq 1/2$, and therefore \bes{ \kappa(\cR_m) \geq \frac12 \alpha^{k^2/m}, } where $\alpha > 1$ is as in \R{coppersmith-rivlin}. Observe that this holds for all $1 < \theta \leq \theta^*$. Now choose $\theta = \theta_m = \epsilon^{-\frac{1}{c m^{\tau}}}$, and observe that $\theta_m \leq \theta^*$ for all large $m$. This value of $\theta$ and the fact that $0 < \epsilon \leq (4 C)^{-2}$ give \bes{ k = \left \lfloor c m^{\tau} \left ( 1 + \frac{\log(\frac{1}{4C})}{\log(1/\epsilon)} \right )\right \rfloor \geq \left \lfloor \frac12 c m^{\tau} \right \rfloor , } and therefore $k \geq \frac13 c m^{\tau}$ for all sufficiently large $m$. We deduce that \bes{ \kappa(\cR_m) \geq \frac12 a^{c^2 m^{2 \tau - 1}/9} \geq \sigma^{m^{2 \tau-1}}, } for some $\sigma > 1$ and all large $m$, as required. } \section{Numerical examples}\label{s:numerical} We conclude this paper with a series of experiments to examine the various theoretical results. Unless otherwise stated, we compute the discrete $L^{\infty}$-norm error of the approximation on a grid of 50,000 equispaced points in $[-1,1]$. Also, we consider the polynomial frame approximation threshold parameter $\epsilon$ rather than $\epsilon'$ (as used in the main theorems). Theoretically, this choice leads to a log-linear sample complexity, but in practice it appears to be adequate. In Fig.\ \ref{f:fig1} we plot the error versus $n$ for different values of the \textit{oversampling} parameter $\eta = m/n$. We compare several different values for the extended domain parameter $\gamma$, and several different values of $\epsilon$. Notice that in all cases, we witness exponential decrease of the error down to some fixed limiting accuracy. The limiting accuracy is related to the stability of the approximation and the parameter $\epsilon$. Observe that it gets smaller with increasing $\eta$, and for sufficiently large $\eta$ it closely tracks the value of $\epsilon$ used. Moreover, it is larger when $\gamma$ is smaller and smaller when $\gamma$ is larger. Both observations are intuitively true. Increasing the number of sample points (i.e.\ larger $\eta$) reduces the maximal growth of a polynomial on $[-1,1]$ relative to its values of the equispaced grid. Similarly, increasing $\gamma$ lengthens the region within which the polynomial cannot exceed the value $1/\epsilon$, and therefore it also cannot grow as large on $[-1,1]$. We also notice that increasing the oversampling parameter makes less difference to the limiting accuracy when $\epsilon$ is larger than it does when $\epsilon$ is smaller. Again, this is intuitively true, since larger $\epsilon$ means the polynomial cannot grow as large on the extended interval. These three observations are also supported by Theorem \ref{t:polynomial-inequality-main}. Here, the sufficient scaling between $m$ and $n$ depends on $\log(1/\epsilon)/\sqrt{\gamma^2-1}$, i.e., it is a decreasing function of both $\epsilon$ and $\gamma$. \begin{figure}[t] \begin{small} \begin{center} \begin{tabular}{ccc} \includegraphics[width = 0.3\textwidth]{fig1_11.eps}& \includegraphics[width = 0.3\textwidth]{fig1_12.eps}& \includegraphics[width = 0.3\textwidth]{fig1_13.eps} \\ \includegraphics[width = 0.3\textwidth]{fig1_21.eps}& \includegraphics[width = 0.3\textwidth]{fig1_22.eps}& \includegraphics[width = 0.3\textwidth]{fig1_23.eps} \\ \includegraphics[width = 0.3\textwidth]{fig1_31.eps}& \includegraphics[width = 0.3\textwidth]{fig1_32.eps}& \includegraphics[width = 0.3\textwidth]{fig1_33.eps} \\ $\gamma = 1.2$ & $\gamma = 1.4$ & $\gamma = 1.8$ \end{tabular} \end{center} \end{small} \caption{Approximation error versus $n$ for approximating the function $f(x) = \frac{1}{1+x^2}$ via $\cP^{\epsilon,\gamma}_{m,n}$, where $m/n = \eta$, using various different values of $\eta$, $\gamma$ and $\epsilon$. The values of $\epsilon$ used are $\epsilon = 10^{-14}$ (top), $\epsilon = 10^{-10}$ (middle) and $\epsilon = 10^{-6}$ (bottom). The dashed line shows the quantity $\theta^{-n}$, where $\theta = \sqrt{2}+1$. } \label{f:fig1} \end{figure} In Fig.\ \ref{f:fig1} the function we consider has poles at $x = \pm \I$, meaning that it is analytic within any Bernstein ellipse $E_{\theta}$ for which $\frac12(\theta - \theta^{-1}) = 1$, i.e.\ $1 < \theta < \sqrt{2}+1 \approx 2.41$. Recall that the interval $[-\gamma,\gamma]$ is contained in the Bernstein ellipse $E_{\tau}$ with $\tau = \gamma + \sqrt{\gamma^2-1}$. In particular, $\tau < \sqrt{2}+1$ for $\gamma = 1.2$ and $\gamma = 1.4$. Our analysis in Theorem \ref{t:possibility-thm} therefore asserts exponential decay of the error with rate roughly $(\sqrt{2}+1)^{-n}$ for these two choices of $\gamma$. This is what we observe in practice in Fig.\ \ref{f:fig1}. On the other hand, when $\gamma = 1.8$, $\tau = \gamma + \sqrt{\gamma^2-1} \approx 3.30 > 2.41 = \sqrt{2}+1$. In this case, our analysis in Theorem \ref{t:possibility-slower-exp} predicts exponential convergence with rate roughly $(\sqrt{2}+1)^{-n}$ down to roughly $\epsilon^{\frac{\log(\sqrt{2}+1)}{\log(\tau)}} \approx \epsilon^{0.74}$, and slower convergence below this level. This is again in agreement with the right column of Fig.\ \ref{f:fig1}. To further investigate this effect of decreasing error decay for less regular functions, in Fig.\ \ref{f:fig2} we plot the error versus $n$ for several different functions and values of $\gamma$. We also plot the theoretical breakpoint described in Theorem \ref{t:possibility-slower-exp}, i.e.\ the value \be{ \label{breakpoint} \epsilon^{\frac{\log(\theta)}{\log(\tau)}},\qquad \tau = \gamma + \sqrt{\gamma^2-1}. } In all cases, we see reasonable agreement between the theoretical results and the empirical performance. First, the error decays with rate $\theta^{-n}$ down to the breakpoint, as in Theorem \ref{t:possibility-slower-exp}, before decreasing more slowly beyond it. This second phase is described by Theorem \ref{t:possibility-algebraic} (since analytic functions are infinitely differentiable). \begin{figure}[t] \begin{small} \begin{center} \begin{tabular}{ccc} \includegraphics[width = 0.3\textwidth]{fig2_11.eps} & \includegraphics[width = 0.3\textwidth]{fig2_12.eps} & \includegraphics[width = 0.3\textwidth]{fig2_13.eps} \\ \includegraphics[width = 0.3\textwidth]{fig2_21.eps} & \includegraphics[width = 0.3\textwidth]{fig2_22.eps} & \includegraphics[width = 0.3\textwidth]{fig2_23.eps} \\ \includegraphics[width = 0.3\textwidth]{fig2_31.eps} & \includegraphics[width = 0.3\textwidth]{fig2_32.eps} & \includegraphics[width = 0.3\textwidth]{fig2_33.eps} \\ $\gamma = 1.25$ & $\gamma = 1.5$ & $\gamma = 2$ \end{tabular} \end{center} \end{small} \caption{Approximation errors versus $n$ for approximating the functions $f_1(x) = \frac{1}{1+4x^2}$ (top), $f_2(x) = \frac{1}{10-9 x}$ (middle) and $f_3(x) = 25 \sqrt{9 x^2-10}$ (bottom) via $\cP^{\epsilon,\gamma}_{m,n}$, where $m/n = 4$, using various different values of $\gamma$ and $\epsilon$. The dot-dashed lines show the breakpoints \R{breakpoint} in each case and dashed line shows the quantity $\theta^{-n}$. In this experiment, the values of $\theta$ are $\theta = \frac12(1+\sqrt{5})$ (top), $\theta = \frac{1}{9}(10 + \sqrt{19})$ (middle) and $\theta = \sqrt{10/9} + 1/3$ (bottom).} \label{f:fig2} \end{figure} It is notable that the error decay rate after the breakpoint is quite different for the functions considered. This effect has also been observed and discussed in the case of Fourier extensions \cite{adcock2014parameter,adcock2014resolution}. It can be understood through Theorem \ref{t:possibility-algebraic}. Recall that this theorem asserts that the error is bounded by \be{ \label{limiting-accuracy-analysis} c g(k,\gamma) \sqrt{m} \left ( n^{-k} + n \epsilon \right ) \nm{f}_{C^k([-1,1])}, } for any $k \in \bbN$ (since all functions considered are infinitely smooth). The derivatives of the first function $f_1(x) = \frac{1}{1+4x^2}$ do not grow too large on $[-1,1]$ with increasing $k$. Hence the constants $ \nm{f}_{C^k([-1,1])}$ in the error term remain reasonably small and we see rapid decrease in $n$. On the other hand, the derivatves of the functions $f_2$ and $f_3$ grow rapidly with $k$, meaning the constants $ \nm{f}_{C^k([-1,1])}$ also grow rapidly with $k$. Thus, \R{limiting-accuracy-analysis} suggests that the error decays progressively more slowly the closer it gets to the limiting value $\epsilon$. This is exactly the effect we observe in practice. \begin{figure}[t] \begin{small} \begin{center} \begin{tabular}{ccc} \includegraphics[width = 0.3\textwidth]{fig3_11.eps} & \includegraphics[width = 0.3\textwidth]{fig3_12.eps} & \includegraphics[width = 0.3\textwidth]{fig3_13.eps} \\ \includegraphics[width = 0.3\textwidth]{fig3_21.eps} & \includegraphics[width = 0.3\textwidth]{fig3_22.eps} & \includegraphics[width = 0.3\textwidth]{fig3_23.eps} \\ \includegraphics[width = 0.3\textwidth]{fig3_31.eps} & \includegraphics[width = 0.3\textwidth]{fig3_32.eps} & \includegraphics[width = 0.3\textwidth]{fig3_33.eps} \\ $\gamma = 1.25$ & $\gamma = 1.5$ & $\gamma = 2$ \end{tabular} \end{center} \end{small} \caption{Approximation errors versus $n$ for approximating the functions $f(x) = \exp(\I \omega \pi x)$via $\cP^{\epsilon,\gamma}_{m,n}$, where $m/n = 4$, using various different values of $\gamma$ and $\epsilon$. The values of $\omega$ are $\omega = 40$ (top), $\omega = 60$ (middle) and $\omega = 80$ (top).} \label{f:fig3} \end{figure} In Fig.\ \ref{f:fig3} we consider approximating the oscillatory function $f(x) = \E^{\I \omega \pi x}$ for various different values of $\omega$. Oscillatory functions are an interesting case study for approximation procedures from equispaced nodes. They are entire functions, yet they grow extremely rapidly along the imaginary axis for large $|\omega|$, meaning that the term $\nm{f}_{E_{\theta},\infty}$ is extremely large unless $\theta \approx 1$. As we see in Fig.\ \ref{f:fig3}, the approximation error is order one until a minimum value of $n = n_0(\omega)$ is met. After this point, the function begins to be resolved and the error decreases rapidly. Determining the behaviour of $n_0(\omega)$ allows us to examine the \textit{resolution power} of the approximation scheme, i.e.\ the number of points needed before decay of the error sets in. For the polynomial frame approximation, we determine this point by recalling the error bound \bes{ \inf_{p \in \bbP_n} \left \{ \nmu{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty} \right \}. } Since $f$ is entire and $|f(x)| = 1$ for real $x$, we can write \bes{ \inf_{p \in \bbP_n} \left \{ \nmu{f - p}_{[-1,1],\infty} + (n+1)\epsilon \nm{p}_{[-\gamma,\gamma],\infty} \right \} \leq 2 \inf_{p \in \bbP_n} \nmu{f - p}_{[-\gamma,\gamma],\infty} + (n+1)\epsilon. } Hence, the resolution power for polynomial approximation is related to the resolution power of best polynomial approximation on the extended interval $[-\gamma,\gamma]$. It is well known (see, e.g., \cite{hale2008new,boyd1989chebyshev,gottlieb1977numerical,adcock2014resolution}) that on the interval $[-1,1]$ an oscillatory function with frequency $\omega$ can be approximated by a polynomial once the degree $n$ exceeds the value $\pi \omega$. Since we consider the extended interval, this implies that $n_0(\omega) = \pi \gamma \omega$. This value is also shown in Fig.\ \ref{f:fig3}. It closely predicts the point at which the error begins to decrease. Note that this suggests choosing a small value of $\gamma$ so as to obtain a higher resolution power. Yet, as seen, this worsens the condition number of the approximation. Or, to put it another way, a smaller $\gamma$ necessitates a more severe oversampling ratio $\eta$ so as to maintain the same condition number, thus worsening the resolution power with respect to the number of equispaced samples $m$. { Finally, we conclude in Fig.\ \ref{f:fig4} with an experiment that compares fixed $\epsilon$ (the main setting in this paper) with varying $\epsilon$, the latter chosen as in Remark \ref{rem:varying-eps} to decay like $\theta^{-n}$ for a given $\theta > 1$. In order to make a fair comparison, we define a maximum allowable condition number $\kappa^* = 100$. Then, given $n$, we find the largest value of $m$ for each scheme such that its condition number is at most $\kappa^*$. In order to do this, we use follow the approach of \cite[\S 8]{adcock2020approximating} and work in the discrete $L^2$-norm (as opposed to the discrete uniform norm considered in previous experiments) over a grid of 50,000 equispaced points in $[-1,1]$. Doing so means that the condition number can be computed as the norm of a certain matrix. For the first function, $f_1$, we observe that varying $\epsilon$ can lead to a benefit whenever the parameter $\theta$ is chosen suitably: specifically, whenever it is chosen close to the value $\theta = \theta^*$, where $\theta^*$ is the largest Bernstein ellipse within which the function is analytic. If $\theta$ is chosen too large, then the scheme behaves similarly to the standard polynomial frame approximation with fixed $\epsilon$. On the other hand, if $\theta$ is chosen too small, then the scheme performs significantly worse. In fact, it can even perform worse that standard polynomial least-squares approximation using orthonormal Legendre polynomials on $[-1,1]$ (see Remark \ref{r:why-not-unit-interval}). For the second function, $f_2$, varying $\epsilon$ conveys no benefit over fixed $\epsilon$, and, as before, it can lead to worse performance if $\theta$ is chosen too small. It is notable that polynomial least-squares approximation outperforms any of the polynomial frame approximations for this function. This is not surprising. The function grows rapidly near the endpoint $x = + 1$. In polynomial frame approximation the approximating polynomial is constrained to be of a finite size on $[-\gamma,\gamma]$. Hence it cannot capture this rapid growth as effectively as in polynomial least-squares approximation, where there is no such constraint. Finally, the third function, $f_3$, is oscillatory and therefore entire. Observe that when $\theta$ is small, the polynomial frame approximation with varying $\epsilon$ initially resolves the function using fewer samples than the scheme with fixed $\epsilon$. Yet, as $m$ increases the error decays more slowly, and is eventually larger than the error for the latter method. Finally, in the second row of Fig.\ \ref{f:fig4} we plot the scaling between $m$ and $n$. As expected, polynomial least-squares approximation exhibits a quadratic scaling, while polynomial frame approximation with fixed $\epsilon$ exhibits a linear scaling. Polynomial frame approximation with varying $\epsilon$ exhibits a quadratic scaling for small $\theta$. When $\theta$ is larger the scaling is at first quadratic and then linear. This arises because $\epsilon$ is constrained to be no smaller than $10^{-14}$ in this experiments, this being done in order to avoid numerical effects in the thresholded SVD. Note that in this regime, the two polynomial frame approximations coincide. The main conclusion from this experiment is that varying $\epsilon$ with $n$ can lead to some benefit (for small $m$), as long as the parameter $\theta$ is chosen carefully. Unfortunately, such a choice is function dependent (compare, for instance, $f_1$ versus $f_3$), and may require knowledge of the domain of analyticity of the (unknown) function. } \begin{figure}[t] \begin{small} \begin{center} \begin{tabular}{ccc} \includegraphics[width = 0.3\textwidth]{fig4_f1_err.eps} & \includegraphics[width = 0.3\textwidth]{fig4_f2_err.eps} & \includegraphics[width = 0.3\textwidth]{fig4_f3_err.eps} \\ \includegraphics[width = 0.3\textwidth]{fig4_f1_mvals.eps} & \includegraphics[width = 0.3\textwidth]{fig4_f2_mvals.eps} & \includegraphics[width = 0.3\textwidth]{fig4_f3_mvals.eps} \\ $f_1(x) = \frac{1}{1+16 x^2}$, $\theta^* = \frac{\sqrt{17}+1}{4}$ & $f_2(x) = \frac{1}{30-29 x}$, $\theta^* = \frac{30 + \sqrt{59}}{29}$ & $f_3(x) = \E^{40 \pi \I x}$, $\theta^* = \infty$ \end{tabular} \end{center} \end{small} \caption{{Comparison of various schemes for approximating the functions $f_1$, $f_2$ and $f_3$. The top row shows the approximation error (measured in the discrete $L^2$-norm) versus $m$. The bottom row shows the relationship between $m$ and $n$ where, given $n$, $m$ is smallest integer such that the (discrete $L^2$-norm) condition number of each scheme is at most $100$. The schemes considered are: polynomial least-squares approximation (``PLS"), which uses orthonormal Legendre polynomials on $[-1,1]$ (see Remark \ref{r:why-not-unit-interval}); polynomial frame approximation with fixed $\epsilon = 10^{-14} $ (``PFF"); and polynomial frame approximation with varying $\epsilon = \max \{ \theta^{-n}, 10^{-14} \}$ (``PFV"). Both PFF and PFV use the value $\gamma = 1.25$. The quantity $\theta^*$ is the `optimal' value of $\theta$, in the sense that it is the parameter of the largest Bernstein ellipse within which the given function is analytic. } } \label{f:fig4} \end{figure} \section{Conclusions and outlook}\label{s:conclusions} In this work, we have shown a positive counterpart to the impossibility theorem of \cite{platte2011impossibility}. Namely, we have shown a possibility theorem (Theorem \ref{t:possibility-thm}) for polynomial frame approximation, which asserts stability and exponential convergence down to a finite, but user-controlled limiting accuracy. This holds for all functions that are analytic in a sufficiently large region -- a condition that must be the case for any scheme achieving this type of error decay, as shown by our extended impossibility theorem (Theorem \ref{t:impossibility-extended}). On the other hand, for insufficiently analytic functions, we have shown exponential decay down to some fractional power of $\epsilon$ (Theorem \ref{t:possibility-slower-exp}), and superalgebraic decay (Theorem \ref{t:possibility-algebraic}) beyond this point. There are several avenues for further investigation. First, recall that our main error bounds involve (albeit small) algebraic factors in $m$ and $n$. It would be interesting to see if such factors could be removed by modifying the approximation scheme. Alternatively, one might instead work in the $L^2$-norm, which is arguably more natural for least-squares approximations. Second, this work was inspired by so-called Fourier extensions \cite{adcock2014parameter,adcock2014numerical,matthysen2017function,lyon2012fast,huybrechs2010fourier,bruno2007accurate,boyd2002fourier,pasquetti1996spectral}, wherein a smooth, nonperiodic function on $[-1,1]$ is approximated by a Fourier series on $[-\gamma,\gamma]$. In practice, linear oversampling appears sufficient for accuracy and stability of $\epsilon$-regularized Fourier extensions, with exponential error decay down to $\epsilon$ \cite{adcock2014parameter,adcock2014numerical}. Proving a similar possibility theorem for this scheme is an open problem. Note that Fourier extension is equivalent to a polynomial approximation problem on an arc of the complex unit circle \cite{geronimo2020fourier,webb2020pointwise}. Fourier extension schemes have several advantages over the polynomial extension scheme studied herein. For example, they generally possess higher resolution power \cite{adcock2014resolution} (recall Fig.\ \ref{f:fig3} and the discussion in \S \ref{s:numerical}). Third, we mention that equispaced points are not special. Similar impossibility theorems have been shown for scattered data, or more generally, any sample points that do not cluster quadratically near the endpoints $x = \pm 1$ \cite{adcock2019optimal}. Beyond pointwise samples, it is notable that an impossibility theorem also holds for reconstructing analytic functions from their Fourier samples \cite{adcock2012stable,adcock2014generalized,adcock2014stability}. The question of whether or not possibility theorems hold for these more general types of samples is also an open problem. Fourth, notice that we have not concerned ourselves with fast computation of the polynomial frame approximation in this work. We anticipate, however, that a fast implementation may be possible, as it is with Fourier extensions \cite{matthysen2015fast,lyon2012fast,matthysen2017function}. One potential idea in this direction is the AZ algorithm \cite{coppe2020AZ}. Fifth and finally, we note that the one-dimensional problem is, in some senses, a toy problem. Polynomial frame approximations we first formalized in \cite{adcock2020approximating} to accurately and stably approximate functions that are defined over general, compact domains in two or more dimensions. Here the domain is embedded in a hypercube, and a tensor-product orthogonal polynomial basis on the hypercube is used to construct the approximation. Such approximations are often used in practice in surrogate model construction problems in uncertainty quantification \cite{adcock2022sparse,adcock2020approximating}. Showing that linear oversampling is sufficient in two or more dimensions and a corresponding possibility theorem for analytic function approximation in arbitrary dimensions would be an interesting and practically-relevant extension. \small \bibliographystyle{plain}
{ "timestamp": "2022-03-08T02:07:01", "yymm": "2110", "arxiv_id": "2110.03755", "language": "en", "url": "https://arxiv.org/abs/2110.03755", "abstract": "We consider approximating analytic functions on the interval $[-1,1]$ from their values at a set of $m+1$ equispaced nodes. A result of Platte, Trefethen \\& Kuijlaars states that fast and stable approximation from equispaced samples is generally impossible. In particular, any method that converges exponentially fast must also be exponentially ill-conditioned. We prove a positive counterpart to this `impossibility' theorem. Our `possibility' theorem shows that there is a well-conditioned method that provides exponential decay of the error down to a finite, but user-controlled tolerance $\\epsilon > 0$, which in practice can be chosen close to machine epsilon. The method is known as \\textit{polynomial frame} approximation or \\textit{polynomial extensions}. It uses algebraic polynomials of degree $n$ on an extended interval $[-\\gamma,\\gamma]$, $\\gamma > 1$, to construct an approximation on $[-1,1]$ via a SVD-regularized least-squares fit. A key step in the proof of our main theorem is a new result on the maximal behaviour of a polynomial of degree $n$ on $[-1,1]$ that is simultaneously bounded by one at a set of $m+1$ equispaced nodes in $[-1,1]$ and $1/\\epsilon$ on the extended interval $[-\\gamma,\\gamma]$. We show that linear oversampling, i.e., $m = c n \\log(1/\\epsilon) / \\sqrt{\\gamma^2-1}$, is sufficient for uniform boundedness of any such polynomial on $[-1,1]$. This result aside, we also prove an extended impossibility theorem, which shows that such a possibility theorem (and consequently the method of polynomial frame approximation) is essentially optimal.", "subjects": "Numerical Analysis (math.NA)", "title": "Fast and stable approximation of analytic functions from equispaced samples via polynomial frames", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.992304351911275, "lm_q2_score": 0.8080672066194946, "lm_q1q2_score": 0.8018486057653119 }
https://arxiv.org/abs/2111.11227
On a discriminator for the polynomial $f(x)=x^3+x$
Let $\Delta(n)$ denote the smallest positive integer $m$ such that $a^3+a(1\le a\le n)$ are pairwise distinct modulo $m$. The purpose of this paper is to determine $\Delta(n)$ for all positive integers $n$.
\section{Introduction} \setcounter{lemma}{0} \setcounter{theorem}{0} \setcounter{corollary}{0} \setcounter{equation}{0} \medskip For a polynomial $f(x)\in \Z[x]$ with all $f(a)(a\in \Z^{+})$ pairwise distinct, we introduce the discriminator $\Delta_f(n)$ defined to be the smallest positive integer $m$ such that $f(a)(1\le a\le n)$ are pairwise distinct modulo $m$. As a simple application of Bertrand's postulate, Arnold, Benkoski and McCabe \cite{ABM} determined $\Delta_{f}(n)$ for $f(x)=x^2$, and they showed that for $n>4$, $\Delta_{f}(n)$ is the smallest positive integer $m\ge 2n$ such that $m$ is $p$ or $2p$ with $p$ an odd prime. Sun \cite{S13} studied $\Delta_f(n)$ for other quadratic polynomials. For example, it was proved in \cite{S13} that if $f(x)=2x(x-1)$ then $\Delta_{f}(n)$ is the least prime number greater than $2n-2$, and in particular $\Delta_{f}(n)$ runs over all prime values. Among other things, Schumer \cite{S} studied $\Delta_f(n)$ with $f(x)=x^3$. For the study of discriminator $\Delta_f(n)$ with other higher degree polynomials $f$, one may refer to \cite{BSW,Moree,MM,Zieve}. In this paper, we focus on $\Delta_f(n)$ with $f(x)=x^3+x$. The main result in this paper is the following. \begin{theorem}\label{theorem1}Let $\Delta(n)=\Delta_f(n)$ with $f(x)=x^3+x$. We have \begin{align*}\Delta(n)=\begin{cases}7\cdot 3^{6s+4} & \ \textrm{ if } n=3^{6s+5}+1 \textrm{ or } n=3^{6s+5}+2 \textrm{ for some } s\in \N, \\ 3^{\lceil \log_3 n\rceil} & \ \textrm{ otherwise},\end{cases}\end{align*} where $\lceil x \rceil$ denotes the smallest integer no less than $x$. \end{theorem} A closely related problem is to determine $D(n)$, which denotes the smallest positive integer $m$ such that $a^3+a(1\le a\le n)$ are pairwise distinct modulo $m^2$. The authors \cite{YZ} proved that $D(n)=3^{\lceil \log_3 \sqrt{n}\rceil}$, which was conjectured by Z.-W. Sun (see Conjecture 6.76 in \cite{S2021}). The present work is motivated by the above original conjecture of Sun. Different from $D(n)$, the discriminator $\Delta(n)$ is not always a power of three. For example, $\Delta(n)=7\cdot 3^{4}$ when $n=244$ or $245$. This was first observed by Sun (see the remark to Conjecture 6.76 in \cite{S2021}). According to Theorem \ref{theorem1}, the third example of $n$ satisfying $\Delta(n)\not=3^{\lceil \log_3 n\rceil}$ is over $10^5$. We prove Theorem \ref{theorem1} by combining methods from elementary number theory and analytic number theory. We point out that in order to deal with $\Delta(n)$ we have to study an incomplete character sum, which is not involved in the work \cite{YZ}. The incomplete character sum will be handed by the elementary method when the length of the summation is about $\frac{p}{4}$, and it will be handed by the analytic method when the length of the summation is about $\frac{p}{6}$. The details will be given in Section 3. Moreover, in the very special case $p=7$, we have to discuss the value of Legendre symbol separately (see Lemma \ref{lemma51} in Section 5). \bigskip We use the following notations in this paper. Let $\Z^{+}$ denote the set of all positive integers and let $\N=\Z^{+}\cup\{0\}$. We use $e(\alpha)$ to denote $e^{2\pi i\alpha}$. The notation $\lceil x \rceil$ denotes the smallest integer no less than $x$, and $\lfloor x\rfloor$ denotes the greatest integer no more than $x$. \bigskip \section{Preparations} \setcounter{lemma}{0} \setcounter{theorem}{0} \setcounter{corollary}{0} \setcounter{equation}{0} We introduce \begin{align*}\mathcal{E}=\{3^{6s+5}+1:\ s\in \N\} \cup \{ 3^{6s+5}+2:\ s\in \N\}.\end{align*} Throughout this paper, we use the letter $k$ to denote \begin{align*}k=\lceil \log_3 n\rceil.\end{align*} We first point out that $a^3+a(1\le a\le n)$ are pairwise distinct modulo $3^{k}$, and therefore $n\le \Delta(n)\le 3^k$. In order to establish Theorem \ref{theorem1}, it suffices to prove the following two results. \begin{lemma}\label{lemma21}Let $n\not\in \mathcal{E}$. Suppose that \begin{align}\label{inequality1} n\le m<3^k<3n.\end{align} Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{lemma}\label{lemma22}Let $n=3^{6s+5}+1$ or $3^{6s+5}+2$ with $s\in \N$. Suppose that \begin{align}\label{inequality2} n\le m<7\cdot 3^{6s+4}.\end{align} Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$. Moreover, $a^3+a(1\le a\le n)$ are pairwise distinct modulo $7\cdot 3^{6s+4}$. \end{lemma} We shall consider the following $8$ cases. (i) $m=\delta p$, where $\delta\ge 6$, $p\ge 5$ is a prime, $p\not=7$ and $p\nmid \delta$. (ii) $m=\delta p^r$, where $\delta\ge 4$, $p\ge 5$ is a prime, $r\ge 2$ is a positive integer. (iii) $m=2^r$, where $r\in \Z^{+}$. (iv) $m=2^rt$, where $t\ge 5$ is an odd number and $r\ge 2$. (v) $m=2^r3^s$, where $r,s\in \Z^{+}$. (vi) $m=3^{r}\cdot 14$, where $r\in \N$. (vii) $m=\delta p^{t}$, where $1\le \delta \le 3$, $p\ge 5$ is a prime, $t\in \Z^{+}$. (viii) $m=3^{r}\cdot 7$, where $r\in \N$. \noindent The letter $\delta$ always denotes a positive integer. Note that \eqref{inequality2} implies \eqref{inequality1}. Throughout this paper, we assume that \eqref{inequality1} holds. \bigskip \section{An incomplete character sum} \setcounter{equation}{0} \medskip For $u\in \Z$, $\delta\in \Z^{+}$ and $p\ge 3$, we introduce \begin{align*}A_p(\delta,u)=\sum_{-\frac{p-1}{2}\le x\le \frac{p-1}{2}}\Big(\frac{\delta^2x^2+4}{p}\Big)e\big(\frac{ux}{p}\big),\end{align*} where $\big(\frac{\cdot}{p}\big)$ denotes the Legendre symbol. \begin{lemma}\label{lemma31}Suppose that $p\ge 3$ is a prime and $p\nmid \delta$. (i) If $p|u$, then $A_p(\delta,u)=-1$. (ii) If $p\nmid u$, then $|A_p(\delta,u)|\le 2\sqrt{p}$.\end{lemma} \begin{proof}For an odd prime $p$, it is well-known that $$\sum_{1\le c\le p-1}\Big(\frac{c}{p}\Big)e(\frac{c}{p})=\sum_{1\le x\le p}e\big(\frac{x^2}{p}\big),$$ and $|\tau_p|=\sqrt{p}$, where $\tau_p$ denotes the above Gauss sum. By $$\sum_{1\le c\le p-1}\Big(\frac{c}{p}\Big)e(\frac{c(\delta^2x^2+4)}{p})=\Big(\frac{\delta^2x^2+4}{p}\Big)\tau_p,$$ we deduce that \begin{align*}A_p(\delta,u)=&\frac{1}{\tau_p}\sum_{-\frac{p-1}{2}\le x\le \frac{p-1}{2}}e(\frac{ux}{p})\sum_{1\le c\le p-1}\Big(\frac{c}{p}\Big)e(\frac{c(\delta^2x^2+4)}{p}) \\= & \frac{1}{\tau_p}\sum_{1\le c\le p-1}e(\frac{4c}{p})\Big(\frac{c}{p}\Big)\sum_{-\frac{p-1}{2}\le x\le \frac{p-1}{2}}e(\frac{c\delta^2x^2+ux)}{p}.\end{align*} Note that \begin{align*}\sum_{-\frac{p-1}{2}\le x\le \frac{p-1}{2}}e(\frac{c\delta^2x^2+ux)}{p}=\Big(\frac{c}{p}\Big)e(\frac{-\overline{4\delta^2c}\, u^2}{p})\tau_p,\end{align*} where $\overline{d}$ means $\overline{d}\cdot d\equiv 1\pmod{p}$. Now we conclude that \begin{align}\label{resultAp}A_p(\delta,u)=\sum_{1\le c\le p-1}e(\frac{-\overline{4\delta^2c}\, u^2+4c}{p}).\end{align} If $p|u$, then the summation in \eqref{resultAp} is a Ramanujan sum and $A_p(\delta,u)=1$. If $p\nmid u$, then by Weil's bound on Kloosterman sums (see (4.19) in \cite{I}) we have $|A_p(\delta,u)|\le 2\sqrt{p}$. This completes the proof. \end{proof} We remark that Lemma \ref{lemma31} (i) is a well-known result. For a prime $p\ge 5$ and $p\nmid \delta$, we define $\ell_p(\delta)$ to be smallest positive integer $x$ such that $$\Big(\frac{-3\delta^2x^2-12}{p}\Big)\in\{0,1\}.$$ We introduce \begin{align*}L_p=\begin{cases}\frac{p+3}{4}\ \ \textrm{ if } p\equiv 1\pmod{12}, \\ \frac{p-1}{4}\ \ \textrm{ if } p\equiv 5\pmod{12}, \\ \frac{p+5}{4}\ \ \textrm{ if } p\equiv 7\pmod{12}, \\ \frac{p+1}{4}\ \ \textrm{ if } p\equiv 11\pmod{12}.\end{cases}\end{align*} We point out that $L_p<\frac{p}{3}$ holds for $p\ge 5$ except $p=7$. \begin{lemma}\label{lemma32}Suppose that $p\ge 5$ is a prime and $p\nmid \delta$. We have $$\ell_p(\delta)\le L_p.$$\end{lemma} \begin{proof}By Lemma \ref{lemma31} (i), we have \begin{align*}A_p(\delta,0)=2\sum_{1\le x\le \frac{p-1}{2}}\Big(\frac{\delta^2x^2+4}{p}\Big)+1=-1,\end{align*} and therefore, \begin{align}\label{halfsum}\sum_{1\le x\le \frac{p-1}{2}}\Big(\frac{\delta^2x^2+4}{p}\Big)=-1.\end{align} We introduce \begin{align*}N_p^{+}=&|\{1\le x\le \frac{p-1}{2}:\ \big(\frac{\delta^2x^2+4}{p}\big)=+1\}|, \\ N_p^{-}=&|\{1\le x\le \frac{p-1}{2}:\ \big(\frac{\delta^2x^2+4}{p}\big)=-1\}|, \\ N_p^{0}=&|\{1\le x\le \frac{p-1}{2}:\ \big(\frac{\delta^2x^2+4}{p}\big)=0\}|.\end{align*} In view of \eqref{halfsum}, we have the following conclusions. If $p\equiv 1\pmod{4}$, then $N_p^{0}=1$, $N_p^{+}=\frac{p-5}{4}$ and $N_p^{-}=\frac{p-1}{4}$. If $p\equiv 3\pmod{4}$, then $N_p^{0}=0$, $N_p^{+}=\frac{p-3}{4}$ and $N_p^{-}=\frac{p+1}{4}$. Case $p\equiv 1\pmod{12}$. We have $\big(\frac{-3}{p}\big)=1$ and $\ell_p(\delta)$ is the smallest positive integer $x$ such that $\big(\frac{\delta^2x^2+4}{p}\big)\in\{0,1\}$. Note that $N_p^{0}+N_p^{+}=\frac{p-1}{4}$. Now we conclude that $\ell_p(\delta)\le \frac{p-1}{2}-(N_p^{0}+N_p^{+})+1=\frac{p-1}{2}-\frac{p-1}{4}+1=L_p$. Case $p\equiv 5\pmod{12}$. We have $\big(\frac{-3}{p}\big)=-1$ and $\ell_p(\delta)$ is the smallest positive integer $x$ such that $\big(\frac{\delta^2x^2+4}{p}\big)\in\{0,-1\}$. Note that $N_p^{0}+N_p^{-}=\frac{p+3}{4}$. Now we conclude that $\ell_p(\delta)\le \frac{p-1}{2}-(N_p^{0}+N_p^{-})+1=\frac{p-1}{2}-\frac{p+3}{4}+1=L_p$. Case $p\equiv 7\pmod{12}$. We have $\big(\frac{-3}{p}\big)=1$ and $\ell_p(\delta)$ is the smallest positive integer $x$ such that $\big(\frac{\delta^2x^2+4}{p}\big)=1$. Note that $N_p^{+}=\frac{p-3}{4}$. Now we conclude that $\ell_p(\delta)\le \frac{p-1}{2}-N_p^{+}+1=\frac{p-1}{2}-\frac{p-3}{4}+1=L_p$. Case $p\equiv 11\pmod{12}$. We have $\big(\frac{-3}{p}\big)=-1$ and $\ell_p(\delta)$ is the smallest positive integer $x$ such that $\big(\frac{\delta^2x^2+4}{p}\big)=-1$. Note that $N_p^{-}=\frac{p+1}{4}$. Now we conclude that $\ell_p(\delta)\le \frac{p-1}{2}-N_p^{-}+1= \frac{p-1}{2}-\frac{p+1}{4}+1=L_p$. We are done. \end{proof} \begin{lemma}\label{lemma33}Suppose that $m=\delta p^{r}$, where $\delta, r\in \Z^{+}$, $p\ge 5$ is a prime and $p\nmid \delta$. (i) If $p^{r}+\delta\frac{p-1}{2}\le n$, then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$. (ii) If $r=1$ and $p+\delta \ell_p(\delta)\le n$, then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$. \end{lemma} \begin{proof}We consider $1\le a\le p^{r}$ and $b=a+\delta c$ with $c\in \Z^{+}$. It suffices to find $a,c\in \Z^{+}$ such that $a+\delta c\le n$ and $$a^2+a(a+\delta c)+(a+\delta c)^2+1\equiv 0\pmod{p^{r}},$$ which is equivalent to \begin{align}\label{congruencequad}(6a+3\delta c)^2\equiv -3\delta^2c^2-12\pmod{p^{r}}.\end{align} In view of \eqref{halfsum}, we conclude that there exists $1\le c\le \frac{p-1}{2}$ such that $-3\delta^4c^2-12$ is a quadratic residue modulo $p$. Then it is easy to deduce that there exists $1\le a\le p^{r}$ such that $(6a+3\delta c)^2\equiv -3\delta^2c^2-12\pmod{p^{r}}$. This completes the proof the conclusion (i). By the definition of $\ell_p(\delta)$, we can find $1\le c\le \ell_p(\delta)$ such that $\big(\frac{-3\delta^4c^2-12}{p}\big)\in \{0,1\}$. Then we can find $1\le a\le p$ such that $(6a+3\delta c)^2\equiv -3\delta^2c^2-12\pmod{p}$. This proves the conclusion (ii). We are done. \end{proof} \begin{lemma}\label{lemma34}Suppose that $m=\delta p$, where $\delta \ge 39$, $p\ge 5$ is a prime, $p\not=7$ and $p\nmid \delta$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}By Lemma \ref{lemma32} and Lemma \ref{lemma33} (ii), we only need to verify $p+\delta L_p\le n$. By \eqref{inequality1}, $n> \frac{\delta p}{3}$ and it suffices to prove $p+\delta L_p\le \frac{\delta p}{3}$. Indeed we can prove $\frac{p}{39}+L_p\le \frac{p}{3}$ for all $p\not=7$. This completes the proof. \end{proof} Similarly, we have the following. \begin{lemma}\label{lemma35}Suppose that $m=\delta p$, where $\delta \ge 13$, $p\ge 165$ is a prime and $p\nmid \delta$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}It suffices to prove $p+\delta L_p\le \frac{\delta p}{3}$. Indeed we can prove $\frac{p}{13}+L_p\le \frac{p}{3}$ for all $p\ge 165$. This completes the proof. \end{proof} \begin{lemma}\label{lemma36}If $p\ge 4000$ and $p\nmid \delta$, then we have $$\ell_p(\delta)\le \frac{p}{6}.$$\end{lemma} \begin{proof}We write \begin{align*}Y=\lfloor\frac{p-1}{6}\rfloor.\end{align*} It suffices to prove \begin{align}\label{sumY}\Big|\sum_{1\le x\le Y}\Big(\frac{\delta^2x^2+4}{p}\Big)\Big|<Y-1,\end{align} since \eqref{sumY} implies that $\Big(\frac{\delta^2x^2+4}{p}\Big)$ can take both $1$ and $-1$ in the range $1\le x\le Y$. We define \begin{align*}A=\sum_{-Y\le x\le Y}\Big(\frac{\delta^2x^2+4}{p}\Big).\end{align*} Note that \eqref{sumY} is equivalent to $|A-1|<2Y-2$, which follows from $|A|<2Y-3$. For $c,x\in \Z$, we have \begin{align*}\frac{1}{p}\sum_{u=1}^{p}e\big(\frac{u(c-x)}{p}\big)=\begin{cases} 1, \ & \textrm{ if } c\equiv x\pmod{p}, \\ 0, \ & \textrm{ if } c\not\equiv x\pmod{p},\end{cases}\end{align*} and therefore, \begin{align*}A=\frac{1}{p}\sum_{1\le u\le p}\sum_{1\le c\le p}\Big(\frac{\delta^2c^2+4}{p}\Big)e(\frac{uc}{p})\sum_{-Y\le x\le Y}e(-\frac{ux}{p}).\end{align*} By Lemma \ref{lemma31}, we obtain \begin{align*}|A|\le \frac{2\sqrt{p}}{p}\sum_{1\le u\le p}\Big|\sum_{-Y\le x\le Y}e(-\frac{ux}{p})\Big|,\end{align*} and by Lemma 4.8 in \cite{YZ} we further have \begin{align*}|A|\le 2\sqrt{p}(2+\ln p).\end{align*} The inequality $|A|<2Y-3$ follows from \begin{align*} 2\sqrt{p}(2+\ln p)<2Y-3.\end{align*} Note that \begin{align*}2Y-3>\frac{p}{3}-5.\end{align*} Now we need to prove \begin{align}\label{check}2\sqrt{p}(2+\ln p)<\frac{p}{3}-5.\end{align} It is easy to prove that \eqref{check} holds for $p\ge 4000$. This completes the proof. \end{proof} \begin{lemma}\label{lemma37}Suppose that $m=\delta p$, where $\delta \ge 6$, $p\ge 4000$ is a prime and $p\nmid \delta$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}By Lemma \ref{lemma33} (ii) and Lemma \ref{lemma36}, it suffices to prove $p+\frac{\delta p}{6}\le \frac{\delta p}{3}$, which holds for $\delta\ge 6$. This completes the proof. \end{proof} \begin{lemma}[Case (i)]\label{lemma38}Let $n\ge 48000$. Suppose that $m=\delta p$, where $\delta \ge 6$, $p\ge 5$ is a prime, $p\not=7$ and $p\nmid \delta$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}In view of Lemma \ref{lemma34}, we only need to consider $\delta<39$. By \eqref{inequality1}, $m\ge 48000$. We deduce that $p=\frac{m}{\delta}\ge \frac{n}{\delta}>165$. By Lemma \ref{lemma35}, we only need to consider $\delta\le 12$. Now we further have $p=\frac{m}{\delta}\ge \frac{n}{\delta}\ge 4000$ and the desired conclusion follows from Lemma \ref{lemma37}. This completes the proof. \end{proof} \begin{remark}\label{remark}One can verify Theorem \ref{theorem1} for $n\le 48000$ with the help of a computer. In fact, Z.-W. Sun has verified the truth of Theorem \ref{theorem1} for $n\le 10^5$. Therefore, the condition $n\ge 48000$ in Lemma \ref{lemma38} can be removed.\end{remark} \bigskip \section{The Cases (ii)-(vii)} \setcounter{lemma}{0} \setcounter{theorem}{0} \setcounter{corollary}{0} \setcounter{equation}{0} \medskip The purpose of this section is to deal with cases (ii)-(vii). \begin{lemma}[Case (ii)]\label{lemma41}Suppose that $m=\delta p^{r}$, where $\delta \ge 4$, $p\ge 5$ is a prime, $r\ge 2$ is a positive integer and $p\nmid \delta$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof} By Lemma \ref{lemma33}, it is sufficient to prove $p^{r}+\delta\frac{p-1}{2}\le n$. By \eqref{inequality1}, $n> \frac{\delta p^{r}}{3}$ and it suffices to prove $p^{r}+\frac{1}{2}\delta p\le \frac{\delta p^{r}}{3}$. This follows from \begin{align}\label{ineqin41}(p^{r-1}-\frac{3}{2})(\delta-3)\ge \frac{9}{2}.\end{align} Since $p^{r-1}-\frac{3}{2}\ge p-\frac{3}{2}\ge \frac{7}{2}$, \eqref{ineqin41} holds if $\delta\ge 5$. In the case $\delta=4$, \eqref{ineqin41} holds if $p^{r-1}\ge 6$. We now only need to consider $\delta=4$, $p=5$, $r=2$, and it is easy to verify that $p^{r}+\delta\frac{p-1}{2}\le \frac{\delta p^{r}}{3}\le n$ holds. This completes the proof. \end{proof} \begin{lemma}[Case (iii)]\label{lemma42}Suppose that $m=2^r$, where $r\in \Z^{+}$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}Note that $2^3+2-1^3-1=2^3$ and $5^3+5-1^3-1=2^7$. For $r\le 3$, we can choose $a=1$ and $b=2$. For $4\le r\le 7$, we can choose $a=1$ and $b=5$. Now we assume that $r\ge 8$. The proof is the same as that of Lemma 3.7 in \cite{YZ}, and thus we explain it briefly. Since $(a+4)^3+(a+4)-a^3-a=4(3(a+2)^2+5)$, it suffices to find $1\le a\le 2^{r-2}-3$ such that $3(a+2)^2+5\equiv 0\pmod{2^{r-2}}$. For $r\ge 8$, we can find $3\le x\le 2^{r-2}-1$ such that $3x^2+5\equiv 0\pmod{2^{r-2}}$. On choosing $a=x-2$, we obtain $3(a+2)^2+5\equiv 0\pmod{2^{r-2}}$. Note that $b=4+a=x+2\le 2^{r-2}+1\le n$. We are done. \end{proof} \begin{lemma}[Case (iv)]\label{lemma43}Suppose that $m=2^{r}t$, where $r\ge 2$ is an integer, $t\ge 5$ is an odd number. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}We consider $1\le a\le 2^{r}$ and $b=a+t$. Note that $2^{r}+t\le \frac{2^rt}{3}$ is equivalent to $(2^r-3)(t-3)\ge 9$, which holds expect that $r=2$ and $t\le 11$. In view of Remark \ref{remark}, we may assume that $n>44$, and by \eqref{inequality1} we have $b\le 2^{r}+t\le n$. It suffices to find $1\le a\le 2^{r}$ such that $$a^2+a(a+t)+(a+t)^2+1\equiv 0\pmod{2^{r}},$$ and the proof of (3.1) in \cite{YZ} also implies the above conclusion. We are done. \end{proof} \begin{lemma}[Case (v)]\label{lemma44}Suppose that $m=2^r3^s$, where $r,s\in \Z^{+}$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}By Lemma \ref{lemma43}, we only need to consider either $r=1$ or $s=1$. We first consider $s=1$. Note that $a^2+a(a+3)+(a+3)^2+1$ is equal to $40=2^3\cdot 5$ when $a=2$. If $r\le 3$, then the desired conclusion follows by choosing $a=2$ and $b=5$. Next we assume $r\ge 4$. Similarly to the proof of (3.1) in \cite{YZ}, we can obtain that for any $j\ge 3$, there exists $1\le a\le 2^{j}-6$ such that \begin{align*}a^2+a(a+3)+(a+3)^2+1\equiv 0\pmod{2^{j}}.\end{align*} In particular, there exists $1\le a\le 2^{r}-6$ such that $a^2+a(a+3)+(a+3)^2+1\equiv 0\pmod{2^{r}}$. The desired conclusion follows by choosing $b=a+3$ and noting that $b\le 2^{r}-3<n$. Now we consider $r=1$. By \eqref{inequality1}, $n>3^s$. We can choose $a=1$ and $b=1+3^s$. The proof is complete. \end{proof} \begin{lemma}[Case (vi)]\label{lemma45}Suppose that $m=3^{r}\cdot 14$, where $r\in \N$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}When $m=14$, it suffices to choose $a=1$ and $b=3$. Now we assume that $r\ge 1$. By Lemma \ref{lemma32} (noting that $L_7=3$) and Lemma \ref{lemma33} (ii) (with $p=7$ and $\delta=2\cdot 3^{r}$), we only need to verify $7+3^{r+1}\cdot 2\le n$. By \eqref{inequality1}, $r=k-3$ and $n>3^{k-1}=3^{r+2}$. Note that $7+3^{r+1}\cdot 2<3^{r+2}$ if $r\ge 1$. This completes the proof. \end{proof} The last task in this section is to consider Case (vii). The proof is as same as that in Section 4 \cite{YZ}. We introduce \begin{align}X:=X_p=\lfloor\frac{p}{3}\rfloor p^{t-1}.\end{align} We aim to find $1\le a\not=b\le \frac{n}{\delta}$ such that $\delta^2(a^2+ab+b^2)+1\equiv 0\pmod{p^{t}}$. Then on choosing $a'=\delta a$, $b'=\delta b$, we obtain $a'^3+a'\equiv b'^3+b'\pmod{m}$. By \eqref{inequality1}, we have $X<\frac{n}{\delta}$. Let \begin{align}\label{definef}f(a,b)=\delta^2(a^2+ab+b^2).\end{align}Now we introduce \begin{align}\label{defineN}\mathcal{N}=\sum_{\substack{1\le a,b\le X \\ f(a,b)+1\equiv 0\pmod{p^{t}}}}1\end{align} and \begin{align*}\mathcal{N}^{\not=}=\sum_{\substack{1\le a\not=b\le X \\ f(a,b)+1\equiv 0\pmod{p^{t}}}}1.\end{align*} Note that $\mathcal{N}^{\not=}\ge \mathcal{N}-2$. The main objective is to prove $\mathcal{N}>2$ (and thus $\mathcal{N}^{\not=}>0$. For $j\ge 1$, we define \begin{align}\label{defineTj}T_j=\sum_{\substack{1\le c\le p^j \\ (c,p)=1}}\sum_{\substack{1\le a,b\le X }}e\big(\frac{cf(a,b)+c}{p^j}\big).\end{align} \begin{lemma}\label{lemma46}Let $\mathcal{N}$ and $T_j$ be given in \eqref{defineN} and \eqref{defineTj} respectively. We have \begin{align*}\mathcal{N}=\frac{X^2}{p^{t}}+\frac{1}{p^{t}}\sum_{j=1}^{t}T_j.\end{align*} If $1\le j\le t-1$, then \begin{align*}T_j=X^2p^{-j}\Big(\frac{-3}{p^j}\Big)\mu(p^j),\end{align*} where $\mu(\cdot)$ is the M\"obius function. Moreover, we also have \begin{align*}|T_t|\le 2p^{\frac{3t}{2}}(2+\ln p^t)^2.\end{align*} \end{lemma} \begin{proof}The three conclusions are corresponding to Lemma 4.2, Lemma 4.4 and Lemma 4.9 in \cite{YZ} respectively. Although we only considered the case $t=2r$ in \cite{YZ}, both the proofs and the conclusions of Lemmas 4.2, 4.4 and 4.7 in \cite{YZ} are valid for all $t\in \Z^{+}$.\end{proof} \begin{lemma}\label{lemma47}If $t\ge 2$, then \begin{align}\label{finalineq1}\mathcal{N}\ge \frac{X^2}{p^{t}}-\Big(\frac{-3}{p}\Big)\frac{X^2}{p^{t+1}}-2p^{t/2}(2+\ln p^{t})^2. \end{align} If $t=1$, then \begin{align}\label{finalineq2}\mathcal{N}\ge \frac{X^2}{p^{t}}-2p^{t/2}(2+\ln p^{t})^2. \end{align} \end{lemma} \begin{proof}The desired conclusions follow from Lemma \ref{lemma46}.\end{proof} \begin{lemma}\label{lemma48}Suppose that $m=\delta p^{t}$, where $1\le \delta \le 3$, $p\ge 5$ is a prime, $t\in \Z^{+}$. Suppose further that $p^t\ge 20000^2$. Then we have \begin{align*}\mathcal{N}^{\not=}>0.\end{align*} \end{lemma} \begin{proof} For $p\ge 7$, we have $\lfloor\frac{p}{3}\rfloor\ge \frac{3}{11}p$ (the equality holds with $p=11$) and $1-\frac{1}{p}\ge \frac{6}{7}$. Thus for $p\ge 7$, we have \begin{align}\label{check1}\lfloor\frac{p}{3}\rfloor^2 p^{-2}\Big(1-\big(\frac{-3}{p}\big)\frac{1}{p}\Big)\ge \lfloor\frac{p}{3}\rfloor^2 p^{-2}(1-\frac{1}{p})\ge \frac{3^2\cdot 6}{11^2 \cdot 7}>\frac{6}{125}.\end{align} For $p=5$, we have \begin{align}\label{check2}\lfloor\frac{p}{3}\rfloor^2 p^{-2}\Big(1-\big(\frac{-3}{p}\big)\frac{1}{p}\Big)=\frac{6}{125}.\end{align} We deduce from \eqref{finalineq1}, \eqref{finalineq2}, \eqref{check1} and \eqref{check2} that (for all $p\ge 5$) \begin{align*}\mathcal{N}\ge \frac{6}{125}p^{t}-2p^{t/2}(2+\ln p^{t})^2. \end{align*} Since $\mathcal{N}^{\not=}\ge \mathcal{N}-2$, we need to prove $$\frac{6}{125}p^{t}>2p^{t/2}(2+\ln p^t)^2+2,$$ which follows from \begin{align}\label{check4}\sqrt{p^t}>\frac{125}{3}(2+\ln p^t)^2+30.\end{align} On writing $q=\sqrt{p^t}$, our task is to prove $q>\frac{500}{3}(1+\ln q)^2+30$. Let $g(x)=\sqrt{x-30}-\sqrt{\frac{500}{3}}(1+\ln x)$. Then $g'(x)=\frac{1}{2\sqrt{x-30}}-\sqrt{\frac{500}{3}}\cdot\frac{1}{x}>\frac{1}{2\sqrt{x}}-\sqrt{\frac{500}{3}}\cdot\frac{1}{x}$ for $x>30$ and $g$ is increasing when $x> \frac{2000}{3}$. Note that $g(20000)>0$. Therefore, $q>\frac{500}{3}(1+\ln q)^2+30$ holds for $q\ge 20000$ and \eqref{check4} holds due to $p^t\ge 20000^2$. The proof is complete. \end{proof} In view of \eqref{inequality1}, for $m=\delta p^t$ (with $1\le \delta \le 3$ and $p\ge 5$ a prime) we have $$3^{k-1}<n\le \delta p^t<3^{k},$$ and we define \begin{align}\label{Nstar}\mathcal{N}^\ast=\sum_{\substack{1\le a<b\le 1+3^{k-1} \\ a^3+a\equiv b^3+b\pmod{\delta p^{t}}}}1.\end{align} We verify $\mathcal{N}^\ast>0$ for $p^t<20000^2$ with the help of a computer. \begin{lemma}\label{lemma49}Let $\mathcal{N}^\ast$ be given in \eqref{Nstar}. Suppose that $m=\delta p^{t}$, where $1\le \delta \le 3$, $p\ge 5$ is a prime, $t\in \Z^{+}$. Suppose further that $p^t<20000^2$. Then we have \begin{align*}\mathcal{N}^\ast>0.\end{align*} \end{lemma} \begin{proof}This is checked by C++.\end{proof} \begin{lemma}[Case (vii)]\label{lemma410}Suppose that $m=\delta p^{t}$, where $1\le \delta \le 3$, $p\ge 5$ is a prime, $t\in \Z^{+}$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}The desired conclusion follows from Lemma \ref{lemma48} and Lemma \ref{lemma49}. \end{proof} \bigskip \section{The Case (viii)} \setcounter{lemma}{0} \setcounter{theorem}{0} \setcounter{corollary}{0} \setcounter{equation}{0} \medskip It is in Case (viii) that we need to distinguish $n\in \mathcal{E}$ or not in the proof. For $m=3^r\cdot 7$, by \eqref{inequality1} we have $r=k-2$ and $$n>3^{r+1}.$$ \begin{lemma}\label{lemma51}(i) If $r\equiv 0\pmod{3}$ or $r\equiv 2\pmod{3}$, then we have $\ell_7(3^r)\le 2$. (ii) If $r\equiv 1\pmod{3}$, then we have $\ell_7(3^r)=3$. \end{lemma} \begin{proof}Since $\big(\frac{-3}{7}\big)=1$, $\ell_7(\delta)$ is the smallest positive integer $x$ such that $\big(\frac{\delta^2x^2+4}{7}\big)=1$. If $r\equiv 0\pmod{3}$, then $3^{2r}x^2+4\equiv x^2+4\equiv 1\pmod{7}$ for $x=2$ and thus $\ell_7(3^r)\le 2$ (indeed $\ell_7(3^r)=2$ in this case). If $r\equiv 2\pmod{3}$, then $3^{2r}x^2+4\equiv 4x^2+4\equiv 1\pmod{7}$ for $x=1$ and thus $\ell_7(3^r)=1$. If $r\equiv 1\pmod{3}$, then $3^{2r}x^2+4\equiv 2x^2+4\pmod{7}$. Note that $\big(\frac{2\cdot 1^2+4}{7}\big)=\big(\frac{2\cdot 2^2+4}{7}\big)=-1$ and $\big(\frac{2\cdot 3^2+4}{7}\big)=1$. Therefore, $\ell_7(3^r)=3$. This completes the proof.\end{proof} \begin{lemma}\label{lemma52}Suppose that $m=3^{r}\cdot 7$, where $r\equiv 0\pmod{3}$ or $r\equiv 2\pmod{3}$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}For $r=0$, we can choose $a=1$ and $b=3$. By Lemma \ref{lemma51} (i), $\ell_7(3^r)\le 2$. For $r\ge 2$, the desired the conclusion follows from Lemma \ref{lemma33} (ii) on noting that $7+3^r\cdot 2\le 3^{r+1}<n$.\end{proof} \begin{lemma}\label{lemma53}Suppose that $m=3^{r}\cdot 7$, where $r\equiv 1\pmod{3}$. Suppose further that either $r\equiv 1\pmod{6}$ or $n\not\in \mathcal{E}$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}Note that $a^2+a(a+3^{r+1})+(a+3^{r+1})^2+1=3a^2+3^{r+2}a+3^{2r+2}+1$. It suffices to find $a\in \Z^{+}$ such that $3a^2+3^{r+2}a+3^{2r+2}+1\equiv 0\pmod{7}$ and $a+3^{r+1}\le n$. Note that for $r\equiv 1\pmod{3}$, we have $3^{2r+2}\equiv 4\pmod{7}$. On writing $r=6s+1+3t$ with $s\in \N$ and $t\in \{0,1\}$, we have \begin{align}\label{conginlemma53}3a^2+3^{r+2}a+3^{2r+2}+1 \equiv 3a^2+3^{3t+3}a+5\equiv 3a^2+(-1)^{t+1}a+5\pmod{7}.\end{align} If $t=0$, then by \eqref{conginlemma53} we can choose $a=1$ such that $3a^2+3^{r+2}a+3^{2r+2}+1\equiv 0\pmod{7}$ and $a+3^{r+1}=1+3^{r+1}\le n$. If $t=1$, then $n\not\in\mathcal{E}$ and $n\ge 3^{r+1}+3$. By \eqref{conginlemma53}, we choose $a=3$ such that $3a^2+3^{r+2}a+3^{2r+2}+1\equiv 0\pmod{7}$. Note that $b=3+3^{r+1}\le n$. We are done. \end{proof} \begin{lemma}[Case (viii), Part 1]\label{lemma54}Let $n\not\in \mathcal{E}$. Suppose that $m=3^{r}\cdot 7$. Then there exist $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$.\end{lemma} \begin{proof}The desired conclusion follows from Lemmas \ref{lemma52}-\ref{lemma53}.\end{proof} \begin{lemma}[Case (viii), Part 2]\label{lemma55}Let $n=3^{6s+5}+1$ or $n=3^{6s+5}+2$. Suppose that $m=3^{r}\cdot 7$. Then $a^3+a(1\le a\le n)$ are pairwise distinct modulo $m$.\end{lemma} \begin{proof}By \eqref{inequality1}, we have $r=6s+4$. Suppose otherwise that we can find $1\le a<b\le n$ such that $b^3+b\equiv a^3+a\pmod{m}$. Note that $3\nmid (a^2+ab+b^2+1)$ for any $a,b\in \Z$, and we conclude that $b=a+3^{6s+4}c$ for some $c\in \Z^{+}$. Since $b=a+3^{3s+4}c<n$, we have $c\le 3$. Write $\delta=3^{6s+4}$. Then $b^3+b\equiv a^3+a\pmod{m}$ implies that $a^2+ab+b^2+1\equiv 0\pmod{7}$, which is equivalent to \begin{align*}(6a+3\delta c)^2\equiv -3\delta^2c^2-12\pmod{7}.\end{align*} Therefore, we have $\Big(\frac{-3\delta^2c^2-12}{7}\Big)\in\{0,1\}$. By Lemma \ref{lemma51} (ii), $\ell_7(\delta)=3$ for $\delta=3^{6s+4}$, and we obtain $c\ge \ell_7(\delta)=3$. Now we conclude that $c=3$ and $b=a+3^{6s+5}$. Then we deduce that $a^2+ab+b^2+1\equiv 3a^2+a+5\equiv 0\pmod{7}$, and which implies $a\ge 3$ and $b=a+3^{3s+5}\ge 3+3^{3s+5}$. This is a contradiction to $b\le n$. The proof is complete. \end{proof} \bigskip \noindent {\it Proof of Lemmas \ref{lemma21}-\ref{lemma22}.} In view of Lemma \ref{lemma38}, Remark \ref{remark}, Lemma \ref{lemma41}, Lemma \ref{lemma42}, Lemma \ref{lemma43}, Lemma \ref{lemma44}, Lemma \ref{lemma45}, Lemma \ref{lemma410}, Lemma \ref{lemma54} and Lemma \ref{lemma55}, we only need to prove that each positive integer $m$ restricted by \eqref{inequality1} must satisfy (at least) one of the $8$ cases in Section 2. By \eqref{inequality1}, $m$ is not a power of $3$. If $m$ has no prime factors greater than $3$, then $m$ belongs to Case (iii) or (v). Next we assume that $m$ has two distinct prime factors greater than $3$. We write $m=m'p_1^{r_1}p_2^{r_2}$, where $p_1\not =p_2$ are two primes, $p_1\nmid m'$, $p_2\nmid m'$ and $r_1,r_2\in \Z^{+}$. Without loss of generality, we further assume that $p_1\not=7$ and $p_2\ge 7$. Let $\delta=m'p_2^{r_2}$. Then $m=\delta p_1^{r_1}$, $\delta\ge 7$ and $p_1\nmid \delta$. We can see that $m$ belongs to either Case (i) or Case (ii). Now we assume that $m$ has only one prime factor greater than $3$. We write $m=2^{i}3^{j}p^{r}$, where $p\ge 5$ is a prime, $r\in \Z^{+}$ and $i,j\in \N$. Note that if $i\ge 2$, then $m$ satisfies the condition of Case (iv). We discuss $i=1$ and $i=0$ below. We first consider $i=1$. If $j=0$, then $m$ belongs to Case (vii). If $j\ge 1$ and $r\ge 2$, then $m$ satisfies the condition of Case (ii). If $j\ge 1$, $r=1$ and $p\not=7$, then $m$ satisfies the condition of Case (i). If $j\ge 1$, $r=1$ and $p=7$, then $m$ satisfies the condition of Case (vi). Now we consider $i=0$. If $0\le j\le 1$, then $m$ belongs to Case (vii). If $j\ge 2$ and $r\ge 2$, then $m$ satisfies the condition of Case (ii). If $j\ge 2$, $r=1$ and $p\not=7$, then $m$ satisfies the condition of Case (i). If $j\ge 2$, $r=1$ and $p=7$, then $m$ satisfies the condition of Case (viii). We have proved that $m$ subject to \eqref{inequality1} must satisfy (at least) one of the $8$ cases in Section 2. According to the remark before Lemma \ref{lemma21}, we also complete the proof of Theorem \ref{theorem1}. \vskip3mm \bigskip
{ "timestamp": "2021-11-25T02:11:59", "yymm": "2111", "arxiv_id": "2111.11227", "language": "en", "url": "https://arxiv.org/abs/2111.11227", "abstract": "Let $\\Delta(n)$ denote the smallest positive integer $m$ such that $a^3+a(1\\le a\\le n)$ are pairwise distinct modulo $m$. The purpose of this paper is to determine $\\Delta(n)$ for all positive integers $n$.", "subjects": "Number Theory (math.NT)", "title": "On a discriminator for the polynomial $f(x)=x^3+x$", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9893474888461861, "lm_q2_score": 0.8104789063814616, "lm_q1q2_score": 0.8018452707913022 }
https://arxiv.org/abs/1201.5989
Nonconvexity of the set of hypergraph degree sequences
It is well known that the set of possible degree sequences for a graph on $n$ vertices is the intersection of a lattice and a convex polytope. We show that the set of possible degree sequences for a $k$-uniform hypergraph on $n$ vertices is not the intersection of a lattice and a convex polytope for $k \geq 3$ and $n \geq k+13$. We also show an analogous nonconvexity result for the set of degree sequences of $k$-partite $k$-uniform hypergraphs and the generalized notion of $\lambda$-balanced $k$-uniform hypergraphs.
\section{Introduction} The \emph{degree sequence} of a graph $G$ on vertices $v_1, v_2, \dots, v_n$ is the sequence $d(G)=(d_1, d_2, \dots, d_n)$, where $d_i$ is the degree of the vertex $v_i$ in $G$. The Erd\H{o}s-Gallai Theorem~\cite{ErdosGallai} states that a sequence $(d_1, d_2, \dots, d_n)$ is the degree sequence of a (simple) graph if and only if $\sum_i d_i$ is even and the $d_i$ satisfy a certain set of inequalities. Koren~\cite{Koren} showed that these inequalities define a convex polytope $D_n(2)$, so that the sequences with even sum lying in this polytope are exactly the degree sequences of graphs on $n$ vertices. (For more on this polytope, see \cite{Stanley}.) We consider the analogous question for $k$-uniform hypergraphs when $k>2$. Klivans and Reiner~\cite{KlivansReiner} verified computationally that the set of degree sequences for $k$-uniform hypergraphs is the intersection of a lattice and a convex polytope for $k=3$ and $n \leq 8$ and asked whether this holds in general. We will show in Section 2 that it does not hold for $k \geq 3$ and $n \geq k+13$. Similarly, we can associate to a bipartite graph a pair of degree sequences giving the degrees of the vertices in each part. The Gale-Ryser Theorem~\cite{Ryser} gives necessary and sufficient conditions in the form of a system of linear inequalities for a pair of degree sequences to arise from a bipartite graph, so that the set of these pairs of degree sequences can again be described as the intersection of a lattice and a convex polytope. We will show in Section 3 that the analogous result does not hold for $k$-partite $k$-uniform hypergraphs if there exist three parts of sizes at least 5, 6, and 6, respectively. We also generalize the notion of $k$-partite $k$-uniform hypergraphs to that of $\lambda$-balanced $k$-uniform hypergraphs and prove a similar statement in this case. \section{Hypergraph degree sequences} A \emph{(simple) $k$-uniform hypergraph} $K$ on the set $[n]=\{1, 2, \dots, n\}$ is a collection of distinct elements (called \emph{hyperedges}) of $\binom{[n]}{k}$, the $k$-element subsets of $[n]$. The \emph{degree sequence} of $K$ is $d(K)=(d_1, d_2, \dots, d_n)$, where $d_i$ is the number of hyperedges in $K$ containing $i$. We consider degree sequences as points in $\mathbf R^n$. Let $e_i$ be the $i$th standard basis vector, and for any $S = \{i_1, \dots, i_k\} \subset [n]$, write $e_S=e_{i_1i_2\dotsm i_k}=e_{i_1}+e_{i_2}+\dotsb+e_{i_k}$. Each degree sequence $d(K)$ is the sum of some subset of the $e_S$'s, so the convex hull of all such degree sequences is the zonotope \[D=D_n(k)=\Big\{ \sum_{S \in \binom{[n]}{k}} c_Se_S \mid 0 \leq c_S \leq 1\Big\}.\] (For more on this polytope, see \cite{BhanuMurthySrinivasan}.) Moreover, if we let $L \subset \mathbf Z^n$ be the lattice generated by the $e_S$ consisting of lattice points whose coordinates have sum divisible by $k$ (as long as $n>k$), then each $d(K)$ lies in $D \cap L$. Our main result will be to show that $D \cap L$ contains a point that is not the degree sequence of a $k$-uniform hypergraph when $k \geq 3$. As a remark, this is closely related to the weaker question of whether every point of $L$ lying in the real cone generated by the $e_S$ lies in the semigroup generated by the $e_S$. This is well known to be the case and is equivalent to normality of the monomial algebra generated by the $\mathbf x^S=x_{i_1}x_{i_2}\cdots x_{i_k}$. (See, for instance, \cite{Sturmfels}.) It is also easy to derive the affirmative answer to this question for $\lambda$-balanced hypergraphs as defined in the next section. The essential difference with the present question is that here we are restricted to using each hyperedge at most once. For a hypergraph $K$, we will define $D(K)$ to be the zonotope generated by the hyperedges in $K$, so \[D(K) = \left\{\sum_{S \in K} c_Se_S \mid 0 \leq c_S \leq 1 \right\}.\] \begin{lemma} \label{face} Let $K$ be a $k$-uniform hypergraph on $n$ vertices. Then any nonempty face of $D(K)$ is a translate of $D(K^0)$ for some $K^0 \subset K$. Moreover, $D(K) \cap L$ contains a point that is not the degree sequence of a subhypergraph of $K$ if and only if $D(K^0) \cap L$ contains a point that is not the degree sequence of a subhypergraph of $K^0$. \end{lemma} \begin{proof} Choose any weight vector $w \in (\mathbf R^*)^n$. To maximize $w\left(\sum c_Se_S\right)=\sum (c_S \cdot w(e_S))$ for $0 \leq c_S \leq 1$, we must take $c_S = 1$ when $w(e_S)>0$ and $c_S=0$ when $w(e_S)<0$, while $c_S$ can be arbitrary if $w(e_S)=0$. Thus the face on which $w$ is maximized is a translate of $D(K^0)$ by $\sum_{S \in K^+} e_S \in L$, where $K^+$ is the set of hyperedges $S$ on which $w$ is positive. The same argument gives the result for degree sequences (simply restrict $c_S$ to be 0 or 1). \end{proof} Therefore it suffices to exhibit a weight vector $w$ to maximize and a point of $D(K^0) \cap L$ that is not the degree sequence of a subhypergraph of $K^0$. \begin{prop} \label{example1} Let $k=3$ and $n=16$. If \[w=(8,6,6,4,1,1,0,0,0,0,-2,-2,-3,-3,-5,-12),\] then \[p=(2,1,1,2,1,1,1,1,1,1,1,1,2,2,2,1)\] lies in $D(K^0)\cap L$ but is not the degree sequence of a subhypergraph of $K^0$. \end{prop} \begin{proof} Since the sum of the entries of $p$ is $21 = 3\cdot 7$, $p$ lies in $L$. Also, \begin{multline*} p = \frac 13 (e_{2,3,16}+e_{4,5,15}+e_{4,6,15}+e_{5,6,11}+e_{5,6,12}+e_{7,8,9}+e_{7,8,10}+e_{7,9,10}+e_{8,9,10})\\ +\frac 23 (e_{1,4,16}+e_{1,13,15}+e_{1,14,15}+e_{2,13,14}+e_{3,13,14}+e_{4,11,12}). \end{multline*} Since $w$ vanishes on each $e_S$ on the right side, it follows that $p \in D(K^0)$. However, $p$ is not the degree sequence of a subhypergraph of $K^0$: since $w_7=w_8=w_9=w_{10}=0$ and otherwise $w_i\neq-w_j$, we have $(e_7+e_8+e_9+e_{10})\cdot e_S$ is 0 or 3 for any $S \in K^0$. But $(e_7+e_8+e_9+e_{10})\cdot p = 4$, which is not divisible by 3, so it cannot be the sum of some $e_S$ for $S \in K^0$. \end{proof} Using this, we can easily derive the following. \begin{thm} \label{main} For $k \geq 3$ and $n \geq k+13$, the set of degree sequences of $k$-uniform hypergraphs on $n$ vertices is not the intersection of a lattice and a convex polytope. \end{thm} \begin{proof} It suffices to show that there is a point in $D \cap L$ that is not a degree sequence (since $D$ and $L$ are the smallest convex polytope and lattice containing all degree sequences). Combining Lemma~\ref{face} and Proposition~\ref{example1} gives the result for $k=3$ and $n=16$. Since $D_n(k)$ is the face of $D_{n+1}(k)$ with last coordinate 0, Lemma~\ref{face} also gives the result for $k=3$ and $n\geq 16$. Consider the map $f\colon (d_1, d_2,\dots, d_n) \mapsto (d_1,d_2,\dots,d_n,\frac{1}{k}(d_1+\dots+d_n))$. Then $d$ is a $k$-uniform hypergraph degree sequence on $n$ vertices if and only if $f(d)$ is a $(k+1)$-uniform hypergraph degree sequence on $n+1$ vertices (simply add vertex $n+1$ to all hyperedges). Since $f$ is linear, it also sends $D_n(k)$ into $D_{n+1}(k+1)$, so any counterexample for $(n,k)$ yields a counterexample for $(n+1,k+1)$. An easy induction completes the proof. \end{proof} It is possible that with additional work or computation the constant 13 may be improved. In the next section, we will prove an analogous result for $k$-partite $k$-uniform hypergraphs as well as the more general $\lambda$-balanced hypergraphs. (Our construction below can also be used to prove Theorem~\ref{main} but with a constant of 14 instead of 13.) \section{$\lambda$-balanced hypergraphs} Let $\lambda=(\lambda_1, \lambda_2, \dots, \lambda_p)$ be a partition of $k$. We say a $k$-uniform hypergraph is \emph{$\lambda$-balanced} if its vertex set can be partitioned into $p$ sets $V_1, \dots, V_p$ such that each hyperedge contains $\lambda_i$ vertices from $V_i$. (We will also call a hyperedge $\lambda$-balanced if it satisfies this property.) A $(1,1,\dots, 1)$-balanced partition is called \emph{$k$-partite}. Note that every $k$-uniform hypergraph is $(k)$-balanced. Let $n_i=|V_i|$, and label the vertices in $V_i$ by $v^i_1, v^i_2, \dots, v^i_{n_i}$. We then associate to a $\lambda$-balanced hypergraph $K$ a degree sequence \[d=(d^1_1, d^1_2, d^1_3, \dots;\quad d^2_1, d^2_2, \dots;\quad \dots;\quad d^p_1, d^p_2, \dots),\] where $d^i_j$ gives the number of hyperedges in $K$ containing vertex $v^i_j$. As before, this degree sequence is $\sum_{S \in K} e_S$, where $e_S$ is the sum of the standard basis vectors in $\mathbf R^{n_1}\times \mathbf R^{n_2} \times \dots \times \mathbf R^{n_p}$ corresponding to vertices in the hyperedge $S$. When $n_i>\lambda_i$ for all $i$, the lattice $L$ generated by all possible $e_S$ consists of all sequences $d$ for which there exists $q\in \mathbf Z$ such that $\sum_{j=1}^{n_i} d^i_j = \lambda_iq$ for all $i$. (In other words, the sum of the degrees of the vertices in $V_i$ must be the same integer multiple of $\lambda_i$.) As before, we let $D$ be the zonotope generated by all $e_S$ for $\lambda$-balanced hyperedges $S$ and ask whether all points in $D \cap L$ are degree sequences for $\lambda$-balanced hypergraphs. We will again find that this is not the case for any $\lambda$ when $k\geq 3$ and the $n_i$ are sufficiently large. We first consider a special case. \begin{prop} \label{example2} Let $\lambda=(1,1,1)$ and $(n_1, n_2, n_3)=(5,6,6)$. Also let \[w = (-7,-7,-7,-7,-7;\quad 1, 1, 2, 2, 3, 3;\quad 6, 6,5,5,4,4)\] and define $K_0$ as in Lemma~\ref{face}. Then \[p=(11,9,6,3,1;\quad 2,4,6,8,3,7;\quad 2,4,6,8,3,7)\] lies in $D(K_0) \cap L$ but is not the degree sequence of a subhypergraph of $K^0$. \end{prop} \begin{proof} Define points \begin{alignat*}{2} p^-&=(10,8,4,2,0;\quad&1,3,5,7,2,6;\quad&1,3,5,7,2,6),\\ p^+&=(12,10,8,4,2;\quad&3,5,7,9,4,8;\quad&3,5,7,9,4,8), \end{alignat*} so $p=\frac12(p^-+p^+)$. Note that the sum of the coordinates of the three parts of $p^-$ are all 24, so $p^- \in L$. Likewise, $p^+$ and $p$ also lie in $L$. Let \[A=(a_{rs})= \begin{pmatrix} 1&2&0&0&0&0\\ 2&3&0&0&0&0\\ 0&0&3&4&0&0\\ 0&0&4&5&0&0\\ 0&0&0&0&1&3\\ 0&0&0&0&3&5\end{pmatrix}.\] Note that $K^0 = \{\{v^1_q,v^2_r,v^3_s\} \mid 1 \leq q \leq 5, 1\leq r,s \leq 6, a_{rs} \neq 0\}$. Then $p^- = \sum e_S$, where the sum ranges over all $S=\{v^1_q, v^2_r, v^3_s\}$ such that $q < a_{rs}$. Likewise $p^+ = \sum e_S$, where instead $q \leq a_{rs}$. Therefore $p^-$, $p^+$, and their midpoint $p$ lie in $D(K_0)$. We will now show that $p$ is not the degree sequence of a hypergraph that uses only hyperedges in $K^0$. Suppose it were, so that we could write $p=\sum_{S \in K} e_S$ for some $K \subset K_0$. Let $B=(b_{rs})$ be the $6 \times 6$ matrix such that $b_{rs}$ counts the number of $q$ for which $\{v^1_q,v^2_r,v^3_s\} \in K$. Then the sequence of row and column sums of $B$ must both be $(2,4,6,8,3,7)$. Since we also know that $0 \leq b_{rs}\leq 5$, this means that: \begin{align*} B_1=\begin{pmatrix} b_{11}&b_{12}\\b_{21}&b_{22} \end{pmatrix} &\in \left\{ \begin{pmatrix}0&2\\2&2\end{pmatrix}, \begin{pmatrix}1&1\\1&3\end{pmatrix}, \begin{pmatrix}2&0\\0&4\end{pmatrix} \right\}\\ B_2=\begin{pmatrix} b_{33}&b_{34}\\b_{43}&b_{44} \end{pmatrix} &\in \left\{ \begin{pmatrix}1&5\\5&3\end{pmatrix}, \begin{pmatrix}2&4\\4&4\end{pmatrix}, \begin{pmatrix}3&3\\3&5\end{pmatrix} \right\}\\ B_3=\begin{pmatrix} b_{55}&b_{56}\\b_{65}&b_{66} \end{pmatrix} &\in \left\{ \begin{pmatrix}0&3\\3&4\end{pmatrix}, \begin{pmatrix}1&2\\2&5\end{pmatrix} \right\}. \end{align*} Moreover, for $1 \leq r,s \leq 6$, the pair $\{v^2_r,v^3_s\}$ can appear in at most $\min \{q, b_{rs}\}$ hyperedges with one of the vertices in $\{v^1_1, \dots, v^1_q\}$. Therefore, if we let $\mu=(11,9,6,3,1)$, then $\mu_1+\dots+\mu_q \leq \sum_{r,s} \min \{q, b_{rs}\}$. In other words, if $\nu = (\nu_1, \dots, \nu_5)$ is the partition such that $\nu_q$ counts the number of $b_{rs}$ that are at least $q$, then $\mu_1 + \dots + \mu_q \leq \nu_1 + \dots + \nu_q$. It is now straightforward to show that there are no possible choices of $B_1$, $B_2$, and $B_3$ satisfying these conditions: if $B_3 = (\begin{smallmatrix}0&3\\3&4\end{smallmatrix})$, we cannot choose $B_1$ such that both $\mu_1 \leq \nu_1$ and $\mu_1+\mu_2 \leq \nu_1+\nu_2$. Similarly if $B_3=(\begin{smallmatrix}1&2\\2&5\end{smallmatrix})$, we cannot choose $B_2$ such that both $\mu_1+\mu_2+\mu_3 \leq \nu_1+\nu_2+\nu_3$ and $\mu_1+\mu_2+\mu_3+\mu_4 \leq \nu_1+\nu_2+\nu_3+\nu_4$. Thus $p \in D(K_0) \cap L$ is not the degree sequence of a hypergraph using only hyperedges in $K_0$. \end{proof} Combining Lemma~\ref{face} and Proposition~\ref{example2} gives our desired result for $3$-partite $3$-uniform hypergraphs, and we can easily extend this result to $k$-partite $k$-uniform hypergraphs. \begin{thm} \label{partite} For $k \geq 3$, consider $k$-partite $k$-uniform hypergraphs with parts of sizes $n_1, n_2, \dots, n_k$ for which $n_1 \geq 5$, $n_2 \geq 6$, $n_3 \geq 6$, and $n_i \geq 1$ otherwise. The corresponding set of degree sequences is not the intersection of a lattice and a convex polytope. \end{thm} \begin{proof} As in Theorem~\ref{main}, combining Lemma~\ref{face} and Proposition~\ref{example2} gives the result for $k=3$ and $(n_1, n_2, n_3)=(5,6,6)$. Also note that the polytopes and lattices for $k\geq 3$ with $(n_1, n_2, n_3, n_4, \dots, n_k) = (5,6,6,1,\dots, 1)$ are all identical to the $k=3$ case (by projecting away the last $k-3$ coordinates) so this also proves those cases. Finally, increasing any $n_i$ but restricting to the face of the zonotope where the new vertices have degree 0 again reduces to the same case by Lemma~\ref{face}, completing the proof. \end{proof} Theorem~\ref{partite} is also easy to extend to $\lambda$-balanced hypergraphs for all $\lambda$ when $k \geq 3$. Consider a $\lambda$-balanced hypergraph on vertex sets $V_1, \dots, V_p$ of sizes $n_1, n_2, \dots, n_p$. We will say that $(n_1, \dots, n_p)$ is a \emph{$\lambda$-coarsening} of $(m_1, \dots, m_k)$ if each $V_i$ can be partitioned into $\lambda_i$ sets such that the sizes of all the resulting sets are $m_1, \dots, m_k$. \begin{thm}\label{balanced} Consider $\lambda$-balanced hypergraphs with parts of sizes $n_1, \dots, n_p$, where $(n_1, \dots, n_p)$ is a $\lambda$-coarsening of $(m_1, m_2, \dots, m_k)$ such that Theorem~\ref{partite} holds for parts of sizes $m_1, \dots, m_k$. (In particular, this will hold whenever the $n_i$ are sufficiently large.) Then the corresponding set of degree sequences is not the intersection of a lattice and a convex polytope. \end{thm} \begin{proof} Let the vertex sets $V_1, \dots, V_p$ have corresponding coarsening $W_1, \dots, W_k$. It suffices to exhibit a weight vector $w$ such that the corresponding $K_0$ as in Lemma~\ref{face} is the complete $k$-partite $k$-uniform hypergraph on $W_1, \dots, W_k$. Indeed, any hyperedge in $K_0$ will be $\lambda$-balanced by the definition of $\lambda$-coarsening, and the lattice generated by hyperedges in $K_0$ is a sublattice of the lattice generated by all $\lambda$-balanced hyperedges. Therefore any counterexample for $K_0$ will yield a counterexample for $\lambda$-balanced hypergraphs as in Lemma~\ref{face}. To exhibit such a weight vector, let $N$ be an integer larger than any $m_i$. Then let the weight of vertices in $W_1$ be $-(1+N+N^2+\dots+N^{k-2})$ and in $W_i$ be $N^{i-2}$ for $2 \leq i \leq k$. Then the only way to pick $k$ vertices the sum of whose weights is 0 is to take one from each $W_i$. In other words, the only hyperedges in $K_0$ are those that have one vertex from each $W_i$, as desired. \end{proof} \section{Acknowledgments} The author would like to thank Victor Reiner for suggesting this direction of study, as well as for useful discussions and overall encouragement. This work was supported by a National Science Foundation Mathematical Sciences Postdoctoral Research Fellowship.
{ "timestamp": "2012-01-31T02:01:49", "yymm": "1201", "arxiv_id": "1201.5989", "language": "en", "url": "https://arxiv.org/abs/1201.5989", "abstract": "It is well known that the set of possible degree sequences for a graph on $n$ vertices is the intersection of a lattice and a convex polytope. We show that the set of possible degree sequences for a $k$-uniform hypergraph on $n$ vertices is not the intersection of a lattice and a convex polytope for $k \\geq 3$ and $n \\geq k+13$. We also show an analogous nonconvexity result for the set of degree sequences of $k$-partite $k$-uniform hypergraphs and the generalized notion of $\\lambda$-balanced $k$-uniform hypergraphs.", "subjects": "Combinatorics (math.CO)", "title": "Nonconvexity of the set of hypergraph degree sequences", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9893474894743616, "lm_q2_score": 0.8104788995148792, "lm_q1q2_score": 0.8018452645069891 }
https://arxiv.org/abs/2105.03086
Pseudorandom sequences derived from automatic sequences
Many automatic sequences, such as the Thue-Morse sequence or the Rudin-Shapiro sequence, have some desirable features of pseudorandomness such as a large linear complexity and a small well-distribution measure. However, they also have some disastrous properties in view of certain applications. For example, the majority of possible binary patterns never appears in automatic sequences and their correlation measure of order 2 is extremely large.Certain subsequences, such as automatic sequences along squares, may keep the good properties of the original sequence but avoid the bad ones.In this survey we investigate properties of pseudorandomness and non-randomness of automatic sequences and their subsequences and present results on their behaviour under several measures of pseudorandomness including linear complexity, correlation measure of order $k$, expansion complexity and normality. We also mention some analogs for finite fields.
\section{Introduction Pseudorandom sequences are sequences generated by deterministic algorithms which shall simulate randomness. In contrast to truly random sequences they are not random at all but guarantee certain desirable features and are reproducible. Automatic sequences, see Section~\ref{sec:automatic_sequences} below for the definition, have some of these desirable features but also some undesirable ones. For example, the Thue-Morse sequence $(t_n)$, defined by \eqref{tmdef} below, \begin{itemize} \item has large $N$th linear complexity, see Section~\ref{sec:linear_complexity}, \item has large $N$th maximum-order complexity, see Section~\ref{sec:max-order_complexity}, \item is balanced and has a small well-distribution measure, see Section~\ref{sec:correlation}. \end{itemize} However, the Thue-Morse sequence \begin{itemize} \item has a very large correlation measure of order $2$, see Section~\ref{sec:correlation}, \item a very small expansion complexity, see Section~\ref{sec:expansion_complexity}, \item and there are short patterns such as $000$ and $111$ which do not appear in the sequence and its subword complexity is only linear, see Section~\ref{sec:normality}. \end{itemize} Hence, despite some nice features this sequence is not looking random at all, see Figure~\ref{tmfig}. The same is true for the Rudin-Shapiro sequence $(r_n)$ defined by~\eqref{rsdef} below and many other related sequences. \begin{figure} \begin{center} \includegraphics[scale=0.15]{TM.png} \qquad \includegraphics[scale=0.15]{RS.png} \end{center} \caption{The first $4096$ elements of the Thue-Morse (left) and Rudin-Shapiro (right) sequence split into $64$ rows of each $64$ sequence elements. Zeros are represented by white, ones are represented by black.} \label{tmfig} \end{figure} Taking suitable subsequences may destroy the non-random structure of the original sequence but may keep the desirable features of pseudorandomness. Promising candidates for such subsequences are \begin{itemize} \item along squares, cubes, bi-squares, ... or any polynomial values for any polynomial $f$ of degree at least $2$ with $f(\mathbb{N}_0)\subset \mathbb{N}_0$, \item along primes, \item along the Piateski-Shapiro sequence $\lfloor n^c\rfloor$, $1<c<2$, \item and along geometric sequences such as $3^n$. \end{itemize} For example, the Thue-Morse sequence and the Rudin-Shapiro sequence along squares still \begin{itemize} \item have a large maximum-order complexity and thus a large linear complexity, see Section~\ref{sec:max-order_complexity}, \item and are asymptotically balanced, see Section~\ref{sec:normality}. \end{itemize} Moreover, in contrast to the original sequence they \begin{itemize} \item have unbounded expansion complexity, see Section~\ref{sec:expansion_complexity}, \item and are normal, that is, asymptotically each pattern appears with the right frequency in the sequence, see Section~\ref{sec:normality}. \end{itemize} Roughly speaking, they look much more random than the original sequences, see Figure~\ref{squarefig}. \begin{figure}[ht] \begin{center} \includegraphics[scale=0.15]{TM_along_squares.png} \qquad \includegraphics[scale=0.15]{RS_along_squares.png} \end{center} \caption{The first $4096$ elements of the Thue-Morse (left) and Rudin-Shapiro (right) sequence along squares split into $64$ rows of each $64$ sequence elements. Zeros are represented by white, ones are represented by black.} \label{squarefig} \end{figure} Still some questions about these sequences remain open such as upper bounds on the correlation measure of order $k$ and on the expansion complexity. We will state explicitly some selected open problems to motivate future research. We also look for further directions in Section~\ref{sec:finite_fields}. In particular, we discuss analogs of the Thue-Morse and Rudin-Shapiro sequence and their subsequences in the setting of finite fields. For general background on automatic sequences and finite automata we refer to the monograph of Allouche and Shallit \cite{alsh2} and also to \cite{alsh,alshya,ev,fo}. For surveys on pseudorandom sequences see \cite{gy,merisa,niwi,sh,towi}. \section{Finite automata and automatic sequences}\label{sec:automatic_sequences} Roughly speaking, a sequence is {\em automatic} if it is generated by a finite automaton, see Definition~\ref{def:sequence} below. \begin{definition Let $k\geq 2$ be an integer. A \emph{finite $k$-automaton} ${\mathcal A}$ is a $6$-tuple $$ {\mathcal A}=(Q,\Sigma, \delta, q_0, \varphi,\Delta), $$ where \begin{itemize} \item $Q$ is a finite set of states, \item $\Sigma=\{0,1,\ldots,k-1\}$ is the input alphabet, \item $\delta: Q \times \Sigma \rightarrow Q$ is the transition function, \item $q_0\in Q$ is the initial state, \item $\Delta$ is the output alphabet \item and $\varphi:Q \rightarrow \Delta$ is the output function. \end{itemize} \end{definition} For example, the \emph{Thue-Morse automaton}, see Figure~\ref{fig:TM}, is a $2$-automaton with $2$ states and the \emph{Rudin-Shapiro automaton}, see Figure~\ref{fig:RS}, is a $2$-automaton with $4$ states, both with inputs and outputs in $\Sigma=\Delta=\{0,1\}$. \begin{figure}[ht] \begin{center} \begin{tikzpicture}[auto,thick] \node (E) at (-2,0) [circle] {}; \node (A) at (0,0) [circle, draw] {$A/0$}; \node (B) at (3,0) [circle, draw] {$B/1$}; \draw [->,bend left] (A) to node {1} (B); \draw [->,bend left] (B) to node {1} (A); \path (B) edge [loop above] node {0} (B); \path (A) edge [loop above] node {0} (A); \draw [->] (E) to node [midway,above,align=center ] {\texttt{start}} (A); \end{tikzpicture} \end{center} \caption{Thue-Morse automaton} \label{fig:TM} \end{figure} \begin{figure}[ht] \begin{center} \begin{tikzpicture}[auto,thick] \node (E) at (-2,0) [circle] {}; \node (A) at (0,0) [circle, draw] {$A/0$}; \node (B) at (3,0) [circle, draw] {$B/0$}; \node (C) at (6,0) [circle, draw] {$C/1$}; \node (D) at (9,0) [circle, draw] {$D/1$}; \draw [->,bend left] (A) to node {1} (B); \draw [->,bend left] (B) to node {0} (A); \draw [->,bend left] (B) to node {1} (C); \draw [->,bend left] (C) to node {1} (B); \draw [->,bend left] (C) to node {0} (D); \draw [->,bend left] (D) to node {1} (C); \path (D) edge [loop above] node {0} (B); \path (A) edge [loop above] node {0} (A); \draw [->] (E) to node [midway,above,align=center ] {\texttt{start}} (A); \end{tikzpicture} \end{center} \caption{Rudin-Shapiro automaton} \label{fig:RS} \end{figure} \begin{definition}\label{def:sequence} Let $\Delta$ be a finite set. A sequence $(s_n)$ over $\Delta$ is called a \emph{$k$-automatic sequence} if there is a $k$-automaton ${\mathcal A}$ such that on input of the digits $n_0,n_1,\ldots$ of the $k$-ary expansion of $n\geq 0$, \begin{equation}\label{eq:k-ary} n=\sum_{i\geq 0} n_i k^i, \quad n_i\in\{0,1,\dots, k-1\}, \end{equation} ${\mathcal A}$ outputs the sequence element $s_n\in \Delta$. Reading of the digits of $n$ starting with the most significant digit is called {\em direct} whereas reading starting with the least significant digit $n_0$ is called {\em reverse}. If not stated otherwise, we use reverse reading. Finally, a sequence is called {\em automatic} if it is $k$-automatic for some $k$. \end{definition} \begin{example}[Thue-Morse sequence] The \emph{Thue-Morse sequence}~$(t_n)$ is a $2$-automatic sequence generated by the Thue-Morse automaton, Figure~\ref{fig:TM}. This sequence is the {\em sequence of the sum of digits modulo $2$}. The sequence begins with $$ 011010011001 \dots, $$ see also Figure~\ref{tmfig} for a picture of the first $4096$ sequence elements. It follows from the defining automaton, see Figure~\ref{fig:TM}, that $(t_n)$ satisfies the following recurrence relation \begin{equation}\label{tmdef} t_n= \left\{ \begin{array}{cl} t_{n/2} & \mbox{if $n$ is even},\\ t_{(n-1)/2}+1 \bmod 2 & \mbox{if $n$ is odd}, \end{array}\right. \quad n=1,2,\ldots \end{equation} with initial value $t_0=0$. \end{example} \begin{example}[Rudin-Shapiro sequence] The \emph{Rudin-Shapiro sequence} $(r_n)$ is a $2$-automatic sequence generated by the Rudin-Shapiro automaton, see Figure~\ref{fig:RS}. The sequence begins with $$ 000100100001 \dots, $$ see also Figure~\ref{tmfig} for a picture of the first $4096$ sequence elements. It follows from the defining automaton, see Figure~\ref{fig:RS}, that $(r_n)$ satisfies the following recurrence relation \begin{equation}\label{rsdef} r_n= \left\{ \begin{array}{cl} r_{\lfloor n/2\rfloor}+1 \bmod 2 & \mbox{if $n\equiv 3 \bmod 4$},\\ r_{\lfloor n/2\rfloor} & \mbox{otherwise}, \end{array}\right. \quad n=1,2,\ldots \end{equation} with initial value $r_0=0$. \end{example} The sequence $((-1)^{r_n})$ over $\{-1,+1\}$ is also called Rudin-Shapiro sequence in the literature. Here we study only the sequence $(r_n)$ over $\{0,1\}$. \begin{example}[Pattern sequences] For a pattern $P\in \Delta^\ell\setminus\{(0,\dots,0)\}$ of length $\ell$ over $\Delta=\{0,1,\dots, k-1\}$ define the sequence $(p_n)$ by \begin{equation*} p_n= e_P(n) \bmod k, \quad 0\leq p_n<k, \quad n=0,1,\dots, \end{equation*} where $e_P(n)$ is the number of occurrences of $P$ in the $k$-ary expansion of $n$. The sequence~$(p_n)$ over~$\Delta$ satisfies the following recurrence relation \begin{equation}\label{eq:recurrence} p_n= \left\{ \begin{array}{cl} p_{\lfloor n/k\rfloor}+1 \bmod k & \text{if } n\equiv a \bmod k^\ell,\\ p_{\lfloor n/k\rfloor} & \text{otherwise,} \end{array} \right. n=1,2,\dots \end{equation} with initial value $p_0=0$, where $a=a(P)$ is the integer $0< a <k^\ell$ such that its $k$-ary expansion corresponds to the pattern $P$. Classical examples for binary pattern sequences are the Thue-Morse sequence with $$k=2,\quad \ell=1, \quad P=1 \quad \mbox{and}\quad a=1,$$ and the Rudin-Shapiro sequence with $$k=2,\quad \ell=2,\quad P=11\quad \mbox{and}\quad a=3.$$ In particular, if $n_0,n_1,\dots$ are the bits of the non-negative integer taken from \eqref{eq:k-ary} with $k=2$, then \begin{equation}\label{sumofdigitsdef} t_n=\sum_{i=0}^\infty n_i \bmod 2 \quad \mbox{and}\quad r_n=\sum_{i=0}^\infty n_in_{i+1} \bmod 2. \end{equation} \end{example} \begin{example}[Rudin-Shapiro-like sequence] Lafrance, Rampersad and Yee \cite{laraye} introduced a {\em Rudin-Shapiro-like sequence} $(\ell_n)$ which is based on the number of occurrences of the pattern $10$ as a scattered subsequence in the binary representation, \eqref{eq:k-ary} with $k=2$, of $n$. That is,~$\ell_n$ is the parity of the number of pairs $(i,j)$ with $i>j$ and $(n_i,n_j)=(1,0)$. See Figure~\ref{fig:RS-like} for its defining automaton. \begin{figure}[ht] \begin{center} \begin{tikzpicture}[auto,thick] \node (E) at (-2,0) [circle] {}; \node (A) at (0,0) [circle, draw] {$A/0$}; \node (B) at (3,0) [circle, draw] {$B/0$}; \node (C) at (6,0) [circle, draw] {$C/1$}; \node (D) at (9,0) [circle, draw] {$D/1$}; \draw [->,bend left] (A) to node {1} (B); \draw [->,bend left] (B) to node {1} (A); \draw [->,bend left] (B) to node {0} (C); \draw [->,bend left] (C) to node {0} (B); \draw [->,bend left] (C) to node {1} (D); \draw [->,bend left] (D) to node {1} (C); \path (D) edge [loop above] node {0} (B); \path (A) edge [loop above] node {0} (A); \draw [->] (E) to node [midway,above,align=center ] {\texttt{start}} (A); \end{tikzpicture} \end{center} \caption{Rudin-Shapiro-like automaton with direct reading} \label{fig:RS-like} \end{figure} This sequence can also be defined by \begin{equation}\label{eq:rslike}\ell_{2n+1}=\ell_n \quad \mbox{and}\quad \ell_{2n}=\ell_n+t_n\bmod 2, \end{equation} see \cite[$(1)$ and $(2)$]{laraye}, with initial value $\ell_0=0$ and where $(t_n)$ is the Thue-Morse sequence. \end{example} \begin{example}[Baum-Sweet sequence] The \emph{Baum-Sweet sequence} $(b_n)$ is a $2$-automatic sequence defined by the rule $b_0=1$ and for $n\ge 1$ $$ b_n= \left\{ \begin{array}{cl} 1& \text{if the binary representation of $n$ contains no block of} \\ & \text{consecutive $0$'s of odd length,}\\ 0& \text{otherwise.} \end{array} \right. $$ Equivalently, we have for $n\geq 1$ of the form $n=4^\ell m$ with $4\nmid m$ that \begin{equation}\label{bsdef} b_n= \left\{ \begin{array}{cl} 0& \text{if $m$ is even}, \\ b_{(m-1)/2}& \text{if $m$ is odd.} \end{array} \right. \end{equation} The sequence $(b_n)$ is generated by the Baum-Sweet automaton in Figure~\ref{fig:BS}. \end{example} \begin{figure}[ht] \begin{center} \begin{tikzpicture}[auto,thick] \node (E) at (-2,0) [circle] {}; \node (A) at (0,0) [circle, draw] {$A/1$}; \node (B) at (3,0) [circle, draw] {$B/1$}; \node (C) at (6,0) [circle, draw] {$C/0$}; \path (A) edge [loop above] node {1} (A); \draw [->] (E) to node [midway,above,align=center ] {\texttt{start}} (A); \draw [->,bend left] (A) to node {0} (B); \draw [->,bend left] (B) to node {0} (A); \draw [->,left] (B) to node [midway,above,align=center ] {1} (C); \path (C) edge [loop above] node {0,1} (C); \end{tikzpicture} \end{center} \caption{Baum-Sweet automaton} \label{fig:BS} \end{figure} \begin{example}[Characteristic sequence of sums of three squares] Consider the {\em characteristic sequence $(c_n)$ of the set of integers which are sums of three squares of an integer}, that is, $$ c_n= \left\{ \begin{array}{cl} 1& \text{if } n=a^2+b^2+c^2 \text{ for some non-negative integers $a,b,c$,} \\ 0& \text{otherwise.} \end{array} \right. $$ By Legendre's three-square theorem, we have the equivalent definition \begin{equation}\label{eq:cndef} c_n= \left\{ \begin{array}{cl} 1& \text{if $n$ is not of the form $n=4^\ell(8k+7)$,} \\ 0& \text{otherwise.} \end{array} \right. \end{equation} See Figure~\ref{fig:3squares} for the defining automaton. \begin{figure}[ht] \begin{center} \begin{tikzpicture}[auto,thick] \node (AA) at (-2,0) [circle] {}; \node (A) at (0,0) [circle, draw] {$A/1$}; \node (B) at (3,0) [circle, draw] {$B/1$}; \node (C) at (0,-3) [circle, draw] {$C/1$}; \node (D) at (3,-3) [circle, draw] {$D/1$}; \node (E) at (6,-1.5) [circle, draw] {$E/1$}; \node (F) at (9,-1.5) [circle, draw] {$F/0$}; \draw [->] (A) to node {1} (B); \draw [->,bend left] (A) to node {0} (C); \draw [->] (B) to node {1} (E); \draw [->] (B) to node {0} (D); \draw [->] (C) to node {1} (D); \draw [->,bend left] (C) to node {0} (A); \draw [->] (E) to node {1} (F); \draw [->] (E) to node {0} (D); \path (D) edge [loop below] node {0,1} (D); \path (F) edge [loop below] node {0,1} (F); \draw [->] (AA) to node [midway,above,align=center ] {\texttt{start}} (A); \end{tikzpicture} \end{center} \caption{Automaton of the characteristic sequence of sums of three squares with reverse reading} \label{fig:3squares} \end{figure} \end{example} \begin{example}[Regular paper-folding sequence] The {\em regular paper-folding sequence} $(v_n)$ with initial value $v_0\in \{0,1\}$ is defined as follows. If $n=2^km$ with an odd $m$, then \begin{equation}\label{pfdef} v_n=\left\{\begin{array}{ll}1,& m\equiv 1\bmod 4,\\ 0, &m\equiv 3\bmod 4,\end{array}\right.\quad n=1,2,\ldots \end{equation} Its defining automaton with four states is given in Figure~\ref{fig:RPF}. \end{example} \begin{figure}[ht] \begin{center} \begin{tikzpicture}[auto,thick] \node (E) at (-2,0) [circle] {}; \node (A) at (0,0) [circle, draw] {$A/v_0$}; \node (B) at (3,0) [circle, draw] {$B/1$}; \node (C) at (6,1.5) [circle, draw] {$C/0$}; \node (D) at (6,-1.5) [circle, draw] {$D/1$}; \path (A) edge [loop above] node {0} (A); \draw [->] (E) to node [midway,above,align=center ] {\texttt{start}} (A); \draw [->] (A) to node {1} (B); \draw [->] (B) to node [midway,above,align=center ] {1} (C); \draw [->] (B) to node [midway,above,align=center ] {0} (D); \path (C) edge [loop above] node {0,1} (C); \path (D) edge [loop above] node {0,1} (D); \end{tikzpicture} \end{center} \caption{Regular paper-folding automaton} \label{fig:RPF} \end{figure} \begin{example}[An automatic apwenian sequence] Any binary sequence $(a_n)$ satisfying $a_0=1$ and $$ a_{2n+2}=a_{2n+1}+a_n \bmod 2,\quad n=0,1,\ldots $$ is called {\em apwenian}, see for example \cite{alhani}. Apwenian sequences which are $2$-automatic are characterized in \cite{alhani}. For example, the sequence $(w_n)$ defined by \begin{equation}\label{apdef} w_{2n}=1\quad \mbox{and}\quad w_{2n+1}=w_n+1\bmod 2,\quad n=0,1,\ldots \end{equation} is apwenian and defined by the automaton in Figure~\ref{fig:apw}. \end{example} \begin{figure}[ht] \begin{center} \begin{tikzpicture}[auto,thick] \node (E) at (-2,0) [circle] {}; \node (A) at (0,0) [circle, draw] {$A/1$}; \node (B) at (3,2) [circle, draw] {$B/1$}; \node (C) at (3,-2) [circle, draw] {$C/0$}; \node (D) at (6,-2) [circle, draw] {$D/0$}; \draw [->] (E) to node [midway,above,align=center ] {\texttt{start}} (A); \draw [->] (A) to node [midway,above,align=center ] {0} (B); \draw [->,bend left] (A) to node [midway,above,align=center ] {1} (C); \draw [->,bend left] (C) to node [midway,above,align=center ] {1} (A); \draw [->] (C) to node [midway,above,align=center ] {0} (D); \path (B) edge [loop above] node {0,1} (B); \path (D) edge [loop above] node {0,1} (D); \end{tikzpicture} \end{center} \caption{Apwenian automaton} \label{fig:apw} \end{figure} In addition to the examples above, all ultimately periodic sequences are $k$-automatic for all integers $k\geq 2$, see \cite[Theorem~5.4.2]{alsh2}. Moreover, by Cobham's theorem \cite[Theorem~11.2.1]{alsh2}, if a sequence $(s_n)$ is both $k$-automatic and $\ell$-automatic and $k$ and $\ell$ are multiplicatively independent,\footnote{Two integers $k$ and $\ell$ are {\em multiplicatively dependent} if $k^r=\ell^s$ for some positive integers $r$ and~$s$. Otherwise they are {\em multiplicatively independent}.} then $(s_n)$ is ultimately periodic. For a prime power $k=q$, $k$-automatic sequences $(s_n)$ over the finite field\footnote{For a prime power $q$ we denote the finite field of size $q$ by $\mathbb{F}_q$.} $\Delta=\mathbb{F}_q$ can be characterized by a result of Christol, see \cite{christol} for prime $q$ and \cite{chka} for prime power $q$ as well as \cite[Theorem~12.2.5]{alsh2}. \begin{theorem}\label{thm:christol} Let \footnote{We denote by $\mathbb{F}_q \llbracket x \rrbracket$ the ring of formal power series over $\mathbb{F}_q$.} $$ G(x)=\sum_{n=0}^{\infty}s_nx^n \in \mathbb{F}_q \llbracket x \rrbracket $$ be the {\em generating function} of the sequence $(s_n)$ over $\mathbb{F}_q$. Then $(s_n)$ is $q$-automatic if and only if $G(x)$ is algebraic over $\mathbb{F}_q(x)$, that is, there is a polynomial $h(x,y)\in \mathbb{F}_q[x,y]\setminus\{0\}$ such that $h(x,G(x))=0$. \end{theorem} Note that for all $m=1,2,\ldots$ a sequence is $k$-automatic if and only if it is $k^m$-automatic by \cite[Theorem~6.6.4]{alsh2} and even a slightly more general version of Christol's result holds: For a prime $p$ and positive integers $m$ and $r$, $(s_n)$ is $p^m$-automatic over $\mathbb{F}_{p^r}$ if and only if $G(x)$ is algebraic over $\mathbb{F}_{p^r}(x)$. \begin{example} The generating function $G(x)$ of the Thue-Morse sequence $(t_n)$ over $\mathbb{F}_2$ satisfies $h(x,G(x))=0$ with \begin{equation}\label{eq:TH-equation} h(x,y)=(x+1)^3y^2 + (x+1)^2y+x. \end{equation} The generating function $G(x)$ of the Rudin-Shapiro sequence $(r_n)$ over $\mathbb{F}_2$ satisfies $h(x,G(x))=0$ with \begin{equation}\label{eq:RS-equation} h(x,y)=(x+1)^{5}y^2 + (x+1)^4 y + x^3. \end{equation} In general, for prime $p$ the generating function $G(x)$ of the $p$-ary pattern sequence $(p_n)$ over $\mathbb{F}_p$ with respect to the pattern $P$ of length $\ell$ satisfies $h(x,G(x))=0$ with \begin{equation}\label{eq:pattern-equation} h(x,y)=(x-1)^{p^\ell +p -1}y^p - (x-1)^{p^\ell} y - x^{a(P)}. \end{equation} The generating function $G(x)$ of the Rudin-Shapiro-like sequence $(\ell_n)$ over $\mathbb{F}_2$ defined by~\eqref{eq:rslike} satisfies $h(x,G(x))=0$ with \begin{equation}\label{eq:rslh} h(x,y)=(x+1)^8y^4+(x^6+x^5+x^2+x)y^2+(x+1)^4y+x^2, \end{equation} see \cite[Proof of Theorem 2]{suzeli}. The generating function $G(x)$ of the Baum-Sweet sequence $(b_n)$ over $\mathbb{F}_2$ satisfies \linebreak[4] $h(x,G(x))=0$ with \begin{equation}\label{eq:BS-equation} h(x,y)=y^3+xy+1. \end{equation} The generating function $G(x)$ of the characteristic sequence $(c_n)$ of sums of three squares~\eqref{eq:cndef} over $\mathbb{F}_2$ satisfies $h(x,G(x))=0$ with \begin{equation}\label{eq:cnh} h(x,y)=(x+1)^8(y+y^4)+x^6+x^5+x^3+x^2+x, \end{equation} see \cite[Equation (7)]{howi}. The generating function $G(x)$ of the regular paper-folding sequence $(v_n)$ over $\mathbb{F}_2$ satisfies $h(x,G(x))=0$ with \begin{equation}\label{pfh} h(x,y)=(x+1)^4(y^2+y)+x. \end{equation} The generating function $G(x)$ of the apwenian sequence $(w_n)$ over $\mathbb{F}_2$ defined by \eqref{apdef} satisfies \begin{equation}\label{aph} h(x,y)=(x+1)(xy^2+y)+1. \end{equation} \end{example} \section{Linear complexity}\label{sec:linear_complexity} The linear complexity is a figure of merit of pseudorandom sequences introduced to capture undesirable linear structure in a sequence. It originates in cryptography and provides a test of randomness which is a standard tool to filter sequences with non-randomness properties and is implemented in many test suites such as NIST and TestU01~\cite{nist,testU01}. \begin{definition} The \emph{$N$th linear complexity} $L(s_n, N)$ of a sequence $(s_n)$ over $\mathbb{F}_q$ is the length~$L$ of a shortest linear recurrence relation satisfied by the first $N$ elements of $(s_n)$, $$ s_{n+L}=c_{L-1}s_{n+L-1}+\dots +c_1s_{n+1}+c_0s_n, \quad 0\leq n\leq N-L-1, $$ for some $c_0,\ldots,c_{L-1}\in \mathbb{F}_q$. We use the convention that $L(s_n,N)=0$ if the first $N$ elements of $(s_n)$ are all zero and $L(s_n,N)=N$ if $s_0=\dots=s_{N-2}=0\ne s_{N-1}$. The sequence~$(L(s_n,N))_{N=1}^\infty$ is called {\em linear complexity profile} of $(s_n)$ and $$ L(s_n)=\sup_{N\ge 1} L(s_n,N) $$ is the {\em linear complexity} of $(s_n)$. \end{definition} Clearly, $0\leq L(s_n,N) \leq N$ and $L(s_n,N)\leq L(s_n,N+1)$. For truly random sequences $(s_n)$ the expected value of its $N$th linear complexity $L$ is $$\frac{N}{2}+O(1), $$ see for example \cite[Theorem~10.4.42]{handbookFF}. Deviations of order of magnitude $\log N$ must appear for infinitely many $N$. More precisely, for a prime power $q$ consider the following probability measure of sequences over $\mathbb{F}_q$ determined by \begin{equation}\label{eq:prob_space} \mathbb{P}\left[(s_n)\in \mathbb{F}_q^{\infty}: (s_0,\dots, s_{\ell-1})= (c_0,\dots, c_{\ell-1}) \right]=q^{-\ell}, \quad c_0,\dots, c_{\ell-1}\in \mathbb{F}_q. \end{equation} Then we have the following result on the deviation from the expected value, see\cite[Theorem~10]{Niederreiter88}. \begin{theorem} We have $$ \limsup_{N\rightarrow \infty }\frac{L(s_n, N) -N/2}{\log N}=\frac{1}{2 \log q}, $$ and $$ \liminf_{N\rightarrow \infty }\frac{L(s_n, N) -N/2}{\log N}=\frac{-1}{2 \log q} $$ with probability one with respect to the probability measure \eqref{eq:prob_space}. \end{theorem} It is well-known \cite[Lemma~1]{Niederreiter88-b} that $L(s_n)<\infty$ if and only if $(s_n)$ is ultimately periodic, that is, its generating function is rational: $G(x)=g(x)/f(x)$ with polynomials $g(x),f(x)\in \mathbb{F}_q[x]$. The $N$th linear complexity is a measure for the unpredictability of a sequence. A large $N$th linear complexity, up to sufficiently large $N$, is necessary, but not sufficient, for cryptographic applications. Sequences of small linear complexity are also weak in view of Monte-Carlo methods, see \cite{do,domewi,dowi,DorferWinterhof}. For more background on linear complexity and related measures of pseudorandomness we refer to \cite[Section~10.4]{handbookFF} and \cite{Niederreiter2003,towi,wi1}. M\'erai and Winterhof \cite{mewi18} showed that automatic sequences which are not ultimately periodic possess large $N$th linear complexity. \begin{theorem}\label{thm:MW-lin-compl-gen} Let $q$ be a prime power and $(s_n)$ be a $q$-automatic sequence over $\mathbb{F}_q$ which is not ultimately periodic. Let $h(x,y)=h_0(x)+h_1(x)y+\dots + h_d(x)y^d\in\mathbb{F}_q[x,y]$ be a non-zero polynomial with $h(G(x),x)=0$ with no rational zero. Put $$ M=\max_{0\leq i\leq d}\{\deg h_i-i\}. $$ Then we have \[ \frac{\displaystyle N-M}{d}\leq L(s_n,N)\leq \frac{\displaystyle (d-1)N+M+1}{d}. \] \end{theorem} See also \cite{XingLam} for the special case $d=2$. The idea of the proof of Theorem \ref{thm:MW-lin-compl-gen} is that small $N$th linear complexity profile gives a good rational approximation to the generating function. However, transcendental elements over $\mathbb{F}_q(x)$ are not well-approximated. Namely, since $(s_n)$ is not ultimately periodic, $G(x)=\sum_{n=0}^\infty s_nx^n\not\in \mathbb{F}_q(x)$ is not rational by \cite[Lemma~1]{Niederreiter88-b}. Let $g(x)/f(x)\in\mathbb{F}_q(x)$ be a rational zero of $h(x,y)$ modulo $x^N$ with $\deg(f)\le L(s_n,N)$ and $\deg(g)<L(s_n,N)$. More precisely, put $L=L(s_n,N)$. Then we have $$ \sum_{\ell=0}^Lc_{\ell}s_{n+\ell}=0 \quad \mbox{for }0\le n\le N-L-1 $$ for some $c_0,\ldots,c_{L}\in \mathbb{F}_p$ with $c_L=-1$. Take \[ f(x)=\sum_{\ell=0}^L c_{\ell}x^{L-\ell} \] and \[ g(x)=\sum_{m=0}^{L-1}\left(\sum_{\ell=L-m}^Lc_\ell s_{m+\ell-L}\right)x^m \] and verify $$ f(x)G(x)\equiv g(x)\bmod x^N. $$ Then \[ h_0(x)f^d(x)+h_1(x)g(x)f^{d-1}(x)+\dots + h_d(x)g(x)^d=K(x) x^N. \] Here $K(x)\neq 0$ since $h(x,y)$ has no rational zero. Comparing the degrees of both sides we get \[ dL+M\geq N \] which gives the lower bound. The upper bound for $N=1$ is trivial. For $N\geq 2$ the result follows from the well-known bound, see for example \cite[Lemma 3]{DorferWinterhof}, \[ L(s_n,N)\leq \max\left\{L(s_n,N-1), N-L(s_n,N-1) \right\} \] by induction. The bound in Theorem \ref{thm:MW-lin-compl-gen} combined with \eqref{eq:TH-equation}-\eqref{aph} gives the following estimates for the $N$th linear complexity of the Thue-Morse sequence $(t_n)$ defined by~\eqref{tmdef} \begin{equation}\label{eq:lin-compl_TM} \left\lceil \frac{N-1}{2} \right\rceil\leq L(t_n,N)\leq \left\lfloor \frac{N}{2} \right\rfloor+1, \end{equation} of the Rudin-Shapiro sequence $(r_n)$ defined by \eqref{rsdef} and the regular paper-folding sequence $(v_n)$ defined by \eqref{pfdef} \begin{equation}\label{eq:lin-compl_RS} \left\lceil\frac{N-3}{2}\right\rceil\leq L(r_n,N), L(v_n,N)\leq \left\lfloor\frac{N}{2}\right\rfloor+2, \end{equation} of the $p$-ary pattern sequence $(p_n)$ defined by \eqref{eq:recurrence} with any pattern $P$ of length $\ell$ $$ \left\lceil\frac{N+1}{p}\right\rceil-p^{\ell-1}\leq L(p_n,N)\leq \left\lfloor\frac{(p-1)N}{p}\right\rfloor+p^{\ell-1}, $$ of the Rudin-Shapiro-like sequence $(\ell_n)$ defined by \eqref{eq:rslike} $$ \left\lceil \frac{N}{4}\right\rceil-1\le L(\ell_n,N)\le \left\lfloor\frac{3N+5}{4}\right\rfloor, $$ of the Baum-Sweet sequence $(b_n)$ defined by \eqref{bsdef} $$ \left\lceil\frac{N}{3}\right\rceil\leq L(b_n,N)\leq \left\lfloor\frac{2N+1}{3}\right\rfloor, $$ of the characteristic sequence $(c_n)$ of sums of three squares defined by \eqref{eq:cndef} $$ \left\lceil\frac{N-7}{4}\right\rceil\le L(c_n,N)\le \left\lfloor\frac{3N}{4}\right\rfloor+2 $$ and of the apwenian sequence $(w_n)$ defined by \eqref{apdef} \begin{equation}\label{aplin} L(w_n,N)=\left\lfloor \frac{N+1}{2}\right\rfloor. \end{equation} Note that the bound \eqref{eq:lin-compl_TM} is also true for the dual $(t'_n)$ of the Thue-Morse sequence, that is, $t'_n=1-t_n$, and apwenian sequences are characterized by the property \eqref{aplin}, see \cite{alhani}. Note that not all apwenian sequences are automatic. The bounds \eqref{eq:lin-compl_TM} for the Thue-Morse sequence and \eqref{eq:lin-compl_RS} for the Rudin-Shapiro sequence are optimal. Using the continued fraction expansions of their generating functions, M\'erai and Winterhof \cite{mewi18} determined the exact value of the $N$th linear complexity profiles of the Thue-Morse and Rudin Shapiro sequence. \begin{theorem} The $N$th linear complexity of the Thue-Morse sequence is $$ L(t_n,N)=2\left \lfloor \frac{N+2}{4}\right\rfloor, \quad N=1,2,\dots $$ and the $N$th linear complexity of the Rudin-Shapiro sequence is $$L(r_n,N)=\left\{\begin{array}{cc} 6\left\lfloor N/12\right\rfloor+4, & N\equiv 4,5,6,7,8,9\bmod 12,\\ 6\left\lfloor (N+2)/12\right\rfloor, & \mbox{otherwise}. \end{array}\right.$$ \end{theorem} The result can be extended to binary pattern sequences $(p_n)$ defined by \eqref{eq:recurrence} with the all one pattern of length $\ell\ge 3$, that is, $a=2^\ell-1$. It follows from Theorem~\ref{thm:MW-lin-compl-gen}, that if an automatic sequence is not ultimately periodic and its generating function has a quadratic minimal polynomial, that is $d=2$ in Theorem~\ref{thm:MW-lin-compl-gen}, then the deviation of the $N$th linear complexity from its expected value $N/2$ is bounded by $(M+1)/2$, $$ \left|L(s_n,N)-\frac{N}{2}\right|\leq \frac{M+1}{2}. $$ Such sequences are said to have \emph{almost perfect} or \emph{$(M+1)$-perfect} linear complexity profile, see \cite{Niederreiter88-b,alhani}. Apwenian sequences are those sequences having $1$-perfect or just {\em perfect} linear complexity profile. The bounds \eqref{eq:lin-compl_TM} and \eqref{eq:lin-compl_RS} imply that the Thue-Morse sequence has $2$-perfect linear complexity profile and the Rudin-Shapiro sequence and the paper-folding sequence both have $4$-perfect linear complexity profile. Although automatic sequences have some good pseudorandom properties including a desirable linear complexity profile, these sequences have also some strong non-randomness properties, see Sections~\ref{sec:correlation}, \ref{sec:expansion_complexity} and \ref{sec:normality} below. Such randomness flaws may be avoided considering subsequences of automatic sequences. For example, the Thue-Morse and Rudin-Shapiro sequences along squares are not automatic, see Section~\ref{sec:normality} below, and seem to have $N$th linear complexity $N/2+O(\log N)$, see Figure~\ref{fig:L}. \begin{figure}[ht] \begin{center} \includegraphics[scale=.51]{TM_along_squares_L.png} \includegraphics[scale=.51]{RS_along_squares_L.png} \end{center} \caption{The $N$th linear complexity of the Thue-Morse (left) and Rudin-Shapiro (right) sequence along squares.} \label{fig:L} \end{figure} \begin{problem Prove that the $N$th linear complexities of the Thue-Morse and Rudin-Shapiro sequences along squares satisfy \footnote{$f(k)=o(g(k))$ is equivalent to $f(k)/ g(k)\rightarrow 0$ as $k\rightarrow \infty$.} $$ L(t_{n^2},N)=\frac{N}{2}+o(N)\quad \mbox{and}\quad L(r_{n^2},N)=\frac{N}{2}+o(N). $$ \end{problem} We remark, that lower bounds on the $N$th linear complexities of $(t_{n^2})$ and $(r_{n^2})$ of order of magnitude $\sqrt{N}$ follow from Theorem~\ref{thm:suwi} and \eqref{eq:ML} in the next section. Additional to these examples, the same problem is also open for other subsequences such as along other polynomial values, along primes etc. \section{Maximum order complexity}\label{sec:max-order_complexity} Maximum order (or nonlinear) complexity is a refinement of the linear complexity considering not only linear but \emph{any} recurrence relation. \begin{definition} The {\em $N$th maximum order complexity} $M(s_n,N)$ is the smallest positive integer~$M$ with $$ s_{n+M}=f(s_{n+M-1},\ldots,s_n),\quad 0\le n\le N-M-1, $$ for some mapping $f:\mathbb{F}_2^M \rightarrow \mathbb{F}_2$. The sequence $(M(s_n,N))_{N=1}^\infty$ is called {\em maximum order complexity profile}. \end{definition} Obviously, we have \begin{equation}\label{eq:ML} M(s_n,N)\le L(s_n,N) \end{equation} and the maximum order complexity is a finer measure for the unpredictability of a sequence than the linear complexity. However, often the linear complexity is easier to analyze both theoretically and algorithmically. Clearly, a sufficiently large maximum order complexity is needed for unpredictability and suitability in cryptography. However, sequences of very large maximum order complexity have also a very large autocorrelation or correlation measure of order $2$, see \eqref{C2M} below, and are not suitable for many applications including cryptography, radar, sonar and wireless communications. The maximum order complexity was introduced by Jansen in \cite[Chapter 3]{ja}, see also~\cite{jabo}. The typical value for the $N$th maximum order complexity is of order of magnitude $\log N$, see \cite{ja,jabo}. An algorithm for calculating the maximum order complexity profile of linear time and memory was presented by Jansen \cite{ja,jabo} using the graph algorithm introduced by Blumer et al.\ \cite{bl}. The maximum order complexity of the Thue-Morse sequence was determined in \cite[Theorem 1]{suwi19}. \begin{theorem}\label{thm:suwi-tm} For $N\ge 4$, the $N$th maximum order complexity of the Thue-Morse sequence~$(t_n)$ satisfies $$M(t_n,N)=2^\ell+1,$$ where $$\ell=\left\lceil \frac{\log (N/5)}{\log 2}\right\rceil.$$ \end{theorem} It is easy to see that \begin{equation}\label{tmmoc}\frac{N}{5}+1\le M(t_n,N)\le 2\frac{N-1}{5}+1 \quad\text{for}\;\; N\ge 4. \end{equation} In Section~\ref{sec:correlation} we will see that such a large maximum order complexity points to undesirable structure in a sequence. The $N$th maximum order complexity of the Rudin-Shapiro sequence and some generalizations is also of order of magnitude $N$, see \cite[Theorem 2]{suwi19}. In particular we have \begin{equation}\label{rsmoc} M(r_n,N)\ge \frac{N}{6}+1,\quad N\ge 4. \end{equation} The maximum order complexity of the subsequences of the Thue-Morse and the Rudin-Shapiro sequence along squares are still large enough, see \cite{suwi}. \begin{theorem}\label{thm:suwi} The $N$th maximum order complexities $M(t_{n^2},N)$ and $M(r_{n^2},N)$ of the subsequences $(t_{n^2})$ and $(r_{n^2})$ of the Thue-Morse and the Rudin-Shapiro sequence along squares satisfy \begin{align*} &M(t_{n^2},N)\geq \sqrt{\frac{2N}{5}},\quad N\ge 21, \quad \mbox{and} \\ &M(r_{n^2},N)\geq \sqrt{\frac{N}{8}},\quad N\ge 64. \end{align*} \end{theorem} We sketch the proof. First, let $t$ be the length of the longest subsequence of~$(t_{n^2})$ that occurs at least twice with different successors among the first $N$ sequence elements. Then $M(t_{n^2},N)\ge t + 1$. Hence the first inequality follows from $$ t_{(i+2^{\ell+1})^2}=t_{(i+2^{\ell+2})^2},\quad i=0,1,\ldots,\left\lfloor \sqrt{2^{\ell+2}-1}\right\rfloor$$ $$\mbox{and}\quad t_{(2^{\ell}+2^{\ell+1})^2}\ne t_{(2^\ell+2^{\ell+2})^2}, $$ which can be shown by induction over $\ell\ge 2$, where $\ell$ is defined by $5\cdot 2^\ell<N\le 5\cdot 2^{\ell+1}$. The second bound follows from $$r_{(i+2^{\ell+3})^2}=r_{(i+2^{\ell+4})^2},\quad i=0,1,\ldots,\left\lfloor\sqrt{2^{\ell+3}-1}\right\rfloor,$$ $$\mbox{and}\quad r_{(2^{\ell+2}+2^{\ell+3})^2}\ne r_{(2^{\ell+2}+2^{\ell+4})^2},$$ where $\ell$ is defined by $2^{\ell+5}\le N<2^{\ell+6}$. Figure \ref{maxordersquare} suggests that $\sqrt{N}$ is the right order of magnitude for the $N$th maximum order complexities of $(t_{n^2})$ and $(r_{n^2})$. For $N\ge 2^{2\ell+2}$ the same lower bound $\sqrt{N/8}$ is true for binary pattern sequences along squares with the all one pattern of length $\ell$, that is, $a=2^\ell-1$ for~$\ell\ge 3$, see \cite{suwi}. \begin{figure}[ht] \begin{center} \includegraphics[scale=.5]{TM_along_squares_M.png} \includegraphics[scale=.5]{RS_along_squares_M.png} \end{center} \caption{The $N$th maximum order complexity of the Thue-Morse (left) and Rudin-Shapiro (right) sequence along squares.} \label{maxordersquare} \end{figure} This result was extended by Popoli \cite{po} to sequences along polynomial values of higher degrees $d$. However, the lower bounds are of order of magnitude $N^{1/d}$. Note that no better lower bounds are known for the $N$th linear complexity of these subsequences of automatic sequences. The problem for other subsequences is still open as for example for subsequences along primes. \begin{problem} Study the maximum-order complexity of the subsequences of the Thue-Morse and the Rudin-Shapiro sequence along primes. \end{problem} The maximum order complexity of other automatic sequences has also been studied. Sun, Zeng and Lin \cite{suzeli} showed that the $N$th maximum order complexity of the Rudin-Shapiro-like sequence $(\ell_n)$ defined by \eqref{eq:rslike} is of order of magnitude~$N$. \bigskip We remark, that in addition to automatic sequences based on the $k$-ary expansion~\eqref{eq:k-ary} of integers, one can consider analogously sequences using other numeration systems. In particular, consider the \emph{Fibonacci numbers} defined by $$ F_0=0,~F_1=1\quad \mbox{and}\quad F_n=F_{n-1}+F_{n-2}\mbox{ for }n\ge 2. $$ Then the unique, see for example \cite[Theorem~3.8.1]{alsh2}, {\em Zeckendorf expansion} or {\em Fibonacci expansion}, of a positive integer~$n$ is $$ n=\sum_{i=0}^\infty e_i F_{i+2},\quad \mbox{where }e_i\in \{0,1\} \mbox{ and } e_ie_{i+1}=0\mbox{ for }i=0,1,\ldots $$ Analogously to the Thue-Morse sum-of-digits sequence $(t_n)$ and the Rudin-Shapiro sequence~$(r_n)$ which can be defined by~\eqref{sumofdigitsdef} we can define and study the {\em Zeckendorf sum-of-digits sequences modulo $2$} $(z_n)$ and $(u_n)$ defined by \begin{equation}\label{zeck} z_n=\sum_{i=0}^\infty e_i \bmod 2 \quad\mbox{and}\quad u_n=\sum_{i=0}^\infty e_ie_{i+2}\bmod 2. \end{equation} Very recently the maximum-order complexity of $(z_n)$ and its subsequences along polynomial values has been studied by Jamet, Popoli and Stoll in \cite{japost}. A lower bound on $M(u_n,N)$ and some generalizations can be obtained along the same lines and will be contained in Popoli's thesis. \section{Well-distribution and correlation measures}\label{sec:correlation} Mauduit and S\'ark\"ozy \cite{masa} introduced two measures of pseudorandomness for finite sequences over $\{-1,+1\}$, the well-distribution measure and the correlation measure of order $k$. We adjust these definitions to infinite binary sequences $(s_n)$ over $\mathbb{F}_2$. \begin{definition} The {\em $N$th well-distribution measure} of $(s_n)$ is defined as $$W(s_n,N)=\max_{a,b,t}\left|\sum_{j=0}^{t-1} (-1)^{s_{a+jb}}\right|,$$ where the maximum is taken over all integers $a,b,t$, $b\ge 1$, such that $0\le a \le a+(t-1)b\le N-1$. \end{definition} The well-distribution measure provides information on the balance, \footnote{Note that the term {\em balanced} is used with a different meaning in combinatorics on words, see for example \cite[Definition~10.5.4]{alsh2}.} that is the distribution of zeros and ones, along arithmetic progressions. For random sequences it is expected to be small. More precisely, Alon et al.\ \cite[Theorem~1]{alko} proved the following result on the typical value of the well-distribution measure. \begin{theorem} For all $\varepsilon>0$, there are numbers $N_0=N_0(\varepsilon)$ and $\delta=\delta(\varepsilon)>0$ such that for $N\ge N_0$ we have $$\delta \sqrt{N}<W(s_n,N) <\frac{\sqrt{N}}{\delta}$$ with probability at least $1-\varepsilon$ with respect to the probability measure \eqref{eq:prob_space}. \end{theorem} Moreover, Aistleitner \cite{aistleitner} showed that there exists a continuous limit distribution of~$\frac{W(s_n,N)}{\sqrt{N}}$. More precisely, for any $t\in \mathbb{R}$ the limit $$ F(t)=\lim_{N\rightarrow \infty} {\mathbb P}\left(\frac{W(s_n,N)}{\sqrt{N}}\le t\right) $$ exists and satisfies $$ \lim_{t\rightarrow \infty}t(1-F(t))e^{t^2/2}=\frac{8}{\sqrt{2\pi}}, $$ with respect to the probability measure~\eqref{eq:prob_space}. \begin{definition} For $k\geq 1$, the {\em $N$th correlation measure of order~$k$} of a binary sequence~$(s_n)$ is $$C_k(s_n,N)=\max_{M,D}\left|\sum^{M-1}_{n=0}(-1)^{s_{n+d_1}}\cdots (-1)^{s_{n+d_k}}\right|,$$ where the maximum is taken over all $D=(d_1,d_2,\ldots,d_k)$ with integers satisfying $0\le d_1<d_2<\cdots<d_k$ and $1\le M\le N-d_k$. \end{definition} The correlation measure of order $k$ provides information about the similarity of parts of the sequence and their shifts. For a random sequence this similarity and thus the correlation measure of order $k$ is expected to be small. More precisely, Alon et al.\ \cite[Theorem 2]{alko} proved the following result on the typical value of the correlation measure of order $k$. \begin{theorem} For any $\varepsilon>0$, there exist an $N_0=N_0(\varepsilon)$ such that for all $N\ge N_0$ we have for a randomly chosen sequence $(s_n)$ and any $k$ with $2\le k\le N/4$, $$ \frac{2}{5}\sqrt{N\log{N\choose k}}<C_k(s_n,N)<\frac{7}{4}\sqrt{N\log{N\choose k}} $$ with probability at least $1-\varepsilon$ with respect to the probability measure \eqref{eq:prob_space}. \end{theorem} Moreover, Schmidt \cite[Theorem 1.1]{schmidt} showed, that for fixed $k$, we have $$ \lim_{N\rightarrow \infty}\frac{C_k(s_n,N)}{\sqrt{2N\log \binom{N}{k-1}}} =1 $$ with probability $1$ with respect to the probability measure \eqref{eq:prob_space}. A large well-distribution measure implies a large correlation measure of order $2$. More precisely we have by \cite[Theorem 1]{ms03} \footnote{$f(k)=O(g(k))$ is equivalent to $|f(k)|\le c g(k)$ for some constant $c>0$.} $$ W(s_n,N)=O\left(\sqrt{NC_2(s_n,N)}\right). $$ Mauduit and Sárközy \cite{masa98} obtained bounds on the well-distribution measure and correlation measure of order $2$ of Thue-Morse sequence $(t_n)$ and Rudin-Shapiro sequence~$(r_n)$. For example, as a consequence of the bound \begin{equation}\label{eq:exp-TM} \left|\sum_{n=0}^{N-1}(-1)^{t_n}z^n\right|\leq (1+\sqrt{3}) N^{\log 3/\log 4},\quad |z|=1, \end{equation} of Gel'fond~\cite[p.~262]{gelfond}, see \cite{foma} for the explicit constant $1+\sqrt{3}$, they obtained a bound on~$W(t_n,N)$. \begin{theorem}\label{thm:masa-thm1} We have $$ W(t_n,N) \leq 2(1+\sqrt{3}) N^{\log 3/\log 4}. $$ \end{theorem} Also, using the bound \begin{equation}\label{eq:exp-RS} \left|\sum_{n=0}^{N-1}(-1)^{r_n}z^n\right|\leq (2+\sqrt{2}) N^{1/2},\quad |z|=1, \end{equation} obtained by Rudin~\cite{rudin} and Shapiro~\cite{shapiro}, see also \cite[Theorem 3.3.2]{alsh2}, they proved a bound on~$W(r_n,N)$. \begin{theorem}\label{thm:masa-thm2} We have $$ W(r_n,N)\leq 2(2+\sqrt{2}) N^{1/2}. $$ \end{theorem} In general, following the proofs of \cite{masa98} we get \begin{equation}\label{eq:W-idea} W(s_n,N) =O\left( \sup_{|z|=1, m\le N}\left|\sum_{n=0}^{m-1} (-1)^{s_n}z^n\right| \right) \end{equation} and thus Theorems \ref{thm:masa-thm1} and \ref{thm:masa-thm2} follow, up to the constant, from \eqref{eq:exp-TM} and \eqref{eq:exp-RS}. However, for $(t_n)$ and $(r_n)$ Mauduit and Sárközy \cite{masa98} detected non-randomness properties by showing that the correlation measure of order $2$ of these sequences is large. \begin{theorem} We have \begin{equation}\label{ctm} C_2(t_n,N)>\frac{N}{12},\quad N\ge 5, \end{equation} and \begin{equation}\label{crs} C_2(r_n,N)>\frac{N}{6},\quad N\ge 4. \end{equation} \end{theorem} Mérai and Winterhof~\cite{mewi2} showed that all automatic sequences share the property of having a large correlation measure of order $2$. They provided the following lower bound in terms of the defining automaton. \begin{theorem}\label{thm:gen_lowe_bound_on_C} Let $(s_n)$ be a $k$-automatic binary sequence generated by the finite automaton $(Q,\Sigma,\delta,q_0,\varphi,\{0,1\})$. Then \[ C_2(s_n,N)\geq \frac{N}{k(|Q|+1)} \quad \text{for } N\geq k(|Q|+1). \] \end{theorem} This result applied to $(t_n)$ and $(r_n)$ gives the following bounds $$ C_2(t_n,N)\geq \frac{N}{6}, \quad N\geq 6, \quad\text{and}\quad C_2(r_n,N)\geq \frac{N}{10}, \quad N\geq 10, $$ which improves \eqref{ctm}. Figures~\ref{fig:well} and \ref{fig:cor} may lead to the conjecture that well-distribution measure and correlation measure of order $2$ of both $(t_{n^2})$ and $(r_{n^2})$ are of order of magnitude $N^{1/2}$ and $(N\log N)^{1/2}$, respectively. \begin{figure}[ht] \begin{center} \includegraphics[scale=.51]{TM_along_squares_W.png} \includegraphics[scale=.51]{RS_along_squares_W.png} \end{center} \caption{The $N$th well-distribution measure of the Thue-Morse (left) and Rudin-Shapiro (right) sequence along squares.} \label{fig:well} \end{figure} \begin{figure}[ht] \begin{center} \includegraphics[scale=.51]{TM_along_squares_C2.png} \includegraphics[scale=.51]{RS_along_squares_C2.png} \end{center} \caption{The $N$th second order correlation measure of the Thue-Morse (left) and Rudin-Shapiro (right) sequence along squares.} \label{fig:cor} \end{figure} \begin{problem} For fixed $k=2,3,\ldots$ show that $$ C_k(t_{n^2},N)=o(N)\quad \mbox{and}\quad C_k(r_{n^2},N)=o(N). $$ \end{problem} Mauduit and Rivat \cite{mari18} showed that $$ \left|\sum_{n=0}^{N-1}(-1)^{r_{n^2}}z^n\right|= O\left( N^{1-\eta}\right),\quad |z|=1,\quad \mbox{for some $\eta>0$}, $$ which, together with \eqref{eq:W-idea}, gives a bound on $W(r_{n^2},N)$ of the same order of magnitude. More precisely, \cite{mari18} deals with the more general case of binary pattern sequences $(p_n)$ defined by \eqref{eq:pattern-equation} with either the all one pattern of length $k\ge 2$, that is, $a=2^k-1$, or the patterns $10\ldots 01$ of length $k\ge 3$, that is, $a=2^{k-1}+1$, and the constants depend on $k$. For the Thue-Morse sequence along squares $(t_{n^2})$ one can easily derive a nontrivial bound on $$ \left|\sum_{n=0}^{N-1}(-1)^{t_{n^2}}z^n\right|,\quad |z|=1, $$ and thus on $W(t_{n^2},N)$ since the proof of \cite[Th\'eor\`eme 1]{mari1} for $z=1$ works also for $z\neq 1$ since after applying a variant of the van der Corput inequality, \cite[Lemma 15]{mari1}, we get an expression which does not depend on the variable $z$ anymore, that is, the same expression as for $z=1$. Theorem~\ref{thm:suwi-tm} in Section~\ref{sec:max-order_complexity} above shows that the Thue-Morse sequence has maximum order complexity $M(t_n,N)$ of order of magnitude $N$. Although a large maximum order complexity is desired it should be not too large since otherwise the correlation measure of order $2$ is large. Namely, we have \begin{equation}\label{C2M} C_2(s_n,N)\ge M(s_n,N)-1 \end{equation} since by \cite[Proposition 3.1]{ja} there exist $0\le n_1<n_2\le N-M(s_n,N)-1$ with $$s_{n_1+i}=s_{n_2+i},\quad i=0,\ldots,M(s_n,N)-2,\quad \mbox{but }s_{n_1+M(s_n,N)-1}\ne s_{n_2+M(s_n,N)-1}$$ and thus $$M(s_n,N)-1=\sum_{i=0}^{M(s_n,N)-2}(-1)^{s_{n_1+i}+s_{n_2+i}}\le C_2(s_n,N).$$ Combining \eqref{tmmoc} and \eqref{C2M} we get for the Thue-Morse sequence $$ C_2(t_n,N)\ge \frac{N}{5},\quad N\ge 4, $$ which further improves the constant in \eqref{ctm}. Combining \eqref{rsmoc} and \eqref{C2M} recovers~\eqref{crs}. The correlation measure of order $2$ with bounded lags of some generalizations of the Rudin-Shapiro sequence has recently been studied in \cite{mastta}. In contrast to the Thue-Morse und Rudin-Shapiro sequence, the well-distribution measure of some other binary automatic sequences is very large. For example, the Baum-Sweet sequence $(b_n)$, the characteristic sequence $(c_n)$ of the sums of three squares, the paper-folding sequence $(v_n)$ and the apwenian sequence $(w_n)$ defined by \eqref{apdef} are very unbalanced and thus have all well-distribution measure of order of magnitude $N$. However, it seems to be interesting to study the well-distribution measure for arbitrary apwenian sequences. For the Rudin-Shapiro like sequence $(\ell_n)$ defined by \eqref{eq:rslike} Lafrance, Rampersad and Yee \cite{laraye} proved $$\liminf_{N\rightarrow \infty} \frac{\sum_{n=0}^{N-1}(-1)^{\ell_n}}{\sqrt{N}}=\frac{\sqrt{3}}{3} \quad \mbox{and}\quad \limsup_{N\rightarrow \infty} \frac{\sum_{n=0}^{N-1}(-1)^{\ell_n}}{\sqrt{N}}=\sqrt{2}. $$ However, a bound on $W(\ell_n,N)$ is not known and in contrast to \eqref{eq:exp-RS} for the Rudin-Shapiro sequence~$(r_n)$, for $(\ell_n)$ the absolute values $$\left|\sum_{n=0}^{N-1}(-1)^{\ell_n} z^n\right|$$ can be of much larger order of magnitude than $\sqrt{N}$ for some $z$ with $|z|=1$, see \cite[Theorem~2]{al16} as well as \cite{chgr}. Finally, we remark that the result of Theorem~\ref{thm:gen_lowe_bound_on_C} provides an estimate on the \emph{state complexity} of sequences in terms of the correlation measure of order $2$. \begin{definition} Let $k\geq 2$. Then the \emph{$N$th state complexity} $SC_k(s_n,N)$ of a sequence~$(s_n)$ over $\mathbb{F}_2$ is the minimum of the number of states of any finite $k$-automaton which generates the first $N$ sequence elements. \end{definition} \begin{cor} Let $(s_n)$ be a binary sequence. Then for all $k\geq 2$ we have \[ SC_k(s_n,N)\geq\frac{N}{k\cdot C_2(s_n,N)}-1 \quad \text{for } N\geq 3. \] \end{cor} \section{Expansion complexity}\label{sec:expansion_complexity} Theorem~\ref{thm:MW-lin-compl-gen} indicates that automatic sequences possess good properties in terms of the linear complexity profile. However, the results of Section~\ref{sec:correlation} show that these sequences have a serious lack of pseudorandomness. Diem~\cite{di} showed that these sequences are not just statistically auto-correlated, but are completely predictable from a relatively short initial segment. He introduced the notion of \emph{expansion complexity} to turn such security flaw into a quantitative form. \begin{definition} Let $(s_n)$ be a sequence over $\mathbb{F}_q$ with generating function $$ G(x)=\sum_{n=0}^{\infty}s_nx^n \in \mathbb{F}_q \llbracket x \rrbracket . $$ For a positive integer $N$, the \emph{$N$th expansion complexity $E(s_n,N)$ of $(s_n)$} is $E(s_n,N)=0$ if $s_0=\dots=s_{N-1}=0$ and otherwise the least total degree of a non-zero polynomial $h(x,y)\in\mathbb{F}_q[x,y]$ such that \begin{equation}\label{eq:h} h(x,G(x))\equiv 0 \bmod x^N. \end{equation} The sequence $(E(s_n,N))_{N=1}^\infty$ is called \emph{expansion complexity profile of $(s_n)$} and $$ E(s_n)=\sup_{N\geq 1} E(s_n,N) $$ is the \emph{expansion complexity of $(s_n)$}. \end{definition} By Christol's Theorem~\ref{thm:christol}, a sequence is automatic if and only if its expansion complexity is finite. For example, we have for the Thue-Morse sequence $(t_n)$, the Rudin-Shapiro sequence $(r_n)$, the $p$-ary pattern sequence $(p_n)$, the Baum-Sweet sequence $(b_n)$, the Rudin-Shapiro like sequence $(\ell_n)$ and the characteristic sequence $(c_n)$ of sums of three squares that $$ E(t_n)= 5, \quad E(r_n)= 7, \quad E(p_n)\le p^\ell+2p-1, \quad E(b_n)= 3, \quad E(\ell_n)\le 12 $$ $$ E(c_n)\le 12, \quad E(v_n)= 6 \quad \mbox{and}\quad E(w_n)= 4, $$ which follows from \eqref{eq:TH-equation}, \eqref{eq:RS-equation}, \eqref{eq:pattern-equation}, \eqref{eq:rslh}, \eqref{eq:BS-equation}, \eqref{eq:cnh}, \eqref{pfh} and \eqref{aph}. The equalities follow from the fact that there is no lower degree polynomial with such property since~$h(x,y)$ is irreducible in these cases, see \cite[Proposition~4]{di}. Diem showed \cite{di} that if a sequence has small expansion complexity, then long parts of such sequences can be computed efficiently from short ones. We summarize his results. \begin{theorem}\label{thm:Diem} Let $(s_n)$ be a sequence over $\mathbb{F}_q$ with expansion complexity $E(s_n)=d$. From the first $d^2$ elements, one can compute an irreducible polynomial $h(x,y)\in \mathbb{F}_q[x,y]$ of degree $\deg h \leq d$ with $h(x,G(x))=0$ in polynomial time in $d\cdot \log q$. Moreover, an initial segment of the sequence of length $M>N$ can be determined from $h$ and the $d^2$ initial values in polynomial time in $d \cdot \log q$ and in linear time in~$M$. \end{theorem} Theorem~\ref{thm:Diem} shows that automatic sequences have a strong non-randomness property. The expansion complexity profile is defined to capture such non-randomness property locally, that is for initial segments of sequences. For the $N$th expansion complexity, we have the trivial bound $E(s_n,N)\leq N-1$ realized by the polynomial $$ h(x,y)=y-\sum_{n=0}^{N-1}s_nx^n. $$ Moreover, one can show the stronger upper bound \begin{equation}\label{eq:exp_upper} \binom{E(s_n,N)+1}{2}\leq N, \end{equation} which holds for all sequence $(s_n)$ and all $N\geq 1$, see \cite[Theorem~1]{GoMeNi}. The $N$th expansion complexity of random sequences is concentrated to its upper bound \eqref{eq:exp_upper}, see \cite[Theorem~2]{GoMe}. \begin{theorem} We have $$\liminf_{N\rightarrow \infty} \frac{E(s_n,N)}{\sqrt{N}} \geq \frac{\sqrt{2}}{2}, $$ with probability one with respect to the probability measure~\eqref{eq:prob_space}. \end{theorem} One can estimate the $N$th expansion complexity $E(s_n,N)$ in terms of the $N$th linear complexity $L(s_n,N)$, see \cite[Theorem~3]{meniwi}. \begin{theorem}\label{thm:MeWiNi} Let $(s_n)$ be a sequence over $\mathbb{F}_q$ and let $G(x)$ bet its generating function. For $N\geq 2$, assume, that $$ G(x)\not\equiv 0\bmod x^N. $$ Let $L_N=L(s_n,N)$ be the $N$th linear complexity and let $$ \sum_{\ell=t_N}^{L_N} c_\ell s_{i+\ell}=0,\quad 0\le i\le N-L_N-1, $$ be a shortest linear recurrence for the first $N$ terms of $(s_n)$, where $c_{L_N}=1$ and $c_{t_N}\ne 0$. Then $$ E(s_n,N)\ge \left\{\begin{array}{ll} L_N-t_N+1 & \mbox{for } N>(L_N-t_N)(L_N-\min\{1,t_N-1\}),\\ \left\lceil \frac{N}{L_N-\min\{1,t_N-1\}}\right\rceil & \mbox{otherwise,} \end{array}\right. $$ and $$ E(s_n, N)\le \min\{L_N+\max\{-1,-t_N+1\},N-L_N+2\}. $$ \end{theorem} The result formulates in a qualitative way that \emph{very} large $N$th linear complexity, that is $N$th linear complexity close to $N$, is a non-randomness property. Moreover, it enables us to estimate the $N$th expansion complexity from below if the $N$th linear complexity is not too close to either $0$ or $N$ (in a logarithmic scale), say, of order of magnitude $\sqrt{N}$. We refer to \cite{meniwi, howi} for applications of Theorem~\ref{thm:MeWiNi} for estimating the $N$th expansion complexity of certain sequences. Certain subsequences of automatic sequences, say, the Thue-Morse and Rudin-Shapiro sequences along squares are not automatic, see Section~\ref{sec:normality} below, and thus have unbounded expansion complexity profile. However, their growth rates are not known. For example, one can study further the Thue-Morse and Rudin-Shapiro sequence along squares. \begin{problem} Estimate the expansion complexity profiles of the subsequences $(t_{n^2})$ and $(r_{n^2})$ of the Thue-Morse and Rudin-Shapiro sequence along squares. \end{problem} Figure~\ref{fig:exp} suggests $E(t_{n^2},N)$ and $E(r_{n^2},N)$ are both of order of magnitude $\sqrt{N}$. \begin{figure}[ht] \begin{center} \includegraphics[scale=.51]{TM_along_squares_E.png} \includegraphics[scale=.51]{RS_along_squares_E.png} \end{center} \caption{The $N$th expansion complexity of the Thue-Morse (left) and Rudin-Shapiro (right) sequence along squares.} \label{fig:exp} \end{figure} Finally, we remark that in order to use the full strength of Theorem~\ref{thm:Diem} for inferring sequence elements, one needs to require the irreducibly of the polynomial $h(x,y)$ in \eqref{eq:h}. In \cite{GoMe, GoMeNi}, the authors studied this variant of the $N$th expansion complexity and the relation between these two complexity measures. \section{Subword complexity and normality}\label{sec:normality} The results of Section~\ref{sec:correlation} show that many automatic sequences, including Thue-Morse and Rudin-Shapiro sequence, are balanced, that is, the frequencies of the symbols are close to the expected values. However, the frequencies of longer patterns are far from uniform. This phenomenon can be made precise by the notion of \emph{subword complexity}. \begin{definition} For a sequence $(s_n)$ over the alphabet $\Delta$ the \emph{subword complexity} $p(s_n,k)$ is the number of distinct subsequences of length $k$. \end{definition} Trivially we have $1\le p(s_n,k)\le |\Delta|^k$ and for ultimately periodic sequences we have $p(s_n,k)=O(1)$. By \cite[Corollary 10.3.2]{alsh2} the subword complexity $p(s_n,k)$ of automatic sequences $(s_n)$ is of order of magnitude $k$. \begin{theorem}\label{thm:subword} If $(s_n)$ is an automatic sequence that is not ultimately periodic, then we have \footnote{$f(k)=\Theta(g(k))$ is equivalent to $c_1g(k)\le f(k)\le c_2g(k)$ for some constants $c_2\ge c_1>0$.} $$p(s_n,k)=\Theta(k).$$ \end{theorem} For the Thue-Morse sequence $(t_n)$, the exact value of its subword complexity $p(t_n,k)$ was independently determined by Brlek \cite[Proposition~4.4]{br} and by de Luca and Varricchio \cite[Proposition~4.4]{deva}, see also \cite[Exercise~10.11.10]{alsh2}. De Luca and Varricchio \cite[Property~3.3]{deva} also showed that patterns such as $000$ and $111$ do not appear in the Thue-Morse sequence and more general the following result. \begin{theorem} The Thue-Morse sequence is cube-free, that is, no pattern of the form $www$ with $w\in \{0,1\}^k$ for some $k\ge 1$ appears in the sequence. \end{theorem} The papers \cite{br,deva} contain also several other results on the non-existence of certain patterns in the Thue-Morse sequence. The subword complexity and the correlation measure of order $\ell$ are related by the following result of Cassaigne et al. \cite[Theorem 6]{cafe}. \begin{theorem} If for some positive integers $k$ and $N$ $$ C_{\ell}(s_n,N)\le \frac{N}{2^{2k+1}},\quad \ell=1,2,\ldots,k, $$ then $$p(s_n,k)=2^k.$$ \end{theorem} For automatic sequences we can have $p(s_n,k)=2^k$ only for finitely many $k$ since $p(s_n,k)=\Theta(k)$. However, certain subsequences of automatic squences are \emph{normal}, that is, all patterns appear in the sequence with the expected frequencies. More formally, a sequence $(s_n)$ is called \emph{normal} if for any fixed length $k$ and any pattern $\mathbf{e}\in \Delta^k$ $$ N_k(s_n, \mathbf{e}, N) =\frac{ \#\{0\le n< N: (s_{n},s_{n+1},\ldots,s_{n+k-1})=\mathbf{e}\}}{N}\rightarrow \frac{1}{|\Delta|^k} \quad \text{as } N \rightarrow \infty. $$ Drmota et al.\ \cite{drmari} and M\"ullner \cite{mu} proved the normality of the Thue-Morse and the Rudin-Shapiro sequences along squares, that is \begin{equation}\label{eq:normality} \lim_{N\rightarrow \infty}N_k(t_{n^2}, \mathbf{e}, N) =2^{-k} \quad \text{and} \quad \lim_{N\rightarrow \infty}N_k(t_{n^2}, \mathbf{e}, N)=2^{-k} \end{equation} for any $\mathbf{e}\in \{0,1\}^k$. The main tool to obtain the results \eqref{eq:normality} is to prove estimates on the sums $$ \sum_{n<N}(-1)^{e_0t_{n^2}+\dots + e_{k-1}t_{(n+k-1)^2} } \quad \text{and} \quad \sum_{n<N}(-1)^{e_0r_{n^2}+\dots + e_{k-1}r_{(n+k-1)^2} } $$ for any $e_0,\dots, e_{k-1}\in\{0,1\}$. These sums can be estimated via a Fourier analytic method of Mauduit and Rivat which has its origin in \cite{mari1,mari2}. For more details we refer to the survey \cite{dr} of Drmota and the original papers \cite{drmari,mu}. In particular, the normality results \eqref{eq:normality} yield the the subword complexities \begin{equation}\label{eq:subword} p(t_{n^2},k)=p(r_{n^2},k)=2^k. \end{equation} It is conjectured but not proved yet that the subsequences of the Thue-Morse sequence~$(t_{f(n)})$ and Rudin-Shapiro sequence~$(r_{f(n)})$ along any polynomial~$f$ of degree $d\ge 3$ are normal, see \cite[Conjecture 1]{drmari}. Even the weaker problem of determining the frequency of $0$ and $1$ in the subsequences $(t_{f(n)})$ and $(r_{f(n)})$ along any polynomial $f(x)$ of degree $d\ge 3$ with $f(\mathbb{N}_0)\subset \mathbb{N}_0$ seems to be very intricate, see \cite[above Conjecture 1]{drmari}. \begin{problem} Show that the subsequences of Thue-Morse and Rudin-Shapiro sequence along cubes, bi-squares, ..., any polynomial values for a polynomial of degree at least~$3$ are normal. \end{problem} However, Moshe \cite{mo} proved the following lower bound on the subword complexity of~$(t_{f(n)})$, \begin{equation}\label{moshe} p(t_{f(n)},k)\ge 2^{k/2^{d-2}}. \end{equation} Stoll \cite{st12,st16} showed that the number of zeros (resp.\ ones) among the first $N$ sequence elements of both, $(t_{f(n)})$ and $(r_{f(n)})$, is at least of order of magnitude $N^{4/(3d+1)}$, $d\ge 3$. For subsequences $(z_{f(n)})$ of the Zeckendorf sum of digits sequence $(z_n)$ defined by \eqref{zeck} the numbers of zeros and ones among the first~$N$ sequence elements are both lower bounded by~$N^{4/(6d+1)}$, see Stoll~\cite{st13}. M\"ullner and Spiegelhofer \cite{musp,sp} addressed the normality problem for the Thue-Morse sequence along the Piateski-Shapiro sequence $\lfloor n^c\rfloor$ for $1<c<3/2$. Moreover, it is asymptotically balanced (or simply normal)\cite[Theorem~1.2]{sp20} for $1<c<2$. For results on the Thue-Morse and Rudin-Shapiro sequence along primes see \cite{bo1,bo2,mari2,mari} and references therein. In particular, the Thue-Morse sequence $(t_p)$ along primes is balanced, see Mauduit and Rivat \cite{mari2}. However, it is not known whether $(t_{f(p)})_{p}$ is normal for any nonconstant polynomial~$f$. From Theorem~\ref{thm:subword} and \eqref{eq:subword} we know that $(t_{n^2})$ and $(r_{n^2})$ are not automatic and by Theorem~\ref{thm:christol} these subsequences are, in contrast to the original sequence, not of bounded expansion complexity, that is, $$ \lim\limits_{N\rightarrow \infty}E(t_{n^2},N)=\lim\limits_{N\rightarrow \infty}E(r_{n^2},N)=\infty. $$ Theorem \ref{thm:subword} combined with \eqref{moshe} implies that $(t_{f(n)})$ is not automatic and $$\lim\limits_{N\rightarrow \infty}E(t_{f(n)})=\infty$$ for any polynomial of degree at least $2$ with $f(\mathbb{N}_0)\subset \mathbb{N}_0$. Note that it was shown in \cite{al82} that $(t_{f(n)})$ is not $2^r$-automatic and in \cite{alsa} that $(r_{f(n)})$ is not $2^r$-automatic and thus we also have $$ \lim\limits_{N\rightarrow \infty}E(r_{f(n)})=\infty. $$ Subsequences of the Thue-Morse sequence along geometric sequences such as $(t_{3^n})$ seem to be even more difficult to analyze. For example, Lagarias \cite[Conjecture~1.12]{la} conjectured that each pattern appears at least once in $(t_{3^n})$. For other related results see\cite{duwe,kast}. For more details on the normality of automatic sequences and their subsequences we refer to \cite{dr}. \section{Analogs for finite fields}\label{sec:finite_fields} An analog for finite fields of the problem on the distribution of automatic sequences and their subsequences was introduced by Dartyge and S\'ark\"ozy \cite{dasa}. It has been further investigated in \cite{damewi,mawi,sw18,sw1,ma}, see also \cite{dielsh,ga17,sw2,damasa,os}. In the finite field setting some problems can be solved although the analog for integers seems to be out of reach including the normality problem for the analog of the Thue-Morse sequence and the frequency problem for the analog of the Rudin-Shapiro sequence both along polynomials. Hence, these analogs for finite fields are further attractive sources of pseudorandomness. For a prime $p$ and $q=p^r$ with $r\geq 2$ let $(\beta_1,\ldots,\beta_r)$ be an ordered basis of $\mathbb{F}_q$ over~$\mathbb{F}_p$. Then one can write all elements $\xi \in \mathbb{F}_q$ as \begin{equation}\label{eq:xi} \xi=\sum_{i=1}^r x_i\beta_i , \quad x_1,\dots, x_r\in \mathbb{F}_p. \end{equation} It is natural to consider the coefficients $x_1,\dots, x_r$ as digits with respect to the basis~$(\beta_1,\dots, \beta_r)$. Then, in analogy to the Thue-Morse and Rudin-Shapiro sequence satisfying \eqref{sumofdigitsdef} we define the {\em Thue-Morse function} $$ T\left(\sum_{i=1}^r x_i\beta_i\right)= \sum_{i=1}^rx_i ,\quad x_1,\ldots,x_r\in \mathbb{F}_p, $$ and {\em Rudin-Shapiro function} $$ R\left(\sum_{i=1}^r x_i\beta_i\right) =\sum_{i=1}^{r-1}x_ix_{i+1} ,\quad x_1,\ldots,x_r\in \mathbb{F}_p, $$ on $\mathbb{F}_q$. Dartyge and S\'ark\"ozy \cite{dasa} studied the balance of the Thue-Morse function along polynomial values. They derived results using the Weil bound \cite[Theorem~5.38]{lini} on additive character sums: \begin{lemma}\label{lemma:weil} Let $f\in\mathbb{F}_q[x]$ be of degree $d\geq 1$ with $\gcd(d,q)=1$ and $\psi$ be a nontrivial additive character of $\mathbb{F}_q$. Then $$ \left|\sum_{\xi\in\mathbb{F}_q}\psi(f(\xi))\right|\leq (d-1)\sqrt{q}. $$ \end{lemma} Put $$ e(\alpha)=\exp(2\pi i \alpha), \quad \alpha\in\mathbb{R}, $$ and note that $\psi(x)=e(T(x)/p)$ is a nontrivial additive character of $\mathbb{F}_q$. Then from $$ \sum_{h=0}^{p-1} e\left(\frac{ha}{p}\right)=\left\{\begin{array}{cc}0,& a\neq 0,\\ p, &a=0,\end{array}\right. \quad a\in \mathbb{F}_p, $$ we get $$ \#\{\xi\in \mathbb{F}_q: T(f(\xi))=c\}= \frac{1}{p}\sum_{h=0}^{p-1}\sum_{\xi\in \mathbb{F}_q}\psi\left(hf(\xi)\right)e\left(\frac{-hc}{p}\right). $$ The contribution of $h=0$ is trivially $p^{r-1}$, which is the expected number of solutions. The other terms for $h\neq 0$ contribute to the error term and can be bounded by Lemma~\ref{lemma:weil}. We immediately get \cite[Theorem~1.2]{dasa}: \begin{theorem}\label{thm:dasa} Let $f\in \mathbb{F}_q[x]$ be of degree $d$ with $\gcd(d, q) = 1$. Then for all $c\in \mathbb{F}_p$, we have $$ \big|\#\{\xi\in \mathbb{F}_q: T(f(\xi))=c\}-p^{r-1}\big|\le (d-1)p^{r/2}. $$ \end{theorem} Later Dartyge, M\'erai and Winterhof \cite{damewi} investigated this problem for the Rudin-Shapiro function. The main difference between the two problems is that the Rudin-Shapiro function is not a linear map contrary to the Thue-Morse function. Standard character sum techniques fail in this situation. Namely, consider $R(f(\xi))$ with $\xi$ having the form \eqref{eq:xi} as a polynomial in the $r$ variables $x_1,\dots, x_r$. Then using Lemma~\ref{lemma:weil} for one coordinate $x_i$ one gets an error term larger than the main term. Stronger results in higher dimension such as the Deligne bound \cite[Th\'eor\`eme 8.4]{De74} also cannot be applied as it needs some more technically intricate conditions which are not satisfied in our situation. However, sacrificing the explicit dependence of the degree $d$, one can use an affine version of the Hooley-Katz Theorem, see \cite{ho} or \cite[Theorem~7.1.14]{handbookFF}. First recall that the \emph{(affine) singular locus} ${\mathcal L}(F)$ of a polynomial $F\in\mathbb{F}_p[x_1,\dots, x_r]$ is the set of common zeros in $\overline{\mathbb{F}_p}^r$ of the polynomials\footnote{$\overline{\mathbb{F}_p}=\bigcup_{n=1}^\infty \mathbb{F}_{p^n}$ denotes the algebraic closure of $\mathbb{F}_p$.} $$ F,\frac{\partial F}{\partial x_1},\ldots,\frac{\partial F}{\partial x_r}. $$ We also recall that the dimension of ${\mathcal L}(F)$ is the largest $d$ for which there exist $1\le i_1<i_2<\ldots<i_d\le r$ such that there is no nonzero polynomial $P$ in $d$ variables with $P(y_{i_1},\ldots,y_{i_d})=0$ for all $(y_1,\ldots,y_r)\in {\mathcal L}(F)$, see \cite[Corollary~9.5.4]{ideal}. \begin{lemma}\label{lemma:HK} Let $Q\in\mathbb{F}_p[x_1,\dots, x_r]$ be of degree~$d\ge 1$ such that the dimensions of the singular loci of $Q$ and its homogeneous part $Q_d$ of degree $d$ satisfy $$ \max\{\dim({\mathcal L}(Q)),\dim({\mathcal L}(Q_d))-1\}\le s. $$ Then the number $N$ of zeros of~$Q$ in $\mathbb{F}_p^r$ satisfies $$ \left|N-p^{r-1}\right|\le C_{d,r}p^{(r+s)/2}, $$ where $C_{d,r}$ is a constant depending only on $d$ and $r$. \end{lemma} Then using Lemma~\ref{lemma:HK}, one can show that the Rudin-Shapiro function is also asymptotically balanced on polynomial values, see \cite[Theorem~1]{damewi}. \begin{theorem}\label{thm:damewi} Let $f\in \mathbb{F}_q[x]$ be of degree $d$ with $\gcd(d, q) = 1$. Then for all $c\in \mathbb{F}_p$, we have $$ \big|\#\{\xi\in \mathbb{F}_q: R(f(\xi))=c\}-p^{r-1}\big|\le C_{d,r} p^{(3r+1)/4}, $$ where the constant $C_{d,r}$ depends only on the degree $d$ of $f$ and $r$. \end{theorem} Theorem \ref{thm:damewi} is nontrivial if $r$ is fixed and $p\rightarrow \infty$. Contrary to Theorem \ref{thm:dasa}, nothing is known for the dual situation. \begin{problem}\label{prob:RS_FF} For fixed prime $p$ show that if $r$ is large enough, then the Rudin-Shapiro function along polynomial values is balanced possibly under some natural restrictions on the polynomial. \end{problem} Analogously to the normality results of Section~\ref{sec:normality}, Makhul and Winterhof \cite{mawi} obtained results on the normality of the Thue-Morse function along polynomial values. For sake of simplicity we state the case when the polynomial $f$ has degree $d$ smaller than the characteristic $p$, \cite[Corollary~1]{mawi}. \begin{theorem Assume $1\leq d < p$ and $s\leq d$. For any polynomial $f\in\mathbb{F}_q[x]$ of degree $d$ and any pairwise distinct $\alpha_1, \dots, \alpha_s\in\mathbb{F}_q$ and any $c_1,\dots, c_s\in \mathbb{F}_p$ we have $$ \big|\#\{\xi\in \mathbb{F}_q: T(f(\xi+\alpha_i))=c_i, 1\leq i\leq s\}-p^{r-s}\big|\le (d-1) p^{r/2}. $$ \end{theorem} Note that the restriction $s\le d$ is natural and counterexamples for $s>d$ are easy to construct. The case of the Rudin-Shapiro function is much more intricate. \begin{problem Study the normality of the Rudin-Shapiro function at $f(x)$. Namely, show that $$ \frac{\#\{\xi\in \mathbb{F}_q: R(f(\xi+\alpha_i))=c_i, 1\leq i\leq s\}}{p^{r-s}} \rightarrow 1 \quad \text{as } p\rightarrow \infty $$ for some $s\geq 2$ and any $f\in\mathbb{F}_q[x]$ of fixed degree. \end{problem} Of course, this problem is also open for fixed $p$ and $r\rightarrow \infty$ even in the simplest case~$s=1$, see Problem~\ref{prob:RS_FF}. It is natural to define the {\em Rudin-Shapiro function} on the polynomial ring~$\mathbb{F}_p[t]$ by assigning the coefficients of the polynomial $f(t)\in\mathbb{F}_p[t]$ to $(x_1,\dots, x_r)$, that is, $$ R(t^r+x_{1}t^{r-1}+\dots + x_r )=\sum_{i=1}^{r-1}x_ix_{i+1}, $$ for $x_1,\dots, x_r\in \mathbb{F}_p$. Analogously to the result of Mauduit and Rivat \cite{mari} on the Rudin-Shapiro sequence along prime numbers, it is natural to investigate the balance and the normality of the Rudin-Shapiro function along irreducible polynomials. As the number of monic irreducible polynomials of degree $r$ is $p^r/r+o(p^r)$, see for example \cite[Theorem~3.25]{lini}, we expect that the frequency of each element $c$ is $\frac{p^{r-1}}{r}+o(p^{r-1})$. For $r=2$ and fixed $c\in \mathbb{F}_p$ we have to count the number of $x_2\in \mathbb{F}_p^*$ such that $t^2+x_2^{-1}ct+x_2$ is irreducible over $\mathbb{F}_p$ or equivalently the discriminant $x_2^{-2}c^2-4x_2$ is a quadratic non-residue modulo~$p$. This number is $$\frac{1}{2}\sum_{x_2\in \mathbb{F}_p^*} \left(1-\left(\frac{c^2-4x_2^3}{p}\right)\right)= \left\{\begin{array}{cc} \frac{p-1}{2}, & c=0,\\ \frac{p-1}{2}+O(p^{1/2}), & c\ne 0,\end{array}\right.$$ by the Weil bound for multiplicative character sums \cite[Theorem 5.41]{lini}, where~$\left(\frac{.}{.}\right)$ is the Legendre symbol. \begin{problem Prove that for all $c\in\mathbb{F}_p$ and $r\ge 3$ we have $$ \lim_{p\rightarrow\infty}\frac{ \#\{f\in \mathbb{F}_p[t]: \deg f=r, f \text{ monic and irreducible over $\mathbb{F}_p$}, R(f)=c \} }{p^{r-1}}=\frac{1}{r}.$$ \end{problem} We remark that one can define the {\em Thue-Morse function} by $$T(f)=T(t^r+x_1t^{r-1}+\ldots+x_r)=x_1+\ldots+x_r=f(1)-1.$$ Note that for irreducible polynomials $f(x)$ we have $T(f)\ne -1$ and for $c\ne -1$ the number of monic irreducible polynomials of degree $r=2$ with $T(f)=c$ is $$ \frac{1}{2}\sum_{u\in \mathbb{F}_p\atop u^2\ne c+1}\left(1-\left(\frac{u^2-c-1}{p}\right)\right) =\frac{p-\left(\frac{c+1}{p}\right)}{2}, $$ where we used a well-known result on sums of Legendre symbols of quadratic polynomials, see for example \cite[Theorem~5.48]{lini}. In general, since $f(x)$ is irreducible whenever $f(x-1)$ is irreducible we have to estimate the number $I_c$ of monic irreducible polynomials with fixed constant term $c\ne 0$ which satisfies $$\frac{1}{r}\left(\frac{p^r-1}{p-1}-2p^{r/2}\right)\le I_c\le \frac{p^r-1}{r(p-1)},$$ see \cite{car} or \cite[Theorem~3.5.9]{handbookFF}, and we get the desired $$ I_c=\frac{p^{r-1}}{r}+o(p^{r-1}) $$ for $r\ge 3$ as well. Moreover, the corresponding normality problem is trivial since for any polynomial~$g(x)$ of degree at most $r-1$ the value $T(f+g)=f(1)+g(1)-1$ is uniquely defined by $T(f)=f(1)-1$ and $g(1)$. For other results on 'digits' along irreducible polynomials see for example \cite[Chapter~3]{handbookFF} and \cite{gakuwa,gr,pol,tuwa,ha,por}. \section*{Acknowledgment} The authors were supported by the Austrian Science Fund FWF grants P 30405 and P~31762. They wish to thank Jean-Paul Allouche, Harald Niederreiter, Igor Shparlinski, Cathy Swaenepoel, Thomas Stoll and Steven Wang for very useful discussions.
{ "timestamp": "2021-05-10T02:08:48", "yymm": "2105", "arxiv_id": "2105.03086", "language": "en", "url": "https://arxiv.org/abs/2105.03086", "abstract": "Many automatic sequences, such as the Thue-Morse sequence or the Rudin-Shapiro sequence, have some desirable features of pseudorandomness such as a large linear complexity and a small well-distribution measure. However, they also have some disastrous properties in view of certain applications. For example, the majority of possible binary patterns never appears in automatic sequences and their correlation measure of order 2 is extremely large.Certain subsequences, such as automatic sequences along squares, may keep the good properties of the original sequence but avoid the bad ones.In this survey we investigate properties of pseudorandomness and non-randomness of automatic sequences and their subsequences and present results on their behaviour under several measures of pseudorandomness including linear complexity, correlation measure of order $k$, expansion complexity and normality. We also mention some analogs for finite fields.", "subjects": "Number Theory (math.NT); Discrete Mathematics (cs.DM)", "title": "Pseudorandom sequences derived from automatic sequences", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9752018390836984, "lm_q2_score": 0.8221891327004133, "lm_q1q2_score": 0.8018003542840739 }
https://arxiv.org/abs/0712.3618
Network Tomography: Identifiability and Fourier Domain Estimation
The statistical problem for network tomography is to infer the distribution of $\mathbf{X}$, with mutually independent components, from a measurement model $\mathbf{Y}=A\mathbf{X}$, where $A$ is a given binary matrix representing the routing topology of a network under consideration. The challenge is that the dimension of $\mathbf{X}$ is much larger than that of $\mathbf{Y}$ and thus the problem is often called ill-posed. This paper studies some statistical aspects of network tomography. We first address the identifiability issue and prove that the $\mathbf{X}$ distribution is identifiable up to a shift parameter under mild conditions. We then use a mixture model of characteristic functions to derive a fast algorithm for estimating the distribution of $\mathbf{X}$ based on the General method of Moments. Through extensive model simulation and real Internet trace driven simulation, the proposed approach is shown to be favorable comparing to previous methods using simple discretization for inferring link delays in a heterogeneous network.
\section*{Appendix: A counter example} Based on the proof of Lemma 1, we can construct a counter example that the distributions of $\{ X_1,X_2,X_3\}$ are not identifiable. Let $c(t;a,\lambda)=e^{-\lambda|t|}I(|t|\leq a)+\lambda e^{-\lambda a}(a+\frac{1}{\lambda}-|t|)I(a<|t|\leq a+\frac{1}{\lambda})$ be a continuous function defined for $t\in\mathcal{R}$. It is easy to check using Polya's condition (\cite{lukacs:1970}) that for any $a\geq0,\lambda>0$, $c(t;a,\lambda)$ is a characteristic function corresponding to a symmetric, non-vanishing, bounded continuous density function. Let $\phi_{X_1}(t)=c(t;2,1)$, $\phi_{X'_1}(t)=c(t;3,1)$, $\phi_{X_2}(t)=\phi_{X'_2}(t)=\phi_{X_3}(t)=\phi_{X'_3}(t)=c(t;0,1)$. Both groups of distributions corresponding to characteristic functions $\{\phi_{X_k}\}$ and $\{\phi_{X'_k}\}$ for $\{ X_{k}\}$ respectively can generate the same joint distribution of $(Y_{1},Y_{2})$. Figure \ref{fig:counter} shows both their characteristic functions and probability density functions on a two-leaf tree: $\phi_{X_1}$ and $\phi_{X_1'}$ are the two curves in the first box of the second row, and $\phi_{X_2}$ and $\phi_{X_3}$ are in the second and third box of the second row respectively, with corresponding density functions plotted in the first row. Notice that these distributions cannot be used for link delays which do not permit negative values. It is an open question whether there exist link delays whose distribution shapes are not identifiable even for the simple two-leaf tree tomography. \begin{figure} \begin{center} \includegraphics[width=3.6in]{counterexample1.eps} \end{center} \caption{A counter example of identifiability for the two-leaf tree model where $(X_1+X_2,X_1+X_3)$ and $(X_1'+X_2',X_1'+X_3')$ have the same joint distribution: The bottom three figures plot the characteristic functions: the first one for $\phi_{X_1}$ and $\phi_{X_1'}$, and the second and third ones for $\phi_{X_2}=\phi_{X_2'}$ and $\phi_{X_3}=\phi_{X_3'}$ respectively; The top three figures plot the corresponding probability density functions.} \label{fig:counter} \end{figure} \section{Fast algorithms derived from the General Method of Moments}\label{sec:algorithm} In this section, we discuss how to estimate the parameters of mixture coefficients. It is worth pointing out that by Theorem \ref{thm:delayiden} the parameter of the link delay mixture models defined above is identifiable when link delays have positive probabilities at zero, which is usually true. \subsection{The General Method of Moments}\label{sec:estimation} Following \cite{bu:2002}, it is not hard to show that the computational complexity of MLE using an EM algorithm for the above flexible mixture model is of order $O(\max_jn_j^J)$, which is too expensive. In this section, we present an estimation approach for network tomography using Fourier transform, following the pioneering work of \cite{feuerverger:1977}. The estimators using this approach can be computed easily as shown below and also exhibit good statistical properties. The motivation is that the characteristic function of $\mathbf{Y}$ is simply the {\em product} of the characteristic function of components of $\mathbf{X}$ as shown in Equation (\ref{eq:CFY}), though the distribution function is a high order convolution of those of $X_j$s. We derive the estimator from the General Method of Moments formally described in \cite{hansen:1982} and \cite{carrasco.florens.2000}. To be self-contained, we give a formal description of our estimator for the tomography model below and discuss its advantages over previous approaches. Suppose that each $X_{j}$ is modeled by a probability density function $f_{X_{j}}(x_j;\theta_{j})$ with an unknown parameter $\theta_j$. Let $\theta=\{\theta_j: j=1,\ldots,J\}$. By Equation~(\ref{eq:CFY}), the joint characteristic function of $\mathbf{Y}$ is \[ \phi_{\mathbf{Y}}(\mathbf{t};\theta)=\prod_{j=1}^J\phi_{X_j}(t_j;\theta_j) \] where $\phi_{X_j}$ is the characteristic function with respect to $f_{X_j}$. Let $\{ \mathbf{Y}(n):1\leq n\leq N\}$ be the independent measurements of $\mathbf{Y}$. The empirical characteristic function of $\mathbf{Y}$ is \[ \hat{\phi}_{\mathbf{Y}}(\mathbf{t})=\frac{1}{N}\sum_{n=1}^{N}\exp(i\mathbf{t}^{T}\mathbf{Y}(n)). \] Similar to the maximum likelihood estimate which is derived by minimizing the Kullback-Leibler divergence between the empirical distribution and the model distribution of $\mathbf{Y}$, an estimate of $\theta$ can be obtained by minimizing an $L_2$ distance between the empirical characteristic function and the model characteristic function of $\mathbf{Y}$, i.e., \begin{eqnarray} \hat{\theta} & = & \arg\min\int\left|\epsilon_{N}(\mathbf{t};\theta)\right|^{2}d\mu(\mathbf{t}),\label{eq:estimator}\end{eqnarray} where \[ \epsilon_{N}(\mathbf{t};\theta)=\sqrt{N}(\hat{\phi}_{\mathbf{Y}}(\mathbf{t})-\phi_{\mathbf{Y}}(\mathbf{t};\theta)), \] and $\mu(\mathbf{t})$ is a specified probability distribution function on $\mathcal{R}^{I}$ (we use the sub script $N$ to show the dependence on the sample size $N$). For a continuous measure $\mu$, the right hand side of (\ref{eq:estimator}) does not have a closed form in general. To evaluate the integral, a Monte Carlo approximation can be used: first randomly draw $K$ samples from $\mu(\mathbf{t})$, say $\{\mathbf{t}_{k}:k=1,2,\cdots, K\}$, and then replace $\mu(\mathbf{t})$ by its empirical distribution based on these samples. Let $\epsilon_{N}(\theta)\equiv(\epsilon_{N}(\mathbf{t}_{k};\theta), k=1,\cdots,K)^T$ be a column vector. We can rewrite (\ref{eq:estimator}) as \begin{eqnarray} \label{eq:CF} \hat{\theta} = \arg\min_{\theta}\epsilon_{N}^{T}(\theta)\epsilon_{N}^{*}(\theta), \end{eqnarray} where $\epsilon_{N}^*(\theta)$ is the conjugate of $\epsilon_{N}(\theta)$. We call it the {\em CF-estimator}, since it is based on characteristic function. The CF-estimator can be considered as a least square estimator based on the residuals evaluated at $\mathbf{t}_1,\ldots,\mathbf{t}_K$, which are obviously correlated. Let $W$ be the covariance matrix of $\epsilon_N(\theta)$, it is easy to show that \[ W_{jk} =\phi_{\mathbf{Y}}(\mathbf{t}_{j}-\mathbf{t}_{k};\theta)-\phi_{\mathbf{Y}}(\mathbf{t}_{j};\theta)\phi_{\mathbf{Y}}^{*}(\mathbf{t}_{k};\theta). \] This motivates a weighted version of the CF-estimator, called {\em WCF}, \begin{eqnarray} \hat{\theta}^{(W)} & = & \arg\min_{\theta}\varepsilon_{N}^{T}(\theta)(W+\delta_N I_{K})^{-1}\varepsilon_{N}^{*}(\theta),\label{eq:WCF} \end{eqnarray} where $I_{K}$ is the $K\times K$ identity matrix and $\delta_N$, a tuning parameter, is used to make sure the inversion is well defined. $\delta$ should be small and we typically choose $\delta_N$ of order $N^{-1/2}$. In practice $W$ cannot be calculated precisely since $\theta$ is unknown. We can either use a $W$ estimated from an initial estimate of $\theta$ such as the CF-estimator, or iterate this process using an iteratively reweighed least squares, which is a common technique used in generalized linear models. {\em 1) Statistical properties}. The characteristic-function based estimators presented in this section fall into the class of Generalized Methods of Moment (GMM) estimators. There is a considerable body of work on their statistical properties, see \cite{feuerverger:1977} and \cite{carrasco.florens.2000}, from which the consistency and asymptotic normality of both CF-estimator \eqref{eq:CF} and WCF-estimator \eqref{eq:WCF}, can be established. In addition, it has been proved by \cite{carrasco.florens.2000} that when the probability measure $\mu$ in Equation~(\ref{eq:estimator}) has a density all over $\mathcal{R}^{I}$, the WCF estimator is asymptotically as efficient as MLE with $K=\infty$ and an appropriate choice of $\delta_N$. {\em 2) Sampling of $\mathbf{t}$}. For both the CF and WCF estimators, the points $\mathbf{t}_k,k=1,\ldots,K$ are sampled based on a probability measure $\mu$. In general, how to choose $\mu$ or sample $\mathbf{t}$ efficiently is a hard problem (\cite{feuerverger:1977}). In the following, we suggest the choice of $\mu$ based on our simulation experiences. Since the scales of components of $\mathbf{Y}$ may be different, we normalize $\mathbf{Y}$ by its empirical covariance matrix and use an elliptic distribution for $\mu$, such as Gaussian. From simulations we notice that sampling $\mathbf{t}$ directly from a probability measure in $\mathcal{R}^I$ does not easily yield good results. This is due to the sparsity of $\mathbf{t}$ in the high dimensional space so that the characteristic functions $\phi_{\mathbf{Y}}(\mathbf{t})$ evaluated at most of the points are close to zero. Since the variance of the residual $\epsilon_N(\mathbf{t};\theta)$ is equal to $1-|\phi_{\mathbf{Y}}(\mathbf{t})|^2$, the closer to zero of the characteristic function $|\phi_{\mathbf{Y}}(\mathbf{t})|^2$, the larger the variance, and the less the information. Although it may lose some efficiency, simulations suggest that better performance can be achieved by sampling $\mathbf{t}$ from lower dimension subspace, for example 2-dim subspaces. When we draw $\mathbf{t}$ from a lower dimensional subspace, it implies that we minimize the residuals (difference between model and empirical characteristic function) only for these subspaces. This may be viewed as a counterpart of the pseudo likelihood approach for network tomography proposed by \cite{liang:2003} but in the Fourier domain. \subsection{A Fast Algorithm by Quadratic Programming} \label{sec:fastalg} With the mixture model described in Section \ref{sec:mixture}, the unknown parameter for the model is $\theta=\{\theta_{j}:j=1,\cdots,J\}$, the mixture coefficients. Now we describe how to estimate $\theta$ iteratively using the approach developed in Section \ref{sec:estimation}. By Equation (\ref{eq:CFY}) and (\ref{eq:cf-mixture}), the model characteristic function of $\mathbf{Y}$ is \[ \phi_{\mathbf{Y}}(\mathbf{t};\theta) = \prod_{j=1}^J\theta_{j}^{T}\Phi_{j}(\mathbf{t}^TA^j), \] where $\Phi_j(t)=(\phi_{j1}(t),\cdots,\phi_{jn_j}(t))^T$. The objective function in obtaining the CF estimate defined in Equation (\ref{eq:CF}) can then be written as \[ \sum_{k=1}^K|\epsilon_N(\mathbf{t}_k;\theta)|^2 = \sum_{k=1}^K\left|\hat{\phi}_{\mathbf{Y}}(\mathbf{t}_k)-\prod_{j=1}^{J}\left(\theta_j^{T}\Phi_{j}(\mathbf{t}_k^TA^j)\right)\right|^{2}. \] It is easy to see that for each $\theta_j$, if the rest of the parameters are known, the optimization function is a quadratic function of $\theta_j$. Specifically, given all other parameters $\{\theta_l:l=1,\cdots,J,l\neq j\}$, the optimal $\theta_j$ can be obtained by minimizing \begin{eqnarray} \label{eq:quadprog} C(\theta_j) & = & \theta_j^{T}D_j\theta_j-2\theta_j^{T}\mathbf{d}_j, \end{eqnarray} where \begin{eqnarray*} D_j & = & \sum_{k=1}^K \left|\prod_{l\neq j}\phi_{X_{l}}(\mathbf{t}_k^{T}A^{l})\right|^{2}Re\{\Phi_{j}^{*}(\mathbf{t}_k^TA^j)\Phi_{j}^T(\mathbf{t}_k^TA^j)\} \end{eqnarray*} is an $n_{j}\times n_{j}$ matrix, and \begin{eqnarray*} \mathbf{d}_j & = & \sum_{k=1}^K Re\{\hat{\phi}_{\mathbf{Y}}^{*}(\mathbf{t}_k)\prod_{l\neq j}\phi_{X_{l}}(\mathbf{t}_k^{T}A^{l})\Phi_{j}(\mathbf{t}_k^TA^j)\} \end{eqnarray*} is an $n_{j}$-dim column vector. This is a standard quadratic programming problem and can be solved quickly. Therefore, estimation of $\theta$ can be obtained by an iterative algorithm as follows. \begin{alg} Iterative quadratic programming\label{alg:quadprog} \begin{itemize} \item[(1)] Choose an initial value for $\theta_j$, $j=1,\ldots, J$. \item[(2)] For each $j=1,\ldots, J$, estimate $\theta_j$ by minimizing (\ref{eq:quadprog}) using quadratic programming. \item[(3)] Repeat Step 2 until convergence. \end{itemize} \end{alg} A nice property of Algorithm \ref{alg:quadprog} is that it always converges to a local solution because the objective function never increases after each iteration and is bounded below by 0. This is similar to EM algorithms, but care is needed in order to obtain the global minimal. Simulations show that $\{ p_{jk}=1/n_{j}\}$ can serve as a good starting value. The computational complexity of each iteration in Algorithm \ref{alg:quadprog} is $O(KIJ\max_j(n_{j})^{3})$ for the {\em CF-estimator}. For the {\em WCF-estimator}, a similar iterative algorithm by quadratic programming can be obtained. Due to the weight matrix, the complexity of each iteration becomes $O(K^3IJ\max_j(n_j)^3)$. \section{Identifiability} In this section, we study the identifiability issue for model \eqref{eq:tomo1} and prove that the distribution of $\mathbf{X}$ is identifiable up to a shift parameter under mild conditions. The main tool we use is characteristic function whose basic properties are reviewed below. \subsection{Characteristic Function} A characteristic function of a univariate random variable $Z$ is defined by \[ \phi_{Z}(t)=E[e^{itZ}]=\int_{-\infty}^{\infty}e^{itz}f_{Z}(z)dz,\,\,\,\, t\in\mathcal{R},\] where $E[\cdot]$ denotes the expectation with respect to $Z$ and $f_{Z}(\cdot)$ is the probability density function of $Z$. By convention, $\phi_Z$ and $f_Z$ denote the characteristic function and probability density function of $Z$, respectively. The characteristic function for a random vector $\mathbf{Z}\in\mathcal{R}^D$ can be defined in a similar manner by considering $\mathbf{t}\in \mathcal{R}^D$ instead of $\mathcal{R}$. It is well known that a probability distribution can be uniquely specified by its characteristic function and vise versa. Suppose $Z_{1}$ and $Z_{2}$ are two independent random variables. Then the joint characteristic function of $\mathbf{Z}=(Z_1,Z_2)$ is \[ \phi_\mathbf{Z}(\mathbf{t}) = \phi_{Z_1}(t_1) \phi_{Z_2}(t_2), \ \ \mathbf{t}=(t_1,t_2), \] which is a product of the marginal characteristic functions of $\mathbf{Z}$. Let $V=Z_1+Z_2$, then the characteristic function of $V$ is simply a product of the characteristic functions of $Z_{1}$ and $Z_{2}$, i.e., \[ \phi_{V}(t)=\phi_{Z_{1}}(t)\phi_{Z_{2}}(t). \] This is much easier to compute than the density function of $V$, say $f_{V}(\cdot)$, which is a convolution of densities of $Z_{1}$ and $Z_{2}$, i.e., \[ f_{V}(v)=\int_{z_{1}\in\mathcal{R}}f_{Z_{1}}(z_{1})f_{Z_{2}}(v-z_{1})dz_{1}. \] For the tomography model \eqref{eq:tomo1}, since the components of $\mathbf{X}$ are mutually independent, it is easy to evaluate the characteristic function of $\mathbf{Y}$ by \begin{eqnarray}\label{eq:CFY} \phi_{\mathbf{Y}}(\mathbf{t}) = E[e^{i\mathbf{t}^{T}\mathbf{Y}}] = E[e^{i(\mathbf{t}^{T}A)\mathbf{X}}] = \prod_{j=1}^{J}\phi_{X_{j}}(\mathbf{t}^{T}A^{j}), \end{eqnarray} where $A^j$ is the $j$th column of $A$. However, it is in general difficult to evaluate the distribution of $\mathbf{Y}$ because it is a high order convolution in terms of the distribution of $\mathbf{X}$. Below we will use the formula \eqref{eq:CFY} for both the identifiability proof and estimation in network tomography. \subsection{Identifiability} By identifiability, we mean that the distribution of $\mathbf{X}$ can be uniquely determined by the distribution of $\mathbf{Y}$. It is important to establish the identifiability. Otherwise, the distribution of $\mathbf{X}$ may not be estimable from the distribution of $\mathbf{Y}$. In the following, we present our general theorems for identifiability and discuss related issues. We assume that $E|X_j|<\infty$, $j=1,\cdots,J$ and that the distribution of $\mathbf{X}$ satisfies one of the two conditions, namely C1 and C2, defined below. \begin{itemize} \item[(C1)] the characteristic function of each $X_j$ is analytic \footnote{An analytic characteristic function corresponds to a distribution function which has moments $m_k$ of all orders $k$ and $\lim\sup_{k\rightarrow\infty}[|m_{k}|/k!]^{1/k}$ is finite}; \item[(C2)] the characteristic function of each $X_j$ has no zeros in $\mathcal{R}$. \end{itemize} We first address the identifiability issue in Lemma 1 for the simple two-leaf tree tomography model described earlier with Figure \ref{fig:twoleaf}. The result will serve as the basis for Theorem 1 and 2 below where the routing topology is not a simple two-leaf tree. \begin{lemma}\label{lemma:lemma} If \(Y_{1}=X_{1}+X_{2}\) and \(Y_{2}=X_{1}+X_{3}\), then the distributions of $X_1,X_2,X_3$ can be identified up to a shift parameter. \end{lemma} \begin{proof} Suppose there exist both $\mathbf{X}=(X_1,X_2,X_3)^T$ and $\mathbf{X}'=(X_1',X_2',X_3')^T$ with mutually independent components that give rise to the same distribution $\mathbf{Y}=(Y_1,Y_2)$, then we show that distributions of $X_j$ and $X_j'$, $j=1,2,3$, are the same up to a shift parameter. By (\ref{eq:CFY}), we have for $t,s\in\mathcal{R}$, \begin{equation} \label{eq:pf} \phi_{X_{1}}(t+s)\phi_{X_{2}}(t)\phi_{X_{3}}(s)=\phi_{X_{1}'}(t+s)\phi_{X_{2}'}(t)\phi_{X_{3}'}(s). \end{equation} Notice that $\varphi_{j}(t)\equiv\log\phi_{X_{j}}(t)/\phi_{X_{j}'}(t)$ is well defined in a neighborhood of the origin with $\varphi_{j}(0)=0$, $j=1,2,3$. Thus for $t$ and $s$ in a neighborhood of zero, \begin{eqnarray*} \varphi_{1}(t+s)+\varphi_{2}(t)+\varphi_{3}(s) & \equiv & 0.\end{eqnarray*} By using the argument of finite differences (c.f. Lemma 1.5.1 of \cite{kagan:1973}), each $\varphi_{j}$ is a linear complex function in a neighborhood of zero and thus in $\mathcal{R}$ with the given condition. That is, there exist complex numbers $a_{j},b_{j}$ such that $\phi_{X_{j}}(t)=\phi_{X_{j}'}(t)e^{a_{j}+ib_{j}t}$ for any $t\in\mathcal{R}$. By evaluating both sides at $t=0$, $a_{k}=0$. By taking the first order derivative on both sides at zero, $iE[X_{j}]=iE[X_{j}']+ib_{j}$ and thus $b_{j}\in\mathcal{R}$, due to $X_{j},X_{j}'\in\mathcal{R}$. Hence $X_{j}$ and $X_{j}'+b_{j}$ have the same distribution. Further, $AE[\mathbf{X}]=AE[\mathbf{X}']$ implies $b_{2}=b_{3}=-b_{1}$ \end{proof} For network delay tomography, as a generalization of the simple two-leaf tree model, let $A$ correspond to a routing matrix derived from a multicast tree (\cite{presti:2002}), where each node, except for the root and leaves, must have at least two children. Let us take the four-leaf tree in Figure \ref{fig:4leaf} as an example of a multicast tree, which will be used for simulation purposes later. Let $X_1,\cdots, X_7$ denote the link delays on the edges from top to bottom and from left to right in the tree, i.e., the link delay on the edge with end node $j$ is denoted by $X_j$. Let $Y_1,\cdots, Y_4$ denote the end-to-end delays from the root node 0 to end node 4, 5, 6 and 7, respectively. Then each element of $\mathbf{Y}=(Y_1,\cdots,Y_4)^T$ is a partial sum of $\mathbf{X}=(X_1,\cdots,X_7)^T$, for example, $Y_1=X_1+X_2+X_4$. This can be written in the form of \eqref{eq:tomo1}, where $A$ is a $4\times 7$ binary matrix and can be derived from the linear equations. From Lemma \ref{lemma:lemma}, the distributions of $X_4,X_5$ are determined up to a shift parameter by the joint distribution of $(Y_1,Y_2)$, so are the distributions of $X_6,X_7$. Using a bottom-up induction on the tree, it follows that the distributions of all components of $\mathbf{X}$ are determined by that of $\mathbf{Y}$ up to shift ambiguity. The same arguments leads to the following theorem. \begin{thm}\label{thm:delayiden} Let \(A\) be the routing matrix derived from a multicast tree, then the distribution of $\mathbf{X}$ is identifiable up to shift ambiguity. \end{thm} Theorem \ref{thm:iden} below provides a general identifiability result for the traffic demand tomography model, where the routing topology is more general than a multicast tree, as studied in \cite{cao:2000a}. \begin{thm}\label{thm:iden} Let \(B\) be the \([I(I+1)/2]\times J\) matrix whose rows consist of the rows of \(A\) and the component-wise products of each different pair of rows from \(A\). If \(B\) has full column rank, then the distributions of \(\mathbf{X}\) are identifiable up to shift ambiguity. The shift ambiguity satisfies the constraint \(E[\mathbf{Y}]=AE[\mathbf{X}]\). \end{thm} \begin{proof} For the convenience of expression, ignore the shift ambiguity. Let $A_iA_k$ be the element-wise product of $A_i$ and $A_k$. Notice that $(A_{i}A_{k})\mathbf{X}$ denotes the common part of $(A_{i}\mathbf{X},A_{k}\mathbf{X})$, i.e. $(Y_{i},Y_{k})$. Since $\{ X_{j}\}$ are mutually independent, by Lemma 1, the distribution of $(A_{i}A_{k})\mathbf{X}$ is identifiable. Thus the distribution of each component of $B\mathbf{X}$ is identifiable. Let $\psi_{k}$ denote the characteristic function of $B_{k}\mathbf{X}$, where $B_{k}$ is the $k$th row of $B$. Then for $k=1,\cdots,I(I+1)/2$ and for $t$ in a neighborhood of zero, \begin{eqnarray*} \log\psi_{k}(t) & = & \sum_{j}B_{kj}\log\phi_{X_{j}}(t),\end{eqnarray*} where $B_{kj}\in\{0,1\}$ is the $(k,j)$th element of $B$. Since $B$ has full column rank, $\{\log\phi_{X_{j}}(t):j=1,\cdots,J\}$ can be uniquely solved from the above linear equations. Then under either (C1) or (C2), $\phi_{X_j}$ is uniquely decided. That is, the distribution of each $X_{j}$ can be uniquely identified. \end{proof} We now discuss three issues related to the above identifiability results. {\em 1) Location ambiguity}. The location ambiguity of the tomography problem has been recognized in previous works. To avoid such ambiguity, \cite{vardi:1996} assumed a Poisson distribution whose mean is the same as its variance, \cite{cao:2000a} used a power relation between mean and variance, and all previous discrete link delay models assume probabilities starting from zero delay. The important message here is that, despite the location ambiguity, Theorem 1 and 2 state that the distributional shape of each $X_j$ can be determined, for example, all orders of central moments that exist are uniquely identified. In practice, to completely identify the distribution including the location, one can bring in some additional information such as the achievable lower bounds of $\mathbf{X}$ for example in delay tomography, and relationship between mean and variance for example in traffic demand estimation. {\em 2) Conditions on the $\mathbf{X}$ distribution}. The distributional assumption on $\mathbf{X}$ is very weak. A lot of well known distributions have analytic characteristic functions, such as Poisson, Gaussian and discrete distributions, which have been used in the literature. A mixture distribution that we later use to model the link delays in Section 4 has an analytic characteristic function. Although, the heavy-tailed distributions do not satisfy (C1), some heavy-tailed distributions such as $\alpha$-stable distributions satisfy (C2). Despite the generality of our conditions, we do note that they are not necessary ones. For theoretical interest, we have constructed a counter example of a distribution $\mathbf{X}=(X_1,X_2,X_3)$ that cannot be identified from $\mathbf{Y}=(Y_1,Y_2)$ for the simple two-leaf tree model $Y_1=X_1+X_2$ and $Y_2=X_1+X_3$, as in the Appendix. {\em 3) Condition on the routing matrix $A$}. \cite{cao:2000a} has shown that the full rank condition in Theorem \ref{thm:iden} is necessary in the context of traffic demand tomography when $\mathbf{X}$ is Gaussian. In practice, such a condition is easily satisfied for routing matrices derived from realistic network topologies. A more general condition of $A$ has been developed to prove the identifiability for Poisson distributions in the context of traffic demand estimation \cite{vardi:1996}. We conjecture that under Vardi's more general condition of $A$, the distribution of $\mathbf{X}$ is identifiable up to mean and variance ambiguity under condition (C1), but we leave the investigation for future work. \section{Introduction} Network performance monitoring and diagnosis is challenging due to the size and decentralized nature of the Internet. The service providers may collect their link level statistics using tools such as Cisco Netflow, whereas the end users can obtain the end-to-end performance by probing the network. Unfortunately, none of them has a global view of the Internet. For instance, when an end-to-end measurement indicates the performance degradation of an Internet path, the exact cause is hard to be uncovered because the path may traverse several autonomous systems (AS) that are often owned by different entities and the service providers generally do not share their internal performance. Even if they do, there is no scalable way to correlate the link level measurements to end-to-end performance in a large network like the Internet. Similarly, the service providers may be interested in the end-to-end path characteristics that they can not observe directly. Network tomography is a technology addressing these issues that infers unobservable characteristics from easily available measurements. There have been two forms of network tomography being studied in the literature. One, called {\it network delay tomography}, estimates the link-level characteristics based on end-to-end measurements, and the other, called {\it traffic demand tomography}, predicts end-to-end path-level traffic intensities based on link-level traffic measurements. The key advantage of network tomography is that it does not require the collaboration between network internal elements and end users. See \cite{castro:2004}, \cite{denby:2007} and references therein for an excellent review. We focus on network delay tomography in this paper, while the proposed approach may also be applied to traffic demand tomography. Network delay tomography aims to estimating network internal characteristics such as loss and delay\footnote{To be precise, the delay here is the queuing delay that excludes the constant link propagation delay, we omit queuing when context is clear}, from end-to-end measurements by exploiting the inherent correlation in performance. Considering a tree spanning a source of probes (root) and a set of receivers (leaves), the packets are potentially subject to queuing delay and loss at each link. The end-to-end (source-to-receiver) measurements may be made passively or actively. The probes for the active measurements can be sent using either multicast or unicast routing\footnote{With multicast, a packet is sent from a source to multiple destinations simultaneously; with unicast, a packet is sent to different destinations separately}. See \cite{lawrence.et.al.2006} and \cite{denby:2007} for examples of how unicast and multicast probes can be designed and sent. Because only one copy of a probe is transmitted on the common links, multicast probing based tomography has the advantage of perfect correlation on the common links, less overhead, and better scalability. Following \cite{presti:2002} and \cite{liang:2003}, we assume that measurements are collected from multicast probes, although the multicast routing is not widely enabled in today's Internet. It has been shown in \cite{bu:2002} on how to apply the tomography algorithms developed for multicast measurements when only unicast measurements are available. The statistical models for both types of network tomography can be unified as follows: \begin{equation}\label{eq:tomo1} \mathbf{Y}=A\mathbf{X}, \end{equation} where $\mathbf{X}=(X_{1},\ldots,X_{J})^T$ is a $J$-dimensional vector of network dynamic parameters, and $\mathbf{Y}=(Y_{1},\ldots,Y_{I})^T$ is an $I$-dimensional vector of measurements and $A$ is an $I\times J$ matrix with elements 0 or 1 which represents the routing topology of the network under consideration. Here we use the superscript $T$ to denote the transpose. In most network tomography scenarios, the components of $\mathbf{X}$ are assumed independent but unobservable. Usually $I$ can be as large as $J^2$ for network demand tomography and as large as $2J-1$ for network delay tomography. In network delay tomography, each component of $\mathbf{X}$ represents an internal link delay and each component of $\mathbf{Y}$ represents a delay measurement from a source to a destination. The objective of network tomography is to estimate the distribution of $\mathbf{X}$ given independent observations from the distribution of $\mathbf{Y}$. As a simple example, Figure \ref{fig:twoleaf} shows a two-leaf tree topology, on which a probing packet is sent from the root node 0 (source) to leaf nodes 2 and 3 (receivers). When the packet arrives at node 1 from the source, it is replicated and transmitted to node 2 and 3 simultaneously as the red arrows show. Let $X_1$ denote the link delay from node 0 to node 1 and let $X_i$ denote the link delay from node 1 to node $i$, $i=2,3$ respectively. Let $Y_1, Y_2$ be the end-to-end delays from node 0 to node 2 and 3 respectively. Then $Y_1=X_1+X_2$ and $Y_2=X_1+X_3$, which can be written in the form of \eqref{eq:tomo1} with $A$ a $2\times 3$ binary matrix, i.e. $A=[1,1,0;1,0,1]$. \begin{figure}[t] \begin{minipage}[t]{.42\textwidth} \begin{center} \epsfig{file=twoleaf1.eps, scale=0.35} \caption{Two-leaf tree} \label{fig:twoleaf} \end{center} \end{minipage} \begin{minipage}[t]{.42\textwidth} \begin{center} \epsfig{file=tree41.eps, scale=0.3} \caption{Four-leaf tree} \label{fig:4leaf} \end{center} \end{minipage \end{figure} There have been significant amount of works on network tomography in recent years. Network tomography was first proposed by \cite{vardi:1996} and then followed by remarkably \cite{tebaldi:1998}, \cite{cao:2000a} and \cite{liang:2003} for traffic matrix estimation, i.e. traffic demand tomography. \cite{caceresIT:1999}, \cite{zhu.geng.2005} and \cite{Xi.et.al.2006} among others studied it for inferring network internal loss. Network delay tomography has also been studied extensively. \cite{presti:2002} developed a fast algebraic algorithm but it is quite inefficient. \cite{bu:2002} showed that the maximum likelihood estimate (MLE) requires exponential computational complexity. \cite{tsang:2003} proposed a penalized maximum likelihood method. \cite{liang:2003} proposed a pseudo-likelihood method with multicast measurements, and recently \cite{lawrence.et.al.2006} proposed local likelihood method with both unicast and multicast measurements, both of which were shown to be fast and quite efficient compared with the MLE. These studies are based on a discrete distribution with equally spaced bins for modeling link delays, where the {\it same} bin width is used for all the links for the ease of computation. \cite{duffield:2001b} pointed out that, however, a single fixed bin width is not appropriate for heterogeneous networks such as the Internet because it does not scale well between both fast links and slow links. They proposed a varying-bin discrete model for estimating link delay distributions based on unicast measurements. Their estimation idea is to use structured bins such that they can iteratively estimate a segment of delay distributions by truncating the delays from both sides, i.e. rounding the left of the segment to zero and the right to infinity. However, the performance of their estimation approach is not better than that using an equal-bin discrete model with an appropriate bin width as they reported, probably due to the bias introduced by their brute-force truncation. Our approach is also based on varying-bin type models but does not suffer from such bias. \cite{shih.hero.2001} proposed to estimate cumulative generating functions (similar to characteristic functions used in this paper) of link delays, but they did not estimate link delay distributions. \cite{shih.hero:2003} also proposed finite mixture models with Gaussian components for link delay distributions based on unicast measurements. There are several previous works that have considered the identifiability issue for the network tomography problem, for example \cite{vardi:1996}, \cite{cao:2000a} and \cite{presti:2002}. These authors considered instances of the tomography problem by assuming specific parametric (such as Poisson and Gaussian) or discrete distributions. We will unify these results and extend the identifiability condition to general distributions under mild assumptions. The contributions of this paper are as follows. First, we prove that the distribution of $\mathbf{X}$ is identifiable up to a shift parameter under general conditions. Second, we propose flexible mixture models of characteristic functions for network delay tomography and develop a fast algorithm for estimation based on the General Method of Moments (GMM). The new approach allows one to model continuous delays on {\it heterogeneous} network links conveniently, where delays may not have the same scale across all network links. Extensive model simulation and real Internet trace-driven simulation suggest that our new approach can yield more accurate estimates of link delay distributions yet is computationally less expensive than previous approaches. The remaining sections of the paper are structured as follows. In Section 2, we address the identifiability issue. We describe the mixture models for link delays in Section 3 and develop a fast algorithm for estimating the delay distributions in Section 4. In Section 5, we present extensive experimental studies for evaluating the proposed method. Section 6 concludes the paper. \section{Network delay tomography using mixture modeling} \label{sec:mixture} Below we focus on network delay tomography and describe a class of flexible mixture models for modeling link delays. It is well known that there does not exist a standard parametric model that can sufficiently model the distributions of network link delays (see \cite{duffield:2001b} and \cite{tsang:2003} among others). But it is possible to define a mixture model which is flexible enough for link delay distributions. Assume that for each link $j$, the link delay $X_j$ follows a mixture density function with $n_j$ components, $X_j \sim f_{X_j}$, defined by : \begin{eqnarray} f_{X_j}(x;\theta_j) = \sum_{l=1}^{n_j} p_{jl}\kappa_{jl}(x), \ \ x>0\label{eq:mixture} \end{eqnarray} where $\theta_j\equiv (p_{jl},\cdots,p_{jn_j})^T$ contains the mixing probabilities with constraint $p_{jl}\geq 0$, $\sum_{l}p_{jl}=1$, and $\{\kappa_{jl}\}$ are some basis density functions. There is another practical reason that we use a mixture model for link delays: The characteristic function of a mixture distribution is a mixture of characteristic functions of the basis distributions and thus can be computed conveniently once the basis distributions are chosen appropriately, as shown later. In this case, the characteristic function of $X_j$ can be expressed as \begin{eqnarray} \phi_{X_j}(t;\theta_j) & = & \sum_{l=1}^{n_j}p_{jl}\phi_{jl}(t), \label{eq:cf-mixture} \end{eqnarray} where $\phi_{jl}$ is the characteristic function of the basis function $\kappa_{jl}$. The basis functions are chosen as follows for modeling link delays. For $j=1,\cdots,J$, let $0=b_{j1}<b_{j2}<\ldots<b_{j(n_j-1)}<\infty$. Define the basis function as \begin{eqnarray}\label{eq:kernel} \left\{ \begin{array}{ll} \kappa_{j1}(x) \ = & \mbox{point mass at zero (for zero)}\\ \kappa_{jk}(x) \ = & \mbox{uniform on $[b_{j(k-1)}, b_{jk}]$} \\ & \hspace*{0in} \mbox{$\ 2 \leq k \leq n_j-1$ (for body)}\\ \kappa_{jn_{j}}(x) \ = & \mbox{exponential with scale $\alpha_j$}\\ & \hspace*{0.1in}\mbox{on $[b_{j(n_j-1)}, \infty]$ (for tail)}\\ \end{array} \right. \end{eqnarray} The point mass at zero link delay is used here because it is well known that for a FIFO queue (First In, First Out), the steady state queuing distribution has zero delay with probability one minus the utilization of the queue. For the body of the distribution, we choose the piecewise uniform model because of its simplicity and flexibility. Finally, an exponential distribution is used to model the tail because it is the right model for the short range dependent traffic model, and for long-range dependent model it represents a trade-off between accuracy and simplicity. In order to reduce the computational complexity, we choose the bin endpoints $\{b_{jk}\}$ in advance. The parameter of interest is composed of the mixture coefficients, denoted as $\theta={\theta_1,\cdots,\theta_J}$. To our advantage, we do not require the bins to be equally spaced. In fact, it is important to choose the bins that are adaptive to individual link delay distributions in order to obtain accurate estimates. Such a varying bin strategy is especially important for a heterogeneous network environment whose link delay distributions vary widely across links, because a single bin width value could be at the same time too coarse grained for a high bandwidth link with small delays but too fine-grained to efficiently capture the essential characteristics of the delay along a low bandwidth link (\cite{presti:2002}). In addition, since a typical delay distribution may have a density varying a lot at different areas, it is important to be able to place more bins in the high density area and fewer bins in the low density area. Note that the equal-bin distribution used by most previous researchers is also a mixture model. Figure \ref{fig:delaypdf} shows two link delay distributions (in solid lines) where one ranges from 0 to 12 (top) and the other from 0 to 240 (bottom). The slashed lines are fitted curves using a equal-bin model with bin-width 1 which accomodates the scales of both links, but with 12 bins for the slow link and as many as 240 bins for the fast link. The dotted lines are fitted curves using varying-bin models where only 10 bins are used for both links. It is clear that the equal-bin model fits the slow link very well, but not the fast link, while the varying-bin model with a small number of bins fits both links well. It is possible to use a very small bin-width for the equal-bin model, but it would require too many bins for the slow link. The varying bins here are chosen based on quantiles of the delay distribution, which works very well in general from our simulation experiences. In reality the quantiles are unknown and we can only obtain an approximation using an initial estimate of the link delay distribution. This process can be iterated until we get a good estimate. The scale parameter $\alpha_j$ for the tail basis in Equation~(\ref{eq:kernel}) is unknown and needs to be estimated. However, the accuracy of the scale estimate is less important if the endpoint $b_{j(n_j-1)}$ of the last bin can be placed at the far end of the tail. For a further simplification, we can fix the tail basis with a crude estimate of $\alpha_j$ for each link $j$ and only estimate the mixing probabilities $\{p_{jk}\}$, which is described next. \begin{figure}[h]\begin{center} \psfig{figure=compevbin.eps,width=3in} \caption{Fitting two link delay distributions using an equal-bin model and a varying-bin model, where the delay on the slow link (bottom) is 20 times in average of that on the fast link (top): the solid lines are for the link delay density functions, slashed lines for the estimated densities using bin-with equal to 1 (12 bins for the fast link and 240 bins for the slow one), and the dotted lines for the estimated densities using varying bin-widths (only 10 bins for each).} \label{fig:delaypdf}\end{center} \end{figure} \section{Simulation and Experimental Studies} \label{sec:simulation} In Section \ref{sec:algorithm}, we have developed simple and fast algorithms using a flexible mixture model for network delay tomography. In this section, we evaluate the performance of the proposed algorithms in terms of statistical efficiency and accuracy. To measure the accuracy of the estimation as compared to the true distributions, we use a $L_1$-distance for discrete link delay distributions and a normalized Mallows distance for continuous link delay distributions. Our evaluation is divided into three pieces. First, we study the efficiency of our estimates by comparing them with that of MLE for a discrete link delay distribution with equally spaced bins. We show that our estimators have comparable efficiency to that of MLE which is statistically efficient and also computable in this setting. Second, we examine the performance of our estimators using model simulations for continuous link delays in an ideal scenario where both temporal and spatial independence hold. Model simulations demonstrate the importance of varying bins selection that should adapt to not only delay distributions of individual links but the different scales of delays across links in order to achieve satisfactory estimates. Finally, we use real trace driven simulations to examine the accuracy of our estimators under more realistic scenarios where the independence assumptions may not be strictly true as appeared in the Internet. Results from our trace driven simulations demonstrate that the estimates made by our algorithms closely match the real distributions. \subsection{Efficiency Evaluation} We study the efficiency of our estimators using a discrete link delay distribution with equally spaced bins on a four-leaf tree (Figure~\ref{fig:4leaf}). For link $j, j=1,\ldots,7$, the link delay has a discrete distribution at $\{0,1,\cdots,5\}$ with probabilities generated uniformly from the space $\sum_{k=1}^6 p_{jk}=1$ with constraints $0<p_{jk}<1$. A total of 500 delay samples are generated for each link from its specified delay distribution and the end-to-end delays are computed according to the model \eqref{eq:tomo1}. The delay distributions of all seven links are estimated using the MLE, the CF-estimator, and the WCF-estimator. We repeat the experiment 100 times with different random seeds. Both the MLE and the CF-estimator use the uniform distribution as starting values whereas WCF uses the CF estimates as starting values. The weight matrix $W$ for WCF is also derived from the CF estimates. For both the CF and WCF estimators, a total of 3000 samples of $\mathbf{t}$ are drawn randomly from the 2-dim subspaces of $I$-dimensional end-to-end delays using a Gaussian distribution with a scale parameter of 5 after normalizing $\mathbf{Y}$. We have also run the recursive algorithm developed in \cite{presti:2002}, but we do not report the result here except to state that it often yields much poorer estimates (similar to observations made by \cite{liang:2003}). Figure~\ref{fig:mn_pdf} shows both the estimated and the simulated seven link delay density functions in one simulation experiment. We observe that all methods give reasonably accurate estimates. To compare errors of the different estimators, we calculate the $L_{1}$ distances between the estimated link delay density functions and the ground truth for each of the 100 experiments. Figure~\ref{fig:mnL1err} reports the 25\%, 75\% quantiles of the $L_{1}$ errors for each link in vertical line segments, whose middle points represent median errors. MLE has the smallest median error, and the median errors of CF and WCF are 50\% and 22\% higher than that of MLE. The results suggest that both CF and WCF are somewhat worse than as expected but comparable to MLE. \subsection{Accuracy Evaluation Using Model Simulation} In this subsection, we investigate the performance of the estimators for continuous delay distributions, which are more realistic than discrete ones since network delays are essentially continuous except at zero. Delay tomography in a heterogeneous network is intrinsically more challenging than in a homogeneous network because links with small delays are not equally represented as links with large delays in the end-to-end delay measurements. In addition, the heterogeneous environment also represents a situation where most of the existing methods such as MLE do not work well because they rely on simple discretization. After all, the real Internet is a heterogeneous network. Thus we report model simulations on a four-leaf tree that resemble a heterogeneous network environment. For simplicity, we do not consider the point mass at zero for model simulations, but we will treat this in later real trace driven simulations. To conduct a comprehensive evaluation, we run simulations for the link delay distributions of different shapes. Due to the space limit, we only report the results for two representative distributions that are i) exponential (uni-modal) and ii) a mixture of an exponential and a Gamma with shape parameter 2 (multi-modal). In both cases, the average link delays on the four-leaf tree are 3, 1, 5, 10, 6, 4 and 20 respectively for link 1 to 7 assigned from top to bottom and left to right, which resembles a heterogeneous network with the average link delays varying by a factor of 20. We generate 2000 delay samples for each link from the specified delay distributions, and we estimate the seven link delay density functions from the resulting end-to-end delays. We use four different estimates of link delay distributions: {\em CF\_equal\_bin, WCF\_equal\_bin, CF\_varying\_bin, WCF\_varying\_bin}. All four estimates are obtained using a mixture model for the link delays of the same form as \eqref{eq:kernel} with $n_j=12$ except removing the point mass at 0. The difference in the mixture model for the estimates lies in the bin placement. For both {\em CF\_equal\_bin} and {\em WCF\_equal\_bin}, the 12 bins are equally spaced using a bin width selected for each link based on variance estimates, which are obtained by solving systems of linear equations, following \cite{duffield:2004}. For both {\em CF\_varying\_bin} and {\em WCF\_varying\_bin}, the bins are located at the quantiles of the delay distributions that corresponds to probabilities $i/13, i=1,\ldots, 12$. Figure \ref{fig:expcdf} and \ref{fig:expgamcdf} plot the estimated cumulative distribution functions for each link delay for case i) and ii) respectively, along with the ground truth in one simulation run. From the figures, we observe that the estimates using varying bins are almost identical to the true distributions. The estimates using equal bins give satisfactory estimates for case i) but not quite as good for the more complex case ii). To measure the accuracy of the estimates, we use the Mallows distance defined for a cumulative distribution $F$ and its estimate $\hat{F}$ by \[ M(F,\hat{F}) = \int_{0}^1 \left|F^{-1}(p) - \hat{F}^{-1}(p)\right| dp, \] where $F^{-1}$ and $\hat{F}^{-1}$ are the inverse cumulative distributions. The Mallows distance can be viewed as the average of absolute difference in quantiles between two distributions. Because the Mallows distance is linear to the scale of distributions, we use $M(F,\hat{F})/\sigma_F$, the normalized Mallows distance, to measure the difference between $F$ and $\hat{F}$, where $\sigma_F$ is the standard deviation of $F$. We repeat the simulation 100 times and compute the normalized Mallows distance between the estimated and true distributions as the error metric for all links. Figure~\ref{fig:expmallow} and \ref{fig:expgammallow} report, corresponding to case i) and ii) respectively, the first and third quartiles of the errors as well as median errors for each link, similar to Figure \ref{fig:mnL1err}. It is clear that the varying bins improve the quality of estimates significantly over equal bins. (Note that the difference between CF and WCF are not significant though.) This suggests that selecting bins based on characteristics of the underlying density distributions is important in improve the accuracy in a heterogeneous network. \subsection{Accuracy Evaluation Using Real Internet Traces} We next investigate how the algorithms perform in a realistic network environment where some of the assumptions may not hold completely. For instance, due to the closed-loop control nature of the TCP protocol, the packets within the same TCP connection have strong temporal dependency. Although the dependency is weakened when many TCP connections are multiplexed as they arrive to a link, the dependency may not be completely gone. We approximate a real scenario by simulating the behavior of a link using the real traces collected from the Internet. Since the traces include the arrival time and the size of each packet, the simulation sees the exact link behaviors as what the original link where the trace collected from seen if we set the bandwidth and the buffers the same as the original link. We use traces from the NLANR web site \footnote{http://pma.nlanr.net/Traces/} that archives packet header traces collected from about ten links at different locations of the Internet. The links differ in both bandwidth and traffic. A 90-second trace is recorded every one (two) hours for each of the links. In our experiment, we first assign traces collected from different sites to the links of the simulated network. We then simulate the links using the assigned traces as input using the standard network simulator tool \cite{ns}. Moreover, we superpose the probes to the traces and record their per-link queuing delays as well as end-to-end delays where the latter is used as input for the estimation whereas the prior is for comparison with the estimates. Notice that the delays on the edge links in a real network may vary more than the core links due to its low bandwidth. In addition, the average delay may also differ dramatically for different links. We resemble a real network in a symmetric binary 8-leaf tree by assuming that both the root and the leaves of the tree are on the edge of the network whereas the interior links are in the core. We assign traces of high rates to links in the core and traces of low rates to the edge links. Figure~\ref{fig:8leaf_cdf} shows both CF and WCF estimate of the delay distribution using a varying bin strategy laid out in Section~\ref{sec:mixture}, along with the simulated distribution. The throughputs across different links vary by a factor of 40. It is easy to see that the estimates are extremely good for most links, except for link 9 that has the smallest average link delay where it shows some marginal error. The average normalized Mallows distance over all links is 0.065 which also suggests a good match between the estimates and simulated results. We have also simulated the four-leaf tree network and a symmetric binary 16-leaf tree network respecitvley, using different traces, and the proposed algorithms give satisfactory results. In addition, our real trace driven simulations suggest that the link delay distributions excluding the tail can be well approximated by a Weibull distribution with a shape parameter slightly smaller than 1. This is not surprising because it has been shown that the queuing delay for a FIFO queue with a Fractional Brownian Motion traffic input has a Weibullian tail. The Weibullian form is also consistent with the finding in \cite{cao:2004}. \section{Conclusion} This paper has presented a general identifiability result and introduced a general estimation approach for the network tomography problem. For network delay tomography, a fast algorithm based on GMM has been developed for estimating the link delay distributions using mixture models of characteristic functions. In comparison with likelihood based approaches, the most significant nature of the new method is that it affords the choice of varying bin widths which adapts to delay variabilities of individual links and has low computational complexity. The new approach can be applied to traffic demand estimation as well. \section*{Acknowledgments} We would like to thank Gang Liang for sharing his simulation codes and Michael Greenwald for helpful discussions. A conference version of the main results has appeared in the Proceeding of IEEE INFOCOM (\cite{chen.et.al.2007}). \input{appendix} \bibliographystyle{Chicago}
{ "timestamp": "2007-12-21T04:47:29", "yymm": "0712", "arxiv_id": "0712.3618", "language": "en", "url": "https://arxiv.org/abs/0712.3618", "abstract": "The statistical problem for network tomography is to infer the distribution of $\\mathbf{X}$, with mutually independent components, from a measurement model $\\mathbf{Y}=A\\mathbf{X}$, where $A$ is a given binary matrix representing the routing topology of a network under consideration. The challenge is that the dimension of $\\mathbf{X}$ is much larger than that of $\\mathbf{Y}$ and thus the problem is often called ill-posed. This paper studies some statistical aspects of network tomography. We first address the identifiability issue and prove that the $\\mathbf{X}$ distribution is identifiable up to a shift parameter under mild conditions. We then use a mixture model of characteristic functions to derive a fast algorithm for estimating the distribution of $\\mathbf{X}$ based on the General method of Moments. Through extensive model simulation and real Internet trace driven simulation, the proposed approach is shown to be favorable comparing to previous methods using simple discretization for inferring link delays in a heterogeneous network.", "subjects": "Methodology (stat.ME); Statistics Theory (math.ST); Applications (stat.AP); Computation (stat.CO)", "title": "Network Tomography: Identifiability and Fourier Domain Estimation", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9752018362008348, "lm_q2_score": 0.8221891305219504, "lm_q1q2_score": 0.8018003497893739 }
https://arxiv.org/abs/1909.06347
Steiner's formula and a variational proof of the isoperimetric inequality
We give a new proof of the isoperimetric inequality in the plane, based on Steiner's formula for the area of a convex neighborhood. This proof establishes the isoperimetric inequality directly, without requiring that we separately establish the existence of an optimal domain. In doing so, this proof bypasses the main difficulty in all of the proofs Steiner outlined for the plane isoperimetric inequality.
\section{Introduction} \label{intro} The classical isoperimetric inequality states that among all simple closed curves of length $L$ in the plane, the unique curve enclosing the largest area is the circle of circumference $L$: \smallskip \begin{theorem}[The Isoperimetric Inequality] \label{ie} Let $\gamma$ be a simple closed curve in the plane of length $L$, enclosing a domain $D$ of area $A$. \medskip Then $L^{2} \geq 4\pi A$, with equality precisely if $\gamma$ is a circle. \end{theorem} \smallskip This paper gives a proof of the isoperimetric inequality based on Steiner's formula, which describes the area of a neighborhood of a convex domain in $\R^{2}$: \smallskip \begin{theorem}[Steiner's Formula, \cite{St}] \label{sf} Let $D$ be a bounded, convex domain in $\R^{2}$, of area $A$ and perimeter $L$, and let $D_{r}$ be the $r$-neighborhood of $D$, i.e. the points in $\R^{2}$ whose distance from $D$ is $r$ or less. Then: \smallskip \begin{itemize} \item[{\bf A.}] $Area(D_{r}) = \pi r^{2} + L r + A$, \medskip \item[{\bf B.}] $Length(\partial D_{r}) = 2\pi r + L$. \end{itemize} \end{theorem} \smallskip Jakob Steiner (March $18^{th}$, 1796 - April $1^{st}$, 1863) proved Theorem \ref{sf} for convex polygons and a similar formula for convex polyhedra in $\R^{3}$. By polygonal approximation, Theorem \ref{sf} then follows for any compact, convex set in $\R^{2}$, and in fact a version of Theorem \ref{sf} holds in much greater generality -- for more about Steiner's formula, see \cite{Gr,Sc}. Steiner was fascinated by the isoperimetric inequality, and he sketched several ideas for proving it -- cf. \cite{Tr,Bl}. The isoperimetric problem was already ancient when Steiner considered it in the nineteenth century, but Theorem \ref{ie} had never been proven rigorously. It remained unproven in Steiner's lifetime, and all of Steiner's ideas for proving the isoperimetric inequality required the same additional step, which he never provided: one must show that the isoperimetric problem has a solution. \\ More precisely, we define the {\bf isoperimetric ratio} of a domain $D$ with area $A$ and perimeter $L$ to be: \begin{equation} \label{ir} \displaystyle {\Huge \frac{L^{2}}{4\pi A}}. \end{equation} \medskip The isoperimetric ratio is scale-invariant -- we formulate the isoperimetric inequality in terms of $L^{2}$ and $A$, as in Theorem \ref{ie}, because $L^{2}$ and $A$ transform the same under rescalings. The isoperimetric inequality then states that the isoperimetric ratio of any plane domain is greater than or equal to $1$, with equality precisely for disks. Steiner developed many proofs that no domain other than a disk could minimize the isoperimetric ratio, but he didn't establish the existence of a domain that minimizes (\ref{ir}). \\ The first proof of the existence of a domain minimizing the isoperimetric ratio seems to have been in unpublished lecture notes of Weierstrass in 1879, cf. \cite{Bl}. The existence of an optimal isoperimetric domain in the plane is now known to be a consequence of several compactness theorems in metric geometry and geometric measure theory, however the proof below does not require that we establish the existence of a minimizer for the isoperimetric ratio -- we show directly that no domain can have an isoperimetric ratio less than $1$. We believe part of the significance of our proof is that it shows how one of Steiner's ideas from convex geometry can be used to prove the isoperimetric inequality without separately establishing the existence of an optimal domain. \\ The basic observation for our proof is the following: if $D$ is a bounded convex domain in $\R^{2}$, we can use Theorem \ref{sf} to calculate the isoperimetric ratio $\mathcal{I}(r)$ of the $r$-neighborhood of $D$ as a function of $r$. Letting $A$ be the area of $D$ and $L$ its perimeter, we have: \begin{equation} \label{sfir} \displaystyle \mathcal{I}(r) = \frac{\left( 2\pi r + L \right)^{2}}{4\pi \left( \pi r^{2} + Lr + A \right)} = \frac{4\pi^{2} r^{2} + 4\pi L r + L^{2}}{4\pi^{2} r^{2} + 4\pi L r + 4\pi A} . \end{equation} \medskip Differentiating with respect to $r$, we have: \begin{equation} \label{sfird} \displaystyle \mathcal{I}'(r) = \frac{\left( 4\pi A - L^{2} \right) \left( 8\pi^{2} r + 4\pi L \right)}{\left( 4\pi^{2} r^{2} + 4\pi L r + 4\pi A \right)^{2}} = \frac{\left( 4\pi A - L^{2} \right) \left( \pi r + L \right)}{4\pi \left( \pi r^{2} + L r + A \right)^{2}}. \end{equation} \medskip This implies that $\mathcal{I}(r)$ is a monotone function of $r$, decreasing if the isoperimetric ratio of $D$ is greater than $1$ and constant if the isoperimetric ratio of $D$ is equal to $1$. If $D$ were a convex domain with an isoperimetric ratio less than $1$, $\mathcal{I}(r)$ would increase monotonically to $1$, the isoperimetric ratio of the disk, as $r$ goes to infinity. As $r$ goes to infinity, the $r$-neighborhoods of any convex domain $D$, when rescaled to have constant area, converge to a disk -- see Proposition \ref{goodvar}. We will see that this gives a variation of the disk, as an argument for the functional on plane domains given by the isoperimetric ratio. We will use Steiner's formula to find its first and second variations -- in particular, we will relate them to the isoperimetric ratio of the domain $D$ in question. We will then be able to deduce Theorem \ref{ie} from the fact that the disk is a critical point, with non-negative second variation, for the isoperimetric ratio on plane domains. For later reference, the quantity $L^{2} - 4\pi A$ whose negative appears in (\ref{sfird}) is called the {\bf isoperimetric deficit} of a domain. \\ It will be important in our proof that, in the plane, the convex hull $conv(D)$ of a non-convex domain $D$ always has a smaller isoperimetric ratio than $D$ itself: $conv(D)$ encloses a larger area than $D$ with a smaller perimeter. Therefore, to prove Theorem \ref{ie}, it is enough to show that the isoperimetric inequality holds for convex domains. Steiner was aware of this fact and used it in several of his ideas for proving the isoperimetric inequality. In dimensions greater than $2$, this is no longer true: the isoperimetric ratio of a $3$-dimensional domain with volume $V$ and surface area $A$ is defined to be $\frac{A^{3}}{36 \pi V^{2}}$. Like (\ref{ir}) for plane domains, the isoperimetric ratio of a domain in $\R^{3}$ is scale-invariant and the ball has isoperimetric ratio equal to $1$. The isoperimetric inequality in $\R^{3}$ states that the isoperimetric ratio of any domain is greater than or equal to $1$, with the ball being the unique minimizer. For a ball with a long spike in $\R^{3}$, both the volume and surface area, and thus the isoperimetric ratio, can be made arbitrarily close to that of the ball by making the spike narrow enough. On the other hand, the convex hull of such a domain will be approximately a cone with a hemispherical cap, with an isoperimetric ratio significantly greater than $1$: for a spike of length $\eta$ on the unit ball, the isoperimetric ratio of its convex hull will be approximately $\frac{\eta + 3}{4}$ for $\eta$ very large. \\ The outline of this paper and our proof of the isoperimetric inequality is as follows: \\ In Section \ref{variations}, we will calculate the first and second variations of the isoperimetric ratio of the disk. We will show that the disk is a stable critical point of the isoperimetric ratio and that any variation has positive second variation unless, to first order, the variation is the sum of a translation and a rescaling of the disk. \\ In Section \ref{monotonicity}, we will use the $r$-neighborhoods of a compact, convex domain $D$ in the plane to construct a variation of the disk of the type analyzed in Section \ref{variations}. We will use Steiner's formula to relate its first and second variations to the isoperimetric deficit of $D$, and in doing so, we will show that the isoperimetric deficit of $D$ is non-negative. \\ Once we know that the isoperimetric inequality $L^{2} - 4\pi A \geq 0$ holds, any of Steiner's arguments then prove that the disk is the only domain for which equality holds. However, we will show in Section \ref{uniqueness} that the uniqueness of the disk as a minimizing domain also follows from our proof. \\ We will prove that the perimeter $L$ and area $A$ of a plane domain $D$ satisfy $L^{2} \geq 4\pi A$ under the assumption that its boundary $\partial D$ is smooth, and we will make the further simplifying assumption that the curvature of $\partial D$ is strictly positive -- that is, the curvature vector of $\partial D$ always points into $D$ and never vanishes. However by approximation (and the reduction to the convex case) this inequality then follows immediately for any plane domain with a rectifiable boundary. The corresponding issue is more difficult in higher dimensions -- this is discussed in Section 2 of \cite{Os}. In all dimensions, however, the boundary of a compact, convex domain can be realized as the Lipschitz image of a round sphere and is therefore rectifiable. \\ Throughout the paper, we will discuss the relationship between this proof and other known proofs of the isoperimetric inequality. Robert Osserman's article \cite{Os} gives an overview of the isoperimetric inequality, its generalizations and their significance in mathematics. Isaac Chavel's \cite{Ch} and Luis Santal\'o's \cite{San} books both discuss many results and questions in geometry and analysis which are based on the isoperimetric inequality and give several proofs of the classical isoperimetric inequality. Bl\r{a}sj\"o discusses the history of the isoperimetric inequality in \cite{Bl}, and Howards, Hutchings and Morgan in \cite{HHM} and Andrejs Treibergs in \cite{Tr} present several proofs of the classical isoperimetric inequality. \\ {\bf Acknowledgments:} I am very happy to thank Christopher Croke, Joseph H.G. Fu and Peter McGrath for their feedback about this work and Isaac Chavel, Frank Morgan and Franz Schuster for their input about the history of the isoperimetric inequality. \section{The First and Second Variations of the Isoperimetric Ratio} \label{variations} We will calculate the first and second variations of the isoperimetric ratio of the disk for variations through families of convex domains -- in particular, we will see that the disk is a critical point of the isoperimetric ratio and, infinitesimally, a minimizer. \\ A compact, convex domain $D$ can be described by its {\bf support function} $p(\theta) : S^{1} \rightarrow \R$, defined as follows: \begin{align*} \displaystyle p(\theta) = \text{max} \left(\lbrace h_{\theta}(x) := x_0 \cos(\theta) + x_1 \sin(\theta) \ | \ x = (x_0, x_1) \in D \rbrace \right). \end{align*} \smallskip If the boundary $\partial D$ of $D$ is smooth and has strictly positive curvature, then $p(\theta) + p''(\theta)$ is its radius of curvature. In this case, the area $A$ and perimeter $l$ of $D$ are given by: \begin{equation} \label{sup_area} \displaystyle A = (\frac{1}{2})\int\limits_{0}^{2\pi} p(\theta) \left( p(\theta) + p''(\theta) \right) d\theta = (\frac{1}{2})\int\limits_{0}^{2\pi} p(\theta)^{2} - p'(\theta)^{2} d\theta, \end{equation} \begin{equation} \label{sup_perim} \displaystyle l = \int\limits_{0}^{2\pi} p(\theta)d\theta. \end{equation} \smallskip This is described in Chapter 1 of \cite{San}. A variation of the unit disk $\mathcal{D}_{0}$ through a family of such domains $\lbrace \mathcal{D}_{t} \rbrace_{t \geq 0}$ can therefore be described by a smooth function $p(\theta, t)$, with $p(\theta, t)$ the support function of the domain $\mathcal{D}_{t}$. In particular, $p(\theta, 0) \equiv 1$. \begin{proposition} \label{12var} Let $\mathcal{D}_{t}$ be a family of compact, convex domains in the plane, with the boundary $\partial \mathcal{D}_{t}$ of each domain smooth and with positive curvature, which give a variation of the disk $\mathcal{D}_{0}$ as above. Let $I(t)$ be the isoperimetric ratio of the domain $\mathcal{D}_{t}$. \medskip Then $I'(0) = 0$ and $I''(0) \geq 0$, with equality if and only if, to first order, the family of domains coincides with a rescaling and translation of the disk. \end{proposition} \begin{proof} Let $p(\theta,t)$ be the support function of $\mathcal{D}_{t}$ as above. Then letting $A(t)$ be the area and $l(t)$ the perimeter of $\mathcal{D}_{t}$, by (\ref{sup_area}) and (\ref{sup_perim}) we have: \begin{equation} \label{var_area} \displaystyle A(t) = (\frac{1}{2})\int\limits_{0}^{2\pi} p(\theta, t)^{2} - \frac{\partial p}{\partial \theta} (\theta, t)^{2} d\theta, \end{equation} \begin{equation} \label{var_perim} \displaystyle l(t) = \int\limits_{0}^{2\pi} p(\theta, t) d\theta. \end{equation} \smallskip Because $p(\theta, 0) \equiv 1$ and $\frac{\partial p}{\partial \theta}(\theta, 0) \equiv 0$, $A'(0)$ and $l'(0)$ are both equal to $\int_{0}^{2\pi} \frac{\partial p}{\partial t}(\theta, 0) d\theta$. \\ We then have that $I'(0) = \frac{2A(0) l(0) l'(0) - A'(0) l(0)^{2}}{4 \pi A(0)^{2}} $ is equal to: \begin{align*} \displaystyle \frac{2 \times \pi \times 2\pi \left( \int\limits_{0}^{2\pi} \frac{\partial p}{\partial t}(\theta, 0) d\theta \right) - 2\pi \times 2\pi \left( \int\limits_{0}^{2\pi} \frac{\partial p}{\partial t}(\theta, 0) d\theta\right)}{4\pi^{3}} = 0. \end{align*} \smallskip $l''(0)$ is equal to $\int_{0}^{2\pi} \frac{\partial^{2} p}{\partial t^{2}}(\theta, 0) d\theta$ and, using again that $p(\theta, 0) \equiv 1$ and $\frac{\partial p}{\partial \theta}(\theta, 0) \equiv 0$, we have: \begin{equation} \label{area_second_derivative} \displaystyle A''(0) = \int\limits_{0}^{2\pi} \left[ \frac{\partial p}{\partial t}(\theta, 0)^{2} + \frac{\partial^{2} p}{\partial t^{2}}(\theta, 0) - \frac{\partial^{2} p}{\partial t \partial \theta}(\theta, 0)^{2} \right] d\theta. \end{equation} \smallskip We then have that $I''(0) = \frac{ \left(2 A'(0) - l'(0) \right)^{2} + 2\pi \left(l''(0) - A''(0) \right)}{2 \pi^{2}}$ is equal to: \begin{equation} \label{wi} \displaystyle \frac{\left(\int\limits_{0}^{2\pi} \frac{\partial p}{\partial t}(\theta, 0) d\theta \right)^{2} + 2\pi \left(\int\limits_{0}^{2\pi} \frac{\partial^{2} p}{\partial \theta \partial t}(\theta, 0)^{2} - \frac{\partial p}{\partial t}(\theta, 0)^{2} d\theta \right)}{2 \pi^{2}}. \end{equation} \smallskip {\bf Wirtinger's inequality} states that if $\varphi(\theta)$ is a $2\pi$-periodic, continuously differentiable function with $\int_{0}^{2\pi} \varphi(\theta) d\theta = 0$, then: \begin{align*} \displaystyle \text{\Large $\int\limits_{0}^{2\pi}$}\varphi'(\theta)^{2} d\theta \geq \text{\Large $\int\limits_{0}^{2\pi}$} \varphi(\theta)^{2} d\theta. \end{align*} \smallskip Equality holds precisely if $\varphi(\theta) = a_{0}\cos(\theta) + a_{1}\sin(\theta)$ for some constants $a_{0}, a_{1}$. Wirtinger's inequality thus implies by (\ref{wi}) that $I''(0) \geq 0$ and is strictly positive unless $\frac{\partial p}{\partial t}(\theta, 0) = a_{0}\cos(\theta) + a_{1}\sin(\theta) + 2\pi \widehat{p}$, where $\widehat{p} = \frac{1}{2\pi}\int_{0}^{2\pi}\frac{\partial p}{\partial t}(\theta, 0) d\theta$. The variation corresponding to $a_{0}\cos(\theta) + a_{1}\sin(\theta)$ gives a translation of the disk, in the direction whose argument is $\arctan(\frac{a_{1}}{a_{0}})$ at speed $\sqrt{a_{0}^{2} + a_{1}^{2}}$, and the variation corresponding to $2\pi \widehat{p}$ rescales the disk, by a factor $1 + t_{0} 2\pi \widehat{p}$ when $t = t_{0}$. \end{proof} Wirtinger's inequality can be proved by comparing the Fourier series of a $2\pi$-periodic function with that of its derivative, cf. \cite{Fo}. Wirtinger's inequality also implies the isoperimetric inequality directly. This was discovered by Hurwitz, who gave the first proof of the isoperimetric inequality based on Fourier analysis and Wirtinger's inequality in \cite{Hu}. A variant of this proof, in which the role of Wirtinger's inequality is made explicit, can be found in \cite{Os} and \cite{BG}. As with our proof, Hurwitz's proof of the isoperimetric inequality does not require that one separately establish the existence of a minimizing domain -- his argument shows directly that $l^{2} \geq 4\pi A$ for any plane domain, with equality precisely when the domain is a disk. \section{Steiner's Formula and the Monotonicity of the Isoperimetric Ratio} \label{monotonicity} To prove Theorem \ref{ie}, we begin by confirming that the $r$-neighborhoods of a bounded, convex domain $D$, when rescaled to have constant area, give a variation of the disk of the type considered in Proposition \ref{12var}: \begin{proposition} \label{goodvar} Let $D$ be a compact, convex domain in the plane whose boundary is smooth and has positive curvature. For $t > 0$, let $\mathcal{D}_{t}$ be the $r = \frac{1}{t}$-neighborhood of $D$, rescaled to have the same area as $D$, and let $\mathcal{D}_{0}$ be a disk with the same area as $D$. \medskip Then $\lbrace \mathcal{D}_{t} \rbrace_{t \geq 0}$ gives a variation of the disk $\mathcal{D}_{0}$, as in Proposition \ref{12var}. More precisely, if $q(\theta)$ is the support function of $D$, this variation is described by: \begin{equation} \label{var_function} \displaystyle p(\theta, t) = \text{\footnotesize $\sqrt{\frac{A}{A t^{2} + lt + \pi}}$} \left(q(\theta) t + 1 \right), \end{equation} \smallskip where $A$ is the area and $l$ is the perimeter of $D$. \end{proposition} \begin{proof} Let $D$ be as above -- without loss of generality, suppose $D$ has area $\pi$. Note first that each $r$-neighborhood of $D$ is also convex, cf. Remark \ref{ms} below, so that the variation in question is through a family of convex sets. If $q(\theta)$ is the support function of $D$, then $q(\theta) + r$ is the support function of $D_{r}$ and, by Theorem \ref{sf}, $\scriptstyle \sqrt{\frac{\pi}{\pi r^{2} + lr + \pi}}$$(q(\theta) + r)$ is the support function of the rescaling of $D_{r}$ whose area is equal to that of $D$. Rewriting this in terms of $t = \frac{1}{r}$ for $r > 0$, we have: \begin{equation} \label{good_var_eqn} \displaystyle p(\theta, t) = \text{\footnotesize $\sqrt{\frac{\pi}{\pi (\frac{1}{t})^{2} + l\frac{1}{t} + \pi}}$} \left( q(\theta) + \frac{1}{t} \right) = \text{\footnotesize $\sqrt{\frac{\pi}{\pi t^{2} + lt + \pi}}$} \left(q(\theta) t + 1 \right). \end{equation} \smallskip We then have: \begin{equation*} \displaystyle p(\theta,t) + \frac{\partial^{2} p}{\partial \theta^{2}}(\theta,t) = \text{\footnotesize $\sqrt{\frac{\pi}{\pi t^{2} + lt + \pi}}$} \left(t(q(\theta) + q''(\theta)) + 1 \right). \end{equation*} \smallskip Since the curvature of $\partial D$ is positive, $q(\theta) + q''(\theta) > 0$, so for all $t > 0$ we also have that $p(\theta,t) + \frac{\partial^{2} p}{\partial \theta^{2}}(\theta,t) > 0$, and that $\partial \mathcal{D}_{t}$ has positive curvature. $p(\theta,t)$ extends smoothly to $t=0$, where it is equal to the support function of the unit disk, and gives a variation of the disk as in Proposition \ref{12var}. \end{proof} \begin{remark} \label{ms} The $r$-neighborhood $D_{r}$ of a compact, convex set $D$ is the {\bf Minkowski sum} of $D$ with a disk of radius $r$ in $\R^{2}$. Minkowski summation of convex sets is discussed extensively in \cite{Sc} and many other texts on convex and integral geometry. \end{remark} We now prove the inequality in Theorem \ref{ie} -- that for a compact domain in $\R^{2}$ with perimeter $l$ and area $A$, $l^{2} \geq 4\pi A$. We will then address the characterization of the equality case in Section \ref{uniqueness}. \begin{proof}[Proof of Theorem \ref{ie}, Part 1] Let $D$ be a compact, convex domain in the plane with area $A$ and boundary length $l$, and suppose $\partial D$ is smooth and has positive curvature as above. By (\ref{sfir}), for $t > 0$, the isoperimetric ratio $I(t)$ of the $(\frac{1}{t})$-neighborhood of $D$ is: \begin{equation} \label{sfir2} \displaystyle I(t) = \frac{l^{2}t^{2} + 4\pi l t + 4\pi^{2}}{4\pi A t^{2} + 4\pi l t + 4\pi^{2}}. \end{equation} \smallskip Letting $\delta$ be the least absolute value of the roots of $f(t) = 4\pi A t^{2} + 4\pi l t + 4\pi^{2}$, the denominator of (\ref{sfir2}) (see Remark \ref{roots} below), the function of $t$ defined by (\ref{sfir2}) extends smoothly to $(-\delta, \infty)$. In particular, (\ref{sfir2}) extends smoothly to $t = 0$ to give the isoperimetric ratio of the variation $\lbrace \mathcal{D}_{t} \rbrace_{t \geq 0}$ of the disk described in Proposition \ref{goodvar}. $I(t)$ is a monotone function of $t \geq 0$, with the sign of $I'(t)$ determined by the isoperimetric deficit of $D$: \begin{equation} \label{sfird2} \displaystyle I'(t) = \frac{\left( l^{2} - 4\pi A \right)\left( lt^{2} + 2\pi t \right)}{4\pi\left(At^{2} + lt + \pi\right)^{2}}. \end{equation} \smallskip Therefore, $I'(0) = 0$ (which also follows from Propositions \ref{12var} and \ref{goodvar}) and for $t > 0$, $I'(t)$ has the same sign as the isoperimetric deficit of $D$. To show that $l^{2} \geq 4\pi A$, we calculate the second derivative of $I(t)$: \begin{equation} \displaystyle I''(t) = \left( \frac{l^{2} - 4\pi A}{2\pi} \right) \left( \frac{\pi^{2} - 3\pi A t^{2} - Alt^{3}}{(At^{2} + lt + \pi)^{3}} \right). \end{equation} \smallskip In particular, $\displaystyle I''(0) = \frac{l^{2} - 4\pi A}{2\pi^{2}}$. The sign of $l^{2} - 4\pi A$ is the same as that of $I''(0)$, which by Proposition \ref{12var} is greater than or equal to $0$. \end{proof} \begin{remark} \label{roots} The roots of the denominator of (\ref{sfir2}), $f(t) = 4\pi A t^{2} + 4\pi l t + 4\pi^{2}$, are: \begin{equation} \displaystyle \frac{-l \pm \sqrt{l^{2} - 4\pi A}}{2A}. \end{equation} \smallskip The isoperimetric inequality is equivalent to the statement that the roots of this polynomial are real, and thus negative, and are distinct unless the domain in question is a disk. For our purposes, it is enough simply to note that any real roots of $f(t)$ are negative since $f(t) \geq 4\pi^{2}$ when $t \geq 0$. The roots of the Steiner polynomial were studied by Green and Osher in \cite{GO} (the Steiner polynomial of a domain with area $A$ and perimeter $l$ is $\pi r^{2} + lr + A$, with roots $\frac{-l \pm \sqrt{l^{2} - 4\pi A}}{2\pi}$). They note that Steiner's formula implies the isoperimetric deficit of the $r$-neighborhood of $D$ is equal to that of $D$. \end{remark} \section{The Uniqueness of the Disk} \label{uniqueness} Once we have shown that $l^{2} \geq 4\pi A$ for all plane domains with perimeter $l$ and area $A$, and thus that the disk minimizes the isoperimetric ratio, any of Steiner's arguments then show that it is the unique minimizer. The uniqueness of the disk as a minimizing domain for the isoperimetric ratio also follows from our argument, subject to some mild technical assumptions: \begin{proof}[Proof of Theorem \ref{ie}, Part 2] Let $D$ be a bounded domain in the plane with smooth (or $C^{2}$) boundary whose area $A$ and boundary length $l$ satisfy $l^{2} = 4\pi A$. We can suppose $A = \pi$ and $l = 2\pi$. Suppose in addition that the curvature of $\partial D$ is positive, as above. By (\ref{sfir2}), in the variation $\lbrace \mathcal{D}_{t} \rbrace_{t \geq 0}$ of the disk constructed from $D$ as in Section \ref{monotonicity}, the isoperimetric ratio of $\mathcal{D}_{t}$ is equal to $1$ for all $t \geq 0$, and therefore $l(t) \equiv 2\pi$. Therefore, \begin{equation} \label{l_deriv} \displaystyle l'(t) = \int\limits_{0}^{2\pi} \frac{\partial p}{\partial t}(\theta,t) d\theta \equiv 0, \end{equation} \begin{equation} \label{l_second_deriv} \displaystyle l''(t) = \int\limits_{0}^{2\pi} \frac{\partial^{2} p}{\partial t^{2}}(\theta,t) d\theta \equiv 0. \end{equation} \smallskip By (\ref{area_second_derivative}) and (\ref{l_second_deriv}), we then have: \begin{equation} \label{a_second_deriv} \displaystyle \int\limits_{0}^{2\pi} \left[ \frac{\partial p}{\partial t}(\theta, t)^{2} - \frac{\partial^{2} p}{\partial t \partial \theta}(\theta, t)^{2} \right] d\theta = A''(t) \equiv 0. \end{equation} \smallskip By (\ref{l_deriv}), (\ref{a_second_deriv}) and Wirtinger's inequality, $\frac{\partial p}{\partial t}(\theta, t) = c_{0}(t) \cos(\theta) + c_{1}(t) \sin(\theta)$ for some functions $c_{0}(t), c_{1}(t)$ of $t$. Letting $q(\theta)$ be the support function of $D$, by (\ref{var_function}), \begin{equation} \displaystyle \frac{q(\theta) - 1}{(t + 1)^{2}} = c_{0}(t) \cos(\theta) + c_{1}(t) \sin(\theta). \end{equation} \smallskip This then implies that $c_{0}(t) = \frac{d_{0}}{(t + 1)^{2}}$, $c_{1}(t) = \frac{d_{1}}{(t + 1)^{2}}$ for some constants $d_{0}, d_{1}$, and that $q(\theta) = d_{0} \cos(\theta) + d_{1} \sin(\theta) + 1$. $D$ is therefore the unit disk centered at $(d_{0}, d_{1})$. \end{proof} We conclude with a few remarks about the technical assumptions in the proof of the characterization of equality above: \\ We have assumed the domain $D$ to be convex, and to have $C^{2}$ boundary whose curvature is strictly positive, so that it can be described by a $C^{2}$ support function $q(\theta)$. However, by the reduction to the convex case, any domain realizing equality in the isoperimetric inequality must be convex. Moreover, for any compact, convex set $D$ and $r >0$, the $r$-neighborhood $D_{r}$ of $D$ has $C^{1,1}$ boundary, which is therefore twice-differentiable almost everywhere. If $D$ realizes equality in the isoperimetric inequality, then by (\ref{sfir}) each of its $r$-neighborhoods $D_{r}$ does as well, and by the convexity of $D_{r}$, the curvature of $\partial D_{r}$ is non-negative at all points where it is defined. Thus, if one can show that a domain which realizes equality in the isoperimetric inequality, whose boundary is twice-differentiable almost everywhere, and has non-negative curvature at all points where its curvature is defined is a disk, one will have shown that $D_{r}$ is a disk for all $r > 0$, and thus that $D$ is a disk as well. \\ The relationship between the regularity of the boundary of a domain and the regularity of its support function and the smoothness properties of $\partial D_{r}$ are both discussed in \cite{Sc}. Osserman discusses the significance of the regularity assumed on the boundaries of domains in the isoperimetric inequality in Section 2 of \cite{Os}. He notes that one can modify a smooth domain by adding ``wiggles" to its boundary, increasing its perimeter while leaving its area unchanged -- thus, ``one has the ironic situation that the more irregular the boundary, the stronger will be the isoperimetric inequality, but the harder it is to prove. The fact is, the isoperimetric inequality holds in the greatest generality imaginable, but one needs suitable definitions even to state it."
{ "timestamp": "2021-01-15T02:23:43", "yymm": "1909", "arxiv_id": "1909.06347", "language": "en", "url": "https://arxiv.org/abs/1909.06347", "abstract": "We give a new proof of the isoperimetric inequality in the plane, based on Steiner's formula for the area of a convex neighborhood. This proof establishes the isoperimetric inequality directly, without requiring that we separately establish the existence of an optimal domain. In doing so, this proof bypasses the main difficulty in all of the proofs Steiner outlined for the plane isoperimetric inequality.", "subjects": "Differential Geometry (math.DG); Classical Analysis and ODEs (math.CA)", "title": "Steiner's formula and a variational proof of the isoperimetric inequality", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9863631671237733, "lm_q2_score": 0.8128673155708975, "lm_q1q2_score": 0.8017823798379101 }
https://arxiv.org/abs/1202.3033
Orange Peels and Fresnel Integrals
There are two standard ways of peeling an orange: either cut the skin along meridians, or cut it along a spiral. We consider here the second method, and study the shape of the spiral strip, when unfolded on a table. We derive a formula that describes the corresponding flattened-out spiral. Cutting the peel with progressively thinner strip widths, we obtain a sequence of increasingly long spirals. We show that, after rescaling, these spirals tends to a definite shape, known as the Euler spiral. The Euler spiral has applications in many fields of science. In optics, the illumination intensity at a point behind a slit is computed from the distance between two points on the Euler spiral. The Euler spiral also provides optimal curvature for train tracks between a straight run and an upcoming bend. It is striking that it can be also obtained with an orange and a kitchen knife.
\section{Summary Paragraph} {\bf There are two standard ways of peeling an orange: either cut the skin along meridians, or cut it along a spiral. We consider here the second method, and study the shape of the spiral strip, when unfolded on a table. We derive a formula that describes the corresponding flattened-out spiral. Cutting the peel with progressively thinner strip widths, we obtain a sequence of increasingly long spirals. We show that, after rescaling, these spirals tends to a definite shape, known as the Euler spiral. The Euler spiral has applications in many fields of science. In optics, the illumination intensity at a point behind a slit is computed from the distance between two points on the Euler spiral. The Euler spiral also provides optimal curvature for train tracks between a straight run and an upcoming bend. It is striking that it can be also obtained with an orange and a kitchen knife.} \section{Outline} Cut the skin of an orange along a thin spiral of constant width (fig.~\ref{fig:orange}) and lay it flat on a table (fig.~\ref{fig:peel}). A natural breakfast question, for a mathematician, is which shape the spiral peel will have, when flattened out. We derive a formula that, for a given cut width, describes the corresponding spiral's shape. For the analysis, we parametrize the spiral curve by a constant speed trajectory, and express the curvature of the flattened-out spiral as a function of time. \begin{figure}[h] \[ \begin{tikzpicture} \node[scale = .925] at (-.4,0) {\includegraphics[scale=0.15]{sphere.jpg}}; \draw[<->,white,very thick] (0.95,1.45) -- node[right, pos=.3, scale=.97] {$1/N$} (1.6,0.4); \end{tikzpicture} \] \caption{An orange, assumed to be a sphere of radius one, with spiral of width $1/N$.\vspace{-1cm}}\label{fig:orange} \end{figure} This is achieved by comparing a revolution of the spiral on the orange with a corresponding spiral on a cone tangent to the surface of the orange (fig.~\ref{fig:cone}, left). Once we know the curvature, we derive a differential equation for our spiral, which we solve analytically (fig.~\ref{fig:spiral}, left). We then consider what happens to our spirals when we vary the strip width. Two properties are affected: the overall size, and the shape. \begin{figure}[h] \[ \begin{tikzpicture} \node at (0,0) {\includegraphics[scale=0.098]{spiral.jpg}}; \draw[very thick] (2.5,-2) -- (2.5,-2.05) -- node[below, scale=.97] {1\,cm} (2.8,-2.05) -- (2.8,-2); \end{tikzpicture} \] \caption{The flattened-out orange peel.}\label{fig:peel} \end{figure} Taking finer and finer widths of strip, we obtain a sequence of increasingly long spirals; rescale these spirals to make them all of the same size. We show that, after rescaling, the shape of these spirals tends to a well defined limit. The limit shape is a classical mathematical curve, known as the \emph{Euler spiral} or the \emph{Cornu spiral} (fig.~\ref{fig:spiral}, right). This spiral is the solution of the \emph{Fresnel integrals}. The Euler spiral has many applications. In optics, it occurs in the study of light diffracting through a slit~\cite{hecht:optics}*{\S10.3.8}. More precisely, the illumination intensity at a point behind a slit is the square of the distance between two points on the Euler spiral, easily determined from the slit's geometry. \begin{figure}[h] \[ \begin{tikzpicture} \draw (0,0) circle (1); \coordinate (c) at (0,2); \begin{scope} \pgftransformcm{1}{0}{0}{.3}{\pgfpoint{0}{0}} \draw (-1,0) arc (-180:0:1); \pgftransformcm{1.07}{0}{0}{1.07}{\pgfpoint{0}{12}} \draw[densely dotted] (-1,0) arc (175:5:1); \foreach \x in {180, 188.5, ..., 360}{\draw[gray!50] (\x:1) -- (c);} \draw (-190:1) arc (-190:10:1) -- (c) -- cycle; \pgftransformcm{.9}{0}{0}{.9}{\pgfpoint{0}{16}} \pgftransformcm{.95}{.1}{0}{1}{\pgfpoint{-.4}{8.2}} \draw[red, densely dotted] (-110:1) arc (-110:-90:1); \draw[red] (-90:1) arc (-90:10:1); \pgftransformcm{.95}{0}{0}{.95}{\pgfpoint{0}{8.2}} \draw[red] (-190:1) arc (-190:10:1); \draw[red] (-190:1) arc (-190:10:1); \pgftransformcm{.95}{0}{0}{.95}{\pgfpoint{0}{8.2}} \draw[red] (-190:1) arc (-190:10:1); \pgftransformcm{.95}{0}{0}{.95}{\pgfpoint{0}{8.2}} \draw[red] (-190:1) arc (-190:-50:1); \draw[red, densely dotted] (-50:1) arc (-50:-30:1); \end{scope} \draw[<->] (0,2)+(30:.1) -- node[scale=.85, right, pos=.45]{$\scriptstyle R$} (30:1.1); \draw[->,decorate,decoration={snake,post length=.5mm,amplitude=.5mm,segment length=2mm}] (0,-1.3) -- (0,-1.75); \begin{scope}[xshift=-3.95cm,yshift=-2.6cm] \foreach \x in {-15, -22.5, ..., -165}{\draw [gray!50] (4,.5) -- +(\x:1.5);} \draw (4,.5) -- +(-10:1.5) arc (-10:-170:1.5) -- cycle; \draw [red] (3.95,.5) +(-80:1.2) arc (-80:-9.5:1.2); \draw [red] (3.95,.5) +(-169.5:1.1) arc (-169.5:-9.5:1.1); \draw [red] (3.95,.5) +(-169.5:1.0) arc (-169.5:-9.5:1.0); \draw [red] (3.95,.5) +(-169.5:.9) arc (-169.5:-60:.9); \draw[red, densely dotted] (3.95,.5) +(-60:.9) arc (-60:-45:.9); \draw[red, densely dotted] (3.95,.5) +(-96:1.2) arc (-96:-80:1.2); \draw[<->] (4,.5) -- node[scale=.85, right, pos=.5]{$\scriptstyle R$} +(-78.7:1.09); \end{scope} \begin{scope}[xshift=3cm,yshift=-1.75cm] \pgftransformcm{.93}{0}{0}{.93}{\pgfpoint{0}{0}} \clip (-1,-2.02) rectangle (2.1,4.2); \draw (0,0) circle (2cm); \draw (0,0) -- (0,1); \draw[<->] (-.14,0) -- node[left, scale=.93] {$s$} (-.14,1); \draw (0,1) -- (0,4); \draw (0,1) -- node[above, xshift=-5,yshift=-2, scale=.9] {$\sqrt{1-s^2}$} (30:2); \draw (0,0) -- node[below, scale=.93] {$1$} (30:2); \draw (30:2) -- (0,4); \filldraw (0,0) circle (.03); \pgftransformxshift{3.6} \pgftransformyshift{1.8} \draw[<->] (30:2) -- node[above,sloped, scale=.9] {$R=\sqrt{1-s^2\,}\!\big/s$} (0,4); \end{scope} \end{tikzpicture} \] \caption{\emph{left:} Spiral on the sphere, transferred to the tangent cone, and developed on the plane, for computing its radius of curvature;\\ \emph{right:} The computation of the radius of curvature $R$ of the flattened spiral.} \label{fig:cone} \end{figure} The same spiral is also used in civil engineering: it provides optimal curvature for train tracks~\cite{profillidis:railway}*{\S14.1.2}. A train that travels at constant speed and increases the curvature of its trajectory at a constant rate will naturally follow an arc of the Euler spiral. The review~\cite{Levien:EECS-2008-111} describes the history of the Euler spiral and its three independent discoveries. \begin{figure}[h] \[ \begin{tikzpicture}[scale=0.8] \useasboundingbox (-2.2,-2.1) rectangle (1.7,1.9); \node at (-0.3,-0.1) {\includegraphics[scale=0.8]{solution.pdf}}; \draw[->] (0,-.3) node[anchor=north, scale=.6] {$t=0$} -- (0,0); \draw[->] (-0.8,-1.24) node[anchor=west, scale=.6] {$t=-2\pi N$} -- (-1.31,-1.24); \draw[->] (0.65,1.21) node[anchor=east, scale=.6] {$t=2\pi N$} -- (1.17,1.21); \end{tikzpicture}\quad\,\,\, \begin{tikzpicture}[scale=0.865] \useasboundingbox (-2.3,-2.0) rectangle (1.6,1.8); \pgftransformyshift {-1.7} \node at (-0.4445,-0.1) {\includegraphics[scale=0.865]{euler.pdf}}; \fill[red] (1.05,1.12) circle (.1) (-1.19,-1.13) circle (.1); \end{tikzpicture} \] \caption{\emph{left:} Maple plot of the orange peel spiral ($N=3$);\\ \emph{right:} the Euler spiral; limit $N \to \infty$.}\label{fig:spiral} \end{figure} \section{Analysis} For the purpose of our mathematical treatment, we shall replace the orange by a sphere of radius one. The spiral on the sphere is taken of width $1/N$, see (fig.~\ref{fig:orange}). The area of the sphere is $4\pi$, so the spiral has a length of roughly $4\pi N$. We describe the flattened-out orange peel spiral by a curve $(x(t),y(t))$ in the plane, parameterized at unit-speed from time $t=-2\pi N$ to $t=2\pi N$. On a sphere of radius one, the area between two horizontal planes at heights $h_1$ and $h_2$ is ${2\pi(h_2-h_1)}$, see (fig.~\ref{fig:area}). It follows that, at time $t$, the point on the sphere has height $s:=t/2\pi N$. \begin{figure}[h] \[ \begin{tikzpicture} \draw (0,0) circle (2cm); \draw (30:2) -- (150:2); \draw (36:2) -- (144:2); \fill[gray!20!white] (150:2) arc (150:144:2cm) -- (36:2) arc (36:30:2cm) -- cycle; \draw[->] (0,2.8) node[anchor=south] {Perimeter $\approx 2\pi\sqrt{1-s^2}$} -- (0,0 |- -36:-2); \draw[->] (-2.5,1.8) node[anchor=south] {Width $\approx \epsilon/\sqrt{1-s^2}$} -- (147:2); \draw[->] (-2.5,0) -- node[left] {$h_1$} (-2.5,0 |- 30:2); \draw[<->] (-2.5,0 |- 30:2) -- node[left] {$\epsilon$} (-2.5,0 |- 36:2); \draw[->] (-2.3,0) -- node[right, yshift=-2.5, xshift=2, fill=white, inner sep=1] {$h_2=h_1+\epsilon$} (-2.3,0 |- 36:2); \draw[dashed] (-2.9,0) -- (1.5,0); \draw[->] (.5,0) --node[anchor=west, fill=white, inner sep=1, xshift=3] {Height $\approx s$} (.5,0 |- 33:2); \node[fill=white, inner sep=2] at (-1.38,-1) {Area = Width $\times$ Perimeter $\approx 2\pi\epsilon$}; \filldraw (0,0) circle (.03); \end{tikzpicture} \] \caption{Area of a thin circular strip on the sphere.} \label{fig:area} \end{figure} Our first goal is to find a differential equation for $(x(t),y(t))$. For that, we compute the radius of curvature $R(t)$ of the flattened-out spiral at time $t$: this is the radius of circle with best contact to the curve at time $t$. For example, $R(-2\pi N)=R(2\pi N)=0$ at the poles, and $R(0)=\infty$ at the equator. For $N$ large, the spiral at time $t$ follows roughly a parallel at height $s$ on the orange. The surface of the sphere can be approximated by a tangent cone whose development on the plane is a disk sector (fig.~\ref{fig:cone}, left). The radius \[ R(t)=\sqrt{1-s^2}/s=\sqrt{(2\pi N)^2-t^2}/t \] of that disk equals the radius of curvature of the spiral at time $t$, and can be computed using Thales' theorem (fig.~\ref{fig:cone}, right). The radius $R(t)$ is in fact only determined up to sign; our choice reflects the NE-SW orientation of the spiral on the sphere. Now, the condition that we move move at unit speed on the sphere --- and on the plane --- is $(\dot x)^2+(\dot y)^2=1$, and the condition that the spiral has a curvature of $R(t)$ is $\dot x\ddot y-\ddot x\dot y=1/R$. Here, $\dot x$ and $\dot y$ are the speeds of $x$ and $y$ respectively, and $\ddot x$ and $\ddot y$ are their accelerations. In fact, introducing the complex path $z(t)=x(t)+iy(t)$, the conditions can be expressed as $|\dot z|^2=1$ and $\ddot z \dot {\bar z}=i/R$. The solution has the general form \[z(t)=\int_0^t\exp(i\phi(u))du,\] for a real function $\phi$; indeed, its derivative is computed as $\dot z=\exp(i\phi(t))$ and has norm $1$. As $\ddot z\dot{\bar z}=i\dot\phi(t)$, we have $\dot\phi(t)=s/\sqrt{1-s^2}$, which has as elementary solution $\phi(t)=-\sqrt{(2\pi N)^2-t^2}$. We have deduced that the flattened-out spiral has parameterization \[\left\{\begin{array}{l} \displaystyle x(t)\,=\,\,\int_0^t\cos\sqrt{(2\pi N)^2-u^2}du,\\ \displaystyle y(t)\,=\,-\int_0^t\sin\sqrt{(2\pi N)^2-u^2}du. \end{array}\right.\] The flattened-out peel of an orange is shown in (fig.~\ref{fig:peel}), and the corresponding analytic solution, computed by \textsc{Maple}~\cite{Maple10}, is shown in (fig.~\ref{fig:spiral}, left). The orange's radius was 3cm, and the peel was 1cm wide, giving $N=3$. \section{Limiting behaviour} What happens if $N$ tends to infinity, that is, if we peel the orange with an ever thinner spiral? For that, we recall the power series approximation \[ \sqrt{a^2-u^2}=a-\frac{u^2}{2a}+\mathcal O(\frac{u^4}{a^3}), \] which we substitute with $a=2\pi N$ in the above expression: \begin{align*} z(t)&=\int_0^t\exp\Big(\!-i\sqrt{(2\pi N)^2-u^2}\,\Big)du\\& \approx\int_0^t\exp \Big(\!-i\Big(2\pi N-\frac{u^2}{2\cdot2\pi N}\Big)\!\Big)du. \end{align*} Taking only values of $N$ that are integers, this simplifies to $\int_0^t\exp(iu^2/4\pi N)du$. We then set $v=u/\sqrt{4\pi N}$ to obtain \[ z(t)\approx \sqrt{4\pi N}\int_0^{t/\sqrt{4\pi N}}\exp(iv^2)dv. \] The approximation error is $\int_0^t\mathcal O(\frac{u^4}{a^3})du=\mathcal O(t^5/N^3)$, which becomes negligible compared to the size $\mathcal O(\sqrt N)$ of the spiral for $|t|\ll N^{0.7}$. The above curve is, up to scaling and parameterization speed, the solution of the classical Fresnel integral \[(X(t),Y(t))=\left(\int_0^t\cos u^2du,\int_0^t\sin u^2du\right),\] defined by the condition that the radius of curvature at time $t$ is $1/2t$; here the parameterization is over $t$ from $-\infty$ to $+\infty$. The corresponding curve is called the Euler spiral and winds infinitely often around the points $\pm(\sqrt{\frac\pi8},\sqrt{\frac\pi8})$. Setting $T:=t/\sqrt{4\pi N}$, the condition $|t|\ll N^{0.7}$ becomes $|T|\ll N^{0.2}$. We have thus proven: \begin{theorem} If $T\ll N^{0.2}$, then the part of the orange peel of width $1/N$ parameterized between $-\sqrt{4\pi N}\,T$ and $\sqrt{4\pi N}\,T$ is a good approximation for the part of the Euler spiral parameterized between $-T$ and $T$. \end{theorem} \noindent Note that for large $N$, the piece of the orange peel parameterized between $-\sqrt{4\pi N}\,T$ and $\sqrt{4\pi N}\,T$ forms a rather thin band around the orange's equator. The contribution of rest of the orange disappears due to the rescaling process. \section{Conclusion} The Euler spiral is a well known mathematical curve. In this article, we explained how to construct it with an orange and a kitchen knife. Flattened fruit peels have already been considered, e.g.\ those of apples~\cite{Turrell}, but were never studied analytically. The Euler spiral that we obtained has had many discoveries across history~\cite{Levien:EECS-2008-111}; ours occurred over breakfast. \begin{bibdiv} \begin{biblist} \bib{hecht:optics}{book}{ author={Hecht, Eugene}, title={Optics}, date={2002}, publisher={Pearson Educat.}, edition={4th} } \bib{Levien:EECS-2008-111}{techreport}{ Author = {Levien, Raph}, Title = {The Euler spiral: a mathematical history}, Institution = {EECS Department, University of California, Berkeley}, Year = {2008}, URL = {http://www.eecs.berkeley.edu/Pubs/TechRpts/2008/EECS-2008-111.html}, Number = {UCB/EECS-2008-111}, } \bib{Maple10}{book}{ author = {Michael B.~Monagan}, author = {Keith O.~Geddes}, author = {K.~Michael Heal}, author = {George Labahn}, author = {Stefan M.~Vorkoetter}, author = {James McCarron}, author = {Paul DeMarco}, title = {Maple~10 Programming Guide}, publisher = {Maplesoft}, year = {2005}, address = {Waterloo ON, Canada}, } \bib{profillidis:railway}{book}{ author={Profillidis, Vassilios A.}, title={Railway management and engineering}, date={2006}, publisher={Ashgate Publishing Ltd.}, pages={469} } \bib{Turrell}{article}{ author={F. M. Turrell}, title={The definite integral symbol}, journal = {Amer. Math. Monthly}, volume = {67}, year = {1960}, number = {7}, pages = {656--658} } \end{biblist} \end{bibdiv} \end{document}
{ "timestamp": "2012-02-15T02:02:37", "yymm": "1202", "arxiv_id": "1202.3033", "language": "en", "url": "https://arxiv.org/abs/1202.3033", "abstract": "There are two standard ways of peeling an orange: either cut the skin along meridians, or cut it along a spiral. We consider here the second method, and study the shape of the spiral strip, when unfolded on a table. We derive a formula that describes the corresponding flattened-out spiral. Cutting the peel with progressively thinner strip widths, we obtain a sequence of increasingly long spirals. We show that, after rescaling, these spirals tends to a definite shape, known as the Euler spiral. The Euler spiral has applications in many fields of science. In optics, the illumination intensity at a point behind a slit is computed from the distance between two points on the Euler spiral. The Euler spiral also provides optimal curvature for train tracks between a straight run and an upcoming bend. It is striking that it can be also obtained with an orange and a kitchen knife.", "subjects": "History and Overview (math.HO)", "title": "Orange Peels and Fresnel Integrals", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9863631639168357, "lm_q2_score": 0.8128673178375734, "lm_q1q2_score": 0.8017823794668609 }
https://arxiv.org/abs/2004.02437
A historical note on the 3/2-approximation algorithm for the metric traveling salesman problem
One of the most fundamental results in combinatorial optimization is the polynomial-time 3/2-approximation algorithm for the metric traveling salesman problem. It was presented by Christofides in 1976 and is well known as "the Christofides algorithm". Recently, some authors started calling it "Christofides-Serdyukov algorithm", pointing out that it was published independently in the USSR in 1978. We provide some historic background on Serdyukov's findings and a translation of his article from Russian into English.
\section{Introduction} \noindent One of the most fundamental problems in combinatorial optimization is the traveling salesman problem, formalized as early as 1832 \citep[cf.][Chapter~1]{ABCC06}: given $n$~cities and their pairwise distances, find a shortest tour{} to visit each city exactly once and return to the starting point. Finding the \emph{shortest} tour{} is a computationally intractable problem even in the special case where the distances between the cities satisfy the triangle inequality \citep{GJ79}. \citet{Chr76} presented an $O(n^3)$-time \emph{3/2\hyp approximation algorithm} for this special case: it yields a tour{} that is at most 3/2 times longer than the shortest one. It % is a prime example for approximation algorithms that entered textbooks and encyclopedias as ``the Christofides algorithm'' or ``the Christofides heuristic'' \citep{GJ79,Chr79,Gut09,WS11,Bla16}. Quite some efforts have been made trying to improve it \citep[cf.\ the surveys of][]{Vyg12,Sve13}. One line of research aims for improving its running time: there are many faster heuristics, which cannot guarantee 3/2-approximate solutions \citep{JM07}, yet $(3/2+\varepsilon)$-approximate solutions for any~$\varepsilon>0$ are computable by a randomized algorithm in $O(n^2\log^4n/\varepsilon^2)$~time \citep{CQ17}. Another line of research aims for improving the approximation factor, which was successful only in special cases: in polynomial time, one can compute 8/7-approximate solutions if the distances are in~$\{1,2\}$ \citep{BK06}, 7/5-approximate solutions if the distances are lengths of shortest paths in an unweighted graph \citep{SV14}, and $(1+\varepsilon)$\hyp approximate solutions for any fixed~$\varepsilon>0$ if the cities are points in fixed\hyp dimensional Euclidean \citep{Aro98,Mit99} or doubling spaces \citep[using a randomized algorithm]{BGK16}, or if the distances are lengths of shortest paths in a graph excluding some fixed minor \citep{DHK11}. For general distances satisfying the triangle inequality, the 3/2 approximation factor of \citeauthor{Chr76}' algorithm remains the state of the art.\footnote{Actually, \citet{Wol80} showed that the length of the computed tours is within a factor 3/2 not only of the optimum, but of a lower bound given by the optimal solution to a relaxation of an integer linear programming model.} Recently, a small but growing group of authors started referring to it as ``the Christofides-Serdyukov algorithm'' \citep{BDW98,DT10,BBMV19,GW19,SZ19,Tar19,TV19},% \footnote{We deliberately omit articles coauthored by Serdyukov's former colleagues from this list.} claiming that it was independently obtained in the USSR by \citet{Ser78}. \looseness=-1 At the one hand, this claim is plausible: in the beginning of the 70s, a~lot of research on computationally intractable problems was carried out parallely in the USSR, leading to independent proofs of seminal results like the Cook-Levin theorem about the NP\hyp completeness of the satisfiability problem for Boolean formulas \citep{Tra84}. Moreover, the submission date given in the journal article of \citet{Ser78}, January 27th, 1976, predates the report of \citet{Chr76}, dated February, 1976. On the other hand, such claims should be treated with caution: for example, the wide\hyp spread claim that Kuratovski's theorem was earlier proved in the USSR has little support \citep{KQS85}. We give some historic background on \citeauthor{Ser78}'s findings, which indeed supports the claim of his independent discovery of the 3/2\hyp approximation algorithm and sheds some light on the timely coincidence of the publications of \citeauthor{Chr76} and \citeauthor{Ser78}. We also provide a translation of \citeauthor{Ser78}'s article in the appendix. \section{Anatoliy I.\ Serdyukov (1951--2001)} \noindent The following information about Serdyukov can be found in \citet{BGS07}, \citet{Dem17},% \footnote{The birth year 1952 given in the book edited by \citet{Dem17} is incorrect.} the archives of the Department of Mechanics and Mathematics of Novosibirsk State University, and the State Public Scientific Technological Library of the Siberian Branch of the Russian Academy of Sciences. Anatoliy Ivanovich Serdyukov was born on October 29th, 1951, in Prokopyevsk, a city in Kemerovo region (Western Siberia), USSR. He graduated from Novosibirsk State University in 1973, after which he was employed in the structures of the Siberian Branch of the Lenin Academy of Agricultural Sciences, then at the Institute of Cytology and Genetics of the Siberian Branch of the Academy of Sciences of the USSR (SB AS USSR), and finally at the Institute of Mathematics of the SB AS USSR (now named the Sobolev Institute of Mathematics, Siberian Branch of the Russian Academy of Sciences), where he was working until his death on February 7th, 2001. In 1980, already working at the Institute of Mathematics, he was awarded the academic degree of candidate of physico\hyp mathematical sciences. \citeauthor{Ser80}'s \citeyearpar{Ser80} thesis is on the complexity of finding Hamiltonian and Eulerian cycles in graphs. His best known results are approximation algorithms for finding \emph{longest} traveling salesman tours \citep[surveyed by][]{BGS07}. Taking into account his graduation year and the submission date of \citeauthor{Ser78}'s \citeyearpar{Ser78} article, January 27th, 1976, \citeauthor{Ser78} must have obtained his 3/2\hyp approximation algorithm as a young graduate student in about 1975. \section{Circulation of Christofides' result between 1976 and 1979} \noindent Authors usually refer to \citeauthor{Chr76}' \citeyear{Chr76} technical report at Carnegie-Mellon University (CMU) as the source of the 3/2\hyp approximation algorithm for the metric traveling salesman problem, which some authors do not consider as published \citep[cf.][who also claims that ``Christofides never published his algorithm'']{Bla16}. \looseness=-1 Apparently, \citeauthor{Chr76}' technical report was not known to a wide audience up to 1978. % For example, \citet{Kar77} and \citet{RSL77} refer to Christofides' abstract in the proceedings of a symposium held at CMU in April 1976. The proceedings were published only in December 1976 \citep{Tra76}. \citet{FHK76} refer to the same abstract, whereas later, in the journal version of their article, \citet{FHK78} refer to the technical report. \citeauthor{Chr76}' technical report could have been popularized in 1977, when its abstract stating the 3/2-approximation was indexed by the NASA abstract journal Scientific and Technical Aerospace Reports \citep{Chr77}. Some authors of that time, for example \citet{LR79}, refer to a journal article of Christofides that is to appear in the journal \emph{Mathematical Programming}. In a combinatorial optimization textbook, \citet{Chr79} describes his algorithm without proving the approximation factor, referring to an article in press in \emph{Mathematical Programming} for the proof, not mentioning his technical report. Interestingly, according to the archives of \emph{Mathematical Programming}, his article was not published. The algorithm with complete proof details was published to a wide audience not later than in the seminal textbook of \citet{GJ79}. Summarizing, \citet{Ser78} submitted his journal article in January 1976, which predates all traces of \citeauthor{Chr76}' publications on this topic. Thus, it is plausible that \citet{Ser78} obtained the result independently. \section{Serdyukov's work between 1974 and 1978} \noindent \looseness=-1 We give some historic background on the findings of \citeauthor{Ser78} to shed some light on the timely coincidence of the publications of \citet{Chr76} and \citet{Ser78}. To this end, it is helpful to interpret the 3/2-approximation algorithm for the traveling salesman problem as follows: A first step computes a minimum\hyp cost spanning tree that connects all the cities. A second step computes a shortest tour{} in the input graph that traverses the edges of the spanning tree. The second step is solved using an approach earlier developed for the closely related Chinese postman problem of computing a shortest tour{} traversing \emph{all} edges of a graph: \citet{Chr73} and \citet{Ser74}, but also \citet{EJ73}, actively study the Chinese postman problem at that time. They all reduce it to the problem of finding a minimum\hyp cost perfect matching on the complete edge\hyp weighted graph on all odd\hyp degree vertices of the input graph.% \footnote{Notably, \cite{Ser74} explicitly introduces the problem that forty years later is intensively studied as the Eulerian extension problem \citep{SBNW11,HJM12,SBNW12,DMNW13,GWY17,BFTxx}.} Surprisingly, while \citeauthor{Chr73}, \citeauthor{EJ73} solve the matching problem using the polynomial\hyp time algorithm of \citet{Edm65b}, \citeauthor{Ser74} reduces it to an exponential number of matching problems in bipartite graphs.% \footnote{In contemporary terms of parameterized complexity theory \citep[cf.][]{CFK+15}, \cite{Ser74} merely describes a fixed\hyp parameter algorithm for the Chinese postman problem parameterized by the number of odd\hyp degree vertices in the input graph.} Apparently, in 1974 neither \citeauthor{Ser74} nor his reviewers were aware of the work of \citet{Chr73}, \citet{EJ73}, or the polynomial\hyp time algorithm for computing maximum\hyp weight matchings in general graphs, published by \citeauthor{Edm65b} nine years earlier. Since \citet{Ser78} uses \citeauthor{Edm65b}' algorithm to solve the matching problem in his 3/2\hyp approximation algorithm for the traveling salesman problem but was unaware of it in 1974, he must have learned about \citeauthor{Edm65b}' algorithm in 1974 or 1975. One scenario is that he learned about it via the article of \citet{Chr73}, which \citet{Ser76} cites in an article studying reductions between matching, covering, the Chinese postman, and the traveling salesman problems. In this scenario, \citeauthor{Ser78} obtained his 3/2\hyp approximation independently of \citeauthor{Chr73} but because of him. Another scenario is that \citet{Ser78} learned about \citeauthor{Edm65b}' algorithm from \citet{Kar76}, whose $O(n^3\log n)$\hyp time implementation of \citeauthor{Edm65b}' algorithm he uses in his 3/2\hyp approximation. \citeauthor{Kar76}'s article was probably not yet published in January 1976, when \citeauthor{Ser78} submitted his article, but he might have had access to a preliminary copy, which is supported by the fact that the titles given by \citeauthor{Ser78} and \citeauthor{Kar76} differ slightly. \section{Conclusion} \noindent Our findings support the claim that \citet{Ser78} discovered the 3/2\hyp approximation algorithm for the metric traveling salesman problem independently of \citet{Chr76}. Concerning the timely coincidence of the publications of \citeauthor{Chr76} and \citeauthor{Ser78}, we conclude that, on the one hand, it was impossible for \citeauthor{Ser78} to find the algorithm much earlier than \citeauthor{Chr76}, being unaware of \citeauthor{Edm65b}' polynomial-time matching algorithm up to 1974. On the other hand, actively working on the Chinese postman before, he found the 3/2\hyp approximation for the traveling salesman problem as soon as he became aware of \citeauthor{Edm65b}' algorithm. \looseness=-1 An English abstract of \citeauthor{Ser78}'s \citeyearpar{Ser78} article was indexed in zbMATH only in 1982 \citep{Ser82}. At~this time, ``the \citeauthor{Chr76} algorithm'' had already entered fundamental textbooks like that of \citet{GJ79}. Moreover, the English abstract does not mention any approximation factors. Thus, it is not surprising that \citeauthor{Ser78}'s result remained largely unknown beyond the USSR. \paragraph{Acknowledgments} We thank Edward Kh.\ Gimadi and Oxana Yu.\ Tsidulko for helpful input. \paragraph{Funding} René van Bevern is supported by the Mathematical Center in Akademgorodok, agreement No.\ 075-15-2019-1675 with the Ministry of Science and Higher Education of the Russian Federation. Viktoriia A.\ Slugina is supported by grant No.\ 19-39-60006 of the Russian Foundation for Basic Research. \bibliographystyle{tsp-history}
{ "timestamp": "2020-04-28T02:07:16", "yymm": "2004", "arxiv_id": "2004.02437", "language": "en", "url": "https://arxiv.org/abs/2004.02437", "abstract": "One of the most fundamental results in combinatorial optimization is the polynomial-time 3/2-approximation algorithm for the metric traveling salesman problem. It was presented by Christofides in 1976 and is well known as \"the Christofides algorithm\". Recently, some authors started calling it \"Christofides-Serdyukov algorithm\", pointing out that it was published independently in the USSR in 1978. We provide some historic background on Serdyukov's findings and a translation of his article from Russian into English.", "subjects": "Data Structures and Algorithms (cs.DS); Discrete Mathematics (cs.DM); History and Overview (math.HO); Optimization and Control (math.OC)", "title": "A historical note on the 3/2-approximation algorithm for the metric traveling salesman problem", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9724147201714922, "lm_q2_score": 0.8244619220634457, "lm_q1q2_score": 0.8017189092353761 }
https://arxiv.org/abs/2101.08898
Consecutive primes which are widely digitally delicate
We show that for every positive integer $k$, there exist $k$ consecutive primes having the property that if any digit of any one of the primes, including any of the infinitely many leading zero digits, is changed, then that prime becomes composite.
\section{\@startsection {section}{1}{\z@} {-30pt \@plus -1ex \@minus -.2ex} {2.3ex \@plus.2ex} {\normalfont\normalsize\bfseries\boldmath}} \renewcommand\subsection{\@startsection{subsection}{2}{\z@} {-3.25ex\@plus -1ex \@minus -.2ex} {1.5ex \@plus .2ex} {\normalfont\normalsize\bfseries\boldmath}} \renewcommand{\@seccntformat}[1]{\csname the#1\endcsname. } \makeatother \newtheorem{theorem}{Theorem} \newtheorem{lemma}{Lemma} \newtheorem{conjecture}{Conjecture} \newtheorem{proposition}{Proposition} \newtheorem{corollary}{Corollary} \newtheorem{definition}{Definition} \setlength{\footskip}{35pt} \allowdisplaybreaks \begin{document} \begin{center} {\Large \bf Consecutive primes} \vskip 5pt {\Large \bf which are widely digitally delicate} \vskip 20pt {\bf Michael Filaseta}\\ {\smallit Dept.~Mathematics, University of South Carolina, Columbia, SC 29208, USA}\\ {\tt filaseta@math.sc.edu}\\ \vskip 20pt {\bf Jacob Juillerat}\\ {\smallit Dept.~Mathematics, University of South Carolina, Columbia, SC 29208, USA}\\ {\tt juillerj@email.sc.edu}\\ \end{center} \vskip 5pt \centerline{\phantom{\smallit Received: , Revised: , Accepted: , Published: }} \vskip 12pt \centerline{\textit{Dedicated to the fond memory of Ronald Graham}} \vskip 15pt \centerline{\bf Abstract} \vskip 5pt\noindent We show that for every positive integer $k$, there exist $k$ consecutive primes having the property that if any digit of any one of the primes, including any of the infinitely many leading zero digits, is changed, then that prime becomes composite. \pagestyle{myheadings} \thispagestyle{empty} \baselineskip=12.875pt \vskip 30pt \section{Introduction} In 1978, M.~S.~Klamkin \cite{klamkin} posed the following problem. \vskip 5pt \centerline{\parbox[t]{12cm}{\textit{Does there exist any prime number such that if any digit (in base $10$) is changed to any other digit, the resulting number is always composite? }}} \vskip 8pt\noindent In addition to computations establishing the existence of such a prime, the published solutions in 1979 to this problem included a proof by P.~Erd\H{o}s \cite{Erd79} that there exist infinitely many such primes. Borrowing the terminology from J.~Hopper and P.~Pollack \cite{hopperpollack}, we call such primes \textit{digitally delicate}. The first digitally delicate prime is $294001$. Thus, $294001$ is a prime and, for every $d \in \{ 0, 1, \ldots, 9 \}$, each of the numbers \[ d\hspace{.1em}94001, \quad 2d4001, \quad 29d\hspace{.1em}001, \quad 294d\hspace{.1em}01, \quad 2940d1, \quad 29400d \] is either equal to $294001$ or composite. The proof provided by Erd\H{o}s consisted of creating a partial covering system of the integers (defined in the next section) followed by a sieve argument. In 2011, T.~Tao \cite{tao} showed by refining the sieve argument of Erd\H{o}s that a positive proportion (in terms of asymptotic density) of the primes are digitally delicate. In 2013, S.~Konyagin \cite{konyagin} pointed out that a similar approach implies that a positive proportion of composite numbers $n$, coprime to $10$, satisfy the property that if any digit in the base $10$ representation of $n$ is changed, then the resulting number remains composite. For example, the number $n=212159$ satisfies this property. Thus, every number in the set \[\{d12159, 2d2159, 21d159, 212d59, 2121d9, 21215d: d\in\{0,1, 2, \dots,9\}\}\] is composite. Later, in 2016, J.~Hopper and P.~Pollack \cite{hopperpollack} resolved a question of Tao's on digitally delicate primes allowing for an arbitrary but fixed number of digit changes to the beginning and end of the prime. All of these results and their proofs hold for numbers written in an arbitrary base $b$ rather than base $10$, though the proof provided by Erd\H{o}s \cite{Erd79} only addresses the argument in base $10$. In 2020, the first author and J.~Southwick \cite{filsou} showed that a positive proportion of primes $p$, are \textit{widely digitally delicate}, which they define as having the property that if any digit of $p$, \textit{including any one of the infinitely many leading zeros of $p$}, is replaced by any other digit, then the resulting number is composite. The proof was specific to base $10$, though they elaborate on other bases for which the analogous argument produces a similar result, including for example base $31$; however, it is not even clear whether widely digitally delicate primes exist in every base. Observe that the first digitally delicate prime, 294001, is not widely digitally delicate since 10294001 is prime. It is of some interest to note that even though a positive proportion of the primes are widely digitally delicate, no specific examples of widely digitally delicate primes are known. Later in 2020, the authors with J.~Southwick \cite{filjuisou} gave a related argument showing that there are infinitely many (not necessarily a positive proportion) of composite numbers $n$ in base $10$ such that when any digit is inserted in the decimal expansion of $n$, including between two of the infinitely many leading zeros of $n$ and to the right of the units digit of $n$, the number $n$ remains composite (see also \cite{Fil10}). In this paper, we show the following. \begin{theorem}\label{maintheorem} For every positive integer $k$, there exist $k$ consecutive primes all of which are widely digitally delicate. \end{theorem} Let $\mathcal P$ be a set of primes. It is not difficult to see that if $\mathcal P$ has an asymptotic density of $1$ in the set of primes, then there exist $k$ consecutive primes in $\mathcal P$ for each $k \in \mathbb Z^{+}$. On the other hand, for every $\varepsilon \in (0,1)$, there exists $\mathcal P$ having asymptotic density $1-\varepsilon$ in the set of primes such that there do not exist $k$ consecutive primes in $\mathcal P$ for $k$ sufficiently large (more precisely, for $k \ge 1/\varepsilon$). Thus, the prior results stated above are not sufficient to establish Theorem~\ref{maintheorem}. The main difficulty in using the prior methods to obtain Theorem~\ref{maintheorem} is in the application of sieve techniques in the prior work. We want to bypass the use of sieve techniques and instead give complete covering systems to show that there is an arithmetic progression containing infinitely many primes such that every prime in the arithmetic progression is a widely digitally delicate prime. This then gives an alternative proof of the result in \cite{filsou}. After that, the main driving force behind the proof of Theorem~\ref{maintheorem}, work of D.~Shiu \cite{shiu}, can be applied. D.~Shiu \cite{shiu} showed that in any arithmetic progression containing infinitely many primes (that is, $an+b$ with $\gcd(a,b) = 1$ and $a > 0$) there are arbitrarily long sequences of consecutive primes. Thus, once we establish through covering systems that such an arithmetic progression exists where every prime in the arithmetic progression is widely digitally delicate, D.~Shiu's result immediately applies to finish the proof of Theorem~\ref{maintheorem}. Our main focus in this paper is on the proof of Theorem~\ref{maintheorem}. However, in part, this paper is to emphasize that the remarkable work of Shiu \cite{shiu} provides for a nice application to a number of results established via covering systems. One can also take these applications further by looking at the strengthening of Shiu's work by J.~Maynard \cite{maynard}. To illustrate the application of Shiu's work in other context, we give some further examples before closing this introduction. A Riesel number is a positive odd integer $k$ with the property that $k \cdot 2^{n}-1$ is composite for all positive integers $n$. A Sierpi\'nski number is a positive odd integer $k$ with the property that $k \cdot 2^{n}+1$ is composite for all nonnegative integers $n$. The existence of such $k$ were established in \cite{riesel} and \cite{sierpinski}, respectively, though the former is a rather direct consequence of P.~Erd{\H o}s's work in \cite{pe} and the latter is a somewhat less direct application of this same work, an observation made by A.~Schinzel (cf.~\cite{FFK}). A Brier number is a number $k$ which is simultaneously Riesel and Sierpi\'nski, named after Eric Brier who first considered them (cf.~\cite{FFK}). The smallest known Brier number, discovered by Christophe Clavier in 2014 (see \cite{sloantwo}) is \[ 3316923598096294713661. \] As is common with all these numbers, examples typically come from covering systems giving an arithmetic progression of examples. In particular, Clavier established that every number in the arithmetic progression \[ 3770214739596601257962594704110\,n + 3316923598096294713661, \quad n \in \mathbb Z^{+} \cup \{ 0 \} \] is a Brier number. Since the numbers $3770214739596601257962594704110$ and $3316923598096294713661$ are coprime, Shiu's theorem gives the following. \begin{theorem} For every positive integer $k$, there exist $k$ consecutive primes all of which are Brier numbers. \end{theorem} Observe that as an immediate consequence the same result holds if Brier numbers are replaced by Riesel or Sierpi\'nski numbers. As another less obvious result to apply Shiu's theorem to, we recall a result of R.~Graham \cite{graham} from 1964. He showed that there exist relatively prime positive integers $a$ and $b$ such that the recursive Fibonacci-like sequence \begin{equation}\label{grahamrecursion} u_{0} = a, \quad u_{1} = b, \quad \text{and} \quad u_{n+1} = u_{n} + u_{n-1} \quad \text{for integers $n \ge 1$}, \end{equation} consists entirely of composite numbers. The known values for admissible $a$ and $b$ have decreased over the years through the work of others including D.~Knuth \cite{knuth}, J.~W.~Nicol \cite{nicol} and M.~Vsemirnov \cite{vsemirnov}, the latter giving the smallest known such $a$ and $b$ (but notably the same number of digits as the $a$ and $b$ in \cite{nicol}). The result has also been generalized to other recursions; see A~Dubickas, A.~Novikas and J.~\v{S}iurys \cite{dns}, D.~Ismailescu, A.~Ko, C.~Lee and J.~Y.~Park \cite{iklp} and I.~Lunev \cite{lunev}. As the Graham result concludes with all $u_{i}$ being composite, the initial elements of the sequence, $a$ and $b$, are composite. However, there is still a sense in which one can apply Shiu's result. To be precise, the smallest known example given by Vsemirnov is done by taking \[ a = 106276436867 \quad \text{and} \quad b = 35256392432. \] With $u_{j}$ defined as above, one can check that each $u_{j}$ is divisible by a prime from the set \[ \mathcal P = \{ 2, 3, 5, 7, 11, 17, 19, 23, 31, 41, 47, 61, 107, 181, 541, 1103, 2521 \}. \] Setting \[ N = \prod_{p \in \mathcal P} p = 1821895895860356790898731230, \] the value of $a$ and $b$ can be replaced by any integers $a$ and $b$ satisfying \[ a \equiv 106276436867 {\hskip -4pt}\pmod{N} \quad \text{and} \quad b \equiv 35256392432 {\hskip -4pt}\pmod{N}. \] As $\gcd(106276436867,N) = 31$ and $\gcd(35256392432,N) = 2$, these congruences are equivalent to taking $a = 31 a'$ and $b = 2 b'$ where $a'$ and $b'$ are integers satisfying \[ a' \equiv 3428272157 {\hskip -4pt}\pmod{58770835350334090028991330} \] and \[ b' \equiv 17628196216 {\hskip -4pt}\pmod{910947947930178395449365615}. \] As a direct application of D.~Shiu's result, we have the following. \begin{theorem} For every $k \in \mathbb Z^{+}$, there are $k$ consecutive primes $p_{1}, p_{2}, \ldots, p_{k}$ and $k$ consecutive primes $q_{1}, q_{2}, \ldots, q_{k}$ such that for any $i \in \{ 1, 2, \ldots, k \}$, the numbers $a = 31 p_{i}$ and $b = 2 q_{i}$ satisfy $\gcd(a,b) = 1$ and have the property that the $u_{n}$ defined by \eqref{grahamrecursion} are all composite. \end{theorem} \vskip 0pt \noindent This latter result is not meant to be particularly significant but rather an indication that Shiu's work does provide information in cases where covering systems are used to form composite numbers. Regarding open problems, given the recent excellent works surrounding the non-existence of covering systems of particular forms (cf.~\cite{BBMST, BBMST2, hough, houghnielsen}), the authors are not convinced that widely digitally delicate primes exist in every base. Thus, a tantalizing question is whether they exist or whether a positive proportion of the primes in every base are widely digitally delicate. In the opposite direction, as noted in \cite{filsou}, Carl Pomerance has asked for an unconditional proof that there exist infinitely many primes which are not digitally delicate or which are not widely digitally delicate. For other open problems in this direction, see the end of the introdutcions in \cite{filjuisou} and \cite{filsou}. \section{The first steps of the argument} As noted in the introduction, to prove Theorem~\ref{maintheorem}, the work of D.~Shiu \cite{shiu} implies that it suffices to obtain an arithmetic progression $An+B$, with $A$ and $B$ relatively prime positive integers, such that every prime in the arithmetic progression is widely digitally delicate. We will determine such an $A$ and $B$ by finding relatively prime positive integers $A$ and $B$ satisfying property ($*$) given by \vskip 5pt \centerline{($*$){\ }\parbox[t]{11cm}{If $d \in \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$, then each number in the set \[ \mathcal A_{d} = \big\{ An+B + d\cdot 10^{k}: n \in \mathbb Z^{+}, k \in \mathbb Z^{+} \cup \{ 0 \} \big\} \] is composite.}} \vskip 8pt\noindent As changing a digit of $An+B$, including any one of its infinitely many leading zero digits, corresponds to adding or subtracting one of the numbers $1, 2, \ldots, 9$ from a digit of $An+B$, we see that relatively prime positive integers $A$ and $B$ satisfying property ($*$) also satisfy the property we want, that every prime in $An+B$ is widely digitally delicate. To find relatively prime positive integers $A$ and $B$ satisfying property ($*$), we make use of covering systems which we define as follows. \begin{definition} A covering system (or covering) is a finite set of congruences \[ x \equiv a_{1} {\hskip -4pt}\pmod{m_{1}}, \quad x \equiv a_{2} {\hskip -4pt}\pmod{m_{2}}, \quad \ldots, \quad x \equiv a_{r} {\hskip -4pt}\pmod{m_{r}}, \] where $r \in \mathbb Z^{+}$, each $a_{j} \in \mathbb Z$, and each $m_{j} \in \mathbb Z^{+}$, such that every integer satisfies at least one congruence in the set of congruences. \end{definition} \noindent In other contexts in the literature, further restrictions can be made on the $m_{j}$, so we emphasize here that we want to allow for $m_{j} = 1$ and for repeated moduli (so that the $m_{j}$ are not necessarily distinct). There will be restrictions on the $m_{j}$ that will arise in the covering systems we build due to the approach we are using. We will see these as we proceed. For each $d \in \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$, we will create a separate covering system to show that the elements of $\mathcal A_{d}$ in ($*$) are composite. Table~\ref{tablenumcong} indicates, for each $d$, the number of different congruences in the covering system corresponding to $d$. \begin{table}[!hbt] \centering \caption{Number of congruences for each covering}\label{tablenumcong} \begin{minipage}{3 cm} \centering \begin{tabular}{|c|c|} \hline $d$ & \# cong. \\ \hline \hline $-9$ & $232$ \\ \hline $-8$ & $441$ \\ \hline $-7$ & $1$ \\ \hline $-6$ & $257$ \\ \hline $-5$ & $268$ \\ \hline $-4$ & $1$ \\ \hline \end{tabular} \end{minipage} \begin{minipage}{3 cm} \centering \begin{tabular}{|c|c|} \hline $d$ & \# cong. \\ \hline \hline $-3$ & $739$ \\ \hline $-2$ & $289$ \\ \hline $-1$ & $1$ \\ \hline $1$ & $37$ \\ \hline $2$ & $1$ \\ \hline $3$ & $203$ \\ \hline \end{tabular} \end{minipage} \begin{minipage}{3 cm} \centering \begin{tabular}{|c|c|} \hline $d$ & \# cong. \\ \hline \hline $4$ & $26$ \\ \hline $5$ & $1$ \\ \hline $6$ & $19$ \\ \hline $7$ & $137$ \\ \hline $8$ & $1$ \\ \hline $9$ & $4$ \\ \hline \end{tabular} \end{minipage} \end{table} The integers we are covering for each $d$ are the exponents $k$ on $10$ in the definition of $\mathcal A_{d}$. In other words, we will want to view each exponent $k$ as satisfying one of the congruences in our covering system for a given $\mathcal A_{d}$. In the end, the values of $A$ and $B$ will be determined by the congruences we choose for the covering systems as well as certain primes that arise in our method. We clarify that the work on digitally delicate primes in prior work mentioned in the introduction used a partial covering of the integers $k$, that is a set of congruences where most but not all integers $k$ satisfy at least one of the congruences, together with a sieve argument. The work in \cite{filsou} on widely digitally delicate primes used covering systems for $d \in \{ 1, 2, \ldots, 9 \}$ and the same approach of partial coverings and sieves for $d \in \{ -9, -8, \ldots, -1 \}$. The work in \cite{filjuisou}, like we will use in this paper, made use of covering systems for all $d \in \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$. For \cite{filjuisou}, some of the covering systems could be handled rather easily by taking advantage of the fact that we were looking for composite numbers satisfying a certain property rather than primes. Next, we explain more precisely how we create and take advantage of a covering system for a given fixed $d \in \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$. We begin with a couple illustrative examples. Table~\ref{tablenumcong} indicates that a number of the $d$ are handled with just one congruence. This is accomplished by taking \[ A \equiv 0 {\hskip -5pt}\pmod{3} \qquad \text{and} \qquad B \equiv 1 {\hskip -5pt}\pmod{3}. \] Observe that each element of $\mathcal A_{d}$ in ($*$) is divisible by $3$ whenever $d \equiv 2 \pmod{3}$. Thus, since $A$ and $B$ are positive, as long as we also have $B > 3$, the elements of $\mathcal A_{d}$ for such $d$ are all composite, which is our goal. Note the crucial role of the order of $10$ modulo the prime $3$. The order is $1$, and the covering system for each of these $d$ is simply $k \equiv 0 \pmod{1}$. Every integer satisfies this congruence, so it is a covering system. The modulus corresponds to the order of $10$ modulo $3$. Note also that we cannot use the prime $3$ in an analogous way to cover another digit $d$ because the choices for $A$ and $B$, and hence the congruences on $A$ and $B$ above, are to be independent of $d$. For example, if $d = 4$, then $An+B + d\cdot 10^{k} \equiv 1 + 4 \equiv 2 \pmod{3}$ and, hence, $An+B + d\cdot 10^{k}$ will not be divisible by $3$. As a second illustration, we see from Table~\ref{tablenumcong} that we handle the digit $d = 9$ with $4$ congruences. The congruences for $d = 9$ are \[ k \equiv 0 {\hskip -5pt}\pmod{2}, \quad \ k \equiv 3 {\hskip -5pt}\pmod{4}, \quad \ k \equiv 1 {\hskip -5pt}\pmod{8}, \quad \ k \equiv 5 {\hskip -5pt}\pmod{8}. \] One easily checks that this is a covering system, that is that every integer $k$ satisfies one of these congruences. To take advantage of this covering system, we choose a different prime $p$ for each congruence with $10$ having order modulo $p$ equal to the modulus. We used the prime $11$ with $10$ of order $2$, the prime $101$ with $10$ of order $4$, the prime $73$ with $10$ of order $8$, and the prime $137$ with $10$ of order $8$. We take $A$ divisible by each of these primes. For ($*$), with $d = 9$, we want $An+B + 9 \cdot 10^{k}$ composite. For $k \equiv 0 \pmod{2}$, we accomplish this by taking $B \equiv 2 \pmod{11}$ and $B > 11$ since then $An+B + 9 \cdot 10^{k} \equiv B + 9 \equiv 0 \pmod{11}$. For $k \equiv 3 \pmod{4}$, we accomplish this by taking $B \equiv 90 \pmod{101}$ and $B > 101$ since then $An+B + 9 \cdot 10^{k} \equiv 90 + 9 \cdot 10^{3} \equiv 9090 \equiv 0 \pmod{101}$. Similarly, for $k \equiv 1 \pmod{8}$ and $B \equiv 56 \pmod{73}$, we obtain $An+B + 9 \cdot 10^{k} \equiv 0 \pmod{73}$; and for $k \equiv 5 \pmod{8}$ and $B \equiv 90 \pmod{137}$, we obtain $An+B + 9 \cdot 10^{k} \equiv 0 \pmod{137}$. Thus, taking $B > 137$, we see that ($*$) holds with $d = 9$. Of some significance to our explanations later, we note that we could have interchanged the roles of the primes $73$ and $137$ since $10$ has the same order for each of these primes. In other words, we could associate $137$ with the congruence $k \equiv 1 \pmod{8}$ above and associate $73$ with the congruence $k \equiv 5 \pmod{8}$. Then for $k \equiv 1 \pmod{8}$ and $B \equiv 47 \pmod{137}$, we would have $An+B + 9 \cdot 10^{k} \equiv 0 \pmod{137}$; and for $k \equiv 5 \pmod{8}$ and $B \equiv 17 \pmod{73}$, we would have $An+B + 9 \cdot 10^{k} \equiv 0 \pmod{73}$. In general, in our construction of widely digitally delicate primes, we want each congruence $k \equiv a \pmod{m}$ in a covering system associated with a prime $p$ for which the order of $10$ modulo $p$ is $m$, but how we choose the ordering of those primes (which prime goes to which congruence) for a fixed modulus $m$ is irrelevant. For each $d \in \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$, we determine a covering system of congruences for $k$, where each modulus $m$ corresponds to the order of $10$ modulo some prime $p$. This imposes a condition on $A$, namely that $A$ is divisible by each of these primes $p$. Fixing $d$, a congruence from our covering system $k \equiv a \pmod{m}$, and a corresponding prime $p$ with $10$ having order $m$ modulo $p$, we determine $B$ such that $A n + B + d\cdot 10^{k} \equiv B + d\cdot 10^{a} \equiv 0 \pmod{p}$. Note that the values of $d$, $a$ and $p$ dictate the congruence condition for $B$ modulo $p$. Each prime $p$ will correspond to a unique congruence condition $B \equiv - d\cdot 10^{a} \pmod{p}$, so the Chinese Remainder Theorem implies the existence of a $B \in \mathbb Z^{+}$ simultaneously satisfying all the congruence conditions modulo primes on $B$. As long as $B$ is large enough, then the condition ($*$) will hold. To make sure that there is a prime of the form $An+B$, we will want $\gcd(A,B) = 1$. For $k \equiv a \pmod{m}$ and a corresponding prime $p$ as above, we will have $A$ divisible by $p$ and $B \equiv - d\cdot 10^{a} \pmod{p}$. Since $d \in \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$, if $p \ge 11$, then we see that $p \nmid B$. We will not be using the primes $p \in \{ 2,5 \}$ as $10$ does not have an order modulo these primes. We have already seen that we are using the prime $p = 3$ for $d \equiv 2 \pmod{3}$, so this ensures that $3 \nmid B$. We will use $p = 7$ for $d \in \{-9, -8, -6, -5, -3, 3, 4 \}$, which then implies $7 \nmid B$. Therefore, the condition $\gcd(A,B) = 1$ will hold. Recall that we used the same congruence and corresponding prime in our covering system for each $d \equiv 2 \pmod{3}$. There is no obstacle to repeating a congruence for different $d$ if the corresponding prime, having $10$ of order the modulus, is different. But in the case of $d \equiv 2 \pmod{3}$, the same prime $3$ was used for different $d$. To illustrate how we can repeat the use of a prime, we return to how we used the prime $p = 11$ above for $d = 9$. We ended up with $A \equiv 0 \pmod{11}$ and $B \equiv 2 \pmod{11}$. In order for us to take advantage of the prime $p = 11$ for $d$, we therefore want $A n + B + d\cdot 10^{k} \equiv 2 + d\cdot 10^{k} \equiv 0 \pmod{11}$. It is easy to check that this holds for $(d,k) \in \{ (-9,1), (-2,0), (2,1), (9,0) \}$. The case $(d,k) = (9,0)$ is from our example with $d = 9$ above. The case $(d,k) = (2,1)$ does not serve a purpose for us as $d = 2$ was covered by our earlier example using the prime $3$ for all $d \equiv 2 \pmod{3}$. The cases where $(d,k) \in \{ (-9,1), (-2,0) \}$ are significant, and we make use of congruences modulo $11$ in the covering systems for $d = -9$ and $d = -2$. Thus, we are able to repeat the use of some primes for different values of $d$. However, this is not the case for most primes we used. A complete list of the primes which we were able to use for more than one value of $d$ is given in Table~\ref{tablerepeatprimes}, together with the list of corresponding $d$'s. The function $\rho(m,p)$ in this table will be explained in the next section. \begin{table}[!hbt] \centering \caption{Primes used for more than one digit $d$}\label{tablerepeatprimes} \begin{minipage}{7 cm} \centering \begin{tabular}{|c|c|c|} \hline prime & $d$'s & $\rho(m,p)$ \\ \hline \hline $3$ & $-7, -4, -1, 2, 5, 8$ & $1$ \\ \hline $7$ & $-9, -8, -6, -5, -3, 3, 4$ & $1$ \\ \hline $11$ & $-9, -2, 9$ & $1$ \\ \hline $13$ & $-9, -3, 3, 4$ & $2$ \\ \hline $17$ & $-8, -6, -3, -2, 7$ & $1$ \\ \hline $19$ & $-6, 4$ & $1$ \\ \hline $23$ & $-9, -8, -6, -3, 3, 7$ & $1$ \\ \hline $29$ & $-9, -8, -6, 1, 3$ & $1$ \\ \hline $31$ & $-8, -2, 6$ & $1$ \\ \hline $37$ & $3, 4$ & $1$ \\ \hline $43$ & $-8, -3, 1$ & $1$ \\ \hline $53$ & $-8, -5, 3$ & $1$ \\ \hline $61$ & $-6, 3, 6$ & $1$ \\ \hline $67$ & $-9, 7$ & $1$ \\ \hline $79$ & $-9, -5$ & $2$ \\ \hline $89$ & $-6, -3, 7$ & $1$ \\ \hline $103$ & $-9, -8, -3$ & $1$ \\ \hline \end{tabular} \end{minipage} \begin{minipage}{5.5 cm} \centering \vspace{-.43cm} \begin{tabular}{|c|c|c|} \hline prime & $d$'s & $\rho(m,p)$ \\ \hline \hline $199$ & $-6, -3, 7$ & $1$ \\ \hline $211$ & $-6, 6$ & $1$ \\ \hline $241$ & $-6, 6$ & $2$ \\ \hline $331$ & $-8, 7$ & $1$ \\ \hline $353$ & $-6, 7$ & $1$ \\ \hline $409$ & $-8, -3$ & $1$ \\ \hline $449$ & $-9, 7$ & $2$ \\ \hline $2161$ & $-6, 6$ & $3$ \\ \hline $3541$ & $-6, 6$ & $1$ \\ \hline $9091$ & $-6, 6$ & $1$ \\ \hline $27961$ & $-6, 6$ & $2$ \\ \hline $1676321$ & $-6, 6$ & $1$ \\ \hline $3762091$ & $-6, 6$ & $2$ \\ \hline $4188901$ & $-6, 6$ & $2$ \\ \hline $39526741$ & $-6, 6$ & $3$ \\ \hline $5964848081$ & $-6, 6$ & $2$ \\ \hline \end{tabular} \end{minipage} \end{table} Recalling that the modulus in a covering system is equal to the order of $10$ modulo a prime $p$, the role of primes and the order of $10$ modulo those primes is significant in coming up with covering systems to deduce ($*$). A modulus $m$ can be used in a given covering system as many times as there are primes with $10$ of order $m$. Thus, for the covering system for $d = 9$, we saw the modulus $8$ being used twice as there are two primes with $10$ of order $8$, namely the primes $73$ and $137$. One can look at a list of primitive prime factors of $10^{k}-1$ such as in \cite{brill}, but we needed much more extensive data than what is contained there. Our approach uses that the complete list of primes for which $10$ has a given order $m$ is the same as the list of primes dividing $\Phi_{m}(10)$ and not dividing $m$ where $\Phi_{m}(x)$ is the $m$-th cyclotomic polynomial (cf.~\cite{brill, filjuisou, filsou}). We used Magma V2.23-1 on a 2017 MacBook Pro to determine different primes dividing $\Phi_{m}(10)$. We did not always get a complete factorization but used that if the remaining unfactored part of $\Phi_{m}(10)$ is composite, relatively prime to the factored part of $\Phi_{m}(10)$ and $m$, and not a prime power, then there must be at least two further distinct prime factors of $\Phi_{m}(10)$. This allowed us then to determine a lower bound on the number of distinct primes of a given order $m$. Though we used most of these in our coverings, sometimes we found extra primes that we did not need to use. In total, we made use of $673$ different moduli $m$ and $2596$ different primes dividing $\Phi_{m}(10)$ for such $m$. Of the $2596$ different primes, there are $590$ which came from $295$ composite numbers arising from an unfactored part of some $\Phi_{m}(10)$, and there are $63$ other composite numbers for which only one prime factor of each of the composite numbers was used. The largest explicit prime (not coming from the $295+63 = 358$ composite numbers) has $1700$ digits, arising from testing what was initially a large unfactored part of $\Phi_{m}(10)$ for primality and determining it is a prime. The largest of the $358$ composite numbers has $17234$ digits. For obvious reasons, we will avoid listing these primes and composites in this paper, though to help with verification of the results, we are providing the data from our computations in \cite{Filweb}; more explicit tables can also be found in \cite{juillerat}. Table~\ref{orderofprimes} in the appendix gives, for each of the $673$ different moduli $m$, the detailed information on the number of distinct primes we used with $10$ of order $m$, which we denote by $L(m)$. Thus, $L(m)$ is a lower bound on the total number of distinct primes with $10$ of order $m$. Note that $L(m)$ is less than or equal to the number of distinct primes dividing $\Phi_{m}(10)$ but not dividing $m$. For each $d \in \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$, the goal is to find a covering system so that ($*$) holds. We have already given the covering systems we obtained for $d \equiv 2 \pmod{3}$ and for $d = 9$. In the next section and the appendix, we elaborate on the covering systems for the remaining $d$. We also explain how the reader can verify the data showing these covering systems satisfy the conditions needed for ($*$). \section{Finishing the argument} To finish the argument, we need to present a covering system for each value of $d$ in $\{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$ as described in the previous section. For the purposes of keeping the presentation of these covering systems manageable, for each $m$ listed in Table~\ref{orderofprimes}, we take the $L(m)$ primes we found with $10$ of order $m$ and order them in some way. Corresponding to the discussion concerning $d = 9$ and the primes $73$ and $137$, the particular ordering is not important to us (for example, increasing order would be fine). Suppose the primes corresponding to $m$ are ordered in some way as $p_{1}, p_{2}, \ldots, p_{L(m)}$. We define $\rho(p_{j},m) = j$. Thus, if $p_{j}$ is the $j$-th prime in our ordering of the primes with $10$ of order $m$, we have $\rho(p_{j},m) = j$. The particular values we used for $\rho(p_{j},m)$ is not important to the arguments. So as to make the entries in Table~\ref{tablerepeatprimes} correct, the entries for $\rho(p,m)$ indicate the values we used for those primes. For example, Table~\ref{orderofprimes} indicates there are $2$ primes of order $6$. One of them is $7$. Table~\ref{tablerepeatprimes} indicates then that $\rho(7,6) = 1$. Thus, we put $7$ as the first of the $2$ primes with $10$ of order $6$. The other prime with $10$ of order $6$ is $13$, and as Table~\ref{tablerepeatprimes} indicates we set $13$ as the second of the $2$ primes with $10$ of order $6$. Tables~\ref{covforminus9}-\ref{covfor9} give the covering systems used for each $d \in \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$ with $d \not\equiv 2 \pmod{3}$. Rather than indicating the prime, which in some cases has thousands of digits, corresponding to each congruence $k \equiv a \pmod{m}$ listed, we simply wrote the value of $\rho(m,p)$. As $m$ corresponds to the modulus used in the given congruence $k \equiv a \pmod{m}$ and the ordering of the primes is not significant to our arguments (any ordering will do), this is enough information to confirm the covering arguments. That said, the time consuming task of coming up with the $L(m)$ primes to order for each $m$ is nontrivial (at least at this point in time). So that this work does not need to be repeated, a complete list of the $L(m)$ primes for each $m$ is given in \cite{Filweb}. Further, the tables in the form of lists can be found there as well, with the third column in each case replaced by the prime we used with $10$ of order the modulus of the congruence in the second column. In the way of clarity, recall that the primes were not explicitly computed in the case that the unfactored part of $\Phi_{m}(10)$ was tested to be composite; instead the composite number is listed in place of both primes in \cite{Filweb}. For the remainder of this section, we clarify how to verify the information in Tables~\ref{covforminus9}-\ref{covfor9}. We address both verification of the covering systems and the information on the primes as listed in \cite{Filweb}. \subsection{Covering Verification.} The most direct way to check that a system $\mathcal C$ of congruences \[ x \equiv a_{1} {\hskip -5pt}\pmod{m_{1}}, \quad x \equiv a_{2} {\hskip -5pt}\pmod{m_{2}}, \quad \ldots, \quad x \equiv a_{s} {\hskip -5pt}\pmod{m_{s}} \] is a covering system is to set $\ell = \text{lcm}(m_{1}, m_{2}, \ldots, m_{s})$ and then to check if every integer in the interval $[0,\ell-1]$ satisfies at least one congruence in $\mathcal C$. If not, then $\mathcal C$ is not a covering system. If on the other hand, every integer in $[0,\ell-1]$ satisfies a congruence in $\mathcal C$, then $\mathcal C$ is a covering system. To see the latter, let $n$ be an arbitrary integer, and write $n = \ell q + r$ where $q$ and $r$ are integers with $0 \le r \le \ell-1$. Since $r \in [0,\ell-1]$ satisfies some $x \equiv a_{j} \pmod{m_{j}}$ and since $\ell \equiv 0 \pmod{m_{j}}$, we deduce for this same $j$ that $n = \ell q + r \equiv a_{j} \pmod{m_{j}}$. The above is a satisfactory approach if $\ell$ is not too large. For the values of $d$ in $ \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$ with $d \not\equiv 2 \pmod{3}$, the least common multiple $\ell$ given by the congruences in Tables~\ref{covforminus9}-\ref{covfor9} are listed in Table~\ref{lcmtable}. The maximum prime divisor of $\ell$ is also listed in the fourth column of Table~\ref{lcmtable}. The value of $\ell$ can exceed $10^{12}$, so we found a more efficient way to test whether one of our systems $\mathcal C$ of congruences, where $\ell$ is large, is a covering system. \begin{table}[!hbt] \centering \caption{Least common multiple of the moduli for the coverings in each table}\label{lcmtable} \begin{minipage}{6.6 cm} \centering \begin{tabular}{|c|c|c|c|} \hline $d$ & Table & $\ell$ & max $p$ \\ \hline \hline $-9$ & $5$ & $14433138720$ & $31$ \\ \hline $-8$ & $6$ & $699847948800$ & $17$ \\ \hline $-6$ & $7$ & $1045044000$ & $29$ \\ \hline $-5$ & $8$ & $56216160$ & $13$ \\ \hline $-3$ & $9$ & $1486147703040$ & $19$ \\ \hline $-2$ & $10$ & $321253732800$ & $23$ \\ \hline \end{tabular} \end{minipage} \begin{minipage}{5.6 cm} \centering \begin{tabular}{|c|c|c|c|} \hline $d$ & Table & $\ell$ & max $p$ \\ \hline \hline $1$ & $11$ & $5040$ & $7$ \\ \hline $3$ & $12$ & $133333200$ & $37$ \\ \hline $4$ & $13$ & $1296$ & $3$ \\ \hline $6$ & $14$ & $360$ & $5$ \\ \hline $7$ &$15$ & $18295200$ & $11$ \\ \hline $9$ & $16$ & $8$ & $2$ \\ \hline \end{tabular} \end{minipage} \end{table} Suppose $\ell > 10^{6}$ in Table~\ref{lcmtable} and the corresponding collection of congruences coming from the table indicated in the second column is $\mathcal C$. Let $q$ be the largest prime divisor of $\ell$ as indicated in the fourth column. Let $w = 4 \cdot 3 \cdot 5 \cdot q$. This choice of $w$ was selected on the basis of some trial and error; other choices are certainly reasonable. We do however want and have that $w$ divides $\ell$. Based on the comments above, we would like to know if every integer in the interval $[0,\ell-1]$ satisfies at least one congruence in $\mathcal C$. The basic idea is to take each $u \in [0,w-1]$ and to consider the integers that are congruent to $u$ modulo $w$ in $[0,\ell-1]$. One advantage of doing this is that not every congruence in $\mathcal C$ needs to be considered. For example, take $d = -3$. Then Table~\ref{lcmtable} indicates $\ell = 1486147703040$ and Table~\ref{tablenumcong} indicates the number of congruences in $\mathcal C$ is $739$. From Table~\ref{covforminus3}, the first few of the congruences in $\mathcal C$ are \[ k \equiv 4 {\hskip -5pt}\pmod{6}, \quad k \equiv 5 {\hskip -5pt}\pmod{6}, \quad k \equiv 0 {\hskip -5pt}\pmod{16}, \quad k \equiv 11 {\hskip -5pt}\pmod{21}. \] Here, $w = 4 \cdot 3 \cdot 5 \cdot 19 = 1140$. If we take $u = 0$, then only the third of these congruences can be satisfied by an integer $k$ congruent to $u$ modulo $w$, as each of the other ones requires $k \not\equiv 0 \pmod{3}$ whereas $k \equiv u \pmod{w}$ requires $k \equiv 0 \pmod{3}$. Let $\mathcal C'$ be the congruences in $\mathcal C$ which are consistent with $k \equiv u \pmod{w}$. One can determine these congruences by using that there exist integers satisfying both $k \equiv a \pmod{m}$ and $k \equiv u \pmod{w}$ if and only if $a \equiv u \pmod{\gcd(m,w)}$. Observe that, with $u \in [0,w-1]$ fixed, we would like to know if each integer $v$ of the form \begin{equation}\label{vequat} v = w t + u, \quad \text{ with } 0 \le t \le (\ell/w)-1 \end{equation} satisfies at least one congruence in $\mathcal C'$. The main advantage of this approach is that, as we shall now see, not all $\ell/w$ values of $t$ need to be considered. First, we note that if $\mathcal C'$ is the empty set, then the integers in \eqref{vequat} are not covered and therefore $\mathcal C$ is not a covering system. Suppose then that $|\mathcal C'| \ge 1$. Let $\ell'$ denote the least common multiple of the moduli in $\mathcal C'$. Let $\delta = \gcd(w,\ell')$. We claim that we need only consider $v = w t + u$ where $0 \le t \le (\ell'/\delta)-1$. To see this, suppose we know that every $v = w t + u$ with $0 \le t \le (\ell'/\delta)-1$ satisfies one of the congruences in $\mathcal C'$. There are integers $q$, $q'$, $r$ and $r'$ satisfying $t = \ell' q' + r'$ where $0 \le r' \le \ell'-1$ and $r' = (\ell'/\delta) q + r$, where $0 \le r \le (\ell'/\delta)-1$. Then \[ v = w t + u = w \ell' q' + w r' + u = w \ell' q' + (w/\delta) \ell' q + w r + u. \] The definition of $\delta$ implies that $w/\delta \in \mathbb Z$. As each modulus in $\mathcal C'$ divides $\ell'$, we see that $v$ satisfies a congruence in $\mathcal C'$ if and only if $w r + u$ does. Here, $w$ and $u$ are fixed and $0 \le r \le (\ell'/\delta)-1$. Thus, we see that for each $u \in [0,w-1]$, we can restrict to determining whether $v$ in \eqref{vequat} satisfies a congruence in $\mathcal C'$ for $0 \le t \le (\ell'/\delta)-1$. Returning to the example of $d = -3$, $\ell = 1486147703040$ and $|\mathcal C| = 739$, where $w = 1140$ and we considered $u = 0$, one can check that $|\mathcal C'| = 19$, $\ell' = 12640320$, $\delta = w$ and $\ell'/\delta = 11088$. Thus, what started out as ominously checking whether over $10^{12}$ integers each satisfy at least one of $739$ different congruences is reduced in the case of $u = 0$ to looking at whether $11088$ integers each satisfy at least one of $19$ different congruences. As $u \in [0,w-1]$ varies, the number of computations does as well. An extreme case for $d = -3$ occurs for $u = 75$, where we get $\ell'/\delta = 14325696$ and $|\mathcal C'| = 47$. As $d$ and $u$ vary, though, this computation becomes manageable for determining that we have covering systems for each $d$ in $ \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$ with $d \not\equiv 2 \pmod{3}$ and $\ell > 10^{6}$. On a 2017 MacBook Plus running Maple 2019 with a 2.3 GHz Dual-Core Intel Core i5 processor, the total cpu time for determining the systems of congruences in Tables~\ref{covforminus9}-\ref{covfor9} are all covering systems took approximately $2.9$ cpu hours, with almost all of this time spent on the case $d = -3$ which took $2.7$ hours. The largest value of $\ell'/\delta$ encountered was $\ell'/\delta = 14325696$ which occurred precisely for $d = -3$ and $u \in \{ 75, 303, 531, 759, 987 \}$. \subsection{Data check.} The most cumbersome task for us was the determination of the data in Table~\ref{orderofprimes}. As noted earlier, although the reader can check the data there directly, we have made the list of primes corresponding to each $m$ available through \cite{Filweb}. With the list of such primes for each $m$, it is still worth indicating how the data can be checked. Recall, in particular, the list of primes is not explicit in the case that there was an unfactored part of $\Phi_{m}(10)$. In this subsection, we elaborate on what checks should be and were done. All computations below were done with the MacBook Pro mentioned at the end of the last subsection and using Magma V2.23-1. For each modulus $m$ used in our constructions (listed in Table~\ref{orderofprimes}), we made a list of primes $p_{1}, p_{2}, \ldots, p_{s}$, written in increasing order, together with up to two additional primes $q_{1}$ and $q_{2}$, included after $p_{s}$ on the list but not written explicitly (as we will discuss). Each prime came from a factorization or partial factorization of $\Phi_{m}(10)$. The primes $p_{1}, p_{2}, \ldots, p_{s}$ are the distinct primes appearing in the factored part of $\Phi_{m}(10)$, and as noted earlier do not include primes dividing $m$. In some cases, a complete factorization was found for $\Phi_{m}(10)$. For such $m$, there are no additional primes $q_{1}$ and $q_{2}$. If $\Phi_{m}(10)$ had an unfactored part $Q > 1$ (already tested to be composite), then we checked that $Q$ is relatively prime to $m p_{1} p_{2} \cdots p_{s}$ and that $Q$ is not of the form $N^{k}$ with $N \in \mathbb Z^{+}$ and $k$ an integer greater than or equal to $2$. As this was always the case for the $Q$ tested, we knew each such $Q$ had two distinct prime factors $q_{1}$ and $q_{2}$. We deduce that there are at least two more primes $q_{j}$, $j \in \{ 1,2 \}$, different from $p_{1}, p_{2}, \ldots, p_{s}$ for which $10$ has order $m$ modulo $q_{j}$. As the data only contains the primes used in the covering systems, we only included the primes $q_{1}$ and $q_{2}$ that were used. Thus, despite $Q$ having at least two distinct prime divisors, we may have listed anywhere from $0$ to $2$ of them. The question arises, however, as to how one can list primes that we do not know; there are primes $q_{1}$ and $q_{2}$ dividing $Q$, but we were unable to (or chose not to) factor $Q$ to determine them explicitly. Instead of listing $q_{1}$ and $q_{2}$ then, we opted to list $Q$. Thus, for each $m$ we associated a list of one of the forms \[ [p_{1}, p_{2}, \ldots, p_{s}], \quad [p_{1}, p_{2}, \ldots, p_{s},Q], \quad [p_{1}, p_{2}, \ldots, p_{s},Q,Q], \] depending on whether $Q$ either did not exist or we used no prime factor of $Q$, we used one prime factor of $Q$, or we used two prime factors of $Q$, respectively. It is possible that $s=0$; for example, the lists associated with the moduli $2888$ and $2976$ each take the middle form with no $p_{j}$ and one composite number. For a fixed $m$, given such a list, say from \cite{Filweb}, one merely needs to check: \begin{itemize} \setlength\itemsep{-0.25em} \item Each element of the list divides $\Phi_{m}(10)$. \item Each element of the list is relatively prime to $m$. \item There is at most one composite number, say $Q > 1$, in the list, which may appear at most twice. The other numbers in the list are distinct primes. \item If the composite number $Q$ exists, then $\gcd(Q, p_{1} p_{2} \cdots p_{s}) = 1$. \item If the composite number $Q$ exists twice, then $Q^{1/k} \not\in \mathbb Z^{+}$ for every integer $k \in [2, \log(Q)/\log(2)]$. \end{itemize} \noindent The upper bond in the last item above is simply because $k > \log(Q)/\log(2)$ implies $1 < Q^{1/k} < 2$ and, hence, $Q^{1/k}$ is not an integer. For each $m$, the value of $L(m)$ in Table~\ref{orderofprimes} is simply the number of elements in the list associated with $m$. With the data from the tables in the Appendix, also available in \cite{Filweb} with the indicated primes $p_{1}, p_{2}, \ldots, p_{s}, q_{1}, q_{2}$ depending on $m$ as above, some further details need to be checked to fully justify the computations. We verified that whenever $m$ is used as a modulus in a table, it was associated with one of the primes dividing $\Phi_{m}(10)$. Furthermore, for any given $d \in \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$, the complete list of primes used as the congruences vary are distinct, noting that $q_{1}$ and $q_{2}$, for a given $m$, will be denoted by the same number $Q$ but represent two distinct prime divisors of $Q$. As $d$ varies, a given modulus $m$ and a prime $p$ dividing $\Phi_{m}(10)$ can be used more than once as indicated in Table~\ref{tablerepeatprimes}. To elaborate, suppose such an $m$ and $p$ is used for each $d \in \mathcal D \subseteq \{ -9, -8, \ldots, -1 \} \cup \{ 1, 2, \ldots, 9 \}$. For each $d \in \mathcal D$, then, there corresponds a congruence $k \equiv a \pmod{m}$, where $a = a(d)$ will depend on $d$, as well as $m$ and $p$. As noted earlier, this is permissible if and only if the values of $d \cdot 10^{a(d)}$ are congruent modulo $p$ for all $d \in \mathcal D$. Thus, for each $p$ that occurs in more than one table, as in Table~\ref{tablerepeatprimes}, a check is done to verify the corresponding values of $d \cdot 10^{a(d)}$ are congruent modulo $p$. The verification of the covering systems needed for Theorem~\ref{maintheorem} is complete, and the work of D.~Shiu \cite{shiu} now implies the theorem.
{ "timestamp": "2021-01-25T02:04:41", "yymm": "2101", "arxiv_id": "2101.08898", "language": "en", "url": "https://arxiv.org/abs/2101.08898", "abstract": "We show that for every positive integer $k$, there exist $k$ consecutive primes having the property that if any digit of any one of the primes, including any of the infinitely many leading zero digits, is changed, then that prime becomes composite.", "subjects": "Number Theory (math.NT)", "title": "Consecutive primes which are widely digitally delicate", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9891815535796247, "lm_q2_score": 0.8104789155369048, "lm_q1q2_score": 0.8017107928143249 }
https://arxiv.org/abs/2108.00987
Threshold Ramsey multiplicity for odd cycles
The Ramsey number $r(H)$ of a graph $H$ is the minimum $n$ such that any two-coloring of the edges of the complete graph $K_n$ contains a monochromatic copy of $H$. The threshold Ramsey multiplicity $m(H)$ is then the minimum number of monochromatic copies of $H$ taken over all two-edge-colorings of $K_{r(H)}$. The study of this concept was first proposed by Harary and Prins almost fifty years ago. In a companion paper, the authors have shown that there is a positive constant $c$ such that the threshold Ramsey multiplicity for a path or even cycle with $k$ vertices is at least $(ck)^k$, which is tight up to the value of $c$. Here, using different methods, we show that the same result also holds for odd cycles with $k$ vertices.
\section{Introduction} The \emph{Ramsey number} $r(H)$ of a graph $H$ is the minimum positive integer $n$ such that any two-coloring of the edges of the complete graph $K_n$ on $n$ vertices contains a monochromatic copy of $H$. Ramsey in 1930 proved that these numbers exist. However, determining or even estimating Ramsey numbers remains a formidable challenge for most graphs. For instance, the Ramsey number of $K_5$ is already not known, while the longstanding bounds $2^{k/2} \leq r(K_k) \leq 4^k$ have only been improved by lower-order factors~\cite{ConlonUp, Sah, Spencer}. To date, there are only a few non-trivial families of graphs for which the Ramsey number is known exactly, including stars, paths, and cycles. Let $P_k$ and $C_k$ denote the path and cycle on $k$ vertices, respectively. In 1967, Gerencs\'er and Gy\'arf\'as~\cite{GG} determined the Ramsey number of paths, namely, \[r(P_k) = k - 1+ \lfloor k/2 \rfloor.\] For cycles, the general case was solved independently by Rosta~\cite{Ros} and by Faudree and Schelp~\cite{cycle2}, who showed that \[ r(C_k) = 3k/2- 1 \text{ if } k \geq 6 \text{ is even \ \ and \ \ } r(C_k)=2k-1 \text{ if } k \geq 5 \text{ is odd}.\] A more general problem than computing Ramsey numbers is to determine the {\it Ramsey multiplicity} $M(H, n)$, the minimum number of monochromatic copies of $H$ guaranteed in any two-edge-coloring of $K_n$. Indeed, it is easy to check that $M(H, n) = 0$ if and only if $n < r(H)$. The asymptotic behaviour of $M(H, n)$ when $H$ is fixed and $n$ tends to infinity has attracted considerable attention. This is in part because of a famous conjecture of Erd\H{o}s~\cite{Erdos62} stating that if $H$ is a clique, then the value of $M(H, n)$ is asymptotically equal to the expected number of monochromatic copies of $H$ in a uniformly random two-edge-coloring of $K_n$. Unfortunately, this conjecture (and a later generalization to all graphs~\cite{BRcommon}) is false already for $H = K_4$, as first shown by Thomason~\cite{Tcommon} (see also~\cite{K4common, Sidcommon}). However, it remains an interesting open problem to determine which graphs satisfy the conjecture, known in the literature as \emph{common graphs}. For instance, the non-three-colorable $5$-wheel is known to be common~\cite{HHKNR} and some hope remains that all bipartite graphs are common because of a connection to a celebrated conjecture of Sidorenko and Erd\H{o}s--Simonovits~\cite{Sidorenko,Sidorenko2, Simon} (see~\cite{CFS, CKLL, CL20, KLL, LS, S1} for some recent results towards this conjecture). We refer the interested reader to~\cite[Section 2.6]{Ramseysurvey} for more on this fascinating subject. Another much-studied problem concerns the value of $M(H, n)$ when it first becomes positive, i.e., when $n = r(H)$. As in our companion paper \cite{CFSW}, we refer to this value as the \emph{threshold Ramsey multiplicity}. \begin{definition} The \emph{threshold Ramsey multiplicity} $m(H)$ of a graph $H$ is the minimum number of monochromatic copies of $H$ in any two-coloring of the edges of $K_n$ with $n=r(H)$. In other words, \[m(H) = M(H, r(H)).\] \end{definition} The threshold Ramsey multiplicity was first studied systematically by Harary and Prins \cite{HP} almost fifty years ago. The exact value of the threshold Ramsey multiplicity is known for all graphs with at most $4$ vertices~\cite{Hsurvey, HP, K4}, but, in general, determining or even providing a non-trivial lower bound on the threshold Ramsey multiplicity appears to be quite challenging. In fact, the behavior of $m(H)$ can be rather erratic. For instance, Harary and Prins \cite{HP} proved that $m(K_2) = 1$ and $m(K_{1,k}) =1$ for $k$ even, but $m(K_{1,k})=2k$ for $k \geq 3$ odd. In the same paper \cite{HP}, Harary and Prins asked for a determination of $m(P_k)$ and $m(C_k)$. It is this question that concerns us in this paper and its companion~\cite{CFSW}. Indeed, in~\cite{CFSW}, not only did we provide the first non-trivial bound for the Ramsey multiplicity of paths and even cycles, but the bound is tight up to a lower-order factor. \begin{theorem}[\cite{CFSW}]\label{thm:even} There is a positive constant $c$ such that, for every positive integer $k$, the threshold Ramsey multiplicity of the path with $k$ vertices satisfies $m(P_k) \geq (ck)^k$ and, if $k$ is even, the threshold Ramsey multiplicity of the cycle on $k$ vertices satisfies $m(C_k) \geq (ck)^k$. \end{theorem} In this paper, we address Harary and Prins' question for odd cycles. Unlike the cases studied in~\cite{CFSW}, this odd-cycle case has received considerable previous attention, with Rosta and Sur\'anyi \cite{RStm} already proving the exponential lower bound $m(C_k)\geq 2^{c k}$ in the 1970's. This was later improved to a superexponential bound in an unpublished work of Rosta (see~\cite{KR}). More recently, K\'arolyi and Rosta~\cite{KR} improved the lower bound to $m(C_k) \geq k^{c k}$. To the best of our knowledge, this was the state-of-the-art prior to our result, which we now state. \begin{theorem}\label{thm:main} There is a positive constant $c$ such that, for every odd positive integer $k$, the threshold Ramsey multiplicity of the cycle on $k$ vertices satisfies $m(C_k) \geq (ck)^k$. \end{theorem} As for paths and even cycles, this bound is tight up to the constant $c$. However, it is proved using rather different methods to those employed in~\cite{CFSW}, because the Ramsey numbers, and the associated extremal colorings, are quite different for odd cycles and for paths and even cycles. To describe the extremal colorings in the odd setting, consider the red/blue edge-coloring $\chi(a,b)$ of the complete graph on $n=a+b$ vertices with vertex set $A \cup B$, $|A|=a$ and $|B|=b$, where $A$ and $B$ form blue cliques and all edges between $A$ and $B$ are red. Let $k \geq 5$ be an odd positive integer. Then $\chi(k-1, k-1)$ is a coloring of the complete graph on $r(C_k) - 1 = 2k - 2$ vertices with no monochromatic $C_k$, while $\chi(k, k-1)$ is a coloring of the complete graph on $r(C_k) = 2k-1$ vertices with exactly $(k-1)!/2$ monochromatic copies of $C_k$, as all monochromatic $C_k$ are in the blue clique of order $k$. This provides an upper bound on $m(C_k)$ showing that the bound in Theorem~\ref{thm:main} is tight apart from a lower-order factor. It also suggests that our bound can be strengthened, as follows. \begin{conjecture} \label{conj:ck} For any sufficiently large odd integer $k$, $m(C_k) = (k-1)!/2$. \end{conjecture} \section{Proof of Theorem \ref{thm:main}} \subsection{Preliminaries} As in our proof of Theorem \ref{thm:even} in \cite{CFSW}, we will use Szemer\'edi's regularity lemma, an important tool which gives a rough structural decomposition for all graphs. Roughly speaking, for any graph, the regularity lemma outputs a vertex partition of the graph into a small number of parts, where the bipartite graph between almost every pair of parts behaves like a random graph. Among its many applications (see, for example,~\cite{Regularity}), this decomposition is useful for embedding and counting copies of sparse graphs, such as the cycles that concern us here. To formally state the regularity lemma, we first need some definitions to quantify what is meant by a ``random-like" bipartite graph. For a pair of disjoint vertex subsets $(X, Y)$ of a graph, let $d(X, Y) = e(X,Y)/|X||Y|$ denote the density of edges between $X$ and $Y$. \begin{definition}[$\epsilon$-regular pair] A pair $(X, Y)$ of disjoint vertex subsets of a graph is said to be $\epsilon$-regular if, for all subsets $U \subset X, V \subset Y$ such that $|U| \geq \epsilon |X|$ and $|V| \geq \epsilon |Y|$, $|d(U, V) - d(X,Y)| \leq \epsilon$. \end{definition} The next lemma sets out two basic facts about $\epsilon$-regular pairs that will be useful later. \begin{lemma}\label{lem:epsregprop} If $(X,Y)$ is an $\epsilon$-regular pair and $d(X, Y)=d$, then the following hold: \begin{enumerate}[(i)] \item If $Y' \subset Y$ satisfies $|Y'| \geq \epsilon |Y|$, then the number of vertices in $X$ with degree in $Y'$ greater than $(d+\epsilon)|Y'|$ is less than $\epsilon |X|$ and the number of vertices in $X$ with degree in $Y'$ less than $(d-\epsilon)|Y'|$ is less than $\epsilon |X|$. \item If $X' \subset X$ and $Y'\subset Y$ are such that $|X'| \geq \alpha |X|$ and $|Y'| \geq \alpha |Y|$, then $(X', Y')$ is $\max(\epsilon/\alpha, 2\epsilon)$-regular. \end{enumerate} \end{lemma} A vertex partition is called {\it equitable} if each pair of parts differ in size by at most one. We are now ready to state the regularity lemma. \begin{lemma}[Szemer\'edi’s regularity lemma]\label{reglem} For every $\epsilon>0$ and positive integer $l$, there are positive integers $n_0$ and $M_0$ such that every graph $G$ with at least $n_0$ vertices admits an equitable vertex partition $V(G) = V_1 \cup \dots \cup V_M$ into $M$ parts with $l \leq M \leq M_0$ where all but at most $\epsilon \binom{M}{2}$ pairs of parts $(V_i, V_j )$ with $1 \leq i < j \leq M$ are $\epsilon$-regular. \end{lemma} In practice, we will use the following standard colored version of the regularity lemma. \begin{lemma}[Colored regularity lemma] \label{reglem-twocolor} For every $\epsilon>0$ and positive integer $l$, there are positive integers $n_0$ and $M_0$ such that every two-edge-coloring of the complete graph $K_n$ with $n \geq n_0$ in red and blue admits an equitable vertex partition $V(G)=V_1 \cup \dots \cup V_M$ into $M$ parts with $l \leq M \leq M_0$ where all but at most $\epsilon{M \choose 2}$ pairs of parts $(V_i,V_j)$ with $1 \leq i < j \leq M$ are $\epsilon$-regular in both the red and blue subgraphs. \end{lemma} Lemmas \ref{reglem} and \ref{reglem-twocolor} are in fact equivalent, since a pair $(V_i,V_j)$ in an edge-coloring of $K_n$ with colors red and blue is $\epsilon$-regular in red if and only if it is $\epsilon$-regular in blue. \subsection{The stability lemma} The main ingredient in the proof of Theorem~\ref{thm:main} is a stability lemma, Lemma \ref{lem:main2} below, which implies that any two-edge-coloring of $K_n$ (where, for us, $n$ will be $r(C_k) = 2k-1$ for some sufficiently large odd $k$) either has a regularity partition whose reduced graph contains a long monochromatic path or the edge-coloring of $K_n$ is close to the coloring $\chi(k, k-1)$ described before Conjecture \ref{conj:ck}. In either case, we can show that the conclusion of Theorem~\ref{thm:main} must hold. Before introducing the stability lemma, we need to make precise what we mean by saying that a two-edge-coloring of the complete graph on $2k-1$ vertices is close to $\chi(k, k-1)$. In the definition, we will refer to the density of a set $X$, given by $d(X,X) = 2e(X)/|X|^2$. \begin{definition}[Extremal coloring with parameter $\lambda$] \label{def:ec2} A two-edge-coloring of the complete graph on $n$ vertices is an {\it extremal coloring with parameter $\lambda$} if there exists a partition $ A \cup B$ of the vertex set such that \begin{itemize} \item $|A| \geq (1/2-\lambda)n$ and $|B| \geq (1/2 - \lambda)n$; \item the graph induced on $A$ has density at least $(1-\lambda)$ in some color, the graph induced on $B$ has density at least $(1-\lambda)$ in the same color, and the bipartite graph between $A$ and $B$ has density at least $(1-\lambda)$ in the other color. \end{itemize} \end{definition} Our key stability lemma is now as follows. \begin{lemma}\label{lem:main2} For any $0<\epsilon < 10^{-20}$, there is a constant $M_0=M_0(\epsilon)$ such that if $\alpha = 20 \sqrt{\epsilon}$, then, for $n$ sufficiently large in terms of $\epsilon$, any two-edge-coloring of the complete graph $K_n$ falls into one of the following two cases: \begin{itemize} \item \textbf{Case 1:} There is a positive integer $\epsilon^{-1} \leq M \leq M_0$ such that if $t$ is the odd integer with \begin{equation} (1/2 + \alpha) M \geq t > (1/2 + \alpha) M - 2, \label{eqn:reducedcycle} \end{equation} then there are disjoint vertex sets $V_0, \dots, V_{t-1}$, indexed by the elements of $\mathbb{Z}/t\mathbb{Z}$, and a color $\chi$ such that, for each $i \in\mathbb{Z} / t \mathbb{Z}$, $|V_i| \geq \lfloor n/M \rfloor$, the pair $(V_i, V_{i+1})$ is $\epsilon$-regular in the color $\chi$, and the edge density between $V_i$ and $V_{i+1}$ in the color $\chi$ is at least $11\epsilon^{1/2}$. \item \textbf{Case 2:} The graph is an extremal coloring with parameter $300 \sqrt{\alpha}$. \end{itemize} \end{lemma} We will hold off on proving Lemma~\ref{lem:main2} until Section~\ref{sec:stab}, first showing, across the next two sections, how Theorem~\ref{thm:main} follows from either of the conclusions in the lemma. \subsection{Theorem~\ref{thm:main} for colorings satisfying Case 1 of Lemma~\ref{lem:main2}} \label{subsec:case1odd} In this section, we prove Theorem~\ref{thm:main} for colorings satisfying Case 1 of Lemma~\ref{lem:main2}. We will repeatedly work in a setting where we have disjoint vertex sets $V_0, \dots, V_{t-1}$ from a graph where the indices of the $V_i$ are the elements of $\mathbb{Z} / t \mathbb{Z}$. We say that a path $P$ of length $\ell$ with vertices $w_0, w_1, \dots, w_\ell$ and edges $w_0 w_1, w_1 w_2, \dots, w_{\ell - 1} w_\ell$ is {\it $(V_0, \dots, V_{t-1})$-transversal} if $w_i \in V_i$ for each $0 \leq i \leq \ell$. Note that we will typically have $\ell > t$, so the path may pass through each of the vertex sets multiple times. \begin{lemma}\label{lem:countpath2} Suppose that $0<\epsilon < 10^{-5}$ and $t$ and $n$ are integers with $t\geq 2$ and $n \geq \epsilon^{-2}$. Suppose also that $V_0, \dots, V_{t-1}$ are disjoint vertex sets in a graph where the indices of the $V_i$ are the elements of $\mathbb{Z} / t \mathbb{Z}$ and, for each $i \in\mathbb{Z} / t \mathbb{Z}$, $|V_i| \geq n$, $(V_i, V_{i+1})$ is $\epsilon$-regular, and $d(V_i,V_{i+1}) \geq d$ for some $d \geq 5\epsilon^{1/2}$. Then the following hold: \begin{enumerate} \item For any integer $\ell$ with $2\leq \ell \leq t (1-\sqrt{\epsilon}) n$ and any vertex $w_0 \in V_0$ with at least $(d-\epsilon)|V_1|$ neighbors in $V_1$, the number of $(V_0, \dots, V_{t-1})$-transversal paths of length $\ell$ starting from $w_0$ is at least $(d-\epsilon - \sqrt{\epsilon})^\ell \prod_{i=1}^\ell (n-\lfloor i/t \rfloor)$. \item For any integer $\ell$ with $4 \leq \ell \leq t (1-3\sqrt{\epsilon}) n$ which is divisible by $t$ and any two (not necessarily distinct) vertices $w_0, w_0' \in V_0$ such that $w_0$ has at least $(d-\epsilon)|V_1|$ neighbors in $V_1$ and $w'_0$ has at least $(d-\epsilon)|V_{t-1}|$ neighbors $V_{t-1}$, the number of $(V_0, \dots, V_{t-1})$-transversal paths of length $\ell$ with end vertices $w_0$ and $w_0'$ is at least $(d-5\sqrt{\epsilon})^{\ell-1} (1-2\sqrt{\epsilon})^{\ell-2} (\epsilon n)\prod_{i=1}^{\ell-2} (n-\lfloor i/t \rfloor).$ \end{enumerate} \end{lemma} \begin{proof} For any integer $0 \leq l\leq \ell$, let $N_l$ be the number of good paths of length $l$ starting from $w_0$, where a $(V_0, \dots, V_{t-1})$-transversal path $w_0, w_1, \dots, w_l$ of length $l$ starting from $w_0$ is {\it good} if there are at least $(d-\epsilon)\big(|V_{l+1}| - \lfloor (l+1)/t \rfloor\big)$ ways to extend the path to $V_{l+1}$. We will prove by induction that $N_l \geq (d-\epsilon - \sqrt{\epsilon})^l \prod_{i=1}^l (n-\lfloor i/t \rfloor)$ for $0 \leq l \leq \ell$, which will settle Part 1. For the base case, note that $N_0=1$, since the path with zero edges starting from $w_0$ is $w_0$ itself and, by assumption, the vertex $w_0 \in V_0$ has at least $(d-\epsilon)|V_1|$ neighbors in $V_1$. Suppose now that the required lower bound holds for $N_{l-1}$ and we wish to deduce the lower bound for $N_l$. Fix any good path $P$ of length $l-1$. By the definition of goodness, there are at least $(d - \epsilon) (|V_{l}| - \lfloor l/t \rfloor)$ choices of $w_l \in V_l$ that extend $P$. To bound $N_l$, we need a lower bound on the number of vertices $w_l \in V_l$ such that the path formed by extending $P$ to $w_l$ is also good. Let $U$ be the set of vertices in $V_{l}$ which have degree less than $(d - \epsilon) |V_{l+1} \setminus V(P)| = (d - \epsilon) (|V_{l+1}| - \lfloor (l+1)/t \rfloor)$ in $V_{l+1} \setminus V(P)$. Note that, since $\ell \leq t (1-\sqrt{\epsilon}) n$, \begin{align*} |V_{l+1} \setminus V(P)| & = |V_{l+1}| - \lfloor (l+1)/t \rfloor \geq |V_{l+1}| - \ell /t - 1 \geq |V_{l+1}| - t(1-\sqrt{\epsilon})n/t - 1\\ & \geq |V_{l+1}| - (1-\sqrt{\epsilon})|V_{l+1}| - 1 = \sqrt{\epsilon} |V_{l+1}| - 1 \geq \epsilon |V_{l+1}|. \end{align*} Together with the fact that $(V_{l}, V_{l+1})$ is $\epsilon$-regular with density at least $d$, Lemma~\ref{lem:epsregprop} (i) implies that $|U| \leq \epsilon |V_{l}|$. Hence, the number of choices for $w_{l}$ such that $P$ extended to $w_l$ is also good is at least \[ (d-\epsilon)(|V_l| - \lfloor l/t \rfloor) - |U| \geq (d-\epsilon)(|V_l| - \lfloor l/t \rfloor) - \epsilon |V_l| \geq (d-\epsilon - \sqrt{\epsilon})(|V_l| - \lfloor l/t \rfloor). \] The last inequality is equivalent to $(\sqrt{\epsilon} - \epsilon) |V_l| \geq \sqrt{\epsilon} \lfloor \ell/t \rfloor$, which again follows from $\ell \leq t (1-\sqrt{\epsilon}) n$. Thus, \begin{equation*} N_l \geq (d-\epsilon-\sqrt{\epsilon}) (n-\lfloor l/t \rfloor) N_{l-1} \geq (d-\epsilon-\sqrt{\epsilon})^l \prod_{i=1}^l (n-\lfloor i/t \rfloor), \end{equation*} establishing Part 1. For Part 2, we pass from the $V_i$ to a collection of subsets $V'_i$. By assumption, $|N_{V_{t-1}}(w_0')|\geq (d-\epsilon) |V_{t-1}|\geq \epsilon |V_{t-1}|$, so we may set aside a subset $W_{t-1}$ of $N_{V_{t-1}}(w_0')$ of size $\epsilon |V_{t-1}|$ and let $V'_{t-1} = V_{t-1} \setminus W_{t-1}$. If $w'_0$ is distinct from $w_0$, we let $V'_0 = V_0 \setminus \{w'_0\}$, while, in all remaining cases, we let $V'_i = V_i$, noting that $|V'_i| \geq (1 - \epsilon) |V_i|$ for all $i$. Therefore, if we set $\epsilon' = 2\epsilon$ and $d' = d-\epsilon$, Lemma~\ref{lem:epsregprop} (ii) now tells us that for each $i \in \mathbb{Z} /t\mathbb{Z}$ the pair of sets $(V_i', V_{i+1}')$ is $\epsilon'$-regular with edge density at least $d'$. As in Part 1, for any positive integer $0 \leq l \leq \ell-3$, let $N_l$ be the number of good paths of length $l$ starting from $w_0$, though with the condition now reading that there are at least $(d'-\epsilon')(|V_{l+1}'| - \lfloor (l+1)/t \rfloor)$ ways to extend the path to a vertex in $V_{l+1}'$. Since $w_0$ has at least $(d-\epsilon)|V_1| - |V_1 \setminus V'_1| \geq (d-2\epsilon) |V_1| > (d'-\epsilon') |V'_1|$ neighbors in $V'_1$ and $\ell \leq t(1-3\sqrt{\epsilon})n \leq t(1-\sqrt{2\epsilon})((1-\epsilon)n) = t(1-\sqrt{\epsilon'}) ((1-\epsilon)n) $, we may apply Part 1 to conclude that \begin{equation} N_{\ell-3} \geq (d'-\epsilon'-\sqrt{\epsilon'})^{\ell-3} \prod_{i=1}^{\ell-3} ((1-\epsilon)n-\lfloor i/t \rfloor). \label{eqn:p1Nl3} \end{equation} Fix any such path $P$ of length $\ell-3$. Suppose its vertices are $w_0, w_1, \dots, w_{\ell-3}$ in order, where $w_j \in V_j'$. By definition, there are at least $(d'-\epsilon')(|V_{\ell-2}'| - \lfloor (\ell-2)/t \rfloor)$ ways to extend the path to a vertex $w_{\ell-2} \in V_{\ell-2}'$. Denote this set of candidates for $w_{\ell-2}$ by $C$. Since $\ell$ is divisible by $t$, we have that $C \subset V_{t-2}$. Using that $d \geq 5\sqrt{\epsilon}$ and $\ell/t \leq (1-3\sqrt{\epsilon})n$, we have that $|C| \geq (d'-\epsilon')(|V_{\ell-2}'|- \lfloor (\ell-2)/t \rfloor) \geq \epsilon' |V_{t-2}'| \geq \epsilon |V_{t-2}|$. Since $|W_{t-1}| \geq \epsilon |V_{t-1}|$ and $(V_{t-2}, V_{t-1})$ is $\epsilon$-regular with density at least $d$, the number of edges between $C$ and $W_{t-1}$ is at least \begin{equation} (d - \epsilon) |C||W_{t-1}| \geq (d-\epsilon) (d'-\epsilon')(|V_{\ell-2}'|- \lfloor (\ell-2)/t \rfloor) \cdot \epsilon |V_{t-1}|. \label{eqn:CVb} \end{equation} Note now that $P$ together with the two end vertices of any edge in $E(C, W_{t-1})$ results in a path of length $\ell-1$ that can be extended to $w_0'$. Therefore, using \eqref{eqn:p1Nl3} and \eqref{eqn:CVb}, we see that the number of paths with end vertices $w_0$ and $w_0'$ is at least \begin{align*} N_{\ell-3}\cdot (d - \epsilon) |C||W_{t-1}| & \geq (d'-\epsilon'-\sqrt{\epsilon'})^{\ell-3} \prod_{i=1}^{\ell-3} ((1-\epsilon)n-\lfloor i/t \rfloor) \\ & \ \ \ \ \ \ \cdot (d-\epsilon) (d'-\epsilon')(|V_{\ell-2}'|- \lfloor (\ell-2)/t \rfloor) \cdot \epsilon |V_{t-1}| \\ & \geq (d-3\epsilon-2\sqrt{\epsilon})^{\ell-1} (\epsilon n) \prod_{i=1}^{\ell-2} ((1-\epsilon)n-\lfloor i/t \rfloor) \\ & \geq (d-3\epsilon-2\sqrt{\epsilon})^{\ell-1}(1-\epsilon - \sqrt{\epsilon} )^{\ell-2} (\epsilon n) \prod_{i=1}^{\ell-2} (n-\lfloor i/t \rfloor) \\ & \geq (d-5\sqrt{\epsilon})^{\ell-1} (1-2\sqrt{\epsilon})^{\ell-2} (\epsilon n)\prod_{i=1}^{\ell-2} (n-\lfloor i/t \rfloor), \end{align*} where the second to last inequality holds because $(1-\epsilon)n - \lfloor i/t \rfloor \geq (1-\epsilon - \sqrt{\epsilon}) (n - \lfloor i/t \rfloor )$ for $i \leq t(1-2\sqrt{\epsilon})n$. \end{proof} Part 2 of Lemma \ref{lem:countpath2} with $w_0 = w_0'$ implies that there are many cycles of length $\ell$ when $\ell$ is divisible by $t$. The next lemma shows that the same result holds even when $\ell$ is not divisible by $t$. \begin{lemma}\label{lem:countcycle1} Suppose that $0< \epsilon <10^{-5}$, $t$ is an odd integer with $t\geq 3$, and $n$ is a positive integer with $n \geq t\epsilon^{-2}$. Suppose also that $V_0, \dots, V_{t-1}$ are disjoint vertex sets in a graph where the indices of the $V_i$ are the elements of $\mathbb{Z}/t\mathbb{Z}$ and, for each $i\in \mathbb{Z}/t\mathbb{Z}$, $|V_i| \geq n$, $(V_i, V_{i+1})$ is $\epsilon$-regular, and $d(V_i,V_{i+1}) \geq d$ for some $d \geq 10\epsilon^{1/2}$. Then, for any odd positive integer $p$ with \begin{equation} 2t+6 \leq p \leq t(1-5\sqrt{\epsilon})n, \label{eqn:q} \end{equation} the number of cycles of length $p$ is at least $\frac{\epsilon^2}{4} n^4 (d-10\sqrt{\epsilon})^{p-2}(1-3\sqrt{\epsilon})^{2p} \prod_{i=1}^{p-4} (n-\lfloor i/t \rfloor)$. \end{lemma} \begin{proof} Suppose $p \equiv r \bmod t$ for some $0 \leq r \leq t-1$. If $r = 0$, then we can choose $w_0=w_0'$ in Part 2 of Lemma \ref{lem:countpath2} in at least $(1-2\epsilon)n$ ways. Then this lemma easily implies that the number of cycles of length $p$ is at least $(1-2\epsilon)\epsilon n^2 (d-5\sqrt{\epsilon})^{p-1} (1-2\sqrt{\epsilon})^{p-2} \prod_{i=1}^{p-2} (n-\lfloor i/t \rfloor)$, which is stronger than the required bound. We may therefore assume that $p$ is not divisible by $t$, i.e., that $r >0$. Define an auxiliary constant $h$ by $h= r+t$ if $r$ is odd, $h = r$ if $r > 2$ and even, and $h =2+2t$ if $r = 2$. Note that $h$ is always a positive even integer with $4 \leq h \leq 2t+ 2$, so a cycle of length $p$ can be constructed by combining an even path $L_1$ of length $h$ alternating between $V_0$ and $V_{t-1}$ with a $(V_0, \dots, V_{t-1})$-transversal path $L_2$ of length $p-h$ with the same end vertices as $L_1$. Since $(p-h)$ is divisible by $t$, the path $L_2$ will use exactly $(p-h)/t$ vertices in each of the $V_i$. By Lemma \ref{lem:epsregprop} (i), the set $U_0$ of vertices in $V_0$ with at least $(d -\epsilon)|V_1|$ neighbors in $V_1$ and at least $(d -\epsilon)|V_{t-1}|$ neighbors in $V_{t-1}$ has size at least $(1-2\epsilon)|V_0|$. We now fix two vertices $u, v \in U_0$ and bound the number of paths of the types $L_1$ and $L_2$ described above with end vertices $u$ and $v$. We first bound the number of paths of type $L_1$ with end vertices $u$ and $v$. Since $h \geq 4$ and $h \leq 2t + 2 \leq 2(1-3\sqrt{\epsilon})n$, we may apply Part 2 of Lemma \ref{lem:countpath2} to $(V_0, V_{t-1})$ to conclude that the number of paths of length $h$ alternating between $V_0$ and $V_{t-1}$ with end vertices $u$ and $v$ is at least \begin{equation} (d-5\sqrt{\epsilon})^{h-1} (1-2\sqrt{\epsilon})^{h-2} (\epsilon n)\prod_{i=1}^{h-2} (n-\lfloor i/2 \rfloor). \label{eq:part1} \end{equation} We now bound the number of paths of type $L_2$ available for each such $L_1$. For a given $L_1$, remove its $h-1$ interior vertices from $V_0$ and $ V_{t-1}$, calling the updated vertex sets $V'_0$ and $V'_{t-1}$, respectively. Since $h \leq 2t+2$, we have $|V_0'| \geq |V_0| - (t+1) \geq (1-\epsilon/2)|V_0|$ and, similarly, $|V_{t-1}'| \geq (1-\epsilon/2)|V_{t-1}|$. Hence, by Lemma \ref{lem:epsregprop} (ii), each of the pairs $(V_0', V_{t-1}')$, $(V_0', V_1)$, and $(V_{t-1}', V_{t-2})$ is $2\epsilon$-regular with density at least $d-\epsilon$. Furthermore, $u$ and $v$ each have at least $(d-\epsilon)|V_{t-1}| - (t+1) \geq (d-2\epsilon)|V_{t-1}| \geq (d-2\epsilon)|V_{t-1}'|$ neighbors in $V_{t-1}'$. Since also \[ 4 \leq p-h \leq t(1-5\sqrt{\epsilon})n \leq t(1-3\sqrt{2\epsilon})(1-\epsilon/2)n,\] we may apply Part 2 of Lemma \ref{lem:countpath2} with $\epsilon$, $\ell$, $d$, and $n$ replaced by $2\epsilon$, $p-h$, $d-\epsilon$, and $(1-\epsilon/2)n$, respectively, to conclude that the number of $(V'_0, V_1, \dots, V_{t-2}, V'_{t-1})$-transversal paths of length $p-h$ with end vertices $u$ and $v$, each of which is a valid choice for $L_2$, is at least \begin{align} & (d-\epsilon - 5\sqrt{2\epsilon})^{p-h-1} (1-2\sqrt{2\epsilon})^{p-h-2} (2\epsilon (1-\epsilon/2)n)\prod_{i=1}^{p-h-2} ((1-\epsilon/2)n-\lfloor i/t \rfloor) \nonumber \\ & \geq (d-10\sqrt{\epsilon})^{p-h-1} (1-3\sqrt{\epsilon})^{p-h-2} (\epsilon n)\prod_{i=1}^{p-h-2} ((1-\epsilon/2)n-\lfloor i/t \rfloor) \nonumber\\ & \geq (d-10\sqrt{\epsilon})^{p-h-1} (1-3\sqrt{\epsilon})^{p-h-2} (\epsilon n) (1-\epsilon/2 - \sqrt{\epsilon/2})^{p-h-2} \prod_{i=1}^{p-h-2} (n-\lfloor i/t \rfloor) \nonumber \\ & \geq (d-10\sqrt{\epsilon})^{p-h-1} (1-3\sqrt{\epsilon})^{p-h-2} (\epsilon n) (1- 2\sqrt{\epsilon})^{p-h-2} \prod_{i=1}^{p-h-2} (n-\lfloor i/t \rfloor) \nonumber \\ & \geq (d-10\sqrt{\epsilon})^{p-h-1} (1-3\sqrt{\epsilon})^{2(p-h-2)} (\epsilon n) \prod_{i=1}^{p-h-2} (n-\lfloor i/t \rfloor), \label{eq:part2} \end{align} where the second inequality holds because $(1-\epsilon/2)n - \lfloor i/t \rfloor \geq (1-\epsilon/2 - \sqrt{\epsilon/2}) (n - \lfloor i/t \rfloor )$ for $i \leq t(1-2\sqrt{\epsilon/2})n$. Therefore, since the number of choices for $u$ and $v$ is at least $\frac{1}{2} (1-2\epsilon)|V_0| ((1-2\epsilon)|V_0|-1)$, we may combine \eqref{eq:part1} and \eqref{eq:part2} to conclude that the number of cycles of length $p$ is at least \begin{align*} & \frac{1}{2} (1-2\epsilon)|V_0| ((1-2\epsilon)|V_0|-1) \cdot (d-5\sqrt{\epsilon})^{h-1} (1-2\sqrt{\epsilon})^{h-2} (\epsilon n)\prod_{i=1}^{h-2} (n-\lfloor i/2 \rfloor) \\ & \ \ \ \ \ \cdot (d-10\sqrt{\epsilon})^{p-h-1} (1-3\sqrt{\epsilon})^{2(p-h-2)} (\epsilon n) \prod_{i=1}^{p-h-2} (n-\lfloor i/t \rfloor) \\ &\geq \frac{1}{4} (1-2\epsilon)^2 n^2 (d-10\sqrt{\epsilon})^{p-2}(1-3\sqrt{\epsilon})^{2p-h-2}(\epsilon n)^2 \prod_{i=1}^{p-4} (n-\lfloor i/t \rfloor) \\ & \geq \frac{\epsilon^2}{4} n^4 (d-10\sqrt{\epsilon})^{p-2}(1-3\sqrt{\epsilon})^{2p} \prod_{i=1}^{p-4} (n-\lfloor i/t \rfloor), \end{align*} as required. \end{proof} Finally, we can show that Theorem~\ref{thm:main} holds for colorings satisfying Case 1 of Lemma \ref{lem:main2}. \begin{proof}[Proof of Theorem \ref{thm:main} for colorings satisfying Case 1 of Lemma \ref{lem:main2}.] Suppose, for concreteness, that $\epsilon = 10^{-30}$ and let $V_0, \dots, V_{t-1}$ be as in Case 1 of Lemma~\ref{lem:main2} with $n = 2k - 1$, where $k$ (and hence $n$) is a sufficiently large odd integer. We wish to apply Lemma~\ref{lem:countcycle1} to show there are many cycles of length $k = (n+1)/2$. To confirm that the conditions of Lemma \ref{lem:countcycle1} hold, we only have to check \eqref{eqn:q}, i.e., that \begin{equation} 2t+6 \leq k \leq t(1-5\sqrt{\epsilon}) \lfloor n/M \rfloor.\label{eqn:knM} \end{equation} By \eqref{eqn:reducedcycle}, $(1/2 + \alpha) M \geq t > (1/2 + \alpha) M - 2$, so it suffices to show that \[k = (n+1)/2 \geq 2(1/2+\alpha)M+6\] and \[ k \leq \left( (1/2 + \alpha) M - 2 \right) (1-5\sqrt{\epsilon})(n/M-1).\] The first inequality easily holds for $n$ sufficiently large in terms of $\epsilon$, while the second inequality holds because \begin{align*} k = (n+1)/2 \leq (1+\epsilon)n/2 \leq (1/2 + \alpha/2)M (1-5\sqrt{\epsilon})(1-\epsilon)(n/M), \end{align*} where we used that $\alpha=20\sqrt{\epsilon}$ and $ (1+\alpha)(1-5\sqrt{\epsilon})(1-\epsilon) \geq (1 + 20\sqrt{\epsilon})(1-6\sqrt{\epsilon}) > 1+ \epsilon $. Hence, since $M \geq 2/\alpha$ and assuming $n \geq M/\epsilon$, \begin{align*} k \leq & (1/2 + \alpha/2)M (1-5\sqrt{\epsilon})(1-\epsilon)(n/M) \leq \left( (1/2 + \alpha) M - 2 \right) (1-5\sqrt{\epsilon})(n/M-1), \end{align*} as required. We may therefore apply Lemma \ref{lem:countcycle1} with parameters $\epsilon$, $p$, $d$, and $n$ replaced by $\epsilon$, $k$, $d$, and $\lfloor n/M \rfloor$, respectively, to conclude that the number of cycles of length $k$ is at least \begin{equation} \frac{\epsilon^2}{4} ( \lfloor n/M \rfloor)^4 (d-10\sqrt{\epsilon})^{k-2}(1-3\sqrt{\epsilon})^{2k} \prod_{i=1}^{k-4} ( \lfloor n/M \rfloor-\lfloor i/t \rfloor). \label {eq:cyclek} \end{equation} Since $ \lfloor n/M \rfloor \geq k/t$ by (\ref{eqn:knM}), the last term in (\ref{eq:cyclek}) satisfies \begin{equation*} \prod_{i=1}^{k-4} ( \lfloor n/M \rfloor-\lfloor i/t \rfloor) \geq \prod_{i=1}^{k-4} ( \lfloor n/M \rfloor- i/t) \geq t^{-(k-4)} (k-4)!. \end{equation*} Therefore, (\ref{eq:cyclek}) is lower bounded by \begin{align*} & \frac{\epsilon^2}{4} ( \lfloor n/M \rfloor)^4 (d-10\sqrt{\epsilon})^{k-2}(1-3\sqrt{\epsilon})^{2k} t^{-(k-4)}(k-4)! \\ & \geq \frac{\epsilon^2}{4} \frac{n^4}{(2M)^4} (d-10\sqrt{\epsilon})^{k-2}(1-3\sqrt{\epsilon})^{2k}t^{-(k-4)} ((k-4)/e)^{k-4} \\ & \geq \frac{\epsilon^2}{4(2M)^4} (d-10\sqrt{\epsilon})^{k-2}(1-3\sqrt{\epsilon})^{2k}t^{-(k-4)} ((k-4)/e)^{k}. \end{align*} Hence, since $\epsilon$ is a constant, $d \geq 11 \sqrt{\epsilon}$, and $M$ and $t$ are bounded in terms of $\epsilon$, there is a constant $c_1$ depending only on $\epsilon$ such that the number of cycles of length $k$ is at least $(c_1 k)^k$, as required. \end{proof} \subsection{Theorem~\ref{thm:main} for colorings satisfying Case 2 of Lemma~\ref{lem:main2}} \label{subsec:case2odd} The proof of Theorem~\ref{thm:main} for colorings satisfying Case 2 of Lemma~\ref{lem:main2} has several cases. To make the presentation cleaner, we first prove some simple claims. \begin{claim}\label{claim:redgeneral} Let $S$ and $T$ be two disjoint vertex sets in a graph $F$ such that any two vertices in $S$ have at least $s$ common neighbors in $T$. If there is an edge within $S$, then the number of cycles of length $l$ is at least $\left( s-l/2 +3/2\right)^{(l-1)/2} (|S| - (l-1)/2)^{(l-3)/2}$ for any odd integer $3 \leq l \leq \min(2s+1, 2|S|-1)$. \end{claim} \begin{proof} Suppose that $(u,u')$ is an edge in $S$. To construct cycles of length $l$, we will find paths $P$ of the form $u, v_1, u_1, v_2, u_2, \dots, v_{(l-1)/2}, u_{(l-1)/2} = u'$ alternating between $S$ and $T$ with end vertices $u$ and $u'$. Each such path together with the edge $(u,u')$ clearly gives rise to a cycle of length $l$. To estimate the number of paths $P$ of the required form, suppose that $u, u_1, u_2, \dots, u_{(l-1)/2-1}$, $u_{(l-1)/2} = u'$ is a fixed sequence of distinct vertices in $S$. Note that any of the $s$ common neighbors of $u$ and $u_1$ in $T$ can be chosen as $v_1$. More generally, given choices for $v_1, \dots, v_{i-1}$, the number of choices for $v_i$ is at least $s - (i-1)$ for $1 \leq i \leq (l-1)/2$, since we may pick any vertex in the common neighborhood of $u_{i-1}$ and $u_{i}$ in $T$ except $v_1, \dots, v_{i-1}$. Therefore, given $u, u_1, u_2, \dots, u_{(l-1)/2-1}$, $u_{(l-1)/2} = u'$, the number of choices for $P$ is at least \[ \prod_{i=1}^{(l-1)/2} (s - (i-1)) \geq \left( s-((l-1)/2-1)) \right)^{(l-1)/2} = \left( s-l/2 +3/2 \right)^{(l-1)/2} .\] Since the number of choices for $u_1, \dots, u_{(l-1)/2-1}$ is \[(|S|-2)(|S|-3) \dots (|S| - (l-1)/2) \geq (|S| - (l-1)/2)^{(l-3)/2},\] the claim follows. \end{proof} \begin{claim}\label{claim:ABedgegeneral} Let $S$ and $T$ be two disjoint cliques in a graph $F$. Suppose that there are two vertex-disjoint paths $P_1$ and $P_2$ such that each path has length at most $2$, has one end vertex in $S$, the other end vertex in $T$, and the rest of the vertices outside $S \cup T$. Then the number of cycles of length $l$ is at least $(((l-1)/2-3)/e)^{l-6}$ for any integer $7 \leq l \leq \min(2|S|-1, 2|T|-1)$. \end{claim} \begin{proof} Suppose that $P_1$ has length $l_1$ and endpoints $a_1 \in S$ and $b_1 \in T$, while $P_2$ has length $l_2$ and endpoints $a_2 \in S$ and $b_2 \in T$. Let $s = \lfloor l/2 \rfloor$. We will construct cycles of length $l$ by concatenating the following four paths: (1) a path $L_1$ of length $s - l_1$ in $S$ with end vertices $a_1$ and $a_2$, (2) the path $P_2$, (3) a path $L_2$ in $T$ of length $l-s-l_2$ with end vertices $b_1$ and $b_2$, and (4) the path $P_1$. This process clearly yields a cycle of length $l$. Since $S$ is a clique, we can always find paths $L_1$ of length $s-l_1$ in $S\setminus \{a_1, a_2\}$ with end vertices $a_1$ and $a_2$. Indeed, the number of such $L_1$ is exactly the number of length-$(s-l_1-1)$ ordered sequences of vertices in $S \setminus \{a_1, a_2\}$. Since $|S| - 2 \geq s-l_1 -1$, such a sequence exists. Furthermore, the number of such sequences is exactly $(|S| - 2)!/(|S| - 2 -(s-l_1-1))! \geq (s-l_1-1)! \geq ((s-l_1-1)/e)^{s-l_1-1}$, where we used the inequality $(x-1)\cdots (x-y) \geq y! \geq (y/e)^y$ for positive integers $x, y$ with $x \geq y+1$. Similarly, the number of choices for $L_2$ is at least $((l-s-l_2-1)/e)^{l-s-l_2-1}$. In total, the number of cycles of length $l$ is at least $ ((s-l_1-1)/e)^{s-l_1-1} ((l-s-l_2-1)/e)^{l-s-l_2-1}$. Since $l_1, l_2 \leq 2$ and $l-s\geq s$, the quantity above is at least $((s-3)/e)^{s-l_1-1} ((s-3)/e)^{l-s-l_2-1}$. If $s-3 \geq e$, then the previous quantity is at least $((s-3)/e)^{l-6}$. Otherwise, we counted a positive integer number of cycles of length $l$, which is at least the bound in the claim. \end{proof} \begin{claim}\label{claim:PbothABgeneral} Let $S$ and $T$ be two disjoint vertex sets in a graph $F$. Suppose that $w \in S$ and there is a complete bipartite graph between $S \setminus \{w\}$ and $T$ and at least one edge between $w$ and $T$ (so, in particular, there may be a complete bipartite graph between $S$ and $T$). If there is a path $P'$ of length two with one end vertex in $S$, the other end vertex in $T$, and the rest of the vertices outside $S \cup T$, then the number of cycles of length $l$ is at least $((l-5)/2e)^{l-5}$ for any odd integer $l$ with $7 \leq l \leq \min(2|S|+1, 2|T|+1)$. \end{claim} \begin{proof} Suppose the two end vertices of $P'$ are $a\in S$ and $b\in T$. We will construct cycles of length $l$ by concatenating $P'$ with paths $P$ of length $l-2$ alternating between $S$ and $T$ with end vertices $a$ and $b$. Fix a neighbor $x$ of $w$ in $T$. If $a = w$, each path $P$ will start with $w$ and then $x$ before returning to some $a' \in S$, while if $a \neq w$, we will avoid $w$ while building our paths. In either case, a lower estimate for the number of cycles of length $l$ is given by estimating the number of paths of length $l - 4$ starting at a fixed $a' \neq w$ and ending at $b$ alternating between $S \setminus \{w\}$ and $T \setminus \{x\}$. Since there is a complete bipartite graph between $S \setminus \{w\}$ and $T \setminus \{x\}$, any length-$(l-5)/2$ sequence of ordered vertices in $S\setminus \{a', w\}$ and any length-$(l-5)/2$ sequence of ordered vertices in $T \setminus \{b, x\}$ give rise to a relevant path by alternating between the two sequences as interior vertices. Such sequences exist because $|S| -2 \geq (l-5)/2$ and $|T| - 2 \geq (l-5)/2$. Thus, the number of choices for the path is at least the product of the number of such sequences in $S \setminus \{a', w\}$ and $T \setminus \{b, x\}$, which is \begin{align*} \frac{(|S|-2)!}{(|S| - (l-5)/2) - 2)!} \cdot \frac{(|T|-2)!}{(|T| - (l-5)/2) - 2)!} & \geq ((l-5)/2)! ((l-5)/2)! \\ & \geq ((l-5)/2e)^{l-5}, \end{align*} where we again used that $(x-1)\cdots (x-k) \geq k! \geq (k/e)^k$ for positive integers $x, k$ with $x \geq k+1$. \end{proof} \begin{claim}\label{claim:tworedABgeneral} Let $S$ and $T$ be two disjoint vertex sets in a graph $F$. If there are no two vertex-disjoint edges between $S$ and $T$, then, by removing at most one vertex from $S\cup T$, there is no edge between $S$ and $T$. \end{claim} \begin{proof} If there is no vertex in $S$ with a neighbor in $T$, the claim trivially holds. If there is exactly one vertex $a \in S$ with neighbors in $T$, then there is no edge between $S \setminus \{a\}$ and $T$. If there is more than one vertex in $S$ with a neighbor in $T$, then all of them have the same neighbor $b \in T$ and there is no edge between $S$ and $T \setminus \{b\}$. In each case, the claim follows. \end{proof} We are now ready to prove Theorem \ref{thm:main} for colorings satisfying Case 2 of Lemma \ref{lem:main2}. \begin{proof}[Proof of Theorem \ref{thm:main} for colorings satisfying Case 2 of Lemma \ref{lem:main2}.] Suppose again that $\epsilon = 10^{-30}$ and $n = 2k - 1$ for $k$ a sufficiently large odd integer, but we now have an extremal coloring of $K_n$ with parameter $\lambda = 300\sqrt{\alpha}$ and vertex partition $A \cup B$, as in Case 2 of Lemma~\ref{lem:main2}. Without loss of generality, we will assume that the red densities within $A$ and $B$ are both at least $1-\lambda$ and the blue density between $A$ and $B$ is at least $1-\lambda$, where $|A|, |B| \geq (1/2 - \lambda)n$. We first conduct a simple cleaning-up procedure. \begin{claim} \label{claim:updateEC} There is a vertex partition of $K_n$ as $A'\cup B' \cup X \cup Y$ satisfying the following conditions: \begin{enumerate} \item $A = A' \cup X$ and $B = B' \cup Y$. \item $|X | \leq 2\sqrt{\lambda}|A|$ and $|Y| \leq 2\sqrt{\lambda}|B|$. \item $|A'| \geq (1/2 - 2\sqrt{\lambda})n$ and $|B'| \geq (1/2 - 2\sqrt{\lambda})n$. \item Each vertex in $A'$ has red degree at least $ (1-3\sqrt{\lambda})|A|$ in $A'$ and blue degree at least $ (1-3\sqrt{\lambda})|B|$ in $B'$. Similarly, each vertex in $B'$ has red degree at least $ (1-3\sqrt{\lambda})|B|$ in $B'$ and blue degree at least $ (1-3\sqrt{\lambda})|A|$ in $A'$. \end{enumerate} \end{claim} \begin{proof} Suppose that there are $x|A|$ vertices in $A$ whose red degree in $A$ is at most $(1-\sqrt{\lambda})|A|$. Then, \[ x|A|(1-\sqrt{\lambda})|A| + (1-x)|A||A| \geq (1-\lambda) |A|^2, \] which implies that $x \leq \sqrt{\lambda}$. Similarly, there are at most $\sqrt{\lambda}|A|$ vertices in $A$ whose blue degree in $B$ is at most $(1-\sqrt{\lambda})|B|$. Letting $X$ be the union of these two bad sets of vertices, we see that $|X | \leq 2\sqrt{\lambda}|A|$. We define $Y \subset B$ similarly, again noting that $|Y| \leq 2\sqrt{\lambda}|B|$. Letting $A' = A \setminus X$ and $B' = B \setminus Y$, we see that items 1 and 2 hold. To verify item 3, note that $|A'| = |A| - |X| \geq (1-2\sqrt{\lambda})|A|$. Since $|A| \geq (1/2 - \lambda) n$, we have $$|A'| \geq (1-2\sqrt{\lambda})(1/2-\lambda)n \geq (1/2 - 2\sqrt{\lambda})n,$$ as required. Similarly, $|B'| \geq (1/2 - 2\sqrt{\lambda})n$. Finally, for item 4, note, for example, that each vertex in $A' = A \setminus X$ has red degree at least $(1-\sqrt{\lambda})|A| - |X| \geq (1-3\sqrt{\lambda})|A|$ in $A'$, while each vertex in $A'$ has blue degree at least $(1-\sqrt{\lambda})|B| - |Y| \geq (1-3\sqrt{\lambda})|B|$ in $B'$. \end{proof} The following claim allows us to assume that all the edges in $A'$ and $B'$ are red, i.e., that $A'$ and $B'$ are both red cliques, as otherwise we would be done. \begin{claim}\label{claim:red} If there is a blue edge within $A'$ or $B'$, then the number of blue cycles of length $k$ with $k = (n+1)/2$ is at least $(n/5)^{k-2}.$ \end{claim} \begin{proof} Without loss of generality, suppose that there is a blue edge $(u,u')$ in $A'$. We will apply Claim \ref{claim:redgeneral} to the blue subgraph with $S = A'$, $T = B'$, and $l = k$. Since, by Claim \ref{claim:updateEC}, every vertex in $A'$ has at least $(1-3\sqrt{\lambda})|B|$ blue neighbors in $B'$, the size of the common blue neighborhood in $B'$ of any two vertices in $A'$ is at least \[2(1-3\sqrt{\lambda})|B| - |B'| \geq (1-6\sqrt{\lambda})|B'|.\] Thus, again in reference to Claim \ref{claim:redgeneral}, we may take $s = (1-6\sqrt{\lambda})|B'|$. It remains to verify that the conditions of Claim \ref{claim:redgeneral} hold, that is, that $k \leq \min(2(1-6\sqrt{\lambda})|B'|+1, 2|A'|-1)$. But this is simple, since \begin{equation} (1-6\sqrt{\lambda})|B'| - (k-1)/2 \geq (1-6\sqrt{\lambda})(1/2 - 2\sqrt{\lambda})n - n/4 > (1/4 - 5\sqrt{\lambda})n \label{eq:eq1} \end{equation} and \begin{equation} |A'| - (k+1)/2 \geq (1/2 - 2\sqrt{\lambda})n - (k+1)/2 \geq (1/4 - 5\sqrt{\lambda})n.\label{eq:eq2} \end{equation} Therefore, by Claim \ref{claim:redgeneral} and the estimates \eqref{eq:eq1} and \eqref{eq:eq2}, the number of cycles of length $k$ is at least \[ ((1/4 - 5\sqrt{\lambda})n)^{(k-1)/2} \cdot ((1/4 - 5\sqrt{\lambda})n)^{(k-3)/2} = ((1/4 - 5\sqrt{\lambda})n)^{k-2} \geq (n/5)^{k-2}, \] as required. \end{proof} Our next claim is as follows. \begin{claim}\label{claim:ABedge} Suppose that $A'$ and $B'$ are both red cliques. If there are two vertex-disjoint red paths $P_1$ and $P_2$ such that each has length at most $2$ and each has one end vertex in $A'$ and the other in $B'$, then there are at least $(n/5e)^{k-6}$ red cycles of length $k$. \end{claim} \begin{proof} We will apply Claim \ref{claim:ABedgegeneral} to the red subgraph with $S = A'$, $T = B'$, and $l = k$. To check that the condition $7 \leq k \leq \min(2|A'| -1, 2|B'|-1)$ holds, note that $$2|A'|-1 \geq (1 - 4\sqrt{\lambda})n-1 > (n+1)/2 = k$$ for $n$ sufficiently large and, similarly, $k \leq 2|B'|-1$. Therefore, we may apply Claim~\ref{claim:ABedgegeneral} to conclude that the number of red cycles of length $k$ is at least $$(((k-1)/2-3)/e)^{k-6}=\left( ((n-1)/4 - 3)/e\right)^{k-6} \geq (n/5e)^{k-6}$$ for $n$ sufficiently large. \end{proof} Therefore, we are done if the assumptions of Claim \ref{claim:ABedge} are satisfied, so we can and will assume that if $A'$ and $B'$ are both red cliques, then there are no two vertex-disjoint red edges between $A'$ and $B'$. By applying Claim \ref{claim:tworedABgeneral} to the red subgraph with $S = A'$ and $T = B'$, we see that we may remove at most one vertex from $A' \cup B'$ to make all the edges between $A'$ and $B'$ blue. Without loss of generality, we may therefore assume that there is a vertex $v$ in $A'$ such that all the edges between $A' \setminus \{v\}$ and $B'$ are blue. In what follows, we let $A'' = A' \setminus \{v\}$. \begin{claim}\label{claim:PbothAB} Suppose that all the edges between $A''$ and $B'$ are blue. If there is a blue path $P'$ of length two with one end vertex in $A''$ and the other in $B'$, then there are at least $(n/8e)^{k-5}$ blue cycles of length $k$. \end{claim} \begin{proof} We will apply Claim \ref{claim:PbothABgeneral} to the blue subgraph with $S = A''$, $T = B'$, and $l = k$, using the fact that the bipartite graph between $A''$ and $B'$ is complete in blue. To check that the condition $7 \leq k \leq \min(2|A''|+1, 2|B'|+1)$, note, for instance, that $$2|A''| + 1 \geq 2|A'| - 1 \geq (1 - 4 \sqrt{\lambda})n -1 \geq (n+1)/2 = k$$ for $n$ sufficiently large. Therefore, by Claim~\ref{claim:PbothABgeneral}, the number of blue cycles of length $k$ is at least $((k-5)/2e)^{k-5} \geq (n/8e)^{k-5}$, as required. \end{proof} Since we are done if the assumptions of Claim \ref{claim:PbothAB} hold, we can now assume that there is no blue path $P'$ of length two with one end vertex in $A''$ and the other in $B'$. This means that any vertex in $\{v\} \cup X \cup Y$ is either completely red to $A''$ or completely red to $B'$. Therefore, there is a vertex partition of $\{v\} \cup X \cup Y$ into two sets $Z_1 \cup Z_2$ such that each vertex in $Z_1$ is completely red to $A''$ and each vertex in $Z_2$ is completely red to $B'$. By another application of Claim \ref{claim:ABedge}, we can also assume that there are no two vertex-disjoint red paths of length at most two each with one end vertex in $A''$ and the other in $B'$. Therefore, if $Z_1$ and $Z_2$ are both non-empty, either $Z_1$ is completely blue to $B'$ or $Z_2$ is completely blue to $A''$. Without loss of generality, suppose that $Z_1$ is completely blue to $B'$. If now $|Z_2| \geq 1$, either there is at most one vertex in $Z_2$ with red neighbors in $A''$ or there is a vertex $a\in A''$ such that this is the only red neighbor of vertices in $Z_2$. Therefore, by removing at most one vertex from $V(G)$, it will also be completely blue between $Z_2$ and $A''$. In summary, we have disjoint sets $Z_1' \subset Z_1$, $Z_2' \subset Z_2$, $A''' \subset A''$, and $B'' \subset B'$ (at most one of which differs from its superset) such that $|Z_1' \cup Z_2'\cup A''' \cup B''| \geq n-1$ and the following conditions hold: \begin{enumerate} \item $Z_1'$ is completely red to $A'''$ and completely blue to $B''$. \item $Z_2'$ is completely blue to $A'''$ and completely red to $B''$. \item It is completely blue between $A'''$ and $B''$. \item $A'''$ and $B''$ are both red cliques. \end{enumerate} Let $\tilde A = A''' \cup Z_1'$ and $\tilde B = B'' \cup Z_2'$. Then it is completely blue between $\tilde A$ and $B''$ and between $\tilde B$ and $A'''$. Furthermore, \[ |\tilde A| \geq |A'''| \geq |A'| - 2 \geq (1/2 -2 \sqrt{\lambda})n - 2 > (1/2 - 3 \sqrt{\lambda})n \] and, similarly, $|\tilde B| \geq (1/2 - 3\sqrt{\lambda})n$. Finally, \begin{equation} |\tilde A| + |\tilde B| \geq n-1 = 2k-2. \label{eqn:almost} \end{equation} By following the proofs of Claims~\ref{claim:red} and~\ref{claim:ABedge}, we obtain the following two results. \begin{claim}\label{claim:tildeABblue} If there is a blue edge within either $\tilde A$ or $\tilde B$, then there are at least $(n/5)^{k-2}$ blue cycles of length $k$. \end{claim} \begin{claim} \label{claim:tildeABred} Suppose that $\tilde A$ and $\tilde B$ are both red cliques, each with $k-1$ vertices. If there are two vertex-disjoint red edges between $\tilde A$ and $\tilde B$, then there are at least $(n/5e)^{k-6}$ red cycles of length $k$. \end{claim} By Claim \ref{claim:tildeABblue}, we can assume that there is no blue edge within $\tilde A$ or $\tilde B$. That is, $\tilde A$ and $\tilde B$ are both red cliques. If either of these cliques has order at least $k$ we are done, as we then get at least $(k-1)!/2 \geq ((k-1)/2e)^{k-1}$ red cycles of length $k$. Hence, by (\ref{eqn:almost}), we can assume that $|\tilde A| = |\tilde B| = k-1$. By Claim \ref{claim:tildeABred}, we can assume that there are no two vertex-disjoint red edges between $\tilde A$ and $\tilde B$. Therefore, applying Claim \ref{claim:tworedABgeneral} to the red subgraph with $S = \tilde A$ and $T = \tilde B$, we see that by removing at most one vertex, say $w$, all the edges between $\tilde A$ and $\tilde B$ are blue. Without loss of generality, we will assume that $w \in \tilde A$, noting that $w$ must have at least one blue neighbor in $\tilde B$, since otherwise $\tilde B \cup \{w\}$ would be a red clique of order $k$, again completing the proof. We require one final observation, proved in the same manner as Claim~\ref{claim:PbothAB}. \begin{claim} \label{claim:tildePbothAB} Suppose that $w \in \tilde A$ and all the edges between $\tilde A \setminus \{w\}$ and $\tilde B$ are blue, while at least one edge between $w$ and $\tilde B$ is blue. If there is a blue path $P'$ of length two with one end vertex in $\tilde A$ and the other in $\tilde B$, then there are at least $(n/8e)^{k-5}$ blue cycles of length $k$. \end{claim} Suppose that $u$ is the single vertex of $K_n$ which is not in $\tilde A \cup \tilde B$. If $u$ has a blue neighbor in both $\tilde A$ and $\tilde B$, then Claim \ref{claim:tildePbothAB} implies that we are done. Therefore, we can assume that $u$ is completely red to either $\tilde A$ or $\tilde B$. If $u$ is completely red to $\tilde A$, then $\tilde A \cup \{ u\}$ is a red clique with $k$ vertices, in which case there are at least $(k-1)!/2$ red cycles of length $k$. Since this is also true if $u$ is completely red to $\tilde B$, this completes the proof. \end{proof} \subsection{Proof of Lemma \ref{lem:main2}} \label{sec:stab} The following stability lemma of Nikiforov and Schelp \cite{NS} is an essential ingredient in our proof. \begin{lemma}[Theorem 13, \cite{NS}]\label{lem:NS} Let $0 < \alpha < 5 \cdot 10^{-6}$, $0 \leq \beta \leq \alpha/25$, and $n \geq \alpha^{-1}$. If $G$ is a graph with $n$ vertices and $e(G) > (1/4 - \beta) n^2$, then one of the following holds: \begin{enumerate} \item There are cycles $C_t \subset G$ for every $t \in [3, \lceil(1/2 + \alpha) n\rceil]$. \item There exists a partition $V(G) = U_0 \cup U_1 \cup U_2$ such that \[|U_0| < 2000\alpha n,\] \[ \left( 1/2 - 10\sqrt{\alpha+\beta}\right) n < |U_1| \leq |U_2| < \left( 1/2 + 10\sqrt{\alpha+\beta}\right) n,\] and the induced subgraph $G- U_0$ on vertex set $V(G) \setminus U_0$ is a subgraph of either the complete bipartite graph between $U_1$ and $U_2$ or its complement. \end{enumerate} \end{lemma} With this preliminary in place, we can begin the proof of Lemma \ref{lem:main2}. We apply the colored regularity lemma, Lemma~\ref{reglem-twocolor}, with parameters $\epsilon$ and $l = \lceil \epsilon^{-1} \rceil$ to the given red/blue coloring of $K_n$. This implies that there exist $n_0(\epsilon)$ and $M_0(\epsilon)$ such that, for any $n \geq n_0$, there is a regular partition of $K_n$ into $M$ parts $V_1, \dots, V_M$ with $\epsilon^{-1} \leq M \leq M_0$. We now consider a reduced graph $H$ with $M$ vertices $v_1, \dots, v_M$ corresponding to $V_1, \dots, V_M$, placing an edge between $v_i$ and $v_j$ if and only if the pair $(V_i, V_j)$ is $\epsilon$-regular. We then color the edge $(v_i, v_j)$ red if the density of red edges between $V_i$ and $V_j$ is at least $d = 12 \epsilon^{1/2}$ and we color an edge blue under the analogous condition with blue in place of red. By the regularity lemma, all but at most $\epsilon \binom{M}{2}$ pairs of distinct vertices of $H$ are edges and, since $d < 1/2$, all edges of $H$ are colored red, blue, or, perhaps, both red and blue. We say an edge is red-only if it is colored in red and not blue, while blue-only is defined similarly. Let the subgraph of $H$ induced by edges containing the color red be $H_R$ and the subgraph induced by edges containing the color blue be $H_B$. Hence, \[ |E(H_R) | + |E(H_B) | \geq (1-\epsilon)\binom{M}{2} > (1 - 2\epsilon) M^2/2,\] where we used that $M \geq \epsilon^{-1}$. Thus, without loss of generality, we can assume that \[ |E(H_R) | > (1-2\epsilon)M^2/4.\] We now apply Lemma~\ref{lem:NS} to $H_R$ with $\beta = \epsilon/2$ and $\alpha = 20 \sqrt{\epsilon}$. There are two cases: \paragraph{Case 1 of Lemma~\ref{lem:NS}.} In this case, we can find a red cycle $C_t$ for every $t \in [3, \lceil(1/2 + \alpha) M\rceil]$. In particular, we can find an odd cycle $C_t$ with \[ (1/2 + \alpha) M \geq t > (1/2 + \alpha) M - 2.\] But this means that there are disjoint vertex sets $V_{k_0}, \dots, V_{k_{t-1}}$ such that, for each $0 \leq i \leq t-1$, $|V_{k_i}| \geq \lfloor n/M \rfloor$ and each pair $(V_{k_i}, V_{k_{i+1}})$ (with addition taken mod $t$) is $\epsilon$-regular in red with red density at least $12 \epsilon^{1/2}$. Thus, we are in Case 1 of Lemma \ref{lem:main2}. \paragraph{Case 2 of Lemma~\ref{lem:NS}.} In this case, there exists a partition $V(H_R) = U_0 \cup U_1 \cup U_2$ such that $|U_0| < 2000\alpha M$ and \begin{equation} \left( 1/2 - 10\sqrt{2\alpha}\right) M < |U_1| \leq |U_2| < \left( 1/2 + 10\sqrt{2\alpha}\right) M. \label{eqn:U1U2} \end{equation} Furthermore, the induced subgraph $H_R-U_0$ is a subgraph of the disjoint cliques on $U_1$ and $U_2$ or a subgraph of the complete bipartite graph between $U_1$ and $U_2$. We will assume that the induced subgraph $H_R-U_0$ is a subgraph of the graph consisting of disjoint cliques on $U_1$ and $U_2$. The other case, where all edges of $H_R - U_0$ are between $U_1$ and $U_2$, can be handled similarly. Thus, by assumption, any edges between $U_1$ and $U_2$ are blue-only. Moreover, since the number of non-adjacent pairs is at most $\epsilon \binom{M}{2}$, the number of blue-only edges between $U_1$ and $U_2$ is at least \begin{equation} |U_1||U_2| - \epsilon \binom{M}{2}. \label{eqn:U1U2blue} \end{equation} Let $U_1' \subset U_1$ be the set of vertices in $U_1$ that have blue degree at least $(1-\sqrt{\epsilon}) |U_2|$ in $U_2$. Suppose $|U_1 \setminus U_1'| = x |U_1|$. Then \[ (1-\sqrt{\epsilon} )|U_2| x |U_1| + |U_2|(1-x)|U_1| \geq |U_1||U_2| - \epsilon M^2/2, \] which implies that $x \leq \sqrt{\epsilon}M^2 / (2|U_1||U_2|)$. Since $|U_1|, |U_2| \geq (1/2 - 10\sqrt{2\alpha}) M$, we have \[ x \leq \sqrt{\epsilon}M^2 / (2|U_1||U_2|) \leq \sqrt{\epsilon}M^2 / (2 (1/2 - 10\sqrt{2\alpha})^2 M^2 )< \sqrt{\epsilon}(2 + 200\sqrt{\alpha}) < 3 \sqrt{\epsilon}, \] where we used that $\alpha < 5\cdot 10^{-6}$. Defining $U_2' \subset U_2$ analogously, we therefore have \begin{equation} |U_1 \setminus U_1'| \leq 3\sqrt{\epsilon}|U_1|, \ \ |U_2 \setminus U_2'| \leq 3\sqrt{\epsilon}|U_2|. \label{eq:U1U1'} \end{equation} Thus, each vertex in $U_1'$ has at least $(1-\sqrt{\epsilon})|U_2| - |U_2 \setminus U_2'| \geq (1-4\sqrt{\epsilon})|U_2|$ blue neighbors in $U_2'$ and, similarly, each vertex in $U_2'$ has at least $(1-4\sqrt{\epsilon})|U_1|$ blue neighbors in $U_1'$. \begin{claim}\label{claim:U1U2red'} If there is a blue edge within $U_1'$ or $U_2'$, then Case 1 of Lemma \ref{lem:main2} holds. \end{claim} \begin{proof} It will suffice to show that there is a blue cycle $C_t$ in $H$, where $t$ is the odd integer with $(1/2 + \alpha)M \geq t > (1/2+\alpha)M-2$. Suppose that there is a blue edge $(u,u')$ in $U_1'$. We will apply Claim \ref{claim:redgeneral} to $H_B$ with $(S, T, l)$ being $(U_1', U_2', t)$. Since each vertex in $U_1'$ has at least $ (1-4\sqrt{\epsilon})|U_2|$ blue neighbors in $U_2'$, any two vertices in $U_1'$ have blue common neighborhood in $U_2'$ of order at least \[ 2(1-4\sqrt{\epsilon})|U_2| - |U_2'| \geq (1-8\sqrt{\epsilon})|U_2'|. \] Thus, we can let $s$ in Claim \ref{claim:redgeneral} be $(1-8\sqrt{\epsilon})|U_2'|$. To check that Claim \ref{claim:redgeneral} applies, we need to show that $(1/2 + \alpha)M \leq \min(2|U_1'| -1, 2s+1)$. First, by (\ref{eqn:U1U2}), (\ref{eq:U1U1'}), and the fact that $M \geq \epsilon^{-1}$, \[ 2|U_1'|-1 \geq 2 (1-3\sqrt{\epsilon})|U_1| -1\geq 2(1-3\sqrt{\epsilon}) (1/2 - 10\sqrt{2\alpha})M -1 > 0.6M > (1/2+\alpha)M. \] Similarly, \[ 2s+1 \geq 2(1-8\sqrt{\epsilon}) (1-3\sqrt{\epsilon})|U_2| \geq 2(1-11\sqrt{\epsilon}) (1/2 - 10\sqrt{2\alpha})M > 0.6M > (1/2+\alpha)M. \] Thus, by Claim \ref{claim:redgeneral}, there is a cycle of length $t$ in $H_B$, as required. \end{proof} We may therefore assume that there is no blue edge inside $U_1'$ or $U_2'$. That is, all the edges within $U_1'$ and $U_2'$ are red-only. We move the vertices in $U_0$ arbitrarily to $U_1$ and $U_2$ to obtain $\tilde U_1$ and $\tilde U_2$. Thus, $\tilde U_1 \cup \tilde U_2$ is a vertex partition of $V(H)$. Let $X_1 \subset V(K_n)$ be the vertices in $K_n$ corresponding to $\tilde U_1$ in $H$ and let $X_2 \subset V(K_n)$ be the vertices corresponding to $\tilde U_2$. We will conclude the proof by showing that the partition $X_1 \cup X_2$ induces an extremal coloring. \begin{claim} The vertex partition $X_1 \cup X_2$ induces an extremal coloring with parameter $\lambda$, where $\lambda = 300 \sqrt{\alpha}$. \end{claim} \begin{proof} By Claim \ref{claim:U1U2red'}, any edge in $U_1'$ is red-only and at most $\epsilon \binom{M}{2}$ pairs of distinct vertices in $U_1'$ are non-adjacent. Moreover, for any red-only edge $(i,j)$ in $U_1'$, the red density between $V_i$ and $V_j$ is at least $1-d$, since otherwise $(i,j)$ would also be colored blue. Since $n/M - 1 < |V_i| \leq n/M+1$, the number of red edges in $X_1$ is at least \[(1-d) (n/M-1)^2 \binom{|U_1'|}{2} - (n/M+1)^2 \epsilon \binom{M}{2}.\] Note also, by (\ref{eqn:U1U2}), that \begin{align*} |X_1| & \leq |\tilde U_1| \cdot (n/M+1) \leq (|U_1|+|U_0|)(n/M+1) \leq (|U_1| + 2000\alpha M) (n/M+1) \nonumber \\ & \leq \left( 1/2 + 10\sqrt{2\alpha} + 2000 \alpha \right) M (n/M+1) \leq (1/2 + 20\sqrt{\alpha}) n. \end{align*} Combining the two inequalities above with (\ref{eqn:U1U2}) and (\ref{eq:U1U1'}), we see that the red density in $X_1$ is at least \begin{align*} \frac{(1-d) \binom{|U_1'|}{2} (n/M-1)^2 - (n/M+1)^2\epsilon M^2/2}{ |X_1|^2/2} & \geq \frac{2(1-d) \binom{(1-3\sqrt{\epsilon})|U_1|}{2} (n/M-1)^2 - (n/M+1)^2\epsilon M^2}{ (1/2 + 20\sqrt{\alpha})^2 n^2 } \\ & \geq 1 - d - 200 \sqrt{\alpha} - 10\sqrt{\epsilon} > 1 - 300 \sqrt{\alpha}. \end{align*} Similarly, $|X_2| \leq (1/2 + 20\sqrt{\alpha}) n$ and the red density in $X_2 \subset V(G)$ is at least $1 - 300 \sqrt{\alpha}$. It only remains to lower bound the blue density between $X_1$ and $X_2$. By (\ref{eqn:U1U2}) and (\ref{eqn:U1U2blue}), the number of blue edges between $X_1$ and $X_2$ is at least \begin{align*} (1-d)\left(|U_1||U_2| - \epsilon \binom{M}{2}\right) (n/M-1)^2 & \geq (1-d)\left((1/2 - 20\sqrt{\alpha})^2 M^2 - \epsilon \binom{M}{2}\right) (n/M - 1)^2 \\ & >(1-d) (1/4 - 25 \sqrt{\alpha})n^2. \end{align*} Thus, by a similar computation to before, the blue density between $X_1$ and $X_2$ is at least \[\frac{(1-d)(1/4 - 25 \sqrt{\alpha})n^2}{|X_1||X_2|} \geq \frac{(1-d)(1/4 - 25 \sqrt{\alpha})n^2}{n^2/4} > 1 - 300\sqrt{\alpha},\] as required. \end{proof}
{ "timestamp": "2021-09-21T02:34:58", "yymm": "2108", "arxiv_id": "2108.00987", "language": "en", "url": "https://arxiv.org/abs/2108.00987", "abstract": "The Ramsey number $r(H)$ of a graph $H$ is the minimum $n$ such that any two-coloring of the edges of the complete graph $K_n$ contains a monochromatic copy of $H$. The threshold Ramsey multiplicity $m(H)$ is then the minimum number of monochromatic copies of $H$ taken over all two-edge-colorings of $K_{r(H)}$. The study of this concept was first proposed by Harary and Prins almost fifty years ago. In a companion paper, the authors have shown that there is a positive constant $c$ such that the threshold Ramsey multiplicity for a path or even cycle with $k$ vertices is at least $(ck)^k$, which is tight up to the value of $c$. Here, using different methods, we show that the same result also holds for odd cycles with $k$ vertices.", "subjects": "Combinatorics (math.CO)", "title": "Threshold Ramsey multiplicity for odd cycles", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9833429614552197, "lm_q2_score": 0.8152324983301568, "lm_q1q2_score": 0.8016531391825138 }
https://arxiv.org/abs/1902.01687
Optimal Nonparametric Inference via Deep Neural Network
Deep neural network is a state-of-art method in modern science and technology. Much statistical literature have been devoted to understanding its performance in nonparametric estimation, whereas the results are suboptimal due to a redundant logarithmic sacrifice. In this paper, we show that such log-factors are not necessary. We derive upper bounds for the $L^2$ minimax risk in nonparametric estimation. Sufficient conditions on network architectures are provided such that the upper bounds become optimal (without log-sacrifice). Our proof relies on an explicitly constructed network estimator based on tensor product B-splines. We also derive asymptotic distributions for the constructed network and a relating hypothesis testing procedure. The testing procedure is further proven as minimax optimal under suitable network architectures.
\section{Introduction} With the remarkable development of modern technology, difficult learning problems can nowadays be tackled smartly via deep learning architectures. For instance, deep neural networks have led to impressive performance in fields such as computer vision, natural language processing, image/speech/audio recognition, social network filtering, machine translation, bioinformatics, drug design, medical image analysis, where they have demonstrated superior performance to human experts. The success of deep networks hinges on their rich expressiveness (see \cite{db11} , \cite{rpkgdj17}, \cite{mpcb14}, \cite{bs14}, \cite{t16}, \cite{l17} and \cite{y17, y18}). Recently, deep networks have played an increasingly important role in statistics particularly in nonparametric curve fitting (see \cite{kk05, hk06, kk17, km11, s17}). Applications of deep networks in other fields such as image processing or pattern recgnition include, to name a few, \cite{lbh15}, \cite{dlhyysszhw13}, \cite{wwhwzzl14}, \cite{gg16}, etc. A fundamental problem in statistical applications of deep networks is how accurate they can estimate a nonparametric regression function. To describe the problem, let us consider i.i.d. observations $(Y_i, \mathbf{X}_i)$, $i=1,2,\ldots, n$ generated from the following nonparametric model: \begin{eqnarray} Y_i=f_0(\mathbf{X}_i)+\epsilon_i, \label{eq:model} \end{eqnarray} where $\mathbf{X}_i \in [0,1]^d$ are i.i.d. $d$-dimensional predictors for a fixed $d\ge1$, $\epsilon_i$ are i.i.d. random noise with $E(\epsilon_i)=0$ and $Var(\epsilon_i)=\tau^2$, $f_0\in\mathcal{H}$ is an unknown function. For any $L\in\mathbb{N}$ and $\mathbf{p}=(p_1,\ldots,p_L)\in\mathbb{N}^L$, let $\mathcal{F}(L,\mathbf{p})$ denote the collection of network functions from $\mathbb{R}^d$ to $\mathbb{R}$ consisting of $L$ hidden layers with the $l$th layer including $p_l$ neurons. The problem of interest is to find an order $R_n$ that controls the $L^2$ minimax risk: \begin{equation}\label{eqn:goal} \inf_{\widehat{f}\in\mathcal{F}(L,\mathbf{p})}\sup_{f_0\in \mathcal{H}}\mathbb{E}_{f_0}\bigg(\|\widehat{f}-f_0\|_{L^2}^2\bigg|\mathbb{X}\bigg)=O_P(R_n), \end{equation} where $\mathbb{X}=\{\mathbf{X}_1,\ldots,\mathbf{X}_n\}$ and the infimum is taken over all estimators $\widehat{f}\in\mathcal{F}(L,\mathbf{p})$. In other words, we are interested in the performance of the ``best'' network estimator in the ``worst'' scenario. Existing results regarding (\ref{eqn:goal}) are sub-optimal. For instance, when $\mathcal{H}$ is a $\beta$-smooth H\"{o}lder class and $L,\mathbf{p}$ are properly selected, it has been argued that $R_n=n^{-\frac{2\beta}{2\beta+d}}(\log{n})^s$ for some constant $s>0$; see \cite{kk05, hk06, kk17, km11, s17, su18, flm18}. Such results are mostly proved based on empirical processes techniques in which the logarithmic factors arise from the entropy bound of the neural network class. The aim of this paper is to fully remove the redundant logarithmic factors, i.e., under proper selections of $L,\mathbf{p}$ one actually has $R_n=n^{-\frac{2\beta}{2\beta+d}}$ in (\ref{eqn:goal}). This means that neural network estimators can exactly achieve minimax estimation rate. Our proof relies on an explicitly constructed neural network which is proven minimax optimal. Some interesting byproducts are worth mentioning. First, the rate $R_n$ can be further improved when $f_0$ satisfies additional structures. Specifically, we will show that $R_n=n^{-\frac{2\beta}{2\beta+1}}$ if $f_0$ satisfies additive structure, i.e., $f_0$ is a sum of univariate $\beta$-H\"{o}lder functions. Such rate is minimax according to \cite{s85}. Second, we will derive the pointwise asymptotic distribution of the constructed neural network estimator which will be useful to establish pointwise confidence interval. Third, the constructed neural network estimator will be further used as a test statistic which is proven optimal when $L,\mathbf{p}$ are properly selected. As far as we know, these are the first provably valid confidence interval and test statistic based on neural networks in nonparametric regression. This paper is organized as follows. Section \ref{sec:prelim} includes some preliminaries on deep networks and function spaces. In Section \ref{sec:optimal:rate:convergence}, we derive upper bounds for the minimax risk and investigate their optimality. Both multivariate regression and additive regression are considered. Section \ref{sec:basis:approximation} contains the main proof strategy, which covers the construction of (optimal) network and relates results on network approximation of tensor product B-splines. As by products, we also provide limit distributions and optimal testing results in Section \ref{sec:byproducts}. The proofs of some of the main results and technical lemmas are deferred to Appendix A-D. \section{Preliminaries}\label{sec:prelim} In this section, we review some notion about deep networks and function spaces. Throughout let $\sigma$ denote the rectifier linear unit (ReLU) activation function, i.e., $\sigma(x)=(x)_+$ for $x\in\mathbb{R}$. For any real vectors $\mathbf{v}=(v_1,\ldots,v_r)^T$ and $\mathbf{y}=(y_1,\ldots,y_r)^T$, define the shift activation function $\sigma_\mathbf{v}(\mathbf{y})=(\sigma(y_1-v_1),\ldots,\sigma(y_r-v_r))^T$. Let $\mathbf{p}=(p_1,\ldots,p_L)\in\mathbb{N}^L$. Any $f\in\mathcal{F}(L,\mathbf{p})$ has an expression \[ f(\mathbf{x})=W_{L+1}\sigma_{\mathbf{v}_{L}}W_{L}\sigma_{\mathbf{v}_{L-1}}\ldots W_2\sigma_{\mathbf{v}_1}W_1 \mathbf{x},\,\,\mathbf{x} \in \mathbb{R}^{d}, \] where $\mathbf{v}_{l}\in \mathbb{R}^{p_l}$ is a shift vector and $W_{l}\in \mathbb{R}^{p_{l}\times p_{l-1}}$ is a weight matrix. Here we have adopted the convention $p_0=d$ and $p_{L+1}=1$. For simplicity, we only consider fully connected networks and do not make any sparsity assumptions on the entries of $\mathbf{v}_l$ and $W_l$. Next let us introduce various function spaces under which the estimation rates will be derived. We will consider two types of function spaces: H\"{o}lder space and additive space. Let $\Omega=[0,1]^d$ denote the domain of the functions. For $f$ defined on $\Omega$, define the supnorm and $L^2$-norm of $f$ by $\|f\|_{\sup}=\sup_{\mathbf{x}\in\Omega}|f(\mathbf{x})|$ and $\|f\|_{L^2}^2=\int_\Omega f(\mathbf{x})^2Q(\mathbf{x})d\mathbf{x}$ respectively. Here $Q(\cdot)$ is the probability density for the predictor $\mathbf{X}_i$'s. For any $\bm{\alpha}=(\alpha_1, \alpha_2, \ldots, \alpha_d)\in \mathbb{N}^d$, define $|\bm{\alpha}|=\sum_{j=1}^d\alpha_j$ and \begin{eqnarray*} \partial^{\bm{\alpha}}f=\frac{\partial^{|\bm{\alpha}|}f}{\partial x_1^{\alpha_1}\ldots \partial x_d^{\alpha_d}}, \end{eqnarray*} whenever the partial derivative exists. For any $\beta>1$ and $F>0$, let $\Lambda^\beta(F, \Omega)$ denote the ball of $\beta$-H\"{o}lder functions with radius $F$, i.e., \begin{eqnarray} \Lambda^\beta(F, \Omega)=\bigg\{f: \Omega\to \mathbb{R} \bigg|\;\sum_{\bm{\alpha}:|\bm{\alpha}|\leq \floor{\beta}}\|\partial^{\bm{\alpha}}f\|_{\sup}+\sum_{\bm{\alpha}:|\bm{\alpha}|=\floor{\beta}}\sup_{\mathbf{x}_1\neq \mathbf{x}_2\in \Omega}\frac{|\partial^{\bm{\alpha}}f(\mathbf{x}_1)-\partial^{\bm{\alpha}}f(\mathbf{x}_2)|}{\|\mathbf{x}_1-\mathbf{x}_2\|_2^{\beta-\floor{\beta}}}\leq F\bigg\},\nonumber \end{eqnarray} in which $\floor{\beta}$ is the largest integer smaller than $\beta$ and $\|v\|_2$ denotes the Euclidean norm of a real vector $v$. For any $F>0$ and $\boldsymbol\beta=(\beta_1, \ldots, \beta_d)\in (1,\infty)^d$, define \begin{eqnarray} \Lambda^{\boldsymbol\beta}_+(F, \Omega)=\left\{f: [0,1]^d \to \mathbb{R}|\; f(\mathbf{x})=a+\sum_{j=1}^dg_{j}(x_j) \textrm{ with } g_j \in \Lambda^{\beta_j}(F, [0,1]), \textrm{ for } j=1,\ldots, d\right\}. \nonumber \end{eqnarray} Clearly, any $f\in\Lambda^{\boldsymbol\beta}_+(F, \Omega)$ has an expression $f(\mathbf{x})=a+\sum_{j=1}^dg_{j}(x_j)$ with the $j$th additive component belonging to the ball of univariate $\beta_j$-H\"{o}lder functions with radius $F$. \section{Minimax Neural Network Estimation}\label{sec:optimal:rate:convergence} In this section, we derive an upper bound for the $L^2$ minimax risk in the problem (\ref{eqn:goal}). The risk bound will be proven optimal under suitable circumstances. To simplify the expressions, we only consider networks with architecture $(L,\mathbf{p}(T))$, where $\mathbf{p}(T):=(T,\ldots,T)\in\mathbb{N}^L$ for any $T\in\mathbb{N}$. In other words, we focus on networks whose $L$ layers each have $T$ neurons. Our results hold under suitable conditions on $L$ and $T$ as well as the following assumption on the design and model error. \begin{Assumption}\label{A0} The probability density $Q(\mathbf{x})$ of $\mathbf{X}$ is supported on $\Omega$. There exists a constant $c>0$ such that $c^{-1}\le Q(\mathbf{x})\le c$ for any $\mathbf{x}\in\Omega$. The error terms $\epsilon_i$'s are independent of $\mathbf{X}_i$'s. \end{Assumption} \begin{theorem}\label{thm:neural:estimator:rate:of:convergence:main:text} Let Assumption \ref{A0} be satisfied and $F>0$ be a fixed constant. Suppose that $T\to\infty$ and $T\log T=o(n)$ as $n\to\infty$, then it follows that \begin{eqnarray}\label{thm1:risk:bound} \inf_{\widehat{f}\in \mathcal{F}(L,\mathbf{p}(T))}\sup_{f_0\in \Lambda^\beta(F, \Omega)} \mathbb{E}_{f_0}\bigg(\|\widehat{f}(\mathbf{x})-f_0(\mathbf{x})\|_{L^2}^2\bigg|\mathbb{X}\bigg)=O_P\bigg(\frac{1}{T^{\frac{2\beta}{d}}}+\frac{T}{n}+\frac{T^2}{2^{\frac{L}{d+k}}}\bigg). \end{eqnarray} As a consequence, if $T\asymp n^{\frac{d}{2\beta+d}}$ and $n^{\frac{2\beta+2d}{2\beta+d}}=o(2^{\frac{L}{d+k}})$, then the following holds: \begin{eqnarray*} \inf_{\widehat{f}\in \mathcal{F}(L,\mathbf{p}(T))}\sup_{f_0\in \Lambda^\beta(F, \Omega)} \mathbb{E}_{f_0}\bigg(\|\widehat{f}(\mathbf{x})-f_0(\mathbf{x})\|_{L^2}^2\bigg|\mathbb{X}\bigg)=O_P(n^{-\frac{2\beta}{2\beta+d}}). \end{eqnarray*} \end{theorem} Proof of Theorem \ref{thm:neural:estimator:rate:of:convergence:main:text} relies on an explicitly constructed network estimator based on tensor product B-splines; see Section \ref{sec:basis:approximation}. The minimax risk bound in (\ref{thm1:risk:bound}) consists of three components ${T^{-\frac{2\beta}{d}}}, n^{-1}T, 2^{-\frac{L}{d+k}}T^2$ corresponding to the bias, variance and approximation error of the constructed network. The optimal risk bound is achieved through balancing the three terms. The approximation error of the constructed network decreases exponentially along with $L$. Networks constructed based on other methods such as local Taylor approximations (\cite{y17}, \cite{y18} and \cite{s17} have similar approximation performance. However, their statistical properties are more challenging to deal with due to the unbalanced eigenvalues of the corresponding basis matrix. In contrast, the eigenvalues of the tensor product B-spline basis matrix are known to have balanced orders, e.g., see \cite{h98}, which plays an important role in deriving the risk bounds. Also notice that the risk bounds will blow out when $L$ is fixed, which partially explains the superior performance of deep networks compared with shallow ones; see \cite{es16}. The optimal rate in Theorem \ref{thm:neural:estimator:rate:of:convergence:main:text} suffers from the `curse' of dimensionality. The following theorem demonstrates that this issue can be addressed when $f_0$ has an additive structure. For $\boldsymbol\beta=(\beta_1, \ldots, \beta_d)\in (1,\infty)^d$, let $\beta^*=\min_{1\le j\le d}\beta_j$. \begin{theorem}\label{thm:rate:convergence:additive:main:text} Let Assumption \ref{A0} be satisfied and $F>0$ be a fixed constant. Suppose that $T\to\infty$ and $T\log T=o(n)$ as $n\to\infty$, then it follows that \begin{eqnarray*} \inf_{\widehat{f}\in\mathcal{F}(L,\mathbf{p}(T))}\sup_{f_0\in \Lambda^{\boldsymbol\beta}_+(F, \Omega)} \mathbb{E}_{f_0}\bigg(\|\widehat{f}(\mathbf{x})-f_0(\mathbf{x})\|_{L^2}^2\bigg|\mathbb{X}\bigg)=O_P\bigg(\frac{1}{T^{2\beta^*}}+\frac{T}{n}+\frac{T^2}{2^{\frac{L}{1+k}}}\bigg). \end{eqnarray*} As a consequence, if $T\asymp n^{\frac{1}{2\beta^*+1}}$ and $n^{\frac{2\beta^*+2}{2\beta^*+1}}=o(2^{\frac{L}{1+k}})$, then \begin{eqnarray*} \inf_{\widehat{f}\in\mathcal{F}(L,\mathbf{p}(T))}\sup_{f_0\in \Lambda^{\boldsymbol\beta}_+(F, \Omega)}\mathbb{E}_{f_0}\bigg(\|\widehat{f}(\mathbf{x})-f_0(\mathbf{x})\|_{L^2}^2\bigg|\mathbb{X}\bigg)=O_P\left(n^{-\frac{2\beta^*}{2\beta^*+1}}\right). \end{eqnarray*} \end{theorem} The rate $n^{-\frac{2\beta^*}{2\beta^*+1}}$ in Theorem \ref{thm:rate:convergence:additive:main:text} is optimal in nonparmetric additive estimation. When $\beta_1=\cdots=\beta_d=\beta$, the rate simply becomes $n^{-\frac{2\beta}{2\beta+1}}$ whose optimality has been proven by \cite{s85}. Otherwise, the optimal rate relies on the least order of smoothness of the $d$ univariate functions. The proof of Theorem \ref{thm:rate:convergence:additive:main:text} is deferred to Appendix C. \section{Construction of Optimal Networks}\label{sec:basis:approximation} In this section, we explicitly construct a network estimator $\widehat{f}_{\textrm{net}}\in\mathcal{F}(L,\mathbf{p}(T))$ and derive its risk bound. Theorems \ref{thm:neural:estimator:rate:of:convergence:main:text} and \ref{thm:rate:convergence:additive:main:text} will immediately follow due to the following trivial fact \begin{equation} \inf_{\widehat{f}\in\mathcal{F}(L,\mathbf{p}(T))}\sup_{f_0} \mathbb{E}_{f_0}\bigg(\|\widehat{f}(\mathbf{x})-f_0(\mathbf{x})\|_{L^2}^2\bigg|\mathbb{X}\bigg)\le \sup_{f_0} \mathbb{E}_{f_0}\bigg(\|\widehat{f}_{\textrm{net}}(\mathbf{x})-f_0(\mathbf{x})\|_{L^2}^2\bigg|\mathbb{X}\bigg).\label{eq:basic:inequality} \end{equation} The construction process starts from a pilot estimator $\widehat{f}_{\textrm{pilot}}$ obtained under tensor product B-splines. The tensor product B-spline basis functions are further approximated through explicitly constructed multi-layer networks, which will be aggregated to obtain the network estimator $\widehat{f}_{\textrm{net}}$. The key step is to show that the discrepancies between the tensor product B-spline basis functions and the corresponding network approximations are reasonably small such that $\widehat{f}_{\textrm{net}}$ will perform similarly as $\widehat{f}_{\textrm{pilot}}$, and thus, optimally. Our construction is different from \cite{y17} and \cite{s17}, where the basis functions are obtained through local Taylor approximation. We find that the eigenvalue performance of the local Taylor basis matrix is difficult to quantify so that the corresponding pilot estimator cannot be used effectively. Instead, the pilot estimator based on tensor product B-splines is more convenient to deal with. Other basis such as wavelets or smoothing splines may also work but this will be explored elsewhere. \subsection{A Pilot Estimator Through Tensor Product B-splines} In this subsection, we review tensor product $B$-splines and construct the corresponding pilot estimator. For any integer $M\geq 2$, let $0=t_0<t_1<\cdots<t_{M-1}<t_M=1$ be knots that form a partition of the unit interval. The definition of univariate B-splines of order $k\ge2$ depends on additional knots $t_{-k+1}<t_{-k+2}<\ldots<t_{-1}<0$ and $1<t_{M+1} <\ldots< t_{M+k-1}$. Given knots $t=(t_{-k+1},\ldots, t_{M+k-1})\in \mathbb{R}^{M+2k-1}$, the univariate $B$-spline basis functions of order $k$, denoted $B_{i,k}(x)$, $i=-k+1,-k+2,\ldots, M-1$, can be defined inductively by $B_{i,s}(x)$ for $s=2,3,\ldots, k$. For $s=2$ and $-k+1\le i\le M+k-3$, define \begin{eqnarray*} B_{i, 2}(x)=\begin{cases} \frac{x-t_i}{t_{i+1}-t_i}, & \textrm{if } x\in [t_i, t_{i+1}]\\ \frac{t_{i+2}-x}{t_{i+2}-t_{i+1}}, & \textrm{if } x\in [t_{i+1}, t_{i+2}]\\ 0,& \textrm{elsewhere} \end{cases}. \end{eqnarray*} Suppose that $B_{i,s}(x)$, $i=-k+1,\ldots, M+k-s-1$ have been defined. Define \begin{eqnarray}\label{induction:formula} B_{i, s+1}=a_{i,s}B_{i,s,t}+b_{i,s}B_{i+1,s, t},\,\,\textrm{for $i=-k+1,-k+2,\ldots, M+k-s-2$,} \end{eqnarray} where \begin{eqnarray*} {a}_{i,s}(x)=\begin{cases} 0, & \textrm{if } x<t_i\\ \frac{x-t_i}{t_{i+s}-t_i}, &\textrm{if } t_i\leq x \leq t_{i+s}\\ 0,&\textrm{if } x> t_{i+s}\\ \end{cases},\;\;\;{b}_{i,s}(x)=\begin{cases} 0, & \textrm{if } x<t_{t+1}\\ \frac{t_{i+s+1}-x}{t_{i+s+1}-t_{i+1}}, &\textrm{if } t_{i+1}\leq x \leq t_{i+s+1}\\ 0,&\textrm{if } x> t_{i+s+1}\\ \end{cases}. \end{eqnarray*} Proceeding with this construction, we can obtain $B_{i,k}(x)$. To approximate a multivariate function, we adopt the tensor product $B$-splines. Define $\Gamma=\{-k+1, -k+2, \ldots, 0, 1, \ldots, M-1\}^d$ and $q=|\Gamma|=(M+k-1)^d$. For $\mathbf{i}=(i_1, i_2, \ldots, i_d)\in \Gamma$, define $D_{\mathbf{i},k}(\mathbf{x})=\prod_{j=1}^dB_{i_j,k}(x_j)$ and obtain the corresponding pilot estimator \begin{eqnarray} \widehat{f}_{\textrm{pilot}}(\mathbf{x})=\sum_{\mathbf{i} \in \Gamma}\widehat{b}_{\mathbf{i}}D_{\mathbf{i}, k}(\mathbf{x}),\label{eq:pilot:estimator} \end{eqnarray} where $\widehat{b}_{\mathbf{i}}, \mathbf{i}\in \Gamma$ are the basis coefficients obtained by the following least square estimation: \begin{equation}\label{LSE:eqn} \widehat{C}:=[\widehat{b}_{\mathbf{i}}]_{\mathbf{i}\in\Gamma}=\argmin_{b_\mathbf{i}\in\mathbb{R}^q}\sum_{i=1}^n \left(Y_i-\sum_{\mathbf{i}\in\Gamma}b_\mathbf{i} D_{\mathbf{i},k}(\mathbf{X}_i)\right)^2. \end{equation} \subsection{Network Approximation of Tensor Product B-splines} In this subsection, we approximate $B_{i,k}$'s through multilayer neural networks. We first construct networks that approximate the univariate B-spline basis functions, and then multiply these networks through a product network $\xmark_s$ introduced by \cite{y17} to approximate the tensor product B-spline basis. Unlike \cite{y17} and \cite{s17}, our construction proceeds in an inductive manner due to the intrinsic induction structure of B-splines. For any $s\ge1$, the product network $\xmark_s(x_1, x_2, \ldots, x_s)$ is constructed to approximate the monomials $\prod_{j=1}^s x_j$. The following Proposition \ref{proposition:appriximation:product:k} which is due to \cite{y17} provides guarantees for $\xmark_s$. \begin{proposition}\label{proposition:appriximation:product:k} For any integers $m\ge1$ and $s\geq 2$, there exists a neural network function $\xmark_s$ with $(s-1)(2m+3)-1$ hidden layers and $10+s$ nodes in each hidden layer such that for all $x_1, x_2, \ldots, x_s\in [0, 1]$, $0\leq \xmark_s(x_1, x_2, \ldots, x_s)\leq 1$ and \begin{eqnarray*} \bigg|\xmark_s(x_1, x_2, \ldots, x_s)-\prod_{j=1}^s x_j\bigg|\leq (s-1)4^{-m+1}. \end{eqnarray*} As a consequence, if $|\widetilde{x}_j-x_j|\leq \delta$ and $\widetilde{x}_j \in [0,1]$ for $j=1,2,\ldots,s$, then \begin{eqnarray*} \bigg|\xmark_s(\widetilde{x}_1, \widetilde{x}_2, \ldots, \widetilde{x}_s)-\prod_{j=1}^s x_j\bigg|\leq (s-1)4^{-m+1}+s\delta. \end{eqnarray*} \end{proposition} In what follows, we will approximate the $k$th order univariate B-spline basis $B_{i,k}$. Fixing integer $m\geq 1$, our method is based on the induction formula (\ref{induction:formula}) which allows us to start from approximating $B_{i,2}$. Specifically, we approximate $B_{i,2}$ by $\widetilde{B}_{i, 2}$ defined as \begin{eqnarray*} \widetilde{B}_{i, 2}(x)=c_1\sigma(x-t_i)+c_2\sigma(x-t_{i+1})+c_3\sigma(x-t_{i+2}), \end{eqnarray*} where \begin{eqnarray} c_1=\frac{1}{t_{i+1}-t_i}, \;\;c_2=-\frac{t_{i+2}-t_{i}}{t_{i+2}-t_{i+1}}c_1, \;\;c_3=-(t_{i+2}-t_{i}+1)c_1-(t_{i+2}-t_{i+1}+1)c_2.\label{eq:definition:c1c2c3} \end{eqnarray} The piecewise linear function $\widetilde{B}_{i,2}$ is exactly a neural network with one hidden layer consisting of three nodes. Suppose that we have constructed $\widetilde{B}_{i,s}(x)$, a neural network approximation of $B_{i,s}$. Next we will approximate $B_{i,s+1}$. For $-k+1\leq i\leq M+k-s-1$, define \begin{eqnarray*} \widetilde{a}_{i,s}(x)=\begin{cases} 0, & \textrm{if } x<t_i\\ \frac{x-t_i}{t_{i+s}-t_i}, &\textrm{if } t_i\leq x \leq t_{i+s}\\ 1,&\textrm{if } x> t_{i+s}\\ \end{cases}, \;\;\;\widetilde{b}_{i,s}(x)=\begin{cases} 1, & \textrm{if } x<t_{i+1}\\ \frac{t_{i+s+1}-x}{t_{i+s+1}-t_{i+1}}, &\textrm{if } t_{i+1}\leq x \leq t_{i+s+1}\\ 0, &\textrm{if } x> t_{i+s+1}\\ \end{cases}. \end{eqnarray*} In terms of ReLU activation function, we can rewrite the above as $\widetilde{a}_{i,s}(x)=\frac{1}{t_{i+s}-t_i}\sigma(x-t_i)-\frac{1}{t_{i+s}-t_i}\sigma(x-t_{i+s})$ and $\widetilde{b}_{i,s}(x)=-\frac{1}{t_{i+s+1}-t_{i+1}}\sigma(x-t_{i+1})+\frac{1}{t_{i+s+1}-t_i}\sigma(x-t_{i+s+1})+1$, which implies that $\widetilde{a}_{i,s}$ and $\widetilde{b}_{i,s}$ are exactly neural networks with one hidden layer consisting of two nodes (see Figure \ref{figure:tildeBi2:to:tildeBi3}). For $i=-k+1,\ldots, M+k-s-2$, define \begin{eqnarray*} \widetilde{B}_{i, s+1}(x)=\frac{\xmark_2(\widetilde{a}_{i,s}(x), \widetilde{B}_{i,s}(x))+\xmark_2(\widetilde{b}_{i,s}(x), \widetilde{B}_{i+1,s}(x))+2\times 4^{-m+1}+\frac{8^{s}}{7}4^{-m}}{1+4\times 4^{-m+1}+\frac{8^{s}}{14}4^{-m+1}}, x\in[0,1]. \end{eqnarray*} The `seemingly strange' normalizing constant forces $\widetilde{B}_{i, s+1}(x)$ to take values in $[0,1]$. We repeat the above steps until we reach the construction of $\widetilde{B}_{i, k}$ (see Figure \ref{figure:tildeBi2:to:tildeBi3} for an illustration of such induction). We then approximate ${B}_{i,k}$ by $\widetilde{B}_{i,k}$. Note that $\widetilde{B}_{i,k}$ has $(2m+4)(k-2)+1$ hidden layers and $8(M+k-3)$ nodes on each hidden layer. \begin{figure}[ht!] \centering \subfigure[]{\includegraphics[width=2 in, height=1.5 in]{p1.pdf}} \hspace{1cm} \subfigure[]{\includegraphics[width=2 in, height=1.5 in]{p2.pdf}} \vspace{0.5cm} \subfigure[]{\includegraphics[width=2 in, height=1.5 in]{p3.pdf}} \hspace{1cm} \subfigure[]{\includegraphics[width=2 in, height=1.5 in]{p4.pdf}} \caption{Construction of $\widetilde{B}_{i,3}$ through induction. (a) and (b) demonstrate the architectures of the networks $\widetilde{a}_{i,2}$ and $\widetilde{b}_{i,2}$. (c) demonstrates the architecture of the network $\widetilde{B}_{i,2}$ with $c_1, c_2, c_3$ defined in (\ref{eq:definition:c1c2c3}). (d) demonstrates the induction relationship between $\widetilde{B}_{i,3}$ and $\widetilde{B}_{i,2}$. } \label{figure:tildeBi2:to:tildeBi3} \end{figure} We next approximate the tensor product B-spline basis $D_{\mathbf{i}}(\mathbf{x})=\prod_{j=1}^dB_{i_j,k}(x_j)$ by \begin{eqnarray*} \widetilde{D}_{\mathbf{i}, k}(\mathbf{x})=\xmark_d(\widetilde{B}_{i_1,k}(x_1),\widetilde{B}_{i_2,k}(x_2),\ldots, \widetilde{B}_{i_d,k}(x_d)), \textrm{ for each } \mathbf{i}=(i_1,\ldots, i_d) \in \Gamma. \end{eqnarray*} Note that $\widetilde{D}_{\mathbf{i}, k}$ has $(2m+3)(k+d-3)+k-1$ hidden layers each consisting of $(d+7)(M+2k-3)^d$ nodes. Finally, paralellizing $\widetilde{D}_{\mathbf{i}, k}(\mathbf{x}), I \in \Gamma$ according to (\ref{eq:pilot:estimator}), we construct $\widehat{f}_{\textrm{net}}$ as \begin{eqnarray} \widehat{f}_{\textrm{net}}(\mathbf{x})=\sum_{\mathbf{i}\in \Gamma}\widehat{b}_{\mathbf{i}}\widetilde{D}_{\mathbf{i}, k}(\mathbf{x}),\,\,\,\, x\in\Omega.\label{eq:optimal:DNN:estimator} \end{eqnarray} In comparing (\ref{eq:pilot:estimator}) with (\ref{eq:optimal:DNN:estimator}), if we can show that ${D}_{\mathbf{i}, k}$ and $\widetilde{D}_{\mathbf{i},k}$ are close enough, then one can expect that $\widehat{f}_{\text{net}}$ performs similarly to $\widehat{f}_{\textrm{pilot}}$. A rich class of statistical results in literature enable us to efficiently analyze $\widehat{f}_{\textrm{pilot}}$. In the rest of our analysis, we focus on cardinal B-splines for convenience. \begin{Assumption}\label{Assumption:A1} \label{A1:c} The knots $\{t_i, i=-k+1,\ldots, M+k-1\}$ have constant separation $h=M^{-1}$. \end{Assumption} \begin{remark} Assumption \ref{Assumption:A1} can be relaxed to $\max_{i}(t_{i+1}-t_{i})/\min_{i}(t_{i+1}-t_{i})\leq c$ for some constant $c>0$, under which one needs to redefine the separation $h=\max_{i}(t_{i+1}-t_{i})$. Results in this section continue to hold. This is a standard assumption for $B$-spline literature; see \cite{h98}. \end{remark} The following Theorem \ref{thm:approximation:sieve:DNN} is the main technical result of this paper, based on which Theorems \ref{thm:neural:estimator:rate:of:convergence:main:text}, \ref{thm:rate:convergence:additive:main:text}, \ref{thm:asymptotic:normality:neural:main:text} and \ref{thm:optimal:test:neural:main:text} will be proved. \begin{theorem}\label{thm:approximation:sieve:DNN} For fixed positive integer $m$, $\widehat{f}_{\textrm{net}}\in \mathcal{F}(L,\mathbf{p}(T))$ with $L=(2m+3)(k+d-1)+1$ and $T=3(M+2k)^d$. Under Assumption \ref{A0} and \ref{Assumption:A1}, if $k>\beta$ and $F>0$, then it holds that \begin{eqnarray*} \sup_{f_0\in\Lambda^\beta(F, \Omega)}\mathbb{E}_{f_0}\left\{\sup_{\mathbf{x} \in \Omega}|\widehat{f}_{\text{net}}(\mathbf{x})-\widehat{f}_{\textrm{pilot}}(\mathbf{x})|^2 \bigg|\mathbb{X}\right\}=O_P(h^{-2d}4^{-2m}). \end{eqnarray*} \end{theorem} Theorem \ref{thm:approximation:sieve:DNN} says that $\widehat{f}_{\text{net}}$ is a neural network with $L=(2m+3)(k+d-3)+k$ hidden layers each consisting of $T=(d+2)(M+k-1)^d$ nodes. The theorem also provides an explicit upper bound in terms of $(h,d,m)$ for the difference between $\widehat{f}_{\textrm{net}}$ and $\widehat{f}_{\textrm{pilot}}$. The proof of Theorem \ref{thm:approximation:sieve:DNN} relies on following Lemma \ref{lemma:approximation:b:spline:d:1}, \ref{lemma:approximation:b:spline:d:d}, \ref{lemma:spline:approximation} and \ref{lemma:difference:c:hat:c:0:main:text}. Let $\mathbf{i}_1, \mathbf{i}_2,\ldots, \mathbf{i}_q$ be the elements of $\Gamma$. Define \begin{eqnarray*} \mathbf{B}_k(x)&=&(B_{-k+1,k}(x), B_{-k+2,k}(x),\ldots, B_{0,k}(x),B_{1,k}(x),\ldots, B_{M-1,k}(x))^T\in \mathbb{R}^{M-k+1},\\ \mathbf{D}_{k}(\mathbf{x})&=&(D_{\mathbf{i}_1, k}(\mathbf{x}),D_{\mathbf{i}_2, k}(\mathbf{x}),\ldots, D_{\mathbf{i}_q, k}(\mathbf{x}))^T\in \mathbb{R}^{q},\\ \widetilde{\mathbf{B}}_k(x)&=&(\widetilde{B}_{-k+1,k}(x), \widetilde{B}_{-k+2,k}(x),\ldots, \widetilde{B}_{M-1,k}(x))^T\in \mathbb{R}^{M-k+1},\\ \widetilde{\mathbf{D}}_k(\mathbf{x})&=&(\widetilde{D}_{\mathbf{i}_1,k}(\mathbf{x}), \widetilde{D}_{\mathbf{i}_2,k}(\mathbf{x}),\ldots, \widetilde{D}_{\mathbf{i}_q,k}(\mathbf{x}))^T\in \mathbb{R}^q. \end{eqnarray*} Lemma \ref{lemma:approximation:b:spline:d:1} quantifies the differences between $\widetilde{\mathbf{B}}_k(\cdot)$ and $\mathbf{B}_k(\cdot)$. For convenience, for $L,p_0,\ldots,p_{L+1}\in\mathbb{N}$, let $\mathcal{N}\mathcal{N}(L,(p_0,p_1,\ldots,p_L,p_{L+1}))$ denote the class of $p_0$-input-$p_{L+1}$-output ReLU neural network functions of $L$ hidden layers, with the $j$th layer consisting of $p_j$ nodes, for $j=1,\ldots,L$. For any $v=(v_1, \ldots, v_p)\in \mathbb{R}^p$, let $\|v\|_\infty=\max_{1\leq i \leq p}|v_i|$. \begin{lemma}\label{lemma:approximation:b:spline:d:1} Given integers $k, M\geq 2$ and knots $t_{-k+1}<t_{-k+2}<\ldots< t_0<t_1< \ldots<t_M< t_{M+1} <\ldots< t_{M+k-1}$ such that $t_0=0, t_M=1$, there exists a $\widetilde{\mathbf{B}}_k\in \mathcal{N}\mathcal{N}(k(2m+3), (1, 3(M+2k), \ldots, 3(M+2k), M+k-1))$ taking values in $[0,1]$, such that \begin{eqnarray*} \sup_{x\in[0,1]}\|\widetilde{\mathbf{B}}_k(x)-\mathbf{B}_{k}(x)\|_\infty\leq \frac{8^{k}}{14}4^{-m}. \end{eqnarray*} \end{lemma} The proof of Lemma \ref{lemma:approximation:b:spline:d:1} is given in Appendix A. Based on Lemma \ref{lemma:approximation:b:spline:d:1}, we can bound the approximation error between $\widetilde{\mathbf{D}}_k$ and $\mathbf{D}_k$, which is summarized as Lemma \ref{lemma:approximation:b:spline:d:d}. \begin{lemma}\label{lemma:approximation:b:spline:d:d} Given integers $k, M\geq 2$ and knots $t_{-k+1}<t_{-k+2}<\ldots< t_0<t_1< \ldots<t_M< t_{M+1} <\ldots< t_{M+k-1}$ with $t_0=0, t_M=1$, there exist a $\widetilde{\mathbf{D}}_k\in \mathcal{N}\mathcal{N}((2m+3)(k+d-1), (d, 3(M+2k)^d, \ldots, 3(M+2k)^d, (M+k-1)^d))$ such that \begin{eqnarray*} \bigg\|\widetilde{\mathbf{D}}_k(\mathbf{x})-\mathbf{D}_{k}(\mathbf{x})\bigg\|_{\infty}\leq [4(d-1)+8^k]4^{-m},\;\textrm{ for all } \mathbf{x} \in \Omega. \end{eqnarray*} Furthermore, each element of $\widetilde{\mathbf{D}}_k$ is in $[0,1]$. \end{lemma} {\it Proof of Lemma \ref{lemma:approximation:b:spline:d:d}.} Let $\widetilde{\mathbf{B}}_k(x_1), \widetilde{\mathbf{B}}_k(x_1), \ldots, \widetilde{\mathbf{B}}_k(x_d)$ be the neural networks provided in Lemma \ref{lemma:approximation:b:spline:d:1}, which satisfy $|\widetilde{B}_{i, k}(x)-{B}_{i, k}(x)|\leq \delta_m$, where $\delta_m=8^k4^{-m}/14$. For each $(i_1, i_2, \ldots, i_d)\in \{-k+1, -k+2, \ldots, 1,2, \ldots, M-1\}^d$, we apply the product network $\xmark_d$ given in Proposition \ref{proposition:appriximation:product:k} to $(\widetilde{B}_{i_1,k}(x_1), \widetilde{B}_{i_2,k}(x_2),\ldots, \widetilde{B}_{i_d,k}(x_d))$. According to Proposition \ref{proposition:appriximation:product:k}, we have \begin{eqnarray*} \bigg|\xmark_d(\widetilde{B}_{i_1,k}(x_1), \widetilde{B}_{i_2,k}(x_2),\ldots, \widetilde{B}_{i_d,k}(x_d))-\prod_{j=1}^d B_{i_j,k}(x_j)\bigg|&\leq& (d-1)4^{-m+1}+d\delta_m\\ &\leq& [4(d-1)+8^k]4^{-m}. \end{eqnarray*} Now we deploy $\xmark_d(\widetilde{B}_{i_1,k}(x_1), \widetilde{B}_{i_2,k}(x_2),\ldots, \widetilde{B}_{i_d,k}(x_d))$ parallelly to construct the network $\widetilde{\mathbf{D}}_k$. Since we apply neural network $X_d$ to output of $\widetilde{\mathbf{B}}_k$, so the total number of hidden layers is at most $k(2m+3)+1+(d-1)(2m+3)-1=(2m+3)(d+k-1)$. Moreover, the number nodes in each hidden layer is not greater than the number of nodes in the output layer, which is further bounded by $3(M+2k)^d$. This completes the proof. The following Lemma \ref{lemma:spline:approximation} is consequence of \cite[Theorem 15.1 and Theorem 15.2]{gkkw06} and \cite[Theorem 12.8 and (13.69)]{s07}, which quantifies an approximation error of tensor product B-spline. \begin{lemma}\label{lemma:spline:approximation} Suppose that Assumption \ref{Assumption:A1} is satisfied. For any $f\in \Lambda^\beta(F, \Omega)$ and any integer $k\geq \beta$, there exists a real sequence $c_i$ such that $\sup_{\mathbf{x} \in \Omega}\bigg|\sum_{\mathbf{i}\in \Gamma}c_i D_{\mathbf{i}}(\mathbf{x})-f(\mathbf{x})\bigg|\leq A_f h^\beta$ where $A_f>0$ which only depends on partial derivatives of $f$ upto order $k$. Moreover, the sequence $c_i$ satisfy $|c_i|\leq A_f$. Moreover, $sup_{f\in \Lambda^\beta(F, \Omega)}A_f<\infty$, where the upper bound only depends on $F$ and $\beta$. \end{lemma} Define $\Phi=(\mathbf{D}_k(\mathbf{x}_1),\ldots, \mathbf{D}_k(\mathbf{x}_n))^T\in \mathbb{R}^{n\times q}$ and $\mathbf{Y}=(Y_1,\ldots Y_n)^T$. The following Lemma \ref{lemma:difference:c:hat:c:0:main:text} quantifies the magnitude of $\widehat{C}=(\Phi^T\Phi)^{-1}\Phi^T\mathbf{Y}$; recalling that such $\widehat{C}$ is the solution to the least square problem (\ref{LSE:eqn}). Its proof is provided in Appendix B. \begin{lemma}\label{lemma:difference:c:hat:c:0:main:text} Under Assumptions \ref{A0} and \ref{Assumption:A1}, if $h=o(1)$ and $\log(h^{-1})=o(nh^d)$, then \[ \sup_{f_0\in\Lambda^\beta(F,\Omega)} \mathbb{E}_{f_0}\left(\widehat{C}^T\widehat{C}\big|\mathbb{X}\right)=O_P(h^{-d}). \] \end{lemma} We are now ready to provide the Proof of Theorem \ref{thm:approximation:sieve:DNN}. {\it Proof of Theorem \ref{thm:approximation:sieve:DNN}.} For any $f_0\in\Lambda^\beta(F, \Omega)$, let $\mathbf{f}_0=(f_0(\mathbf{x}_1),\ldots,f_0(\mathbf{x}_n))^T$. Also let $\bm{\epsilon}=(\epsilon_1,\ldots,\epsilon_n)^T$, $\widehat{\mathbf{f}}_{\textrm{pilot}}=(\widehat{f}_{\textrm{pilot}}(\mathbf{x}_1), \ldots,\widehat{f}_{\textrm{pilot}}(\mathbf{x}_n))^T$. According to Lemma \ref{lemma:spline:approximation} and by $k\geq \beta$, there exists a $C=(c_1, c_2,\ldots, c_q)^T\in \mathbb{R}^q$ such that for any $\mathbf{x}\in\Omega$, $|C^T\mathbf{D}_k(\mathbf{x})-f_0(\mathbf{x})|\leq A_{f_0}h^\beta$. By least square algorithm (\ref{LSE:eqn}), we have \begin{eqnarray*} \widehat{\mathbf{f}}_{\textrm{pilot}}=\Phi(\Phi^T \Phi)^{-1}\Phi^T \mathbf{Y}&=&\Phi(\Phi^T \Phi)^{-1}\Phi^T (\Phi C+\mathbf{E}+\bm{\epsilon})\\ &=&\Phi C+\Phi(\Phi^T \Phi)^{-1}\Phi^T \mathbf{E}+\Phi(\Phi^T \Phi)^{-1}\Phi^T \bm{\epsilon}\\ &=&\mathbf{f}_0-(I-\Phi(\Phi^T \Phi)^{-1}\Phi^T)\mathbf{E}+\Phi(\Phi^T \Phi)^{-1}\Phi^T \bm{\epsilon}, \end{eqnarray*} where $\mathbf{E}=(E_1, E_2, \ldots, E_n)^T\in \mathbb{R}^n$ with $E_i=f_0(\mathbf{x}_i)-C^T\mathbf{D}_k(\mathbf{x}_i)$. It follows from (\ref{eq:pilot:estimator}), (\ref{eq:optimal:DNN:estimator}) and (\ref{LSE:eqn}) that $\widehat{f}_{\textrm{pilot}}(\mathbf{x})=\widehat{C}^T\mathbf{D}_k(\mathbf{x})$ and $\widehat{f}_{\textrm{net}}(\mathbf{x})=\widehat{C}^T\widetilde{\mathbf{D}}_k(\mathbf{x})$. Therefore, for any $\mathbf{x}\in\Omega$, we have \begin{eqnarray*} |\widehat{f}_{\textrm{pilot}}(\mathbf{x})-\widehat{f}_{\textrm{net}}(\mathbf{x})|^2&=&\big\|\widehat{C}^T\left(\mathbf{D}_k(\mathbf{x})-\widetilde{\mathbf{D}}_k(\mathbf{x})\right)\big\|_2^2\\ &=& \widehat{C}^T\widehat{C}\left(\mathbf{D}_k(\mathbf{x})-\widetilde{\mathbf{D}}_k(\mathbf{x})\right)^T\left(\mathbf{D}_k(\mathbf{x})-\widetilde{\mathbf{D}}_k(\mathbf{x})\right)\\ &\leq&q\widehat{C}^T\widehat{C} \sup_{\mathbf{x} \in [0,1]^d}\big\|\mathbf{D}_k(\mathbf{x})-\widetilde{\mathbf{D}}_k(\mathbf{x})\big\|_{\infty}^2 \le q\widehat{C}^T\widehat{C}[4(d-1)+8^k]^24^{-2m}, \end{eqnarray*} where the last inequality follows from Lemma \ref{lemma:approximation:b:spline:d:d}. Following Lemma \ref{lemma:difference:c:hat:c:0:main:text} and the fact $q\asymp h^{-d}$, we have \begin{eqnarray*} \sup_{f_0\in\Lambda^\beta(F,\Omega)} \mathbb{E}_{f_0}\bigg(\sup_{\mathbf{x} \in \Omega}|\widehat{f}_{\textrm{pilot}}(\mathbf{x})-\widehat{f}_{\textrm{net}}(\mathbf{x})|^2 \bigg|\mathbb{X}\bigg)&\leq& q[4(k-1)+8k]^2 4^{-2m}\sup_{f_0\in\Lambda(F,\Omega)}\mathbb{E}\left(\widehat{C}^T\widehat{C}\big|\mathbb{X}\right)\\ &=&O_P(h^{-2d}4^{-2m}), \end{eqnarray*} which completes the proof. Theorem \ref{thm:neural:estimator:rate:of:convergence:main:text} is a simple consequence of Theorem \ref{thm:approximation:sieve:DNN} and Lemma \ref{lemma:rate:of:convergence:sieve} below with $h\asymp T^{-1/d}$ and $m\asymp \frac{L}{3(k+d)}$. Proof of Lemma \ref{lemma:rate:of:convergence:sieve} is deferred to Appendix B. \begin{lemma}\label{lemma:rate:of:convergence:sieve} Under the Assumptions \ref{A0} and \ref{Assumption:A1}, if $h=o(1)$ and $\log(h^{-1})=o(nh^d)$ hold, then it holds that \begin{eqnarray*} \sup_{f_0\in\Lambda^\beta(F,\Omega)} \mathbb{E}_{f_0}\bigg\{\|\widehat{f}_{\textrm{pilot}}-f_0\|_{L^2}^2\bigg|\mathbb{X}\bigg\}=O_P\bigg(h^{2\beta}+\frac{1}{nh^d}\bigg). \end{eqnarray*} \end{lemma} \section{Asymptotic Distribution and Optimal Testing}\label{sec:byproducts} In this section, we derive asymptotic distributions for $\widehat{f}_{\textrm{net}}$ and a corresponding hypothesis testing procedure. The results are simply byproducts of Theorem \ref{thm:approximation:sieve:DNN}. Theorem \ref{thm:asymptotic:normality:neural:main:text} below establishes a pointwise asymptotic distribution for $\widehat{f}_{\textrm{net}}(\mathbf{x})$ for any $\mathbf{x} \in \Omega$. The result is a direct consequence of Theorem \ref{thm:approximation:sieve:DNN} and the asymptotic distribution of $\widehat{f}_{\textrm{pilot}}(\mathbf{x})$. \begin{theorem}\label{thm:asymptotic:normality:neural:main:text} Under the Assumptions \ref{A0} and \ref{Assumption:A1}, if $hn^{\frac{1}{2\beta+d}}=o(1)$, $\log(h^{-1})=o(nh^d)$ and $n^{1/2}h^{-d/2}=o(4^{m})$, then for any $\mathbf{x} \in \Omega$, we have \begin{eqnarray*} \frac{\widehat{f}_{\textrm{net}}(\mathbf{x})-f_0(\mathbf{x})}{\sqrt{\mathbf{D}^T_k(\mathbf{x})(\Phi^T\Phi)^{-1}\mathbf{D}_k(\mathbf{x})}}\xrightarrow[]{D}N(0,1), \end{eqnarray*} where $\Phi=(\mathbf{D}_k(\mathbf{x}_1),\mathbf{D}_k(\mathbf{x}_2),\ldots, \mathbf{D}_k(\mathbf{x}_n))^T\in \mathbb{R}^{n\times q}$. \end{theorem} {\it Proof of Theorem \ref{thm:asymptotic:normality:neural:main:text}.} For fixed $\mathbf{x}\in [0,1]^d$, let $V(\mathbf{x})=\mathbf{D}^T_k(\mathbf{x})(\Phi^T\Phi)^{-1}\mathbf{D}_k(\mathbf{x})$. By \cite[Theorems 3.1 and 5.2]{h03}, it follows that \begin{eqnarray}\label{thm11:eqn:1} \frac{\widehat{f}_{\textrm{pilot}}(\mathbf{x})-f_0(\mathbf{x})}{\sqrt{V(\mathbf{x})}}\xrightarrow[]{D}N(0,1). \end{eqnarray} By Lemma \ref{lemma:empirical:eigen:value} in Appendix B, with probability approaching 1, we have $V(\mathbf{x})\geq \frac{2}{a_1nh^d}\mathbf{D}_k(\mathbf{x})^T\mathbf{D}_k(\mathbf{x})\geq \frac{2b^2}{a_1nh^d}$. By the proof of Theorem \ref{thm:approximation:sieve:DNN} we get that $|\widehat{f}_{\textrm{pilot}}(\mathbf{x})-\widehat{f}_{\textrm{net}}(\mathbf{x})|^2=O_P(h^{-2d}4^{-2m})$. Therefore, \begin{equation}\label{thm11:eqn:2} \frac{\widehat{f}_{\textrm{pilot}}(\mathbf{x})-\widehat{f}_{\textrm{net}}(\mathbf{x})}{\sqrt{V(\mathbf{x})}}=O_P(n^{1/2}h^{-d/2}4^{-m})=o_P(1). \end{equation} Theorem \ref{thm:asymptotic:normality:neural:main:text} follows by (\ref{thm11:eqn:1}) and (\ref{thm11:eqn:2}). This completes the proof. In what follows, we consider a hypothesis testing problem: $H_0: f_0=0$ vs. $H_1: f\neq0$. Consider a test statistic $T_n=\|\widehat{f}_{\textrm{net}}\|_n^2$, where $\|f\|_n^2=\sum_{i=1}^n f(\mathbf{x}_i)^2/n$ is the empirical norm. The following Theorem \ref{thm:optimal:test:neural:main:text} derives null distribution of $T_n$ and analyzes its power under a sequence of local alternatives. Again, this result is a byproduct of Theorem \ref{thm:approximation:sieve:DNN}. \begin{theorem}\label{thm:optimal:test:neural:main:text} Suppose $n^{\frac{4\beta+2d}{4\beta+d}}=O(4^m)$ and $h\asymp n^{-\frac{2}{4\beta+d}}$, then the following hold: \begin{enumerate} \item Under $H_0: f_0=0$, it follows that \begin{eqnarray}\label{thm:11:1} \frac{nT_n-q}{\sqrt{2q}}\xrightarrow[]{D}N(0,1). \end{eqnarray} \item For any $\delta>0$, there exists a $C_\delta>0$ such that, under $H_1: f=f_0$ with $\|f_0\|_n\geq C_\delta n^{-\frac{2\beta}{4\beta+d}}$, it holds that \begin{eqnarray}\label{thm:11:2} \mathbb{P}\bigg(\bigg|\frac{nT_n-q}{\sqrt{2q}}\bigg|>z_{\alpha/2}\bigg)\geq 1-\delta, \end{eqnarray} where $z_{\alpha/2}$ is the $1-\alpha/2$ upper percentile of standard normal variable. \end{enumerate} \end{theorem} Part (\ref{thm:11:1}) of Theorem \ref{thm:optimal:test:neural:main:text} suggests a testing rule at significance $\alpha$: reject $H_0$ if and only if \[ \bigg|\frac{nT_n-q}{\sqrt{2q}}\bigg|\ge z_{\alpha/2}. \] Part (\ref{thm:11:2}) of Theorem \ref{thm:optimal:test:neural:main:text} says that the power of $T_n$ is at least $1-\delta$ provided that the null and alternative hypotheses are separated by $C_\delta n^{-\frac{2\beta}{4\beta+d}}$ in terms of $\|\cdot\|_n$-norm. The separation rate is optimal in the sense of \cite{Ingster93}. {\it Proof of Theorem \ref{thm:optimal:test:neural:main:text}.} Observe that \begin{eqnarray}\label{eq:them:optimal:test:neural} \frac{n\|\widehat{f}_{\textrm{net}}\|_n^2-q}{\sqrt{2q}}=\frac{n\|\widehat{f}_{\textrm{pilot}}\|_n^2-q}{\sqrt{2q}}+\frac{n\|\widehat{f}_{\textrm{net}}\|_n^2-n\|\widehat{f}_{\textrm{pilot}}\|_n^2}{\sqrt{2q}}. \end{eqnarray} By Theorem \ref{thm:approximation:sieve:DNN} and Lemma \ref{lemma:rate:of:convergence:sieve} (see Appendix B for its proof), both $\|\widehat{f}_{\textrm{net}}-\widehat{f}_{\textrm{pilot}}\|_n$ and $\|\widehat{f}_{\textrm{pilot}}-f_0\|_n$ are $O_P(1)$, we have \begin{eqnarray*} |\|\widehat{f}_{\textrm{net}}\|_n^2-\|\widehat{f}_{\textrm{pilot}}\|_n^2|&=& |\|\widehat{f}_{\textrm{net}}\|_n-\|\widehat{f}_{\textrm{pilot}}\|_n|\times \left(\|\widehat{f}_{\textrm{net}}\|_n+\|\widehat{f}_{\textrm{pilot}}\|_n\right)\\ &\leq&\|\widehat{f}_{\textrm{net}}-\widehat{f}_{\textrm{pilot}}\|_n\times \left(\|\widehat{f}_{\textrm{net}}-\widehat{f}_{\textrm{pilot}}\|_n+2\|\widehat{f}_{\textrm{pilot}}\|_n\right)\\ &\leq&\|\widehat{f}_{\textrm{net}}-\widehat{f}_{\textrm{pilot}}\|_n\times \left(\|\widehat{f}_{\textrm{net}}-\widehat{f}_{\textrm{pilot}}\|_n+2\|\widehat{f}_{\textrm{pilot}}-f_0\|_n+2\|f_0\|_n\right)\\ &=&\|\widehat{f}_{\textrm{net}}-\widehat{f}_{\textrm{pilot}}\|_n\times O_P(1)\\ &=&O_P(h^{-d}4^{-m}). \end{eqnarray*} Therefore, the second term in (\ref{eq:them:optimal:test:neural}) is of order $O_P(nh^{-d}4^{-m}q^{-1/2})=O_P\left(n^{\frac{4\beta+2d}{4\beta+d}}4^{-m}\right)=o_P(1)$, where we have used the fact $q\asymp h^{-d}$. The result then follows by Lemma \ref{lemma:optimal:test:sieve} in Appendix D. This completes the proof. \newpage
{ "timestamp": "2019-02-06T02:12:29", "yymm": "1902", "arxiv_id": "1902.01687", "language": "en", "url": "https://arxiv.org/abs/1902.01687", "abstract": "Deep neural network is a state-of-art method in modern science and technology. Much statistical literature have been devoted to understanding its performance in nonparametric estimation, whereas the results are suboptimal due to a redundant logarithmic sacrifice. In this paper, we show that such log-factors are not necessary. We derive upper bounds for the $L^2$ minimax risk in nonparametric estimation. Sufficient conditions on network architectures are provided such that the upper bounds become optimal (without log-sacrifice). Our proof relies on an explicitly constructed network estimator based on tensor product B-splines. We also derive asymptotic distributions for the constructed network and a relating hypothesis testing procedure. The testing procedure is further proven as minimax optimal under suitable network architectures.", "subjects": "Machine Learning (cs.LG); Machine Learning (stat.ML)", "title": "Optimal Nonparametric Inference via Deep Neural Network", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.983342957061873, "lm_q2_score": 0.8152324983301567, "lm_q1q2_score": 0.8016531356009147 }
https://arxiv.org/abs/1509.02945
Low-Dimensional Galerkin Approximations of Nonlinear Delay Differential Equations
This article revisits the approximation problem of systems of nonlinear delay differential equations (DDEs) by a set of ordinary differential equations (ODEs). We work in Hilbert spaces endowed with a natural inner product including a point mass, and introduce polynomials orthogonal with respect to such an inner product that live in the domain of the linear operator associated with the underlying DDE. These polynomials are then used to design a general Galerkin scheme for which we derive rigorous convergence results and show that it can be numerically implemented via simple analytic formulas. The scheme so obtained is applied to three nonlinear DDEs, two autonomous and one forced: (i) a simple DDE with distributed delays whose solutions recall Brownian motion; (ii) a DDE with a discrete delay that exhibits bimodal and chaotic dynamics; and (iii) a periodically forced DDE with two discrete delays arising in climate dynamics. In all three cases, the Galerkin scheme introduced in this article provides a good approximation by low-dimensional ODE systems of the DDE's strange attractor, as well as of the statistical features that characterize its nonlinear dynamics.
\section{Introductuion} Systems of delay differential equations (DDEs) are widely used in many fields such as the biosciences, climate dynamics, chemistry, control theory, economics, and engineering \cite{Bhattacharya_al82, Diaz2014, Diekmann_al95, GCStep15, Ghil_Childress'87, GZT08, Hale_Lunel93, Sieber2014, Kuang93, LS10, MacDonald89, Michiels_al07, Roques_al15, Smith11, Stepan89}. In particular, certain DDEs or more general differential equations with retarded arguments can be derived from hyperbolic partial differential equations that support wave propagation \cite{chekroun_glatt-holtz,Galanti_al00,Hale_Lunel93}. In contrast to ordinary differential equations (ODEs), the state space associated even with a scalar DDE is infinite-dimensional, due to the presence of time-delay terms, which require providing initial data over an interval $ [-\tau,0],$ where $\tau >0$ is the delay. It is often desirable, though, to have low-dimensional ODE systems that capture qualitative features, as well as approximating certain quantitative aspects of the DDE dynamics. The derivation of ODE approximations of DDEs involves, in general, two types of function spaces as state space: that of continuous functions $C([-\tau, 0]; \mathbb{R}^d)$, and the Hilbert space $L^2([-\tau, 0); \mathbb{R}^d)$. {\mg The former spaces have} been extensively used in the case of bifurcation analysis \cite{Casal_al80,Chow_al77,Das_al02,Faria_al95,Kazarinoff_al78,Nayfeh08,wischert1994delay}, {\mg while the latter are} typically adopted in situations where quantitative accuracy is an important factor, such as in optimal control \cite{Banks_al78,kappel1978autonomous,Banks_al84,Kappel86,Kappel_al87,Kunisch82,Ito_Teglas86,banks1979spline}. Within the Hilbert {\mg space} setting, different basis functions have been proposed to decompose the state space; {\mg these} include, among others, step functions \cite{Banks_al78,kappel1978autonomous}, splines \cite{banks1979spline,Banks_al84}, and orthogonal polynomial functions, such as Legendre polynomials \cite{Kappel86,Ito_Teglas86}. Compared with step functions or splines, the use of orthogonal polynomials leads typically to ODE approximations with lower dimensions, for a given precision \cite{banks1979spline,Ito_Teglas86}. On the other hand, classical polynomial basis functions do not live in the domain of the linear operator underlying the DDE, which leads to technical complications in establishing convergence results \cite{Kappel86,Ito_Teglas86}; see Remark~\ref{Rmk_problems_to_overcome}(iii) below. In the present article, we propose to avoid these technical difficulties in approximating DDEs as systems of ODEs by using an alternative polynomial basis: the elements of this basis belong naturally to the domain of the underlying linear operator, but they have not been used in the DDE literature so far. The polynomials we shall use are named after Koornwinder~\cite{Koo84}, who investigated polynomials that are orthogonal with respect to weight functions adjoining point masses, as discussed in Section~\ref{sect:basis} below. This polynomial basis turns out to be particularly useful not only for the rigorous analysis of polynomial-based Galerkin approximations of nonlinear systems of DDEs, as shown in Section~\ref{Sec_Galerkin_approx}, but also for their numerical treatment, cf. Section~\ref{Sect_Numerics}. Useful new properties of the Koornwinder polynomials are identified in Lemma ~\ref{Fundamental_lemma} for the scalar case, and a generalization of these polynomials to the vector case is given in Section~\ref{Sec_Vectorization}; the latter includes the multi-dimensional extension of Lemma ~\ref{Fundamental_lemma}, namely Lemma \ref{Super_Fundamental_lemma}. We show that these properties are essential for checking key stability and convergence conditions in {\mg Lemmas~\ref{Lem_A2} and \ref{Lem_A1}. Standard Galerkin approximation results for abstract nonlinear ODEs} in Hilbert spaces are recalled in Theorem \ref{ParisVI_thm} and the rest of Section~\ref{Subsect_ODE_Galerkin}. They are then applied, with the help of Lemmas~\ref{Lem_A2} and \ref{Lem_A1}, to nonlinear systems of DDEs in Section~\ref{Subsect_DDE_Galerkin}. Finite-time uniform convergence results are then derived for the proposed Galerkin approximations of nonlinear systems of DDEs, subject to simple and checkable conditions on the nonlinear term. These conditions are identified in Section~\ref{Subsect_DDE_Galerkin}; see Corollaries~\ref{Cor_DDE_local_Lip_Case1} and \ref{Cor_DDE_local_Lip_Case2}. The results apply to a broad class of nonlinear systems of DDEs, as discussed in Section~\ref{Sec_examples}. The proposed framework yields a simple numerical calculation of the corresponding Galerkin approximations. Their coefficients are easily computable from the original system of DDEs by relying on simple recurrence formulas, cf.~Proposition~\ref{thm:Pn}, and by solving upper triangular systems of linear equations, cf.~Proposition~\ref{prop:dPn}; see Section \ref{Sec_Galerkin_analytic} and Appendix~\ref{Appendix_systems}. Finally, we outline here a useful interpretation of our proposed scheme regarding the finite-dimensional approximation of the linear part $\mathcal{A}$ of general systems of DDEs, when considered in the framework of Hilbert spaces, cf.~\eqref{Def_A2}. This interpretation relies on a formulation of the evolution in time of the initial state $\{x(\theta): \theta \in [-\tau, 0]\}$ as the solution of a partial differential equation (PDE); see also Remark \ref{PDE_rem}. To do so, we first distinguish between the {\it historic part} of the evolving state, $\{x(t + \theta): \theta \in [-\tau, 0)\}$, and the {\it state part}, $\{x(t)\}$. Denoting by $u(t,\theta)$ the historic part, one can then rewrite, for instance, the simple linear DDE \begin{equation}\label{lin_case} \dot{x}=x(t-\tau), \quad \tau>0, \end{equation} as the linear PDE \begin{equation}\label{lin_PDE} \partial_t u = \partial_{\theta}u, \quad -\tau \le \theta < 0, \end{equation} with the boundary condition \begin{equation}\label{PDE_BC} \partial_{t} u|_{\theta = 0} = u(t,-\tau), \;\; t \geq 0. \end{equation} The key point is that, roughly speaking, the local differential operator $v\mapsto \partial_{\theta} v$ --- obtained as the history component of $\mathcal{A}$, and written out explicitly in~Eq.~\eqref{Def_A}, for instance --- is approximated here by the nonlocal differential operator \begin{equation}\label{nonlocal_PDE_intro} v \mapsto \partial_{\theta} v +b_N(\theta)\Big(v(-\tau)- \partial_{\theta} v\big\vert_{\theta=0}\Big), \quad \theta \in[-\tau,0), \end{equation} and that $b_N(\theta)$ --- expressed by means of Koornwinder polynomials, cf.~\eqref{bN-coeff} --- is a bounded oscillatory coefficient that vanishes in $L^2$ as $N\rightarrow \infty$; see Lemma \ref{Fundamental_lemma}. This nonlocal operator is the PDE representation for the history component of our Koornwinder-based Galerkin approximation $\mathcal{A}_N$ given in \eqref{Eq_AN}. Note that terms such as $v(-\tau) - \partial_{\theta} v\big\vert_{\theta=0}$, which is responsible for the nonlocal aspect of \eqref{nonlocal_PDE_intro}, play an important role in the theory of numerical approximation of DDEs; see, for instance,\cite[Appendix A]{GZT08} and references therein. In particular, this term provides the exact value of the jump associated with the boundary condition \eqref{PDE_BC}. The first such jump occurs at $t=0$ in the derivative of solutions of Eq.~\eqref{lin_case} that emanate from a constant history\footnote{\ For instance, if $\tau =1$ and the history is given by $\{\psi(\theta)\equiv c, \; - \tau \le \theta < 0\}$, then the solution $x(t)$ to Eq.~\eqref{lin_case} is equal to $c(t+1)$ on $[0,1]$. This discontinuity leads to a jump $\psi(-1)-\partial_{\theta} \psi\big\vert_{\theta=0} = c$ in its time derivative at $t=0$.}; this discontinuity propagates to higher-order derivatives at subsequent, integer multiples of the delay, $t=k\tau$, $k\in \mathbb{Z}^+$. The fact that this jump term is weighted by a vanishing term suggests that, for a given degree of accuracy, good approximation can be expected when using relatively low-dimensional Koornwinder-based approximations, as long as $b_N$ vanishes sufficiently quickly. We do not address such numerical considerations here; see, however, Table \ref{table} in Remark \ref{PDE_rem} for results in the case of Eq.~\eqref{lin_case}. Instead, in Section~\ref{Sect_Numerics}, we provide several applications that show the proposed approximation to be not only rigorously justified, but very effective in nonlinear cases that yield quasi-periodic and chaotic, as well as nearly Brownian dynamics. In each case, low-dimensional ODE systems succeed in approximating important topological as well as statistical features of the corresponding DDE's nonlinear dynamics. The article is organized as follows. In Section~\ref{sect:preliminaries}, we introduce the functional framework that will be adopted in Section \ref{Subsect_DDE_Galerkin} to recast a system of nonlinear DDEs into an abstract ODE. This framework relies on Hilbert spaces endowed with a natural inner product with a point mass. Koornwinder polynomials are then introduced in Section~\ref{sect:basis}. The convergence of the Galerkin ODE systems built by projecting onto these polynomials to the original DDEs is proven in Section \ref{Sec_Galerkin_approx}. We provide explicit expressions of the Galerkin approximation in Section~\ref{Sec_Galerkin_analytic} for the scalar case, and in Appendix~\ref{Appendix_systems} for nonlinear systems of DDEs. Finally, numerical applications to three nonlinear DDEs are provided in Section~\ref{Sect_Numerics}. These applications involve: (i) a simple DDE with distributed delays whose solutions recall Brownian motion \cite{sprott2007simple}; (ii) a DDE with a discrete delay that illustrates bimodal, as well as chaotic dynamics \cite{sprott2007simple}; and (iii) a periodically forced DDE with two discrete delays as a highly idealized model of the El Ni\~no-Southern Oscillation (ENSO: \cite[and references therein]{GZT08}). In all three examples, it is shown that our Galerkin scheme provides a good approximation by low-dimensional ODE systems of the DDE's strange attractor, as well as the statistical features that characterize the associated nonlinear dynamics. \section{Background and motivation} \label{sect:preliminaries} We introduce in this section the functional framework that will be adopted in Section \ref{Subsect_DDE_Galerkin} for the derivation of Galerkin approximations of a given nonlinear system of DDEs. Several function spaces can be used as a state space for the reformulation of a system of DDEs into an abstract ODE where among the most standard ones, those built-up out of the space of continuous functions on the interval $[-\tau,0]$ play an important role in the DDE theory; e.g.~\cite{Diekmann_al95,Hale_Lunel93}. In this article, we adopt instead the use of Hilbert spaces which are more classically used in control or approximation theory of DDEs; see e.g., \cite{Banks_al78,burns1983linear,curtain1995,Kappel86,kappel1986equivalence,Kappel_al87,nakagiri1989controllability}. For a didactic expository of the associated theory of semigroups for (linear) systems of DDEs in this functional setting we refer to \cite[Sect.~2.4]{curtain1995}; see also \cite{burns1983linear}. More precisely, the following Hilbert product space \begin{equation} \label{H_space} \mathcal{H} := L^2([-\tau,0); \mathbb{R}^d) \times \mathbb{R}^d, \end{equation} will serve as our state space, and will be endowed with the inner product defined for any $(f_1, \gamma_1),\, (f_2, \gamma_2) \in \mathcal{H}$, as: \begin{equation} \label{H_inner} \langle (f_1, \gamma_1), (f_2, \gamma_2) \rangle_{\mathcal{H}} := \frac{1}{\tau} \int_{-\tau}^0\langle f_1(\theta), f_2(\theta) \rangle \mathrm{d} \theta + \langle \gamma_1,\gamma_2\rangle, \end{equation} where $\langle \cdot, \cdot \rangle$ denotes the Euclidean inner product of $ \mathbb{R}^d$. We will also make use of the following subspace of $\mathcal{H}$: \begin{equation} \mathcal{V} := H^1([-\tau,0); \mathbb{R}^d) \times \mathbb{R}^d, \end{equation} where $H^1([-\tau,0); \mathbb{R}^d)$ denotes the standard Sobolev subspace of $L^2([-\tau,0); \mathbb{R}^d)$; see, e.g.~ \cite[Chap.~8]{brezis_book}. This space consists of functions that are square integrable and whose first-order weak derivatives exist in a distributional sense and are also square integrable. Instead of presenting the general nonlinear systems of DDEs considered in this article (see Section \ref{Subsect_DDE_Galerkin}), we introduce below a class of scalar DDEs that will serve us to identify within a simple context, the issues inherent to the Galerkin approximation of DDEs; see Remark \ref{Rmk_problems_to_overcome} hereafter. \begin{ex}\label{Ex_DDE_into_ODE} In this example, we recall how a scalar DDE can be recast into an abstract ODE. For simplicity, we will focus on the following autonomous scalar DDE ($d=1$): \begin{equation} \label{Eq_DDE} \frac{\mathrm{d} x(t)}{\mathrm{d} t} = a x(t) + bx(t-\tau) + c \int_{t-\tau}^t x(s)\mathrm{d} s + F\Big(x(t), \int_{t-\tau}^t x(s) \mathrm{d} s \Big), \end{equation} where $a$, $b$ and $c$ are real numbers, $\tau> 0$ is the delay parameter, and $F$ is a given scalar nonlinear function. The case of nonlinear systems of DDEs will be dealt with in Section \ref{Subsect_DDE_Galerkin}. The reformulation of Eq.~\eqref{Eq_DDE} into an abstract ODE is classical and proceeds as follows. Let us denote by $x_t$ the time evolution of the history segments of a solution to Eq.~\eqref{Eq_DDE}, namely \begin{equation} \label{shift} x_t(\theta):=x(t+\theta), \qquad t \ge 0, \qquad \theta \in [-\tau, 0]. \end{equation} Now, by introducing the new variable \begin{equation} u(t) := (x_t,x(t))=(x_t, x_t(0)), \end{equation} Eq.~\eqref{Eq_DDE} can be rewritten as the following abstract ODE: \begin{equation} \label{eq:DDE_abs} \frac{\mathrm{d} u}{\mathrm{d} t} = \mathcal{A} u + \mathcal{F}(u), \end{equation} where the linear operator $\mathcal{A} \colon D(\mathcal{A}) \subset \mathcal{V} \rightarrow \mathcal{H}$ is defined by \begin{equation} \begin{aligned} \label{Def_A} \lbrack \mathcal{A} \Psi \rbrack (\theta) & := \begin{cases} {\displaystyle \frac{\mathrm{d}^+ \Psi^D}{\mathrm{d} \theta}}, & \theta \in[-\tau, 0), \vspace{0.4em}\\ {\displaystyle a \Psi^S + b\Psi^D(-\tau) + c \int_{-\tau}^0 \Psi^D(s)\mathrm{d} s}, & \theta = 0, \end{cases} \end{aligned} \end{equation} with the domain $\mathcal{A}$ given by (cf. \cite[Prop.~2.6]{Kappel86}) \begin{equation} \label{D_of_A} D(\mathcal{A}) = \Big \{(\Psi^D, \Psi^S) \in L^2([-\tau, 0); \mathbb{R})\times \mathbb{R} : \Psi^D \in H^1([-\tau, 0); \mathbb{R}), \lim_{\theta \rightarrow 0^-} \Psi^D(\theta) = \Psi^S \Big \}; \end{equation} and with the nonlinearity $\mathcal{F} \colon \mathcal{H} \rightarrow \mathcal{H}$ defined by \begin{equation} \begin{aligned} \label{Def_F} [\mathcal{F} (\Psi) ](\theta) & := \begin{cases} 0, & \theta \in[-\tau, 0), \vspace{0.4em}\\ F \Big(\Psi^S, \int_{-\tau}^0 \Psi^D(s) \mathrm{d} s \Big), & \theta = 0, \end{cases} \quad \text{ } \forall \: \Psi = (\Psi^D, \Psi^S) \in \mathcal{H}. \end{aligned} \end{equation} With $D(\mathcal{A})$ such as given in \eqref{D_of_A}, the operator $\mathcal{A}$ generates a linear $C_0$-semigroup on $\mathcal{H}$ so that the Cauchy problem associated with the linear equation $\dot{u}=\mathcal{A} u$ is well-posed in the Hadamard's sense; see e.g~\cite[Thm.~2.4.6]{curtain1995}. The well-posedness problem for the nonlinear equation depends obviously on the nonlinear term $\mathcal{F}$ and we refer to Section \ref{Subsect_DDE_Galerkin} for a solution to this problem within our functional framework; see also \cite{Webb76}. \end{ex} \needspace{1\baselineskip} \begin{rem}\label{Rmk_problems_to_overcome} \hspace*{2em} \vspace*{-0.4em} \begin{itemize} \item[{\mg (i)}] It is important to note that when instead of $L^2([-\tau, 0); \mathbb{R}^d)$, the space of continuous functions $X=C([-\tau, 0]; \mathbb{R}^d)$ endowed with the supremum norm is retained \cite{Hale_Lunel93}, the continuity requirement at $0$ in \eqref{D_of_A} is naturally satisfied. On the other hand, $X$ is not a Hilbert space and the analysis of the {\it adjoint eigenvalue problem} \cite[Sect.~7.5]{Hale_Lunel93} is required for the derivation of low-dimensional ODE systems which no longer contain memory terms \cite{wischert1994delay}. By working within the framework of Hilbert spaces we avoid technicalities inherent to the analysis of this adjoint problem. \item[(ii)] When we consider the Hilbert space $\mathcal{H}$, a natural choice of set of functions to decompose the solutions of \eqref{eq:DDE_abs} is constituted by the eigenfunctions of the operator $\mathcal{A}$ with domain $D(\mathcal{A})$. When $\mathcal{A}$ does not contain distributed delay terms, these eigenfunctions are well-known and can be found in e.g.~\cite[Thm.~2.4.6]{curtain1995}. In case where the eigenvalues of $\mathcal{A}$ are all simple, this set of eigenfunctions actually correspond to the set $\mathcal{E}$ of eigenfunctions in $C([-\tau, 0]; \mathbb{R}^d)$ \cite[Thm.~4.2, p.~207]{Hale_Lunel93}. The latter set may fail however in approximating continuous functions \cite[Cor.~8.1, p. 222]{Hale_Lunel93} and can be even finite-dimensional \cite[p.~220]{Hale_Lunel93} which limits seriously its usage in practice\footnote{for the purpose of low-dimensional approximations.} if for instance snippets of solutions to Eq.~\eqref{eq:DDE_abs} are spanned by elements outside of $\mathcal{E}$.\footnote{See however \cite[Thm.~2.5.10]{curtain1995} for a sufficient condition for the set of (generalized) eigenfunctions to be dense in $\mathcal{H}$ still for the case when $\mathcal{A}$ does not contain distributed delay terms.} \item[(iii)] Due to the aforementioned limitations of the eigenfunctions, other basis functions are often used for the derivation of ODE systems to approximate the dynamics of the underlying DDE. Choices proposed in the literature include step functions \cite{Banks_al78,kappel1978autonomous}, splines \cite{banks1979spline,Banks_al84}, and orthogonal polynomial functions such as Legendre polynomials \cite{Kappel86,Ito_Teglas86}.\footnote{It is also worth mentioning the more recent works \cite{Vyasarayani12,Wahi_al05}, in which interesting approximation schemes based on linear and sine functions have been proposed for the case of state dependent delays, and for which successful numerical performances have been reported although rigorous convergence results seem still to be lacking, within this approach.} In most of the cases, a version of the Trotter-Kato theorem (see e.g.~\cite[Thm.~4.5, p.~88]{Pazy83}) is typically used to obtain finite-time uniform approximation results of the semigroup generated by $\mathcal{A}$. In the cases of step functions and splines, the conditions required in the Trotter-Kato theorem (see e.g.~Conditions {\bf (A1)} and {\bf (A2)} in Theorem~\ref{ParisVI_thm} below) have been analyzed in \cite{Banks_al78} and \cite{banks1979spline}. For the case of Legendre polynomials, technical complications have been encountered to check these conditions either in the setting of Galerkin approximation \cite{Kappel86} or in the setting of tau-method \cite{Ito_Teglas86} largely due to the fact that the basis functions do not live in the domain of $\mathcal{A}$. As noted in \cite[p.~168]{Kappel86} or in \cite[Sect.~5]{Ito_Teglas86}, either $X_N\not\subset D(\mathcal{A})$ or $\Pi_N$ is not orthogonal for the polynomial functions considered in \cite{Kappel86} and \cite{Ito_Teglas86}, respectively. On the other hand, at a given precision, the use of polynomial basis leads typically to ODE approximations with lower dimensions when compared with those built out of step functions or splines \cite{banks1979spline,Ito_Teglas86}. \qed \end{itemize} \end{rem} The problems discussed in (iii) of the above remark already encountered in the linear case, have limited the applications of polynomial basis for the approximation of nonlinear systems of DDEs. The above discussion leads naturally to the question whether there exists an orthogonal polynomial basis for which standard approximation results for abstract nonlinear systems such as recalled in Theorem \ref{ParisVI_thm} below, could be applied to the case of nonlinear systems of DDEs. The next section introduces orthogonal polynomials that will allow us to answer this question by the affirmative, leading to direct and explicit formulas for the rigorous Galerkin approximations of a broad class of nonlinear systems of DDEs; see Sections~\ref{Subsect_DDE_Galerkin} and \ref{Sec_Galerkin_analytic}. As explained next, the key is to seek for polynomials that live in the domain of $\mathcal{A}$, which is achieved here by seeking for polynomials to be orthogonal for the inner product \eqref{H_inner} with a point mass. \section{Orthogonal polynomials for inner products with a point mass} \label{sect:basis} The inner product given in \eqref{H_inner} is naturally associated with the measure \begin{equation}\label{Eq_dx+point-mass} \nu(\mathrm{d} \theta)=\mathrm{d} \theta +\delta_0, \end{equation} where $\delta_0$ denotes the Dirac measure concentrated at $\theta=0$. Orthogonal polynomials with respect to the Lebesgue measure $\mathrm{d} \theta$ or measures having a smooth density with respect to it, has a long history \cite{Szego75}. The study of orthogonal polynomials with respect to a measure including a point mass such as given by \eqref{Eq_dx+point-mass} has been studied only lately \cite[Chap.~2.9]{Ismail05}. It was in particular noticed that orthogonal polynomials with respect to such a measure can be expressed in terms of polynomials orthogonal with respect to the smooth part of the measure; see \cite{Uvarov69} for an early contribution on the topic. Koornwinder in \cite{Koo84} dealt with the case of orthogonal polynomials on $[-1,1]$ associated with measures given by \begin{equation}\label{Eq_Koorn_dx} \nu(\mathrm{d} x)= \frac{\Gamma(\alpha+\beta+2)}{2^{\alpha+\beta+1}\Gamma(\alpha+1)\Gamma(\beta+1)}(1-x)^\alpha (1+x)^\beta \mathrm{d} x + M \delta_{-1}+ N\delta_{1}, \; \alpha, \beta > -1, \end{equation} i.e.~associated with measures having a Jacobi weight on $[-1,1]$ with two point-masses added to the extremities of the interval. Although many properties --- such as three-term recurrence relationships or differential equations satisfied by such polynomials --- remain valid in the case of a measure with a point mass, subtle but important qualitative and quantitative differences arise. For instance, \cite[Thm.~ 3 c)]{Alfaro_al97} shows that the zeroes closest to $1$ of polynomials orthogonal with respect to the measure $\nu$ given in \eqref{Eq_Koorn_dx} converge to $1$ faster than those associated with the standard Jacobi polynomials. It is our goal to show that orthogonal polynomials with respect to the measure $\nu$ in \eqref{Eq_Koorn_dx} allows us to work within a simple and more direct framework than those found in the literature, for building Galerkin approximations of DDEs. Indeed, the approximation of DDEs by systems of ODEs built from orthogonal polynomials were not, so far, relying on classical Galerkin schemes as noted in \cite[p.~168]{Kappel86} or in \cite[Sect.~5]{Ito_Teglas86}. The results given below correspond to the case $\alpha = \beta = M=0$, and $N = 1$ considered in \cite{Koo84}. \subsection{Koornwinder polynomials} We recall next from \cite[Eq.~(2.1)]{Koo84} the following sequence of Koornwinder polynomials $\{K_n\}$ that can be built from the Legendre polynomials $L_n$ according to \begin{equation} \label{eq:Pn} K_n(s) := -(1+s)\frac{\mathrm{d}}{\mathrm{d} s} L_n(s) +( n^2 + n + 1) L_n(s), \; s \in [-1, 1], \; n \in \mathbb{N}. \end{equation} As recalled above, this polynomial sequence is known to be orthogonal when a Dirac point-mass at the right endpoint, $\delta_{1}$, is adjoined to the Lebesgue measure \cite{Koo84}, in other words \begin{equation} \begin{aligned} \int_{-1}^{1} K_n(s) K_m(s) \mathrm{d} \mu (s)& = \frac{1}{2} \int_{-1}^{1} K_n(s) K_m(s) \mathrm{d} x + K_n(1) K_m(1)\\ &=0, \, \mbox{ if } m\neq n. \end{aligned} \end{equation} This orthogonality property and the main properties satisfied by $\{K_n\}$ on which we will rely on, are summarized from \cite{Koo84} in the proposition below. \begin{prop} \label{thm:Pn} The polynomial $K_n$ defined in \eqref{eq:Pn} {\color{black} is of degree $n$} and admits the following expansion in terms of the Legendre polynomials: \begin{equation} \label{eq:Pn2} K_n(s) = - \sum_{j = 0}^{n-1} (2j+1)L_j(s) + (n^2 + 1) L_n(s), \qquad n \in \mathbb{N}; \end{equation} and the following normalization property holds: \begin{equation} \label{eq:Pn_normalization} K_n(1) = 1, \qquad n \in \mathbb{N}. \end{equation} Moreover, the sequence given by \begin{equation} \label{eq:Pn_prod} \{\mathcal{K}_n := (K_n, K_n(1)) : n \in \mathbb{N}\} \end{equation} forms an orthogonal basis of the product space \begin{equation} \label{eq:E} \mathcal{E} := L^2([-1,1); \mathbb{R}) \times \mathbb{R}, \end{equation} where $\mathcal{E}$ is {\color{black} endowed} with the following inner product: \begin{equation} \label{eq:inner_E} \langle (f, a), (g, b) \rangle_{\mathcal{E}} = \frac{1}{2} \int_{-1}^1 f(s)g(s) \mathrm{d} s + ab, \quad (f,a), (g, b) \in \mathcal{E}. \end{equation} {\color{black} Moreover $\Big\{\frac{\mathcal{K}_n}{\|\mathcal{K}_n\|_{\mathcal{E}}}\Big\}$ forms a Hilbert basis of $\mathcal{E}$ where the norm $\|\mathcal{K}_n\|_{\mathcal{E}}$ of $\mathcal{K}_n$ induced by $\langle \cdot, \cdot \rangle_{\mathcal{E}}$ possesses the following analytic expression:} \begin{equation} \label{eq:Pn_norm} \|\mathcal{K}_n\|_{\mathcal{E}} = \sqrt{\frac{(n^2+1)((n+1)^2+1)}{2n+1}}, \qquad n \in \mathbb{N}. \end{equation} \end{prop} \begin{proof} Based on \eqref{eq:Pn}, the proof consists essentially of algebraic manipulations relying on the following standard properties of the Legendre polynomials \cite[Sect.~3.3]{Shen_al11}: \begin{itemize} \item Orthogonality: \begin{equation} \label{eq:Ln_orth} \int_{-1}^1 L_m(s) L_n(s) \mathrm{d} x = \frac{2}{2n+1} \delta_{mn}, \qquad m,n \in \mathbb{N}, \end{equation} where $\delta_{mn}$ denotes the Kronecker delta. \medskip \item Normalization: \begin{equation} \label{eq:Ln_normalization} L_n(1) = 1, \qquad n \in \mathbb{N}. \end{equation} \item Three-term recurrence relation: \begin{equation} \label{eq:Ln_recur} (n+1) L_{n+1}(s) = (2n+1) s L_n(s) - n L_{n-1}(s), \; s\in [-1,1], \; n \ge 1, \end{equation} {\color{black} where} the first two Legendre polynomials are {\color{black} given by} \begin{equation} L_0 \equiv 1 \qquad \text{and} \qquad L_1(s) = s. \end{equation} \item First order derivative recurrence relation: \begin{equation} \label{eq:dLn} \frac{\mathrm{d} L_n}{\mathrm{d} s}(s) = \sum_{k \in I_n} (2k+1)L_k(s), \end{equation} where \begin{equation} \label{eq:idx_n} I_n:=\{k \in \mathbb{N} : 0\le k \le n-1, k+n \text{ is odd}\}. \end{equation} \end{itemize} Standard density arguments, outlined in Appendix~\ref{sect:thm_Pn_proof}, allow us then to conclude the proof. \end{proof} \subsection{Rescaled Koornwinder basis} \label{Sect_rescaled_basis} From the original Koornwinder basis given on the interval $[-1, 1]$, orthogonal polynomials on the interval $[-\tau, 0]$ for the inner product \eqref{H_inner} can now be easily obtained by using a simple linear transformation $\mathcal{T}$ defined by: \begin{equation} \label{eq:linear_transf} \mathcal{T} \colon [-\tau, 0] \rightarrow [-1, 1], \qquad \theta \mapsto 1 + \frac{2 \theta }{\tau}. \end{equation} Indeed, for $K_n$ given in \eqref{eq:Pn2}, let us define the polynomial $K_n^\tau$ by \begin{equation} \begin{aligned} \label{eq:Pn_tilde} K^\tau_n\colon [-\tau, 0] & \rightarrow \mathbb{R}, \\ \theta & \mapsto K_n \Bigl( 1 + \frac{2 \theta }{\tau} \Bigr), \qquad n \in \mathbb{N}. \end{aligned} \end{equation} Since the sequence $\{\mathcal{K}_n = (K_n, K_n(1)) : n \in \mathbb{N}\}$ forms an orthogonal basis for $\mathcal{E}$ (cf.~Proposition~\ref{thm:Pn}), it follows then that the polynomial sequence \begin{equation} \label{eq:Pn_tilde_prod} \{\mathcal{K}_n^\tau := (K_n^\tau, K_n^\tau(0)) : n \in \mathbb{N}\} \end{equation} forms an orthogonal basis for the space $\mathcal{H} = L^2([-\tau,0); \mathbb{R}) \times \mathbb{R}$ endowed with the inner product $\langle \cdot, \cdot \rangle_{\mathcal{H}}$ given in \eqref{H_inner} for $d=1$. Since $K_n(1)=1$ from \eqref{eq:Pn_normalization}, we have \begin{equation}\label{Eq_normalization} K_n^\tau(0) = 1. \end{equation} Moreover, by applying the transformation $\mathcal{T}$, we get trivially that \begin{equation}\label{Eq_inv_innerproduct} \|\mathcal{K}_n^\tau \|_{\mathcal{H}} = \| \mathcal{K}_n \|_{\mathcal{E}}. \end{equation} We have then the following fundamental lemma. \begin{lem}\label{Fundamental_lemma} The rescaled Koornwinder polynomials $\{K_j^{\tau}\}_{j\geq 0}$ satisfy the following properties: \begin{equation} \label{Eq_identity1_0} \boxed{ \sum_{j=0}^{\infty} \frac{K^\tau_j}{\|\mathcal{K}_j^\tau \|_{\mathcal{H}}^2}=0, \quad \text{ in the $L^2$ sense},} \end{equation} and \begin{equation} \label{Eq_identity2_0} \boxed{ \sum_{j=0}^{\infty} \frac{1}{\|\mathcal{K}_j^\tau \|_{\mathcal{H}}^2}=1.} \end{equation} Moreover, each function in $L^2([-\tau, 0]; \mathbb{R})$ enjoys the following decomposition in terms of the Koornwinder polynomials $K_j^\tau$: \begin{equation} \label{L2_decomp_0} \boxed{ f = \sum_{j=0}^\infty \frac{\langle f, K_j^\tau \rangle_{L^2} }{\tau \|\mathcal{K}_j^\tau \|_{\mathcal{H}}^2 } K_j^\tau,\qquad \text{ } \forall \: f \in L^2([-\tau, 0]; \mathbb{R}).} \end{equation} \end{lem} \begin{proof} For any $\Psi=(\Psi^D,\Psi^S) \in \mathcal{H}$, we have\footnote{Note that the equality in \eqref{Eq_decomp} holds in the sense that $ \Big \|\Psi - \sum_{j=0}^{\infty} \frac{\langle \Psi,\mathcal{K}^\tau_j \rangle_{\mathcal{H}}}{\|\mathcal{K}^\tau_j\|_{\mathcal{H}}^2}\mathcal{K}_j^\tau \Big\|_\mathcal{H}$ = 0, which is equivalent to $\Big \|\Psi^D - \sum_{j=0}^{\infty} \frac{\langle \Psi,\mathcal{K}^\tau_j \rangle_{\mathcal{H}}}{\|\mathcal{K}^\tau_j\|_{\mathcal{H}}^2} K_j^\tau \Big\|_{L^2} = 0$ and $\Big |\Psi^S - \sum_{j=0}^{\infty} \frac{\langle \Psi,\mathcal{K}^\tau_j \rangle_{\mathcal{H}}}{\|\mathcal{K}^\tau_j\|_{\mathcal{H}}^2}\Big| = 0$.} \begin{equation} \begin{aligned}\label{Eq_decomp} \Psi &=\sum_{j=0}^{\infty} \frac{\langle \Psi,\mathcal{K}^\tau_j \rangle_{\mathcal{H}}}{\|\mathcal{K}^\tau_j\|_{\mathcal{H}}^2}\mathcal{K}_j^\tau\\ & =\sum_{j=0}^{\infty} \Big(\frac{1}{\tau}\langle \Psi^D,K_j^\tau \rangle_{L^2}+\Psi^S K_j^\tau(0)\Big) \frac{\mathcal{K}_j^\tau }{\|\mathcal{K}_j^\tau \|_{\mathcal{H}}^2}. \end{aligned} \end{equation} Now, let $\Psi^D$ to be the zero-function on $[-\tau,0]$ and $\Psi^S$ to be 1. For such a $\Psi$, by equalizing respectively the $D$-components and $S$-components of the RHS and LHS of \eqref{Eq_decomp}, one then obtains from \eqref{Eq_normalization} that \begin{equation} \sum_{j=0}^{\infty} \frac{K^\tau_j}{\|\mathcal{K}_j^\tau \|_{\mathcal{H}}^2}=0, \quad \text{ in the $L^2$ sense}, \end{equation} and \begin{equation} \sum_{j=0}^{\infty} \frac{1}{\|\mathcal{K}_j^\tau \|_{\mathcal{H}}^2}=1. \end{equation} The decomposition of $L^2$ functions given in \eqref{L2_decomp_0} follows directly from \eqref{Eq_decomp} by considering $\Psi := (f, 0) \in \mathcal{H}$. Again, the equality holds in the $L^2$ sense. \end{proof} As an illustration of the identities \eqref{Eq_identity1_0} and \eqref{Eq_identity2_0}, Figure~\ref{fig:Cancel_prop} displays numerical computations of the partial sum $\sum_{j=0}^{N-1} \frac{K^\tau_j}{\|\mathcal{K}_j^\tau \|_{\mathcal{H}}^2}$ for $N=20$ and $N=60$, in the case $\tau = 0.5$. \begin{figure}[hbtp] \centering \includegraphics[height=0.4\textwidth,width=.75\textwidth]{Cancel_prop.pdf} \caption{{\footnotesize Sum of the first $N$ rescaled Koornwinder polynomials: blue curve corresponds to $N=20$, and red curve to $N=60$.}} \label{fig:Cancel_prop} \end{figure} \needspace{1\baselineskip} \begin{rem} \label{Rmk_KoornwinderBasis} \hspace*{2em} \vspace*{-0.4em} \begin{itemize} \item[(i)] Note that the continuity condition, $\lim_{\theta \rightarrow 0^-} \Psi^D(\theta) = \Psi^S $, required in \eqref{D_of_A} in order that $\Psi \in D(\mathcal{A})$, is here naturally satisfied by the Koornwinder basis function $\mathcal{K}_n^\tau = (K_n^\tau, K_n^\tau(0))$. It constitutes thus, for the inner product \eqref{H_inner}, an orthogonal polynomial basis whose elements live in the domain of the linear operator $\mathcal{A}$ given in \eqref{Def_A}. As explained at the beginning of Section~\ref{sect:basis}, the key element for such a construction relies on the inclusion of a point mass adjoined to the continuous part of the measure. When this point mass is absent, the corresponding orthogonal polynomials are (rescaled) Legendre polynomials. The associated basis in this latter case is given by (cf.~\cite{Kappel86}) \begin{equation} \label{Legendre-based basis} \mathfrak{B} := \{ \psi_1:= (0_{\mathcal{H}}, 1) \} \cup \{\psi_n:= (L^\tau_{n-2}, 0) \ \vert \ n=2, 3, \cdots \}, \end{equation} where $0_\mathcal{H}$ denotes the zero function on $\mathcal{H}$, and $L^\tau_n$ is the Legendre polynomial of degree $n$ defined on the interval $[-\tau, 0]$. Clearly, none of the elements in $\mathfrak{B}$ belongs to $D(\mathcal{A})$ since $\lim_{\theta \rightarrow 0^-} L^\tau_n(\theta) \neq 0$. \item[(ii)] The fact that the Koornwinder basis functions live in $D(\mathcal{A})$ allows us to construct {\it standard} Galerkin approximations; whereas, extra correction terms are required in the Galerkin approximation built from the Legendre-based basis given in \eqref{Legendre-based basis} (see e.g.~\cite[p.~169]{Kappel86}). Moreover, technical complications such as pointed out in Remark~\ref{Rmk_problems_to_overcome} iii) do not take place for the case of Koornwinder basis. Indeed, it will be shown in Section~\ref{Subsect_DDE_Galerkin} that the properties of the Koornwinder polynomials such as summarized in Corollary~\ref{Fundamental_lemma} as well as the vectorized version given by Corollary~\ref{Super_Fundamental_lemma} below, turn out to be sufficient to obtain finite-time uniform approximation results of the semigroup generated by the linear operator $\mathcal{A}$; see Lemmas~\ref{Lem_A2} and \ref{Lem_A1}. \item[(iii)] It is also worth mentioning that Corollary~\ref{Fundamental_lemma} and Corollary~\ref{Super_Fundamental_lemma} as well as the rigorous convergence results presented in Section~\ref{Subsect_DDE_Galerkin} are not limited to the case of Koornwinder basis constructed here. Given any polynomial basis on $[-\tau, 0]$ that are orthogonal with respect to a measure of the form $\nu(\mathrm{d} \theta) = \mathrm{d} \rho(\theta) + \delta_0$ with $\rho$ being a positive non-decreasing function on $[-\tau, 0]$, the aforementioned results would still hold. We refer to \cite{Uvarov69} for the construction of such polynomials when orthogonal polynomials with respect to $\widetilde{\nu}(\mathrm{d} \theta) = \mathrm{d} \rho(\theta)$ are known. \qed \end{itemize} \end{rem} \subsection{Vectorization of Koornwinder polynomials}\label{Sec_Vectorization} We introduce here a generalization of the Koornwinder polynomials that will turn out to be useful for the approximation of nonlinear DDE systems. The purpose is here to build from the Koornwinder polynomials introduced above, linear subspaces $\mathcal{H}_{N}$ that approximate $\mathcal{H} := L^2([-\tau,0); \mathbb{R}^d) \times \mathbb{R}^d$, for $d>1$. Each function $\Psi$ in $\mathcal{H}$ has here $d$-components that can be, each, approximated by a series of Koornwinder polynomials as in \eqref{Eq_decomp}. If we restrict such an approximation to the first $N$ Koornwinder polynomials, $\mathcal{H}_{N}$ becomes then an $N\times d$-dimensional subspace; see \eqref{subspace_HNd} below. Our goal is also here to introduce a vectorization of Koornwinder polynomials which allows for a natural extension of Lemma \ref{Fundamental_lemma} in the case $d>1$. This extension of Lemma \ref{Fundamental_lemma} will be particularly useful to provide finite-dimensional approximation of the linear part of systems of DDEs; see Lemma \ref{Lem_A2}. To do so, given $j\in\{1,\cdots,Nd\}$, we can associate a Koornwinder polynomial of degree $j_q \in\{0,\cdots,N-1\}$, as follows \begin{equation}\label{index_relation} j=d j_q +j_r, \end{equation} where $j_r \in \{1,\cdots,d\}$ is given by \begin{equation}\label{index_relationb} j_r := \begin{cases} \mathrm{mod}(j, d), & \text{if } \mathrm{mod}(j, d) \neq 0, \\ d, & \text{otherwise.} \end{cases} \end{equation} Let us introduce now the following $d$-dimensional mapping from $[-\tau,0]$ to $\mathbb{R}^d$ \begin{equation} \label{vectorized_K} \mathbf{K}^\tau_{j}(\theta):= (\underbrace{0, \cdots, 0}_{\text{$j_r-1$ entries}}, K^\tau_{j_q}(\theta), \underbrace{0, \cdots, 0}_{\text{$d-j_r$ entries}})^\mathrm{tr}, \qquad \theta \in [-\tau,0]. \end{equation} The vector $\mathbf{K}^\tau_{j}(\theta)$ is then nothing else than a $d$-dimensional canonical vector whose $j_r^{th}$-entry is given by the value at $\theta$ of the (rescaled) Koornwinder polynomial of degree $j_q$; the integers $j_q$ and $j_r$ being related to $j$ according to \eqref{index_relation}-\eqref{index_relationb}. Based on these vectorized (rescaled) Koornwinder polynomials $\mathbf{K}^\tau_{j}$, we also introduce \begin{equation}\label{Eq_superKoor} \mathbb{K}^\tau_{j}:= \big( \mathbf{K}^\tau_{j}, \mathbf{K}^\tau_{j}(0) \big), \qquad j \ge 1. \end{equation} In the remaining part of this section, we summarize some key properties of $\mathbf{K}^\tau_{j}$ and $\mathbb{K}^\tau_{j}$ for later usage. Hereafter, we use $\mathcal{H}_1$ to denote the space $\mathcal{H}$ defined in \eqref{H_space} for the case $d=1$, i.e., \begin{equation} \mathcal{H}_1 = L^2([-\tau,0); \mathbb{R}) \times \mathbb{R}, \end{equation} which is again endowed with the inner product given in \eqref{H_inner} (still with $d=1$). Since the sequence $\{\mathcal{K}^\tau_j = (K_j^\tau, K_j^\tau(0)) : j \in \mathbb{N}\}$ forms an orthogonal basis for $\mathcal{H}_1$ (cf.~Section~\ref{Sect_rescaled_basis}), one can readily check that $\{\mathbb{K}^\tau_j : j \in \mathbb{N}^*\}$ forms an orthogonal basis for the space $\mathcal{H}$. Note also that \begin{equation}\label{Eq_norm_superKoor} \|\mathbb{K}^\tau_{j}\|_\mathcal{H} = \|\mathcal{K}^\tau_{j_q}\|_{\mathcal{H}_1}, \qquad j \in \mathbb{N}^*. \end{equation} Given this vectorization of Koornwinder polynomials, we can now formulate the following extension of Lemma \ref{Fundamental_lemma} that summarizes the key properties of the $\mathbf{K}_j$'s which will be used for the rigorous approximation of nonlinear systems of DDEs such as described in Section \ref{Subsect_DDE_Galerkin}. \begin{lem}\label{Super_Fundamental_lemma} The vectorized rescaled Koornwinder polynomials $\{\mathbf{K}^\tau_j \}_{j\geq 1}$ satisfy the following properties: \begin{equation} \label{Eq_identity1} \boxed{ \sum_{j=1}^{\infty} \frac{ \langle \alpha, \mathbf{K}^\tau_j(0) \rangle }{\|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2} \mathbf{K}^\tau_j =0 \quad \text{ in the $L^2([-\tau,0); \mathbb{R}^d)$ sense}, \quad \text{ } \forall \: \alpha \in \mathbb{R}^d,} \end{equation} and \begin{equation} \label{Eq_identity2} \boxed{ \sum_{j=1}^{\infty} \frac{ \langle \alpha, \mathbf{K}^\tau_j(0) \rangle }{\|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2} \mathbf{K}^\tau_j(0)= \alpha, \quad \text{ } \forall \: \alpha \in \mathbb{R}^d.} \end{equation} Moreover, each function in $L^2([-\tau, 0]; \mathbb{R}^d)$ enjoys the following decomposition in terms of the vectorized Koornwinder polynomials $\mathbf{K}_j^\tau$: \begin{equation} \label{L2_decomp} \boxed{ f = \sum_{j=1}^\infty \frac{\langle f,\mathbf{K}_j^\tau \rangle_{L^2} }{\tau \|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2 } \mathbf{K}_j^\tau,\qquad \text{ } \forall \: f \in L^2([-\tau, 0]; \mathbb{R}^d);} \end{equation} and the following identity holds: \begin{equation} \label{Eq_identity3} \boxed{ \sum_{j=1}^\infty \frac{\langle f,\mathbf{K}_j^\tau \rangle_{L^2} }{\tau \|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2 } \mathbf{K}_j^\tau(0) = 0,\qquad \text{ } \forall \: f \in L^2([-\tau, 0]; \mathbb{R}^d).} \end{equation} \end{lem} \begin{proof} The above identities can be obtained by using the same type of reasoning as given in the proof of Lemma~\ref{Fundamental_lemma} for the scalar case. Indeed, by noting that $\{\mathbb{K}^\tau_j \,:\, j \in \mathbb{N}^*\}$ forms an orthogonal basis of $\mathcal{H}$, any $\Psi \in \mathcal{H}$ admits the following decomposition: \begin{equation} \begin{aligned}\label{Eq_decomp_vec} \Psi &=\sum_{j=1}^{\infty} \frac{\langle \Psi,\mathbb{K}^\tau_j \rangle_{\mathcal{H}}}{\|\mathbb{K}^\tau_j\|_{\mathcal{H}}^2}\mathbb{K}_j^\tau\\ & =\sum_{j=1}^{\infty} \Big(\frac{1}{\tau}\langle \Psi^D,\mathbf{K}_j^\tau \rangle_{L^2}+ \langle \Psi^S, \mathbf{K}_j^\tau(0) \rangle \Big) \frac{\mathbb{K}_j^\tau}{\|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2}, \end{aligned} \end{equation} where we have used the identity \eqref{Eq_norm_superKoor} in the last equality above. Now, let $\Psi^D \in L^2([-\tau, 0]; \mathbb{R}^d)$ to be the zero-function and $\Psi^S$ to be an arbitrary vector $\alpha \in \mathbb{R}^d$. For such a $\Psi$, by equalizing respectively the $D$-components and $S$-components of the RHS and LHS of \eqref{Eq_decomp_vec}, we obtain respectively \eqref{Eq_identity1} and \eqref{Eq_identity2}. The identities \eqref{L2_decomp} and \eqref{Eq_identity3} also follow directly from \eqref{Eq_decomp_vec} by considering $\Psi^D = f$ and $\Psi^S = 0 \in \mathbb{R}^d$. \end{proof} \section{Galerkin approximation: Rigorous results}\label{Sec_Galerkin_approx} In this section, we establish the convergence of the Galerkin scheme based on the rescaled and vectorized Koornwinder polynomials of Section~\ref{Sec_Vectorization}. These convergence results apply, as we shall see, to a broad class of nonlinear systems of DDEs. As mentioned in the Introduction and in Remark \ref{Rmk_problems_to_overcome}-(iii), the advantage of the Koornwinder basis relies on the facts that the constitutive basis functions are orthogonal and belong each to the domain of the linear operator associated with a given DDE. In particular, there is no discontinuity at the right end point for each basis element, by construction; see Section \ref{sect:basis}. Thanks to these properties of the basis functions, convergence results for the associated Galerkin systems can be derived in a straightforward fashion (see Corollary \ref{Cor_DDE_global_Lip}) and under useful criteria on the nonlinear terms (see Corollaries \ref{Cor_DDE_local_Lip_Case1} and \ref{Cor_DDE_local_Lip_Case2}), compared to other Galerkin schemes built from other bases; see, e.g., \cite{Kappel86,Kappel_al87, Vyasarayani12,Wahi_al05} and references therein. In the following, we first present in Section \ref{Subsect_ODE_Galerkin} a general convergence result for Galerkin approximations of abstract nonlinear ODEs in Hilbert spaces by relying essentially on the theory of semigroups and the Trotter-Kato theorem \cite[Thm.~4.5, p.~88]{Pazy83}. The result is then applied to the DDE context in Section \ref{Subsect_DDE_Galerkin}. General examples are provided in Section \ref{Sec_examples}. \subsection{Galerkin approximations of nonlinear ODEs in Hilbert spaces} \label{Subsect_ODE_Galerkin} We first present a general convergence result for Galerkin approximations of abstract nonlinear differential equations in a Hilbert space $X$, endowed with a norm $\|\cdot\|_X$. The mathematical setting is somewhat classical but we recall it below for the reader's convenience and later use. In that respect, we assume in this Section the linear operator $\mathcal{L}$ to be the infinitesimal generator of a $C_0$-semigroup of bounded linear operators $T(t)$ on $X$. Recall that in that case the domain $D(\mathcal{L})$ of $\mathcal{L}$ is dense in $X$ and that $\mathcal{L}$ is a closed operator; see \cite[Cor.~2.5, p.~5]{Pazy83}. Under these assumptions, recall that there exists $M\geq 1$ and $\omega \geq 0$ \cite[Thm.~2.2, p.~4]{Pazy83} such that \begin{equation}\label{Eq_control_T_t} \|T(t)\| \le M e^{\omega t}, \qquad t \ge 0, \end{equation} where $\|\cdot \|$ denotes the operator norm subordinated to $\|\cdot\|_X$. We are concerned with finite-dimensional approximations of the following initial-value problem: \begin{equation} \begin{aligned} \label{ODE} \frac{\mathrm{d} u}{\mathrm{d} t} &= \mathcal{L} u + \mathcal{G}(u), \\ u(0) &= u_0, \end{aligned} \end{equation} where $u_0 \in X$. A {\it mild solution} of \eqref{ODE} over $[0,T]$, will be any function $u\in C([0,T],X)$ such that for $u_0\in X$, \begin{equation}\label{Eq_mild} u(t)=T(t)u_0 + \int_0^t T(t-s) \mathcal{G}(u(s)) \mathrm{d} s. \end{equation} Let $\{X_N: N \in \mathbb{N}\}$ be a sequence of subspaces of $X$ associated with {\it orthogonal projectors} \begin{equation} \Pi_N: X \rightarrow X_N, \end{equation} such that \begin{equation}\label{Eq_identity_approx} \|\Pi_N-\mbox{Id}_X\|\underset{N\rightarrow \infty}\longrightarrow 0, \end{equation} and \begin{equation}\label{Eq_XN_in_domain} X_N\subset D(\mathcal{L}), \; \forall \, N. \end{equation} The corresponding Galerkin approximation of \eqref{ODE} associated with $X_N$ is then given by: \begin{equation} \begin{aligned} \label{ODE_Galerkin} \frac{\mathrm{d} u_N}{\mathrm{d} t} &= \mathcal{L}_N u_N + \Pi_N \mathcal{G}(u_N), \\ u_N(0) &= \Pi_N u_0, \; u_0\in X, \end{aligned} \end{equation} where \begin{equation}\label{Def_LN} \mathcal{L}_N := \Pi_N \mathcal{L} \Pi_N : X \rightarrow X_N. \end{equation} In particular, the domain $D(\mathcal{L}_N)$ of $\mathcal{L}_N$ is $X$, because of \eqref{Eq_XN_in_domain}. As we will see in Section \ref{Subsect_DDE_Galerkin}, the choice of vectorized Koornwinder polynomials as a basis function will allow us to define subspaces $X_N$ naturally associated with orthogonal projectors $\Pi_N$ that satisfy the above properties in contrast to other polynomial functions used for (non-standard) Galerkin approximation or other Legendre-tau approximations of systems of DDEs used so far; see e.g.~\cite{Ito_Teglas86,Kappel86}. See also Remark~\ref{Rmk_problems_to_overcome}-iii). These nice properties will allow us also to rely on the following general convergence result regarding {\it standard} Galerkin schemes, for the case of nonlinear systems of DDEs; see Section \ref{Subsect_DDE_Galerkin} below. \vspace{1ex} \begin{thm} \label{ParisVI_thm} Let $\mathcal{L}$ and $\{X_N\}_{N\geq 0}$ be as described above. Assume furthermore the following set of assumptions: \begin{itemize} \item[{\bf (A1)}] For each $N\in \mathbb{N}$, the linear flow $e^{\mathcal{L}_N t}:X_N \rightarrow X_N$ extends to a $C_0$-semigroup $T_N(t)$ on $X$. Furthermore the following uniform bound is satisfied by the family $\{T_N(t)\}_{N\geq 0, t\geq0}$ \begin{equation} \label{Eq_control_linearflow} \quad \|T_N(t)\| \le M e^{\omega t}, \quad N\geq 0, \; \quad t \ge 0, \end{equation} where $\|T_N(t)\|=\sup\{\|T_N(t)x\|_X, \; \|x\|_X\leq 1, x\in X\}$. \item[{\bf (A2)}] The following convergence holds \begin{equation} \label{Eq_L_Approx} \lim_{N \rightarrow \infty} \|\mathcal{L}_N \phi - \mathcal{L} \phi \|_X = 0, \quad \text{ } \forall \: \phi \in D(\mathcal{L}). \end{equation} \item[{\bf (A3)}] $\mathcal{G}$ is globally Lipschitz. \end{itemize} Then for any $u_0 \in X$, there exists a unique mild solution of \eqref{ODE} and such a solution can be approximated uniformly on each bounded interval $[0, T]$ by the sequence $\{t\mapsto u_N(t; \Pi_N u_0)\}_{N\geq 0}$ of mild solutions obtained from \eqref{ODE_Galerkin}, i.e.: \begin{equation} \label{uniform_conv_est_ODE} \lim_{N\rightarrow \infty} \sup_{t \in [0, T]} \|u_N(t; \Pi_N u_0) - u(t; u_0)\|_X = 0, \qquad \text{ } \forall \: T > 0. \end{equation} \end{thm} \vspace{1ex} \begin{proof} Recall that the existence and uniqueness of solutions to Eq.~\eqref{Eq_mild} emanating from any initial data $u_0 \in X$ can be proved by a fixed point argument in $C([0,T],X)$ as in the proof of e.g.~\cite[Prop.~4.3.3]{Cazenave_al98}, by relaxing the semigroup of contractions requirement therein to the $C_0$-semigroup setting adopted here; see also \cite[Thm.~6.1.1]{Lunardi04}. Given $u_0\in X$, let $u$ be thus the unique mild solution of Eq.~\eqref{ODE}. By the variation-of-constants formula applied to Eq.~\eqref{ODE_Galerkin} we have on the other hand, for $0\leq t\leq T$, \begin{equation} \begin{aligned} u_N(t) = e^{\mathcal{L}_N t} \Pi_N u_0 + \int_0^t e^{\mathcal{L}_N (t -s )} \Pi_N \mathcal{G}(u_N(s)) \mathrm{d} s. \end{aligned} \end{equation} Then $v_N(t)=u(t) - u_N(t)$ satisfies \begin{equation} \begin{aligned} \label{eq:residual} v_N(t)&= T(t) u_0 - e^{\mathcal{L}_N t} \Pi_N u_0 + \int_0^t T(t-s) \mathcal{G}(u(s)) \mathrm{d} s - \int_0^t e^{\mathcal{L}_N (t -s )} \Pi_N \mathcal{G}(u_N(s)) \mathrm{d} s \\ & = T(t) u_0 - e^{\mathcal{L}_N t} \Pi_N u_0 + \int_0^t \big( T(t-s) - e^{\mathcal{L}_N (t -s )} \Pi_N \big) \mathcal{G}(u(s)) \mathrm{d} s \\ & \hspace{10em} + \int_0^t e^{\mathcal{L}_N (t -s )} \Pi_N \big( \mathcal{G}(u(s)) - \mathcal{G}(u_N(s)) \big ) \mathrm{d} s. \end{aligned} \end{equation} Let us introduce \begin{equation} \begin{aligned} r_N(s) & := \|u(s) - u_N(s) \|_X, \\ \epsilon_N(u_0) & := \sup_{t\in[0, T]} \|T(t) u_0 - e^{\mathcal{L}_N t} \Pi_N u_0\|_X, \\ d_N(s) & := \sup_{t\in[s, T]} \| \big( T(t-s) - e^{\mathcal{L}_N (t -s )} \Pi_N \big) \mathcal{G}(u(s)) \|_X. \end{aligned} \end{equation} We obtain then from \eqref{eq:residual} that \begin{equation} \begin{aligned} r_N(t) & \le \epsilon_N(u_0) + \int_0^t d_N(s) \mathrm{d} s + \int_0^t \|e^{\mathcal{L}_N (t -s )} \Pi_N \big( \mathcal{G}(u(s)) - \mathcal{G}(u_N(s)) \big )\|_X \mathrm{d} s \\ & \le \epsilon_N(u_0) + \int_0^t d_N(s) \mathrm{d} s + M \mathrm{Lip}(\mathcal{G}) \int_0^t e^{\omega (t -s )} r_N(s) \mathrm{d} s \\ & \le \epsilon_N(u_0) + \int_0^T d_N(s) \mathrm{d} s + M \mathrm{Lip}(\mathcal{G}) e^{\omega T} \int_0^t r_N(s) \mathrm{d} s, \end{aligned} \end{equation} where we have used the global Lipschitz condition on $\mathcal{G}$ and \eqref{Eq_control_linearflow} to derive the second inequality. It follows then from Gronwall's inequality that \begin{equation} \label{rN_est} r_N(t) \le \Big( \epsilon_N(u_0) + \int_0^T d_N(s) \mathrm{d} s \Big) e^{M \mathrm{Lip}(\mathcal{G}) T e^{\omega T} }, \qquad \text{ } \forall \: t \in [0, T]. \end{equation} We are thus left with the estimation of $\epsilon_N(u_0)$ and $\int_0^T d_N(s) \mathrm{d} s $ as $N \rightarrow \infty$. The assumptions {\bf (A1)}--{\bf (A2)} allow us to use a version of the Trotter-Kato theorem \cite[Thm.~4.5, p.88]{Pazy83}\footnote{Recall that because $\mathcal{L}$ is the generator of a $C_0$-semigroup $T(t)$ on $X$, it satisfies $\|T(t)\| \leq Me^{\omega t},$ and as a consequence the resolvent set of $\mathcal{L}$ contains the interval $]\omega,\infty[$; see \cite[Thm.~5.3 p.~20]{Pazy83}. In particular, for any $f\in \mathcal{H}$ and any $\lambda >\omega$, the equation $(\lambda I -\mathcal{L}) x=f$ possesses a unique solution $x \in D(\mathcal{L})$, which implies in particular that $(\lambda I -\mathcal{L}) D(\mathcal{L})$ is dense in $X$ as required by the version of the Trotter-Kato theorem used here. This explains why this density requirement, consequence of our working assumptions, is omitted in the formulation of Theorem \ref{ParisVI_thm}.} which implies together with \eqref{Eq_identity_approx} that \begin{equation} \lim_{N\rightarrow \infty} e^{\mathcal{L}_N t} \Pi_N \phi = T(t) \phi, \quad \text{ } \forall \: \phi \in X, \end{equation} uniformly for $t$ in bounded intervals. It follows that \begin{equation} \label{epsN_est} \lim_{N\rightarrow \infty} \epsilon_N(u_0) = 0, \quad \text{ } \forall \: u_0 \in X, \end{equation} and that $d_N$ converges point-wisely to zero on $[0,T]$, i.e. \begin{equation} \label{dN_est1} \lim_{N\rightarrow \infty} d_N(s) = 0, \quad \text{ } \forall \: s \in [0, T]. \end{equation} On the other hand, from \eqref{Eq_control_T_t}, \eqref{Eq_control_linearflow}, and {\bf (A3)}, we get \begin{equation} \begin{aligned} \| \big( T(t-s) - e^{\mathcal{L}_N (t -s )} \Pi_N \big) &\mathcal{G}(u(s)) \|_X \le 2 M e^{\omega (t-s)} \|\mathcal{G}(u(s)) \|_X \\ & \le 2 M e^{\omega (t-s)} \big(\mathrm{Lip}(\mathcal{G}) \|u(s)\|_X + \|\mathcal{G}(0)\|_X\big), \end{aligned} \end{equation} which implies \begin{equation} \label{dN_est2} d_N(s) \le 2 M e^{\omega T} \big(\mathrm{Lip}(\mathcal{G}) \|u(s)\|_X+ \|\mathcal{G}(0)\|_X\big), \quad \text{ } \forall \: s \in [0, T]. \end{equation} Since $u\in C([0,T],X)$, $s\mapsto \| u(s)\|_X$ is integrable on $[0,T]$, and the Lebesgue's dominated convergence theorem allows us then to conclude from \eqref{dN_est1} and \eqref{dN_est2} that \begin{equation} \label{dN_est3} \lim_{N\rightarrow \infty} \int_{0} ^T d_N(s) \mathrm{d} s = 0. \end{equation} The desired uniform convergence estimate \eqref{uniform_conv_est_ODE} is then trivially obtained from \eqref{rN_est}. \end{proof} \subsection{Galerkin approximations of nonlinear systems of DDEs}\label{Subsect_DDE_Galerkin} In this section, given the Hilbert product space \begin{equation*} \mathcal{H}:=L^2([-\tau, 0); \mathbb{R}^d)\times \mathbb{R}^d, \;\; d\geq 1, \end{equation*} endowed with the inner product \eqref{H_inner}, we restrict our attention to the following abstract ODE: \begin{equation} \begin{aligned} \label{Eq_abstract_ODE_DDE} \frac{\mathrm{d} u}{\mathrm{d} t} & = \mathcal{A} u + \mathcal{F}(u), \\ \end{aligned} \end{equation} where $\mathcal{F}$ is a nonlinear operator that will be specified later on, and where \textemdash\, given $L_D$, a bounded linear operator from $H^1([-\tau, 0); \mathbb{R}^d)$ to $ \mathbb{R}^d$ and, $L_S$, a bounded linear operator from $\mathbb{R}^d$ to $\mathbb{R}^d$ \textemdash\, the linear operator $\mathcal{A}$ is given by \begin{equation} \begin{aligned} \label{Def_A2} \lbrack \mathcal{A} \Psi \rbrack (\theta) & := \begin{cases} {\displaystyle \frac{\mathrm{d}^+ \Psi^D}{\mathrm{d} \theta}}, & \theta \in[-\tau, 0), \vspace{0.4em}\\ {\displaystyle L_S\Psi^S + L_D\Psi^D}, & \theta = 0, \end{cases} \end{aligned} \end{equation} for any $\Psi = (\Psi^D, \Psi^S)$ that lives in the domain, $D(\mathcal{A})$, defined as \begin{equation} \label{D_of_A2} D(\mathcal{A}): = \Big \{\Psi \in \mathcal{H} : \Psi^D \in H^1([-\tau, 0); \mathbb{R}^d), \lim_{\theta \rightarrow 0^-}\Psi^D(\theta) =\Psi^S \Big \}. \end{equation} Such an abstract setting arises naturally in the reformulation of a broad class of nonlinear systems of DDEs as an abstract ODE in $\mathcal{H}$; see e.g.~\cite{burns1983linear,curtain1995}. Examples of operators $L_D$ depending explicitly on the delay $\tau$ are given below; see \eqref{Def_LD}. It is well-known that under these assumptions, the operator $\mathcal{A}$ defines a $C_0$-semigroup on $\mathcal{H}$ \cite[Thm.~2.3]{burns1983linear}, and in particular $\mathcal{A}$ is dense in $\mathcal{H}$ and is a closed operator. We turn now to the definition of the subspaces $X_N$ and $\Pi_N$ of the previous section. For each positive integer $N$, we define the $Nd$-dimensional subspace $\mathcal{H}_{N} \subset \mathcal{H}$ to be spanned by the first $Nd$ vectorized Koornwinder polynomials introduced in \eqref{Eq_superKoor}, namely \begin{equation} \label{subspace_HNd} \mathcal{H}_{N} = \mathrm{span} \Big \{ \mathbb{K}^\tau_{1}, \cdots, \mathbb{K}^\tau_{Nd} \Big\}. \end{equation} As noted in Section \ref{Sec_Vectorization}, these polynomials are orthogonal for the inner product with a point mass such as defined in \eqref{H_inner}. The subspace $\mathcal{H}_{N}$ is thus naturally associated with an orthogonal projector $\Pi_N$, as required in the previous section. The approximation property \eqref{Eq_identity_approx} is satisfied due to the density arguments outlined in Appendix~\ref{sect:thm_Pn_proof}. Recall finally that by construction $\mathbb{K}^\tau_{j} \in D(\mathcal{A})$ for any $j\in \mathbb{N}^*$, and therefore \begin{equation}\label{Eq_inclusion} \mathcal{H}_{N}\subset D(\mathcal{A}). \end{equation} The corresponding $N$-dimensional Galerkin approximation of Eq.~\eqref{Eq_abstract_ODE_DDE} reads then: \begin{equation}\label{Eq_DDE_Galerkin} \frac{\mathrm{d} u_N}{\mathrm{d} t} = \mathcal{A}_N u_N + \Pi_N \mathcal{F}(u_N), \end{equation} with \begin{equation}\label{Eq_AN} \mathcal{A}_N := \Pi_N \mathcal{A} \Pi_N, \end{equation} which is therefore well defined on $\mathcal{H}$ because of \eqref{Eq_inclusion}. We are now in position to check Conditions {\bf (A1)} and {\bf (A2)} of Theorem \ref{ParisVI_thm}. To check Condition {\bf (A1)}, we will make usage of the following extension of the linear flow $e^{\mathcal{A}_N t}$: \begin{equation}\label{Eq_extension} T_N(t) u=e^{\mathcal{A}_N t} \Pi_N u +(I-\Pi_N) u, \; u\in \mathcal{H}. \end{equation} Such an extension leads naturally to a $C_0$-semigroup on $\mathcal{H}$. The stability condition \eqref{Eq_control_linearflow}, will require however some specifications of the operator $L_D$ that will be made clear later. Condition {\bf (A2)} can be however checked in the general setting by making an appropriate use of the properties of the vectorized Koornwinder polynomials summarized in Lemma \ref{Super_Fundamental_lemma}. More precisely, \begin{lem} \label{Lem_A2} Let $\mathcal{H}_N$ be the subspace defined in \eqref{subspace_HNd}. Then for $\mathcal{A}$ defined in \eqref{Def_A2} and $\mathcal{A}_N$ defined in \eqref{Eq_AN} associated with the orthogonal projector $\Pi_N$ onto $\mathcal{H}_N$, we have \begin{equation} \lim_{N \rightarrow \infty} \|\mathcal{A}_{N} \Psi - \mathcal{A} \Psi \|_\mathcal{H} = 0, \qquad \text{ } \forall \: \Psi \in D(\mathcal{A}). \end{equation} \end{lem} \begin{proof} By construction $\mathbb{K}^\tau_{j} \in D(\mathcal{A})$ for each $j\in \mathbb{N}^*$. Since $\big\{ \frac{\mathbb{K}^\tau_{j}}{\|\mathbb{K}^\tau_{j}\|_\mathcal{H}} : j \in \mathbb{N}^*\big\}$ forms a Hilbert basis of $\mathcal{H}$, it suffices to show that \begin{equation} \label{Goal_convergence} \lim_{N \rightarrow \infty} \|\mathcal{A}_{N} \Psi - \mathcal{A} \Psi \|_\mathcal{H} = 0, \;\; \Psi \in \underset{k\geq 1}\bigcup \mathcal{H}_k. \end{equation} We recall from \eqref{Eq_norm_superKoor} that $\|\mathbb{K}^\tau_{j}\|_{ \mathcal{H}} = \|\mathcal{K}^\tau_{j_q}\|_{\mathcal{H}_1}$ for all $j \in \mathbb{N}$. It follows then that the orthogonal projector $\Pi_{N}$ associated with the subspace $\mathcal{H}_{N}$ takes the following explicit form: \begin{equation} \begin{aligned} \label{DDE_projector} \Pi_{N} \Psi & = \sum_{j = 1}^{Nd} \frac{ \big \langle \Psi, \mathbb{K}^\tau_{j} \big \rangle_{\mathcal{H}}}{\|\mathcal{K}^\tau_{j_q}\|_{\mathcal{H}_1}^2} \mathbb{K}^\tau_{j} \\ & = \sum_{j = 1}^{Nd} \Big(\frac{1}{\tau}\langle \Psi^D, \mathbf{K}_j^\tau \rangle_{L^2}+ \langle \Psi^S, \mathbf{K}_j^\tau(0) \rangle \Big) \frac{\mathbb{K}_j^\tau }{\|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2} \\ & = \begin{pmatrix} p_N & q_N \\ p'_N & q'_N \end{pmatrix} \begin{pmatrix} \Psi^D \\ \Psi^S \end{pmatrix}, \end{aligned} \end{equation} where the operators $p_N, p'_N, q_N, q'_N$ are defined as following: \begin{subequations} \begin{eqnarray} \hspace{-3em} p_N: L^2([-\tau, 0]; \mathbb{R}^d) \rightarrow L^2([-\tau, 0]; \mathbb{R}^d), & & \Psi^D \mapsto \sum_{j = 1}^{Nd} \frac{\langle \Psi^D, \mathbf{K}_j^\tau \rangle_{L^2}}{\tau \|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2} \mathbf{K}_j^\tau; \label{Def_pN} \\ p'_N: L^2([-\tau, 0]; \mathbb{R}^d) \rightarrow \mathbb{R}^d, & & \Psi^D \mapsto \sum_{j = 1}^{Nd} \frac{\langle \Psi^D, \mathbf{K}_j^\tau \rangle_{L^2}}{ \tau \|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2} \mathbf{K}_j^\tau(0); \label{Def_pN'} \\ q_N: \mathbb{R}^d \rightarrow L^2([-\tau, 0]; \mathbb{R}^d), & & \Psi^S \mapsto \sum_{j = 1}^{Nd} \frac{ \langle \Psi^S, \mathbf{K}_j^\tau(0) \rangle}{\|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2} \mathbf{K}_j^\tau; \label{Def_qN}\\ q'_N: \mathbb{R}^d \rightarrow \mathbb{R}^d, & & \Psi^S \mapsto \sum_{j = 1}^{Nd} \frac{ \langle \Psi^S, \mathbf{K}_j^\tau(0) \rangle}{\|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2}\mathbf{K}_j^\tau(0). \label{Def_qN'} \end{eqnarray} \end{subequations} In the following, we arbitrarily fix $\Phi \in \mathcal{H}_k$ for some integer $k>0$. Now let us choose $N$ such that $Nd \ge k$, then $\Pi_{Nd} \Phi = \Phi$, and we get for each such $N$ \begin{equation}\label{Eq_abstract_AN} \mathcal{A}_{N} \Phi = \Pi_{N} \mathcal{A} \Pi_{N} \Phi = \Pi_{N} \mathcal{A} \Phi = \begin{pmatrix} p_N \frac{\mathrm{d}^+}{\mathrm{d} \theta}\Phi^D + q_N (L_S \Phi^S + L_D \Phi^D) \vspace{1em}\\ p'_N \frac{\mathrm{d}^+}{\mathrm{d} \theta}\Phi^D + q'_N (L_S \Phi^S + L_D \Phi^D) \end{pmatrix}. \end{equation} We obtain then \begin{equation} \label{eq:AN_residual} (\mathcal{A} - \mathcal{A}_{N}) \Phi = \begin{pmatrix} (I^D - p_N) \frac{\mathrm{d}^+}{\mathrm{d} \theta}\Phi^D - q_N \big(L_S \Phi^S + L_D \Phi^D\big) \vspace{1em}\\ - p'_N \frac{\mathrm{d}^+}{\mathrm{d} \theta}\Phi^D + (I^S - q'_N) \big(L_S \Phi^S + L_D \Phi^D \big) \end{pmatrix}, \; \text{ if } Nd \ge k, \end{equation} where $I^D$ and $I^S$ denote the identity maps on $L^2([-\tau, 0]; \mathbb{R}^d)$ and $\mathbb{R}^d$, respectively. We show below that the RHS of \eqref{eq:AN_residual} converges to zero. Let us begin with the term $(I^D - p_N) \frac{\mathrm{d}^+}{\mathrm{d} \theta}\Phi^D$. Note that by comparing the definition of $p_N$ give by \eqref{Def_pN} and the decomposition of $L^2([-\tau, 0]; \mathbb{R}^d)$ functions given by \eqref{L2_decomp}, we see that for each $f \in L^2([-\tau,0]; \mathbb{R}^d)$, the term $p_N f$ is the partial sum of the first $N$ terms in the corresponding decomposition. It follows then that \begin{equation} \label{Eq_pN_residual} \lim_{N \rightarrow \infty} \|(I^D - p_N) f\|_{L^2} = 0, \quad \text{ } \forall \: f \in L^2([-\tau,0]; \mathbb{R}^d). \end{equation} Since $\Phi \in \mathcal{H}_k \subset D(\mathcal{A})$, it holds that $\frac{\mathrm{d}^+}{\mathrm{d} \theta}\Phi^D \in L^2([-\tau,0]; \mathbb{R}^d)$. We obtain then from \eqref{Eq_pN_residual} that \begin{equation} \label{pN_est} \lim_{N \rightarrow \infty}\Big\|(I^D - p_N) \Big(\frac{\mathrm{d}^+}{\mathrm{d} \theta}\Phi^D \Big) \Big\|_{L^2} = 0. \end{equation} \medskip We turn now to the estimates for $q_N (L_S \Phi^S + L_D \Phi^D)$. By the definition of $q_N$ in \eqref{Def_qN}, we get \begin{equation} q_N \big(L_S \Phi^S + L_D \Phi^D \big) = \sum_{j = 1}^{Nd} \frac{ \langle L_S \Phi^S + L_D \Phi^D, \mathbf{K}_j^\tau(0)\rangle}{\|\mathcal{K}_{j_q}^\tau \|_{\mathcal{H}_1}^2} \mathbf{K}_j^\tau. \end{equation} It follows then from the identity \eqref{Eq_identity1} that \begin{equation} \label{qN_est} \lim_{N \rightarrow \infty}\Big\|q_N \big( L_S \Phi^S + L_D \Phi^D \big)\|_{L^2} = 0. \end{equation} For the term $p'_N \frac{\mathrm{d}^+}{\mathrm{d} \theta}\Phi^D$, since $\frac{\mathrm{d}^+}{\mathrm{d} \theta}\Phi^D \in L^2([-\tau,0); \mathbb{R}^d)$, it follows from the definition of $p'_N$ given in \eqref{Def_pN'} and the identity \eqref{Eq_identity3} that \begin{equation} \label{pN'_est} \lim_{N \rightarrow \infty} \Big| p'_N \frac{\mathrm{d}^+}{\mathrm{d} \theta}\Phi^D\Big| = 0, \end{equation} where $\vert\cdot\vert$ denotes the Euclidean norm of $\mathbb{R}^d$. By using the identity \eqref{Eq_identity2}, we also get \begin{equation} \label{qN'_est} \lim_{N \rightarrow \infty} \big| (I^S - q'_N) \big(L_S \Phi^S + L_D \Phi^D \big) \big| = 0. \end{equation} Now, by using the estimates \eqref{pN_est}, \eqref{qN_est}, \eqref{pN'_est}, and \eqref{qN'_est}, we get from \eqref{eq:AN_residual} that \begin{equation} \lim_{N \rightarrow \infty} \| (\mathcal{A} - \mathcal{A}_{N}) \Phi \|_\mathcal{H} = 0, \end{equation} and \eqref{Goal_convergence} follows. \end{proof} \vspace{1ex} \begin{rem}\label{PDE_rem} We explain here how the truncated linear operator $\mathcal{A}_N$ defined in \eqref{Eq_abstract_AN} is related to an interesting class of nonlocal linear PDEs. For the sake of clarity, we discuss the case $d=1$. For convenience, let us write $v(\theta)=\Phi^D(\theta)$ and, recalling e.g.~Eq.~\eqref{lin_PDE} in the Introduction, replace $\mathrm{d}^+ v/\mathrm{d} \theta$ by $v_{\theta}$ in \eqref{Eq_abstract_AN}. One then obtains that, when $v\in \mathcal{H}_N$, \begin{equation}\label{Eq_abstract_AN2} \begin{pmatrix} p_N v_{\theta}\\ p'_N v_{\theta} \end{pmatrix}=\begin{pmatrix} v_{\theta}\\ v_{\theta}(0) \end{pmatrix}-\sum_{n=0}^{N-1} \frac{v_{\theta}(0)}{\|\mathcal{K}^\tau_n\|_{\mathcal{H}}^2}\mathcal{K}^\tau_n. \end{equation} Next, we use the expressions of $\mathcal{A}_N$, $p_N$ and $q_N$ --- given, respectively, in \eqref{Eq_abstract_AN}, \eqref{Def_pN} and \eqref{Def_qN} --- to note that the $D$-component $v$ of any solution $u$ of \begin{equation*} \frac{\mathrm{d} u} {\mathrm{d} t} = \mathcal{A}_N u, \end{equation*} which emanates from initial data taken\footnote{For such initial data, the solution stays in $\mathcal{H}_N$, by the definition of $\mathcal{A}_N$.} in $\mathcal{H}_N$, satisfies the following {\it nonlocal linear PDE}: \begin{equation}\label{nonlocal_PDE} \partial_t v =\partial_{\theta} v +b_N(\theta)\Big(L_S v(t,0) + L_D v- \partial_{\theta} v\big\vert_{\theta=0}\Big); \end{equation} here \begin{equation}\label{bN-coeff} \displaystyle b_N(\theta)=\sum_{n=0}^{N-1} \frac{K_n^\tau (\theta)}{\|\mathcal{K}_n^{\tau}\|_{\mathcal{H}}^2}, \end{equation} and the rescaled Koornwinder polynomials $K_n^\tau$ are given by \eqref{eq:Pn_tilde}, since $d=1$. We see therewith that the Galerkin scheme used herein introduces a nonlocal perturbation term with respect to the $D$-component of $\mathcal{A}$ given in \eqref{Def_A2}. This perturbation term results from the difference between the $S$-component of $\mathcal{A}$ and the derivative at $0$, when applied to functions in $\mathcal{H}_N$. From Lemma \ref{Lem_A2} above and Lemmas \ref{Lemma_stability_prep} and \ref{Lem_A1} below, one can infer by the Trotter-Kato theorem that the effects on the solutions of Eq.~\eqref{nonlocal_PDE} of this nonlocal perturbation --- which vanishes as $N\rightarrow \infty$, due to \eqref{Eq_identity1_0} --- do not lead to a degenerate situation and that actually these solutions converge over finite intervals to those of the local PDE, $\partial_t v=\partial_\theta v$. This nice convergence is due to the key properties of the Koornwinder polynomials, as summarized in Lemma \ref{Fundamental_lemma} for $d=1$, and in Lemma \ref{Super_Fundamental_lemma} for the multidimensional case; see also Fig.~\ref{fig:Cancel_prop} for the nature of the approximation at $\theta = 0$. To conclude this remark, we return now to the discussion in the Introduction concerning the approximation of discontinuities that arise, for instance, in the first-order derivative of a DDE's solution, cf.~\cite[Appendix A and references therein]{GZT08}. In Table \ref{table}, we report the corresponding differences over the interval $[0,2]$ between the analytic solution of Eq.~\eqref{lin_case} with $\tau=1$ and history $x(t)\equiv 1$, on the one hand, and low-dimensional Galerkin approximations obtained by application of the formulas derived hereafter in Section~\ref{Sec_Galerkin_analytic}, on the other. \begin{table}[h] \caption{Errors in Galerkin approximation of Eq.~\eqref{lin_case} \label{table}} \centering \begin{tabular}{ccc} \toprule\noalign{\smallskip} $N$ & Max.~error in Galerkin solution & Max.~error in 1$^{\textrm{st}}$-order derivative\\ & & of Galerkin solution\\ \noalign{\smallskip}\hline\noalign{\smallskip} 4 & $6.9 \times 10^{-3}$ & $5.9\times 10^{-2}$\\ 8 & $9.3\times 10^{-4}$ & $2.2\times 10^{-2}$\\ 16 & $2.3\times 10^{-4}$ & $1.1\times 10^{-2}$\\ 32 & $9.0\times 10^{-5}$ & $5.4\times 10^{-3}$\\ \noalign{\smallskip} \bottomrule \end{tabular} \end{table} The second column of this table is obviously consistent with the rigorous convergence result of Corollary \ref{Cor_DDE_global_Lip} proved below. The third column shows that, furthermore, the aforementioned discontinuities present in the derivative of the DDE's solutions are well captured by the proposed methodology as well. \qed \end{rem} \medskip In the following, we restrict the linear operator $\mathcal{A}$ defined in \eqref{Def_A2} to be such that $L_S: \mathbb{R}^d \rightarrow \mathbb{R}^d$ is a bounded linear operator, and $L_D$ is defined by \begin{equation} \begin{aligned} \label{Def_LD} L_D : H^1([-\tau,0); \mathbb{R}^d) & \rightarrow \mathbb{R}^d, \\ \Psi^D & \mapsto B \Psi^D(-\tau) + \int_{-\tau}^0 C(s) \Psi^D(s) \mathrm{d} s, \end{aligned} \end{equation} with $B: \mathbb{R}^d \rightarrow \mathbb{R}^d$ being a bounded linear operator\footnote{Note that $\Psi^D(-\tau)$ is well-defined for $\Psi^D\in H^1([-\tau,0); \mathbb{R}^d)$, since the latter Sobolev space is continuously embedded in the space of continuous functions $C([-\tau,0]; \mathbb{R}^d)$; see \cite[Thm.~8.8]{brezis_book}.}, and $C(\cdot) \in L^2([-\tau, 0); \mathbb{R}^{d\times d})$. In the following preparatory lemma, Lemma \ref{Lemma_stability_prep}, we recall by means of basic estimates, that the existence of $\omega>0$ such that $\mathcal{A}-\omega I$ is dissipative in $\mathcal{H}$, can be guaranteed. This fact will be used to establish a stability condition of type \eqref{Eq_control_linearflow} (with $M=1$) for the semigroups $T(t)$ and $T_N(t)$, generated respectively by $\mathcal{A}$ and its finite-dimensional approximation $\mathcal{A}_N$; see Lemma \ref{Lem_A1}. The proofs of these Lemmas are quite straightforward but are reproduced below for the sake of completeness. \begin{lem} \label{Lemma_stability_prep} Let $\mathcal{A}$ be defined such as in \eqref{Def_A2} with $L_D$ such as specified in \eqref{Def_LD} and $L_S: \mathbb{R}^d \rightarrow \mathbb{R}^d$ to be a bounded linear operator. Then \begin{equation} \langle \mathcal{A} \Psi, \Psi \rangle_{\mathcal{H}} \le \omega \|\Psi\|_{\mathcal{H}}^2, \quad \forall\; \Psi \in D(\mathcal{A}), \end{equation} with\footnote{Throughout this article, we will denote indistinguishably by $|\cdot|$, either the Euclidean norm of a vector in $\mathbb{R}^d$, or its subordinated (operator) norm, in the case of a $d \times d$ matrix. It should be clear from the context which norm is used.} \begin{equation} \label{omega} \omega = \Big(1 + \frac{1}{2 \tau} + |L_S| + \frac{\tau}{2} |B|^2 + \frac{\tau}{4} \|C\|^2_{L^2} \Big). \end{equation} \end{lem} \begin{proof} Let $\Psi \in D(\mathcal{A})$, then by the definition of $ \mathcal{A}$ given in \eqref{Def_A2}, we have \begin{equation} \label{Stab_est0} \langle \mathcal{A} \Psi, \Psi \rangle_{\mathcal{H}} = \frac{1}{\tau} \int_{-\tau}^0 \Big\langle \frac{\mathrm{d}^+ \Psi^D}{\mathrm{d} \theta}(\theta), \Psi^D(\theta) \Big\rangle \mathrm{d} \theta + \langle L_S \Psi^S, \Psi^S \rangle + \langle L_D \Psi^D, \Psi^S \rangle. \end{equation} Note that \begin{equation} \begin{aligned} \label{Stab_est1} \frac{1}{\tau} \int_{-\tau}^0 \Big\langle \frac{\mathrm{d}^+ \Psi^D}{\mathrm{d} \theta}(\theta), \Psi^D(\theta) \Big\rangle \mathrm{d} \theta & = \frac{1}{2 \tau} \int_{-\tau}^{0} \mathrm{d} |\Psi^D(\theta)|^2 \\ & = \frac{1}{2 \tau} \Big(|\Psi^S|^2 - |\Psi^D(-\tau)|^2 \Big). \end{aligned} \end{equation} Using the definition of $L_D$ given in \eqref{Def_LD}, we obtain \begin{equation} \begin{aligned} \langle L_D \Psi^D, \Psi^S \rangle & = \Big\langle B \Psi^D(-\tau) + \int_{-\tau}^0 C(\theta) \Psi^D(\theta) \mathrm{d} \theta, \Psi^S \Big\rangle \\ & \le |B| |\Psi^D(-\tau)| | \Psi^S| + \|C\|_{L^2} \|\Psi^D\|_{L^2}|\Psi^S|. \end{aligned} \end{equation} It follows then from Young's inequality that \begin{equation} \begin{aligned} \label{Stab_est2} \langle L_D \Psi^D, \Psi^S \rangle & \le \Big( \frac{1}{2 \tau} |\Psi^D(-\tau)|^2 + \frac{\tau}{2} |B|^2 | \Psi^S|^2 \Big) \\ & \quad + \Big( \frac{1}{\tau}\|\Psi^D\|_{L^2}^2 + \frac{\tau}{4} \|C\|^2_{L^2} |\Psi^S|^2 \Big). \end{aligned} \end{equation} Note also that \begin{equation} \label{Stab_est3} \langle L_S \Psi^S, \Psi^S \rangle \le |L_s| |\Psi^S|^2. \end{equation} Now, by using \eqref{Stab_est1}, \eqref{Stab_est2}, and \eqref{Stab_est3} in \eqref{Stab_est0}, we get \begin{equation} \begin{aligned} \langle \mathcal{A} \Psi, \Psi \rangle_{\mathcal{H}} & \le \Big( \frac{1}{2 \tau} + |L_S| + \frac{\tau}{2}|B|^2 + \frac{\tau}{4} \|C\|^2_{L^2} \Big) |\Psi^S|^2 + \frac{1}{\tau}\|\Psi^D\|_{L^2}^2 \\ & \le \Big(1 + \frac{1}{2 \tau} + |L_S| + \frac{\tau}{2}|B|^2 + \frac{\tau}{4} \|C\|^2_{L^2} \Big) \Big(| \Psi^S|^2 + \frac{1}{\tau}\|\Psi^D\|_{L^2}^2 \Big) \\ & = \Big(1 + \frac{1}{2 \tau} + |L_S| + \frac{\tau}{2}|B|^2 + \frac{\tau}{4} \|C\|^2_{L^2} \Big) \|\Psi\|_{\mathcal{H}}^2, \end{aligned} \end{equation} leading thus to the desired estimate. \end{proof} \begin{lem} \label{Lem_A1} Let $\mathcal{A}$ be defined such as in \eqref{Def_A2} with $L_D$ such as specified in \eqref{Def_LD} and $L_S: \mathbb{R}^d \rightarrow \mathbb{R}^d$ to be a bounded linear operator. Then, the linear semigroups $T(t)$ and $T_N(t)$ generated respectively by $\mathcal{A}$ and $\mathcal{A}_N$ defined in \eqref{Eq_AN}, satisfy \begin{equation}\label{stable_estimates} \|T(t)\| \le e^{\omega t} \quad \text{ and } \quad \|T_N(t)\| \le e^{\omega t}, \qquad t \ge 0, \end{equation} with $\omega$ given by \eqref{omega}. \end{lem} \begin{proof} Since $T(t)$ is a $C_0$-semigroup with infinitesimal generator $\mathcal{A}$, we have that $T(t) u_0 \in D(\mathcal{A})$ for all $u_0 \in D(\mathcal{A})$, and that \begin{equation} \frac{\mathrm{d}}{\mathrm{d} t} T(t)u_0 = \mathcal{A} T(t) u_0, \qquad \text{ } \forall \: u_0 \in \mathcal{A}, \; t \ge 0; \end{equation} cf.~\cite[Thm.~2.4 c) p.5]{Pazy83}. We obtain thus \begin{equation} \label{T_est1} \frac{\mathrm{d} }{\mathrm{d} t }\| T(t) u_0 \|^2_{\mathcal{H}} = 2 \langle \mathcal{A} T(t) u_0, T(t) u_0 \rangle_{\mathcal{H}} \le 2 \omega \|T(t) u_0 \|^2_{\mathcal{H}}, \qquad \text{ } \forall \: u_0 \in D(\mathcal{A}), \end{equation} where we have used Lemma~\ref{Lemma_stability_prep} to obtain the last inequality above with $\omega$ given by \eqref{omega}. It follows then from Gronwall's inequality that \begin{equation} \label{T_est2} \| T(t) u_0 \|^2_{\mathcal{H}} \le e^{2\omega t} \|u_0\|^2_{\mathcal{H}}, \qquad \text{ } \forall \: u_0 \in D(\mathcal{A}). \end{equation} Since $D(\mathcal{A})$ is dense in $\mathcal{H}$ and $T(t)$ are bounded operators on $\mathcal{H}$, the estimate \eqref{T_est2} still holds for general initial data in $\mathcal{H}$, leading in turn to \begin{equation} \label{stability_A} \| T(t)\| \le e^{\omega t}, \qquad t \ge 0. \end{equation} The estimate for $T_N$ is also trivial and proceeds as follows. First note that by the definition of $T_N$ given by \eqref{Eq_extension}, we have \begin{equation} \begin{aligned} \|T_N(t) u_0\|^2_\mathcal{H} & = \big\langle e^{\mathcal{A}_N t} \Pi_N u_0 + (I - \Pi_N) u_0, e^{\mathcal{A}_N t} \Pi_N u_0 + (I - \Pi_N) u_0 \big\rangle_{\mathcal{H}} \\ & = \big\langle e^{\mathcal{A}_N t} \Pi_N u_0, e^{\mathcal{A}_N t} \Pi_N u_0 \big\rangle_{\mathcal{H}} + \big\langle (I - \Pi_N) u_0, (I - \Pi_N) u_0 \big\rangle_{\mathcal{H}} \\ & = \| e^{\mathcal{A}_N t} \Pi_N u_0 \|^2_{\mathcal{H}} + \| (I - \Pi_N) u_0\|^2_{\mathcal{H}}. \end{aligned} \end{equation} It follows that \begin{equation} \begin{aligned} \label{TN_est1} \frac{\mathrm{d} }{\mathrm{d} t }\| T_N(t) u_0 \|^2_{\mathcal{H}} &= \frac{\mathrm{d} }{\mathrm{d} t }\| e^{\mathcal{A}_N t} \Pi_N u_0 \|^2_{\mathcal{H}} \\ & = 2 \langle \mathcal{A}_N e^{\mathcal{A}_N t} \Pi_N u_0, e^{\mathcal{A}_N t} \Pi_N u_0 \rangle_{\mathcal{H}}, \qquad \text{ } \forall \: u_0 \in \mathcal{H}. \end{aligned} \end{equation} Note also that \begin{equation} \begin{aligned} \label{TN_est2} \langle \mathcal{A}_N \Psi, \Psi \rangle_{\mathcal{H}} & = \langle \Pi_N \mathcal{A} \Pi_N \Psi, \Psi \rangle_{\mathcal{H}} \\ & = \langle \Pi_N \mathcal{A} \Pi_N \Psi, \Pi_N \Psi \rangle_{\mathcal{H}} + \langle \Pi_N \mathcal{A} \Pi_N \Psi, (I- \Pi_N) \Psi \rangle_{\mathcal{H}}\\ & = \langle \Pi_N \mathcal{A} \Pi_N \Psi, \Pi_N \Psi \rangle_{\mathcal{H}} \\ & = \langle \mathcal{A} \Pi_N \Psi, \Pi_N \Psi \rangle_{\mathcal{H}} \\ & \le \omega \|\Pi_N \Psi\|_{\mathcal{H}} \\ & \le \omega \|\Psi\|_{\mathcal{H}}. \end{aligned} \end{equation} We obtain then from \eqref{TN_est1} that \begin{equation} \begin{aligned} \label{TN_est3} \frac{\mathrm{d} }{\mathrm{d} t }\| T_N(t) u_0 \|^2_{\mathcal{H}} & \le 2 \omega \|e^{\mathcal{A}_N t} \Pi_N u_0\|^2_{\mathcal{H}} \\ & \le 2\omega \big( \|e^{\mathcal{A}_N t} \Pi_N u_0\|^2_{\mathcal{H}} + \| (I - \Pi_N) u_0\|^2_{\mathcal{H}} \big) \\ & = 2\omega \|e^{\mathcal{A}_N t} \Pi_N u_0 +(I - \Pi_N) u_0\|^2_{\mathcal{H}} \\ & = 2\omega \|T_N(t) u_0 \|^2_{\mathcal{H}}, \qquad \text{ } \forall \: u_0 \in \mathcal{H}. \end{aligned} \end{equation} The desired estimate for $\| T_N(t)\|$ can be derived now from \eqref{TN_est3} by using Gronwall's inequality. \end{proof} \begin{rem} Note that the estimate about $T_N$ in \eqref{stable_estimates} shows in particular that solutions of \eqref{nonlocal_PDE} grow at most exponentially with a rate independent of $N$, and stay uniformly bounded over finite intervals. \qed \end{rem} With these preparatory lemmas, we are now in position to obtain as corollaries of Theorem~\ref{ParisVI_thm}, the convergence results for the Galerkin approximation \eqref{Eq_DDE_Galerkin} of $d$-dimensional nonlinear systems of DDEs of the form \begin{equation}\label{Eq_nln_sys} \frac{\mathrm{d} \mathbf{x}}{\mathrm{d} t}=L_S \mathbf{x}(t)+ B \mathbf{x}(t-\tau) + \int_{t-\tau}^t C(s-t) \mathbf{x}(s) \mathrm{d} s + \mathbf{F}\Big(\mathbf{x}(t),\int_{t-\tau}^t \mathbf{x}(s) \mathrm{d} s\Big), \end{equation} where $\mathbf{F}:\mathbb{R}^d\times \mathbb{R}^d \rightarrow \mathbb{R}^d$, and $L_S$, $B$, and $C$ are as given in \eqref{Def_LD}. We first sate the result for the case of global Lipschitz nonlinearity, keeping in mind that already for the case of scalar DDEs ($d=1$), chaotic dynamics can take place under such a simple nonlinear setting; see Section \ref{Sec_nearly-brownian} for a numerical illustration. \vspace{1ex} \begin{cor} \label{Cor_DDE_global_Lip} Let $\mathcal{A}$ be defined such as in \eqref{Def_A2} with $L_D$ such as specified in \eqref{Def_LD} and $L_S: \mathbb{R}^d \rightarrow \mathbb{R}^d$ to be a bounded linear operator. Assume that the nonlinearity $\mathcal{F} \colon \mathcal{H} \rightarrow \mathcal{H}$ defined by \begin{equation} \begin{aligned} \label{Def_F_sys} [\mathcal{F} (\Psi) ](\theta) & := \begin{cases} 0, & \theta \in[-\tau, 0), \vspace{0.4em}\\ \mathbf{F} \Big(\Psi^S, \int_{-\tau}^0 \Psi^D(s) \mathrm{d} s\Big), & \theta = 0, \end{cases} \quad \text{ } \forall \: \Psi = (\Psi^D, \Psi^S) \in \mathcal{H}, \end{aligned} \end{equation} is globally Lipschitz. Then, for each $u_0 \in \mathcal{H}$, the mild solution of \eqref{Eq_DDE_Galerkin} emanating from $\Pi_N u_0$ converges uniformly to the mild solution of \eqref{Eq_abstract_ODE_DDE} emanating from $u_0$ on each bounded interval $[0, T]$, i.e.: \begin{equation} \label{uniform_conv_est} \lim_{N\rightarrow \infty} \sup_{t \in [0, T]} \|u_N(t; \Pi_N u_0) - u(t; u_0)\|_{\mathcal{H}} = 0, \qquad \text{ } \forall \: T > 0, \; u_0 \in \mathcal{H}. \end{equation} \end{cor} \vspace{1ex} \begin{proof} This corollary is a direct consequence of Lemma~\ref{Lem_A1} and Lemma~\ref{Lem_A2}, ensuring respectively, Conditions {\bf (A1)} and {\bf (A2)} of Theorem~\ref{ParisVI_thm}. \end{proof} \vspace{1ex} In the next two corollaries, we relax the global Lipschitz condition assumed in Corollary~\ref{Cor_DDE_global_Lip} to a local Lipschitz condition in addition to either a sublinear growth for $\mathcal{F}$ (see Corollary~\ref{Cor_DDE_local_Lip_Case1}) or an energy inequality satisfied by $\mathcal{F}$; see Corollary~\ref{Cor_DDE_local_Lip_Case2}. \begin{cor} \label{Cor_DDE_local_Lip_Case1} Let $\mathcal{A}$ be defined in \eqref{Def_A2} with $L_D$ specified in \eqref{Def_LD} and $L_S: \mathbb{R}^d \rightarrow \mathbb{R}^d$ to be a bounded linear operator. Assume that the nonlinearity $\mathcal{F}$ given by \eqref{Def_F_sys} is locally Lipschitz in the sense that for all $r > 0$ there exists $L(r)>0$ such that for any $\Psi_1$ and $\Psi_2$ in $\mathcal{H}$, we have \begin{equation} \label{Local_Lip_cond} \|\Psi_1\|_{\mathcal{H}}<r \mbox{ and } \|\Psi_2\|_{\mathcal{H}} < r \Longrightarrow \|\mathcal{F}(\Psi_1) - \mathcal{F}(\Psi_2)\|_{\mathcal{H}} \le L(r) \|\Psi_1 - \Psi_2\|_{\mathcal{H}}. \end{equation} Assume also that $\mathcal{F}$ satisfies the following sublinear growth: \begin{equation} \label{Sublinear_onF} \|\mathcal{F}(\Psi)\|_{\mathcal{H}} \le \gamma_1 \|\Psi\|_{\mathcal{H}} + \gamma_2, \qquad \text{ } \forall \: \Psi \in \mathcal{H}, \end{equation} where $\gamma_1>0$ and $\gamma_2\geq 0$. Then, for each $u_0 \in \mathcal{H}$, the mild solution $u_N(t; \Pi_N u_0)$ of \eqref{Eq_DDE_Galerkin} emanating from $\Pi_N u_0$ and, the mild solution $u(t; u_0)$ of \eqref{Eq_abstract_ODE_DDE} emanating from $u_0$, do not blow up in any finite time. Moreover, $u_N(t; \Pi_N u_0)$ converges uniformly to $u(t; u_0)$ on each bounded interval $[0, T]$, i.e.: \begin{equation} \label{uniform_conv_est_Case1} \lim_{N\rightarrow \infty} \sup_{t \in [0, T]} \|u_N(t; \Pi_N u_0) - u(t; u_0)\|_{\mathcal{H}} = 0, \qquad \text{ } \forall \: T > 0, \; u_0 \in \mathcal{H}. \end{equation} \end{cor} \begin{proof} Recall that the local Lipschitz condition \eqref{Local_Lip_cond} on $\mathcal{F}$ ensures the existence and uniqueness of a local mild solution $u(t; u_0)$ to \eqref{Eq_abstract_ODE_DDE} emanating from any $u_0 \in \mathcal{H}$; see e.g. \cite[Prop.~4.3.3]{Cazenave_al98}.\footnote{\cite[Prop.~4.3.3]{Cazenave_al98} is derived for the case of contraction semigroups. However, the proof can be easily adapted to the case of more general $C_0$-semigroups $T(t)$ for which $\| T(t) \| \le M e^{\omega t}$.} By recalling that $\|T(t)\|_{\mathcal{H}} \le e^{\omega t}$ (see Lemma~\ref{Lem_A1}) and by using the sublinear growth assumption \eqref{Sublinear_onF} on $\mathcal{F}$, we obtain for mild solutions \begin{equation} \begin{aligned} \label{Energy_est_for_u} \|u(t)\|_{\mathcal{H}} & \le \| T(t) u_0\|_{\mathcal{H}} + \int_{0}^t \|T(t-s)\mathcal{F}(u(s))\|_{\mathcal{H}} \mathrm{d} s \\ & \le e^{\omega t} \|u_0\|_{\mathcal{H}} + \int_{0}^t e^{\omega(t-s)} \Big( \gamma_1 \|u(s)\|_{\mathcal{H}} + \gamma_2 \Big) \mathrm{d} s \\ & \le e^{\omega t} \|u_0\|_{\mathcal{H}} + \frac{\gamma_2 (e^{\omega t} -1)}{\omega} + \gamma_1 \int_{0}^t e^{\omega(t-s)} \|u(s)\|_{\mathcal{H}} \mathrm{d} s, \end{aligned} \end{equation} where the positive constant $\omega$ is given by \eqref{omega}. A simple multiplication by $e^{-\omega t}$ to both sides of \eqref{Energy_est_for_u}, leads then trivially to \begin{equation} \begin{aligned} e^{-\omega t}\|u(t)\|_{\mathcal{H}} & \le \|u_0\|_{\mathcal{H}} + \frac{\gamma_2 (1 - e^{-\omega t})}{\omega} + \gamma_1 \int_{0}^t e^{-\omega s} \|u(s)\|_{\mathcal{H}} \mathrm{d} s \\ & \le \|u_0\|_{\mathcal{H}} + \frac{\gamma_2}{\omega} + \gamma_1 \int_{0}^t e^{-\omega s} \|u(s)\|_{\mathcal{H}} \mathrm{d} s. \end{aligned} \end{equation} An application of the Gronwall's inequality to $v(t):=e^{-\omega t}\|u(t)\|_{\mathcal{H}}$, gives then \begin{equation} \label{Eq_bound_for_u} \|u(t)\|_{\mathcal{H}} \le \Big( \|u_0\|_{\mathcal{H}} + \frac{\gamma_2}{\omega} \Big) e^{(\omega + \gamma_1) t}, \end{equation} preventing thus the blow up of a mild solution in finite time. Similarly, for mild solutions $u_N$ of \eqref{Eq_DDE_Galerkin}, we have \begin{equation} \label{Eq_bound_for_uN} \|u_N(t)\|_{\mathcal{H}} \le \Big( \|u_0\|_{\mathcal{H}} + \frac{\gamma_2}{\omega} \Big) e^{(\omega + \gamma_1) t}. \end{equation} by noting that $\|\Pi_N\| < 1$ for all $N \ge 1$, $\|T_N(t)\|_{\mathcal{H}} \le e^{\omega t}$ (see Lemma~\ref{Lem_A1}), and by using the sublinear growth assumption \eqref{Sublinear_onF} on $\mathcal{F}$, preventing also the blow up of any mild solution $u_N$ of \eqref{Eq_DDE_Galerkin}, in finite time. Finally, \eqref{Eq_bound_for_u} and \eqref{Eq_bound_for_uN} lead to \begin{equation} \begin{aligned} & \|u(t; u_0)\|_{\mathcal{H}} \le C(T, \|u_0\|_{\mathcal{H}}), && \text{ } \forall \: t \in [0, T], \\ & \|u_N(t; \Pi_N u_0)\|_{\mathcal{H}} \le C(T, \|u_0\|_{\mathcal{H}}), && \text{ } \forall \: t \in [0, T] \text{ and } N \in \mathbb{N^*}, \end{aligned} \end{equation} where \begin{equation*} C(T, \|u_0\|_{\mathcal{H}}) := \Big( \|u_0\|_{\mathcal{H}} + \frac{\gamma_2}{\omega} \Big) e^{(\omega + \gamma_1) T}. \end{equation*} Now, we can follow the proof of Theorem \ref{ParisVI_thm} to obtain the desired convergence result \eqref{uniform_conv_est_Case1}, with the only difference consisting of the global Lipschitz estimates used therein, by the local Lipschitz condition \eqref{Local_Lip_cond} applied to $\Psi$ such that \begin{equation*} \| \Psi\|_{\mathcal{H}} \leq2 C(T, \|u_0\|_{\mathcal{H}}). \end{equation*} \end{proof} \begin{cor} \label{Cor_DDE_local_Lip_Case2} Let $\mathcal{A}$ be defined in \eqref{Def_A2} with $L_D$ specified in \eqref{Def_LD} and $L_S$ to be a bounded linear operator from $\mathbb{R}^d$ to $\mathbb{R}^d$. Assume that the nonlinearity $\mathcal{F}$ given by \eqref{Def_F_sys} is locally Lipschitz in the sense of \eqref{Local_Lip_cond}. Assume also that the following energy inequality holds for $\mathcal{F}$ \begin{equation} \label{Energy_ineq_onF} \langle \mathcal{F}(\Psi), \Psi \rangle_{\mathcal{H}} \le \gamma_1 \|\Psi\|^2_{\mathcal{H}} + \gamma_2, \qquad \text{ } \forall \: \Psi \in \mathcal{H}, \end{equation} where $\gamma_1 \in \mathbb{R}$ and $\gamma_2 \geq 0$. Then, for each $u_0 \in D(\mathcal{A})$, the strong solution $u_N(t; \Pi_N u_0)$ of \eqref{Eq_DDE_Galerkin} emanating from $\Pi_N u_0$, and the strong solution\footnote{By strong solutions of \eqref{Eq_abstract_ODE_DDE}, we mean a solution in $C([0,T],D(\mathcal{A}))\cap C^1([0,T],\mathcal{H})$ of \eqref{Eq_abstract_ODE_DDE}.} $u(t; u_0)$ of \eqref{Eq_abstract_ODE_DDE} emanating from $u_0$, do not blow up in any finite time. Moreover, $u_N(t; \Pi_N u_0)$ converges uniformly to $u(t; u_0)$ on each bounded interval $[0, T]$, i.e.: \begin{equation} \label{uniform_conv_est_engest_onF} \lim_{N\rightarrow \infty} \sup_{t \in [0, T]} \|u_N(t; \Pi_N u_0) - u(t; u_0)\|_{\mathcal{H}} = 0, \qquad \text{ } \forall \: T > 0, \; u_0 \in D(\mathcal{A}). \end{equation} Furthermore, any strong solutions $v=u$ of \eqref{Eq_abstract_ODE_DDE} or $v=u_N$ of \eqref{Eq_DDE_Galerkin}, emanating respectively from $v(0)=u_0 \in D(\mathcal{A})$ or $v(0)=\Pi_N u_0$, have their $\mathcal{H}$-norm controlled as follows: \begin{equation} \|v(t)\|^2_{\mathcal{H}}\leq e^{\kappa t} \|v(0)\|^2_{\mathcal{H}} + \frac{2 \gamma_2}{\kappa} (e^{\kappa t}-1), \; t>0, \end{equation} where $\kappa=2 (\omega+ \gamma_1)$, with $\omega$ given in \eqref{omega}. \end{cor} \begin{proof} Let $u_0\in D(\mathcal{A})$, and let $u$ be the mild solution of \eqref{Eq_abstract_ODE_DDE} emanating from $u_0$, such as ensured by the local Lipschitz condition on $\mathcal{F}$. Then by adapting the proof of e.g.~\cite[Prop.~4.3.9]{Cazenave_al98},\footnote{The regularity result \cite[Prop.~4.3.9]{Cazenave_al98} is stated for the case of contraction semigroups. However, the proof can be adapted to the case of $C_0$-semigroups for which $\|T(t)\| \le e^{\omega t}$ (i.e. with $M=1$) such as encountered here when $L_D$ is as specified in \eqref{Def_LD}.} we have that there exists a map $T:D(\mathcal{A})\rightarrow (0,\infty]$, for which $u\in C([0,T(u_0)),D(\mathcal{A}))\cap C^1([0,T(u_0)),\mathcal{H})$ and $u$ solves the initial-value problem \begin{subequations} \begin{eqnarray} \label{DDE_IVP} & \displaystyle{\frac{\mathrm{d} u}{\mathrm{d} t}} \hspace{-1.5em} & =\mathcal{A} u +\mathcal{F}(u), \label{DDE_IVP_eq1}\\ & u(0) \hspace{-0.5em} & =u_0. \end{eqnarray} \end{subequations} By taking the $\mathcal{H}$-inner product on both sides of \eqref{DDE_IVP_eq1} with the solution $u\in \mathcal{D}(\mathcal{A})$, and using the energy inequality \eqref{Energy_ineq_onF} and the stability property $\langle \mathcal{A} \Psi, \Psi \rangle_{\mathcal{H}} \le \omega \| \Psi\|^2_{\mathcal{H}}$ from Lemma~\ref{Lemma_stability_prep}, we obtain \begin{equation} \label{Energy_est_for_u_2} \frac{1}{2} \frac{\mathrm{d}}{\mathrm{d} t}\|u\|^2_{\mathcal{H}} = \langle \mathcal{A} u, u \rangle_{\mathcal{H}} + \langle \mathcal{F}(u), u \rangle_{\mathcal{H}} \le \omega \|u\|^2_{\mathcal{H}} + \gamma_1 \|u\|^2_{\mathcal{H}} + \gamma_2, \end{equation} where the positive constant $\omega$ is given by \eqref{omega}. It follows then from Gronwall's inequality that \begin{equation} \label{Eq_bounds_u_Case2} \|u(t; u_0)\|^2_{\mathcal{H}} \leq e^{\kappa t} \|u_0\|^2_{\mathcal{H}} + \frac{2 \gamma_2}{\kappa} (e^{\kappa t}-1), \quad t \in [0, T(u_0)), \; u_0 \in D(\mathcal{A}), \end{equation} where $\kappa = \omega + \gamma_1$. Similarly, by noting that \begin{equation*} \langle \Pi_N \mathcal{F}(\Psi), \Psi \rangle_{\mathcal{H}} = \langle \mathcal{F}(\Psi), \Psi \rangle_{\mathcal{H}}, \qquad \text{ } \forall \: \Psi \in \mathcal{H}_N, \end{equation*} we have \begin{equation} \|u_N(t; \Pi_N u_0)\|^2_{\mathcal{H}} \leq e^{\kappa t} \|u_0\|^2_{\mathcal{H}} + \frac{2 \gamma_2}{\kappa} (e^{\kappa t}-1), \quad t \in [0, T(u_0)), \; u_0 \in \mathcal{H}. \end{equation} We have thus shown that for any initial data $u_0 \in D(\mathcal{A})$, the strong solutions $u(t; u_0)$ and $u_N(t; \Pi_N u_0)$ do not blow up in any finite time. The convergence result \eqref{uniform_conv_est_engest_onF} can then be deduced as in the proof of Corollary~\ref{Cor_DDE_local_Lip_Case1}. \end{proof} \vspace{1ex} \begin{rem} \label{Rmk_forcing} It is worth mentioning that the conclusions of Corollaries \ref{Cor_DDE_global_Lip}, \ref{Cor_DDE_local_Lip_Case1} and \ref{Cor_DDE_local_Lip_Case2} still hold when the underlying system of DDEs \eqref{Eq_nln_sys} is perturbed by a suitable time-dependent forcing, $\mathbf{g}(t)$. For instance, it suffices to assume that $\mathbf{g}(t) \in L^2_{\mathrm{loc}}([0,\infty); \mathbb{R}^d)$ to still get the convergence results. \qed \end{rem} \subsection{Examples}\label{Sec_examples} In this section we provide some class of nonlinear scalar DDEs of the form \eqref{Eq_DDE} that fit with the assumptions of Corollary \ref{Cor_DDE_local_Lip_Case2}. In that respect we restrict our attention to the case of $\mathcal{A}$ such as defined in \eqref{Def_A} for $d=1$. We discuss below some classes of nonlinearities that verify the local Lipschitz condition \eqref{Local_Lip_cond} and the energy inequality \eqref{Energy_ineq_onF}. Extension to systems can be easily built out of these examples and are thus left to the reader. \subsubsection{Delay equations with a global Lipschitz nonlinearity}\label{sec_ex1} Let $\mathcal{F}$ be given such as in \eqref{Def_F}, and for which $F$ is assumed to be of the form \begin{equation}\label{F_example0} F \Big(\Psi^S,\int_{-\tau}^0 \Psi^D(\theta) \mathrm{d} \theta \Big)=g\Big(\int_{-\tau}^0 \Psi^D(\theta) \mathrm{d} \theta\Big) +h(\Psi^S), \end{equation} with $g$ and $h$ to be global Lipschitz maps from $\mathbb{R}$ to $\mathbb{R}$, of constants $L_1$ and $L_2$, respectively. Then, we have \begin{equation*} \langle \mathcal{F}(\Psi), \Psi \rangle_{\mathcal{H}} =g \Big(\int_{-\tau}^0 \Psi^D(\theta) \mathrm{d} \theta \Big) \Psi^S +h(\Psi^S) \Psi^S, \end{equation*} which gives \begin{equation*} \begin{aligned} \langle \mathcal{F}(\Psi), \Psi \rangle_{\mathcal{H}} &\leq L_1 \Big|\int_{-\tau}^0 \Psi^D(\theta) \mathrm{d} \theta\Big|\Big|\Psi^S\Big|+ (|g(0)|+|h(0)|)|\Psi^S| + L_2 |\Psi^S|^2\\ & \leq \gamma_1 \| \Psi \|^2_{\mathcal{H}} +\gamma_2, \end{aligned} \end{equation*} with $\gamma_1>0$ and $\gamma_2\geq 0$, and thus $F$ satisfies the energy inequality \eqref{Energy_ineq_onF}. The local Lipschitz condition \eqref{Local_Lip_cond} for $\mathcal{F}$ is trivially satisfied under the assumptions on $F$. Note that such nonlinear equations arise in many applications where a delayed monotone feedback mechanism is naturally involved in the description of the system's evolution; see \cite{GZT08,krisztin2008,Mallet_Sell96}. It is also interesting to mention that such seemingly simple scalar DDEs with global Lipschitz nonlinearity can also support chaotic dynamics as illustrated in Section \ref{Sec_nearly-brownian} below. \subsubsection{Delay equations with locally Lipschitz nonlinearity}\label{sec_ex2} We relax now the global Lipschitz requirement. In that respect, we consider \begin{equation} \label{F_example} F\Big(\Psi^S, \int_{-\tau}^0 \Psi^D(\theta) \mathrm{d} \theta \Big)=- g_1 \Big(\int_{-\tau}^0 \Psi^D(\theta) \mathrm{d} \theta\Big) g_2\big(\Psi^S \big), \end{equation} and assume that \begin{itemize} \item[(i)] $g_1: \mathbb{R} \rightarrow \mathbb{R}^+$ is locally Lipschitz; \item[(ii)] $g_2: \mathbb{R} \rightarrow \mathbb{R}$ is locally Lipschitz and verifies the condition \begin{equation}\label{Eq_pos_g2} g_2(x)x \ge 0, \qquad x \in \mathbb{R}. \end{equation} \end{itemize} These assumptions allow us to consider a broad class of nonlinear effects that are not necessarily bounded or polynomial. We check below that the abstract nonlinear map $\mathcal{F}$ defined in \eqref{Def_F} with $F$ given by \eqref{F_example} and that satisfy (i) and (ii), satisfies also the conditions of Corollary \ref{Cor_DDE_local_Lip_Case2}. We first check the local Lipschitz condition. Trivially, let us first remark that \begin{equation} \label{F_Lip_est0} \|\mathcal{F}(\Psi_1) - \mathcal{F}(\Psi_2) \|_{\mathcal{H}} = \left |F \Big(\Psi_1^S, \int_{-\tau}^0 \Psi^D_1(\theta) \mathrm{d} \theta \Big) - F \Big(\Psi^S_2, \int_{-\tau}^0 \Psi^D_2(\theta) \mathrm{d} \theta \Big) \right|. \end{equation} Let us introduce the notations $\alpha_i := \Psi_i^S$ and $\beta_i := \int_{-\tau}^0 \Psi^D_i(\theta) \mathrm{d} \theta$, $i = 1, 2$. Let $\Psi_i$ be chosen such that for $i = 1, 2$, $\|\Psi_i\|_{\mathcal{H}} \le R$ for some $R>0$. It follows that \begin{equation} |\alpha_i| \le R, \quad |\beta_i| = \left| \int_{-\tau}^0 \Psi^D_i(\theta) \mathrm{d} \theta \right| \le \sqrt{\tau} \|\Psi^D_i\|_{L^2}, \end{equation} and thus by definition of the $\mathcal{H}$-inner product \eqref{H_inner} \begin{equation} |\beta_i| \le \tau \|\Psi_i\|_{\mathcal{H}} \le \tau R, \end{equation} which leads to \begin{equation} \begin{aligned} \label{F_Lip_est1} & |F (\alpha_1, \beta_1) - F (\alpha_2, \beta_2) |\\ & \le |g_1(\beta_1) g_2(\alpha_1) - g_1(\beta_2) g_2(\alpha_2)| \\ & \le \Big|g_1(\beta_1) \big (g_2(\alpha_1) - g_2(\alpha_2) \big) \Big| + \Big|\big( g_1(\beta_1) - g_1(\beta_2) \big) g_2(\alpha_2) \Big| \\ & \le \mathrm{L}_2(R) |g_1(\beta_1)| |\alpha_1 - \alpha_2| + \mathrm{L}_1(\tau R) |g_2(\alpha_2)| |\beta_1 - \beta_2|, \end{aligned} \end{equation} where $\mathrm{L}_1(r)$ (resp.~$\mathrm{L}_2(r)$) denotes the local Lipschitz constant associated with $g_1(x)$ (resp.~$g_2(x)$) for $|x|<r$. On the other hand, \begin{equation*} \begin{aligned} |g_1(\beta_1)| & \le |g_1(\beta_1) - g_1(0)| + |g_1(0)| \\ & \le \mathrm{L}_1(\tau R) |\beta_1| + |g_1(0)| \\ & \le \tau R \mathrm{L}_1(\tau R) + |g_1(0)|, \end{aligned} \end{equation*} and \begin{equation*} \begin{aligned} |g_2(\alpha_2)| \le R \mathrm{L}_2(R) + |g_2(0)|. \end{aligned} \end{equation*} Note also that \begin{equation*} |\alpha_1 - \alpha_2| \le \| \Psi_1 - \Psi_2 \|_{\mathcal{H}}, \end{equation*} and that \begin{equation*} |\beta_1 - \beta_2| \le \sqrt{\tau} \|\Psi^D_1 - \Psi^D_2\|_{L^2} \le \tau \|\Psi_1 - \Psi_2\|_{\mathcal{H}}. \end{equation*} We obtain then from \eqref{F_Lip_est1} that \begin{equation} \label{F_Lip_est2} |F (\alpha_1, \beta_1) - F (\alpha_2, \beta_2) | \le L(R) \|\Psi_1 - \Psi_2\|_{\mathcal{H}}, \end{equation} where \begin{equation*} L(R):= \mathrm{L}_2(R) \Big(\sqrt{\tau} R \mathrm{L}_1(\tau R) + |g_1(0)| \Big) + \tau \mathrm{L}_1(\tau R) \Big(R \mathrm{L}_2( R) + |g_2(0)| \Big), \end{equation*} which gives the local Lipschitz property of $\mathcal{F}$ as a map from $\mathcal{H}$ to $\mathcal{H}$. The energy inequality \eqref{Energy_ineq_onF} is here readily satisfied since \begin{equation} \langle \mathcal{F}(\Psi), \Psi \rangle_{\mathcal{H}} = F \Big(\Psi^S, \int_{-\tau}^0 \Psi^D(\theta) \mathrm{d} \theta \Big) \Psi^S \leq 0, \end{equation} because of \eqref{Eq_pos_g2}. \needspace{1\baselineskip} \begin{rem}\label{Rmk_ex} \hspace*{2em} \vspace*{-0.4em} \begin{itemize} \item[(i)] Note that famous delayed models from population dynamics are covered by Corollary \ref{Cor_DDE_local_Lip_Case2}, although not satisfying \eqref{F_example}. For instance, delayed logistic equations of the form \begin{equation} \frac{\mathrm{d} x}{\mathrm{d} t}=r x(t)\Big(1-K^{-1}\int_{t-\tau}^t \omega(s) f(x(s))\mathrm{d} s\Big), \; r, K, \tau>0, \end{equation} with $\omega \in L^{\infty} (\mathbb{R},\mathbb{R}^+)$ and $f\in L^1(\mathbb{R},\mathbb{R}^+)$ satisfying the inequality \begin{equation*} |f(x)-f(y)| \leq \gamma |x-y|, \end{equation*} for some $\gamma > 0$ and for almost every $x,y \in \mathbb{R}$, are still covered by Corollary \ref{Cor_DDE_local_Lip_Case2}. \item[(ii)] Note also that many other nonlinear effects could have been considered in \eqref{Eq_DDE} for which the convergence result of Corollary \ref{Cor_DDE_local_Lip_Case2} would hold. For instance we could have considered \begin{equation}\label{other_nonl} F\Big(\Psi^S, \int_{-\tau}^0 \Psi^D(\theta) \mathrm{d} \theta\Big)= \sum_{j=1}^{2p-1} b_j \Big(\int_{-\tau}^0 \Psi^D(\theta) \mathrm{d} \theta\Big) (\Psi^S)^j, p\geq 2, \end{equation} where each $b_j$ is a local Lipschitz function in $L^{\infty}(\mathbb{R})$ and $b_{2p-1} (\cdot )\leq \beta <0$ for some constant $\beta$. Under these assumptions, the nonlinearity \eqref{other_nonl} satisfies the energy inequality \eqref{Energy_ineq_onF} with $\gamma_1 = 0$ and with $\gamma_2$ sufficiently large, depending on the $L^\infty$ norm of the functions $b_j$'s. This is because the term with the largest power $(\Psi^S)^{2p-1}$ is strictly negative by assumption on $b_{2p-1}$, and the other terms with a lower degree can be controlled by using Young's inequality. \qed \end{itemize} \end{rem} \section{Galerkin approximation: Analytic formulas for scalar DDEs}\label{Sec_Galerkin_analytic} This section is devoted to the derivation of explicit expressions of the Galerkin approximation \eqref{Eq_DDE_Galerkin} associated with nonlinear DDEs. For simplicity, we focus on the case of a scalar DDE taking the form given by \eqref{Eq_DDE}. The more general case of nonlinear systems of DDEs is dealt with in Appendix \ref{Appendix_systems}; see also Appendix \ref{Subsect_var_C} for the case where the linear part of \eqref{Eq_DDE} involves a distributed-delay term as in \eqref{Eq_nln_sys}. As a preparation for the forthcoming analytic derivations, we need to express the derivative of the Koornwinder polynomials in terms of the polynomials themselves. This is the content of the following proposition. \begin{prop} \label{prop:dPn} The Koornwinder polynomial $K_n$ of degree $n\in \mathbb{N}$ defined in \eqref{eq:Pn} satisfies the differential relation \begin{equation} \label{eq:dPn} \frac{\mathrm{d} K_n}{\mathrm{d} s}(s) = \sum_{k = 0}^{n-1} a_{n,k} K_k(s), \quad s \in (-1,1), \end{equation} where the coefficients $\boldsymbol{a}_n:=(a_{n,0}, \cdots, a_{n,n-1})^\mathrm{tr}$, satisfy the upper triangular system of linear equations \begin{equation} \label{eq:algebraic} \mathbf{T}\boldsymbol{a}_n = \boldsymbol{b}_n, \end{equation} with $\mathbf{T}:=(\mathbf{T}_{i,j})_{n\times n}$ and $b_{n}:=(b_{n,0}, \cdots, b_{n,n-1})^\mathrm{tr}$ given by \begin{equation} \begin{aligned} \label{eq:algebraic_def} \mathbf{T}_{i,j} & = \begin{cases} 0, & \; \text{ if } j < i,\\ i^2 + 1, & \; \text{ if } j = i,\\ -(2i+1), & \; \text{ if } j > i, \end{cases} \; \qquad \text{ where } \qquad 0 \le i, j \le n-1, \\ b_{n,i} & = \begin{cases} \frac{1}{2}(2i+1)(n+i+1)(n-i), & \text{ if $n+i$ is even}, \vspace{1em}\\ (n^2 + n)(2i+1) - \frac{i}{2}(n+i)(n-i+1) & \\ \hspace{2em} -\frac{ i}{2}(i+1)(n-i-1)(n+i+2), & \text{ if $n+i$ is odd}. \end{cases} \end{aligned} \end{equation} For the rescaled version $K^\tau_n$ defined by \eqref{eq:Pn_tilde}, it holds that \begin{equation} \label{eq:dKn_tau} \frac{\mathrm{d} K^\tau_n}{\mathrm{d} \theta}(\theta) = \frac{2}{\tau} \sum_{k = 0}^{n-1} a_{n,k} K^\tau_k(\theta), \qquad n \in \mathbb{N}, \quad \theta\in (-\tau,0). \end{equation} \end{prop} \begin{proof} See Appendix \ref{sect:coef_matrix_proof}. \end{proof} Let us now rewrite the unknown $u_N$ in the Galerkin system \eqref{Eq_DDE_Galerkin} in terms of the first $N$ rescaled Koornwinder polynomials, i.e.: \begin{equation} \label{x_t expand} u_N(t) = \sum_{n=0}^{N-1} y_n(t) \mathcal{K}^\tau_n, \qquad t \ge 0, \end{equation} where \begin{equation} y_n(t) = \frac{\langle u_N(t), \mathcal{K}^\tau_n \rangle_{\mathcal{H}}}{\| \mathcal{K}^\tau_n \|^2_{\mathcal{H}}}. \end{equation} We then replace $u_N$ in Eq.~\eqref{Eq_DDE_Galerkin} by the expansion given in \eqref{x_t expand}, and take the $\mathcal{H}$-inner product on both sides with $\mathcal{K}^\tau_j$ for each $j \in \{0, \cdots, N-1 \}$ to obtain: \begin{equation} \label{GalerkinCalc_v1} \|\mathcal{K}^\tau_j\|_{\mathcal{H}}^2 \frac{\mathrm{d} y_j}{\mathrm{d} t} = \sum_{n=0}^{N-1} y_n(t) \left \langle \mathcal{A}_N \mathcal{K}^\tau_n, \mathcal{K}^\tau_j \right \rangle_{\mathcal{H}} + \Bigg \langle \Pi_N \mathcal{F} \Bigg ( \sum_{n=0}^{N-1} y_n(t) \mathcal{K}^\tau_n \Bigg), \mathcal{K}^\tau_j \Bigg \rangle_{\mathcal{H}}. \end{equation} Recall that the linear operator $\mathcal{A}$ here is defined by \eqref{Def_A}. Then, for each $n \in \{0, \cdots, N-1 \}$, it holds that \begin{equation} \begin{aligned} \mathcal{A}_N \mathcal{K}^\tau_n = \Pi_N \mathcal{A} \mathcal{K}^\tau_n & = \sum_{l = 0}^{N-1} \Big(\frac{1}{\tau} \Big \langle \frac{\mathrm{d}^+}{\mathrm{d} \theta} K_n^\tau, K_l^\tau \Big \rangle_{L^2} \\ & \qquad + \Big(a K^\tau_n(0) + b K_n^\tau(-\tau) + c \int_{-\tau}^0 K_n^\tau(\theta) \mathrm{d} \theta \Big) \Big) \frac{\mathcal{K}_l^\tau }{\|\mathcal{K}_l^\tau \|_{\mathcal{H}}^2}. \end{aligned} \end{equation} We obtain then that \begin{equation} \begin{aligned} \label{GalerkinCalc_part1_1} \left \langle \mathcal{A}_N \mathcal{K}^\tau_n, \mathcal{K}^\tau_j \right \rangle_{\mathcal{H}} & = \frac{1}{\tau} \Big \langle \frac{\mathrm{d}^+}{\mathrm{d} \theta} K_n^\tau, K_j^\tau \Big \rangle_{L^2} \\ & \quad + \Big( a K^\tau_n(0) + b K^\tau_n(-\tau) + c \int_{-\tau}^0 K_n^\tau(\theta) \mathrm{d} \theta\Big) K^\tau_j(0). \end{aligned} \end{equation} It follows from the expression of $\frac{\mathrm{d} K^\tau_n}{\mathrm{d} \theta}$ in \eqref{eq:dKn_tau} that \begin{equation} \begin{aligned} \label{GalerkinCalc_part1_2} \frac{1}{\tau} \int_{-\tau}^0 \frac{\mathrm{d}^+ K^\tau_n}{\mathrm{d} \theta}(\theta) K^\tau_j(\theta)\mathrm{d} \theta & = \frac{2}{\tau}\sum_{k=0}^{n-1} a_{n,k} \left( \frac{1}{\tau} \int_{-\tau}^0 K^\tau_k(\theta) K^\tau_j(\theta)\mathrm{d} \theta \right) \\ & = \frac{2}{\tau}\sum_{k=0}^{n-1} a_{n,k} \left( \langle \mathcal{K}^\tau_k, \mathcal{K}^\tau_j \rangle_{\mathcal{H}} - K^\tau_k(0) K^\tau_j(0) \right ) \\ & = \frac{2}{\tau} \sum_{k=0}^{n-1} a_{n,k} \left( \delta_{j,k} \|\mathcal{K}^\tau_j\|^2_{\mathcal{H}} - 1 \right ), \end{aligned} \end{equation} where $\delta_{j,k}$ denotes the Kronecker delta, and the last equality above follows from the orthogonal property of the Koornwinder polynomials as well as the normalization property $K^\tau_n(0) = 1$; cf.~\eqref{Eq_normalization}. Note that $K^\tau_0 \equiv 1$, which follows from the definition of the Koornwinder polynomials $K_n$ given by \eqref{eq:Pn} and the fact that the first Legendre polynomial $L_0$ is identically $1$. We get then $\|\mathcal{K}^\tau_0\|^2_{\mathcal{H}} = 2$. Note also that \begin{equation} \langle \mathcal{K}^\tau_n, \mathcal{K}^\tau_0 \rangle_{\mathcal{H}} = \frac{1}{\tau} \int_{-\tau}^0 K^\tau_n(\theta) \mathrm{d} \theta + 1 = \delta_{n,0} \|\mathcal{K}^\tau_0\|^2_{\mathcal{H}} = 2 \delta_{n,0}, \end{equation} leading thus to \begin{equation} \label{Eq_int_Kn} \int_{-\tau}^0 K^\tau_n(\theta) \mathrm{d} \theta = \tau (2 \delta_{n,0} - 1), \quad n \in \mathbb{N}. \end{equation} By using \eqref{Eq_int_Kn}, the normalization property $K^\tau_j(0) = 1$ and the identity $K^\tau_j(-\tau) = K_j(-1)$ (valid for any $j \ge 0$), we obtain \begin{equation} \label{GalerkinCalc_part1_3} \Big( a K^\tau_n(0) + b K^\tau_n(-\tau) + c \int_{-\tau}^0 K_n^\tau(\theta) \mathrm{d} \theta \Big) K^\tau_j(0) = a + b K_n(-1) + c \tau (2 \delta_{n,0} - 1). \end{equation} Now, by using \eqref{GalerkinCalc_part1_2} and \eqref{GalerkinCalc_part1_3} in \eqref{GalerkinCalc_part1_1}, we obtain \begin{equation} \begin{aligned} \label{GalerkinCalc_part1} \sum_{n=0}^{N-1} y_n(t) \left \langle \mathcal{A}_N \mathcal{K}^\tau_n, \mathcal{K}^\tau_j \right \rangle_{\mathcal{H}} & = \sum_{n=0}^{N-1} \Bigl ( a + b K_n(-1) + c \tau (2 \delta_{n,0} - 1) \\ & \hspace{5em} + \frac{2}{\tau} \sum_{k=0}^{n-1} a_{n,k} \left( \delta_{j,k} \|\mathcal{K}_j\|^2_{\mathcal{H}} - 1 \right ) \Bigr) y_n(t). \end{aligned} \end{equation} For the nonlinear part, since $\langle \Pi_N \Phi, \mathcal{K}^\tau_n \rangle_{\mathcal{H}} = \langle \Phi, \mathcal{K}^\tau_n \rangle_{\mathcal{H}}$ for all $\Phi \in \mathcal{H}$ and all $n \in \{0, \cdots, N-1\}$, together with the definition of $\mathcal{F}$ given in \eqref{Def_F}, we obtain \begin{equation} \label{GalerkinCalc_part2a} \Bigl \langle \Pi_N \mathcal{F} \Bigl ( \sum_{n=0}^{N-1} y_n(t) \mathcal{K}^\tau_n \Bigr ), \mathcal{K}^\tau_j \Bigr \rangle_{\mathcal{H}} = F \Biggl( \sum_{n=0}^{N-1} y_n(t), \int_{-\tau}^0 \sum_{n=0}^{N-1} y_n(t) K^\tau_n(\theta) \mathrm{d} \theta \Biggr). \end{equation} From \eqref{Eq_int_Kn}, it also follows that \begin{equation} \label{GalerkinCalc_part2b} \int_{-\tau}^0 \sum_{n=0}^{N-1} y_n(t) K^\tau_n(\theta) \mathrm{d} \theta = \tau y_0(t) - \tau \sum_{n=1}^{N-1} y_n(t). \end{equation} By using \eqref{GalerkinCalc_part2b} in \eqref{GalerkinCalc_part2a}, we get \begin{equation} \label{GalerkinCalc_part2} \Bigl \langle \Pi_N \mathcal{F} \Bigl ( \sum_{n=0}^{N-1} y_n(t) \mathcal{K}^\tau_n \Bigr ), \mathcal{K}^\tau_j \Bigr \rangle_{\mathcal{H}} = F \Biggl( \sum_{n=0}^{N-1} y_n(t), \tau y_0(t) - \tau \sum_{n=1}^{N-1} y_n(t) \Biggr). \end{equation} Now, by using \eqref{GalerkinCalc_part1} and \eqref{GalerkinCalc_part2} and recalling that $\|\mathcal{K}^\tau_n\|_{\mathcal{H}} = \|\mathcal{K}_n\|_{\mathcal{E}}$, we obtain from Eq.~\eqref{GalerkinCalc_v1} the following explicit form of the $N$-dimensional Galerkin system \eqref{Eq_DDE_Galerkin}: \begin{equation} \label{Galerkin_AnalForm} \begin{aligned} \frac{\mathrm{d} y_j}{\mathrm{d} t} & = \frac{1}{\|\mathcal{K}_j\|_{\mathcal{E}}^2 } \sum_{n=0}^{N-1} \Big( a + b K_n(-1) + c \tau (2 \delta_{n,0} - 1) \\ & \hspace{8em} + \frac{2}{\tau}\sum_{k=0}^{n-1} a_{n,k} \left( \delta_{j,k} \|\mathcal{K}_j\|^2_{\mathcal{E}} - 1 \right) \Big) y_n(t) \\ & \hspace{1em} + \frac{1}{\|\mathcal{K}_j\|_{\mathcal{E}}^2} F \left( \sum_{n=0}^{N-1} y_n(t), \tau y_0(t) - \tau \sum_{n=1}^{N-1} y_n(t) \right), \;\; 0\leq j\leq N-1. \end{aligned} \end{equation} For later usage, we rewrite the above Galerkin system into the following compact form: \begin{equation} \label{Galerkin_cptForm} \boxed{\frac{\mathrm{d} \boldsymbol{y}}{\mathrm{d} t} = A \boldsymbol{y} + G (\boldsymbol{y}),} \end{equation} where $A \boldsymbol{y}$ denotes the linear part of Eq.~\eqref{Galerkin_AnalForm}, and $G(\boldsymbol{y})$ the nonlinear part. Namely, $A$ is the $N\times N$ matrix whose elements are given by \begin{equation} \label{eq:A} \boxed{ \begin{aligned} (A)_{j,n} & = \frac{1}{\|\mathcal{K}_j\|_{\mathcal{E}}^2 }\Big(a + b K_n(-1) + c \tau (2 \delta_{n,0} - 1) \\ & \hspace{8em}+ \frac{2}{\tau}\sum_{k=0}^{n-1} a_{n,k} \left( \delta_{j,k} \|\mathcal{K}_j\|^2_{\mathcal{E}} - 1 \right ) \Big), \end{aligned} } \end{equation} where $j, n = 0, \cdots, N-1$, and the nonlinear vector field $G \colon \mathbb{R}^N \rightarrow \mathbb{R}^N$, is given component-wisely by \begin{equation} \label{eq:G} \boxed{G_j(\boldsymbol{y}) = \frac{1}{\|\mathcal{K}_{j}\|_{\mathcal{E}}^2} F \left( \sum_{n=0}^{N-1} y_n(t), \tau y_0(t) - \tau \sum_{n=1}^{N-1} y_n(t) \right), \; 0\leq j\leq N-1.} \end{equation} \medskip \begin{rem} \label{Rmk_Galerkin_forcing} When a time-dependent forcing $g(t)$ is added to the RHS of the DDE \eqref{Eq_DDE}, the only change in the corresponding Galerkin system is to add a term $\|\mathcal{K}_{j}\|_{\mathcal{E}}^{-2} g(t)$ to each $G_j$ in \eqref{eq:G}. Recall also from Remark~\ref{Rmk_forcing} that it is sufficient to require that $g\in L_{\mathrm{loc}}^2(\mathbb{R}; \mathbb{R})$ in order to ensure the convergence result of the Galerkin system over any finite interval. \qed \end{rem} \section{Approximation of chaotic dynamics: Numerical results} \label{Sect_Numerics} In this section, we report on the performance of our Galerkin schemein approximating quasi-periodic and chaotic dynamics. In the case of the latter, it is well known that any finite-time uniform convergence result --- such as the one given by \eqref{uniform_conv_est_engest_onF} and obtained in Corollary \ref{Cor_DDE_local_Lip_Case2} --- becomes less useful, due to sensitivity to initial data. In this case, when individual solutions diverge exponentially, it is natural to consider instead the approximation of the strange attractor or of the statistics of the dynamics. More generally, the approximation of meaningful invariant probability measures supported by the strange attractor is of primary interest in chaotic dynamics. Nonlinear DDEs, like those considered here, are known to support such probability measures once a global attractor is known to exist \cite{chekroun_glatt-holtz}. If, in addition to the finite-time uniform convergence \eqref{uniform_conv_est_engest_onF}, a uniform dissipativity assumption is satisfied, then any sequence of invariant measures associated with the Galerkin approximation converges weakly to an invariant measure of the full system; see \cite[Thm.~2.2]{wang2009approximating}. The uniform dissipativity assumption referred to in \cite{wang2009approximating} is satisfied if one can establish, uniformly in $N$, the existence of an absorbing ball for the Galerkin reduced systems in another separable Hilbert space $\mathcal{V}$, which is compactly imbedded in $\mathcal{H}$, in the case of strongly dissipative systems. We show in Sections~\ref{Sec_nearly-brownian} and \ref{Sect_bimodal} below that, for two simple nonlinear scalar DDEs that --- even in instances in which the uniform dissipativity assumption of \cite[Thm.~2.2]{wang2009approximating} is not guaranteed --- our Galerkin scheme is still able to approximate significant statistical properties of the chaotic dynamics, or the strange attractor itself. Finally, we illustrate in Section~\ref{Sect_ENSO}, {\mg for a highly idealized ENSO model from climate dynamics} that our approach also works in the case of periodically forced DDEs with multiple delays. \subsection{``Nearly-Brownian'' chaotic dynamics.}\label{Sec_nearly-brownian} We consider here a modified version of the DDE analyzed in \cite{sprott2007simple} with a sinusoidal nonlinearity. The modification consists of replacing the discrete delay by distributed delays. As we will see, this modification does not affect the main dynamical properties identified in \cite{sprott2007simple} in the case of a discrete delay, once the proper parameter values are chosen. More precisely, we consider the following DDE \begin{equation}\label{Eq_sindel} \dot{x}=a\sin\Big(\int_{t-\tau}^tx(s) \mathrm{d} s\Big), \end{equation} with $a=0.5$ and $\tau=5.5$. This example fits within the general class discussed in Section \ref{sec_ex1}, to which the rigorous convergence results described in Sections \ref{Subsect_DDE_Galerkin} in an abstract setting, and in Section \ref{Sec_Galerkin_analytic} more concretely, do apply. As pointed out in the introduction of this section, such finite-time uniform convergence results are essential in general, but they are not the ones we are necessarily looking for in approximating chaotic dynamics. We rely, therefore, on careful numerical simulations to explore the performance of our Koornwinder-polynomial--based Galerkin systems to approximate the statistical features of the chaotic dynamics in these examples. A sample trajectory of Eq.~\eqref{Eq_sindel} is shown in black in Fig.~\ref{Trajectories_sindel}. While the governing DDE is perfectly deterministic, the visual resemblance of this trajectory to a sample path of Brownian motion is obvious. A trajectory with the same constant initial history over the interval $[-\tau, 0)$, but obtained by solving a 10D-approximation by our Galerkin scheme of the DDE, is shown as the red curve in the figure. It is clear that individual trajectories do have the same overall behavior, which is nearly Brownian, but the pointwise approximation of the exact trajectory by the approximate one is not good. To study instead the statistics of the solutions, we have estimated --- for a collection of $n=10^4$ initial histories of constant value drawn uniformly in $[-1,1]$ --- the empirical probability density function (PDF) of the corresponding $x(t)$-values at a given time $t$. Figure \ref{Sine_model_statistics} reports on the numerical results at $t=2680$ when $x(t)$ is simulated from Eq.~\eqref{Eq_sindel} by forward Euler integration (black curve) with a time step of $\Delta t = \tau/2^{10}$, and when $x(t)$ is approximated by a 10D-Galerkin approximation (red curve) obtained from the analytic formulas \eqref{Galerkin_cptForm}--\eqref{eq:G} applied to Eq.~\eqref{Eq_sindel}. The Galerkin system {\mg of ODEs} is integrated using a semi-implicit Euler method that still uses $\Delta t = \tau/2^{10}$, but in which the linear part $A y$ is treated implicitly, {\mg while} the nonlinear term $G(y)$ is treated explicitly. The approximate solution $x_N(t)$, provided by an $N$-dimensional Galerkin system of the form \eqref{Galerkin_cptForm}, is obtained by using the expansion \eqref{x_t expand} into Koornwinder polynomials. More precisely, $x_N(t)$ is obtained as the state part of $u_N$ given by \eqref{x_t expand} which, thanks to the normalization property $K_n^\tau(0) = 1$ given in ~\eqref{Eq_normalization}, reduces to \begin{equation} x_N(t) = \sum_{j = 1}^N y_j(t), \end{equation} where $y:=(y_0, \cdots, y_{N-1})$ is the vector solution of \eqref{Galerkin_cptForm}. Also shown in blue in panels (a, c) of Fig.~\ref{Sine_model_statistics}, is a Gaussian distribution with the same mean and standard deviation, both of which are estimated from the $x(t)$-values at $t=2680$ of the simulated DDE solutions. In both cases, the empirical distributions, as obtained from the simulated DDE solution or as obtained by using a Galerkin approximation with $N=10$, closely follow a Gaussian law. \begin{figure}[hbtp] \centering \includegraphics[height=0.4\textwidth,width=.9\textwidth]{Trajectories_sindel.pdf} \caption{{\footnotesize {\mg Trajectory $x(t)$ simulated by} the DDE \eqref{Eq_sindel} (black curve), and its 10D-Galerkin approximation {\mg $x_N(t)$, with $N = 10$} (red curve), obtained by the method described in Section~\ref{Sec_Galerkin_analytic}.}} \label{Trajectories_sindel} \end{figure} \begin{figure}[hbtp] \centering \includegraphics[height=0.7\textwidth,width=.9\textwidth]{sine_statistics_1E4runs.pdf} \caption{{\footnotesize {\mg (a, c):} Probability density functions (PDFs); and (b, d): standard deviations $\sigma$, as estimated from the simulations of the DDE \eqref{Eq_sindel} (black curve {\mg and black circles, in the left panels and the right panels, respectively),} and its 10D-Galerkin approximation (red curve {\mg and red circles)} by the method described in Section~\ref{Sec_Galerkin_analytic}, respectively. In each of the panels (a) and (b), the results are compared with a Gaussian distribution estimated by standard analytic formulas (blue curves), and in each of the panels (c) and (d), the results are compared with the best linear regression fit (blue curve) of $\log(\sigma)$ versus $\log(t)$, providing the corresponding slope and the exponent reported therein.}} \label{Sine_model_statistics} \end{figure} For both $x(t)$, as simulated by numerically solving Eq.~\eqref{Eq_sindel}, and $x_N(t)$, as obtained by using a Galerkin approximation with $N=10$, the standard deviations of the corresponding collection of trajectories are shown versus time in panels (b) and (d) of Fig.~\ref{Sine_model_statistics}. The best linear regression fit of $\log(\sigma)$ versus $\log(t)$ gives $\sigma=0.195 t^{0.479}$, in the case of Eq.~\eqref{Eq_sindel}, and $\sigma=0.190 t^{0.480}$, in the case of the 10D-Galerkin approximation. In both cases, the slopes of the fitted lines indicate that the {\mg deterministic dynamics of Eq.~\eqref{Eq_sindel}} mimics that of Brownian motion, for which the slope would be $0.5$, with a diffusion coefficient $D=\sigma^2/t$. In particular, the solutions do not stay within any bounded subset; the dynamics thus violates any dissipation criterion, while still possessing an attractor with strange behavior (not shown). To the best of our knowledge, the 10D-Galerkin approximation computed here {\mg thus provides} the first example of a chaotic system of ODEs whose solutions exhibit a statistical behavior close to that of Brownian motion. \subsection{Bimodal chaotic dynamics with low-frequency variability.} \label{Sect_bimodal} The model studied in this subsection, \begin{equation}\label{Eq_DDE_chafee} \dot{x}=a x(t-\tau) -b x(t-\tau)^3, \end{equation} is also based on \cite{sprott2007simple}; the parameter values used here are $a=0.5$, $b=20$ and $\tau=3.35$. Remark~\ref{Rmk_Chafee} at the end of this subsection discusses similarities between the dynamics of the model above and important aspects of ENSO dynamics. \begin{figure}[hbtp] \centering \includegraphics[height=0.5\textwidth,width=.9\textwidth]{Super_chafee_6D_6e6_half.pdf} \caption{{\footnotesize The attractor associated with Eq.~\eqref{Eq_DDE_chafee} (left panel) and its approximation obtained from a 6D-Galerkin approximation (right panel).}} \label{Super_chafee} \end{figure} \begin{figure}[hbtp] \centering \includegraphics[height=0.4\textwidth,width=.8\textwidth]{PSD_chafee.pdf} \caption{{\footnotesize Power spectral density of $x(t)$ as simulated from Eq.~\eqref{Eq_DDE_chafee}, and from a 6D-Galerkin approximation derived from the analytic formulas described in Section \ref{Sec_Galerkin_analytic} and applied to Eq.~\eqref{Eq_DDE_chafee}.}} \label{PSD_chafee} \end{figure} In contrast to the DDE considered in the previous {\mg subsection,} Eq.~\eqref{Eq_DDE_chafee} does not fit directly within the general framework of Section \ref{Subsect_DDE_Galerkin}, for which rigorous convergence results are available. The discrete lag present in the cubic term leads to complications for a rigorous analysis. Replacing this lag effect by a distributed one, as in Eq.~\eqref{Eq_sindel}, would place the DDE of Eq.~\eqref{Eq_DDE_chafee} into the class considered in Section~\ref{sec_ex2}, for which finite-time uniform convergence results are guaranteed. But even then, we cannot be assured {\it a priori} of an effective approximation of statistical features of the dynamics, as discussed above. The purpose of this subsection is to show that, even in such a borderline case with respect to the theory presented in this article, statistical and even topological features can still be remarkably well approximated by the Galerkin systems of Section \ref{Sec_Galerkin_analytic}, when appropriately modified to handle the case of discrete delay in the nonlinear terms.\footnote{This modification consists just of noting that, by replacing the nonlinear term with distributed delays in \eqref{Eq_DDE} by $F(x(t), x(t-\tau))$, the identity \eqref{GalerkinCalc_part2a} becomes \begin{equation} \Bigl \langle \Pi_N \mathcal{F} \Bigl ( \sum_{n=0}^{N-1} y_n(t) \mathcal{K}^\tau_n \Bigr ), \mathcal{K}^\tau_j \Bigr \rangle_{\mathcal{H}} = F \Biggl( \sum_{n=0}^{N-1} y_n(t), \sum_{n=0}^{N-1} y_n(t) K_n(-1) \Biggr); \end{equation} {\mg the latter, in turn,} leads to the corresponding change in the formula \eqref{eq:G} for the nonlinear vector field in Eq.~\eqref{Galerkin_cptForm}. \label{footnote_discrete_delay}} As shown in the left panel of Figure \ref{Super_chafee} in natural delayed coordinates, the strange attractor associated with Eq.~\eqref{Eq_DDE_chafee} (black) has a nearly symmetric topological shape and is constituted by two fairly high-density``islands'' connected by a foliation of heteroclinic-like orbits. These properties of the attractor are very well captured by a 6D-Galerkin approximation (right panel, red). These two high-density islands, along with the lower-density areas of heteroclinic-like connections {\mg between the two,} give rise to a bimodal chaotic behavior, which gives rise, in turn, to an interesting time variability. Figure ~\ref{PSD_chafee} plots the results of a standard numerical estimation of the spectral density --- also known in the engineering literature as the power spectrum --- associated with $x(t)$, as directly simulated using Eq.~\eqref{Eq_DDE_chafee} (black curve) and as approximated by a 6D-Galerkin approximation (red curve). In both cases, the power spectra are estimated from the corresponding autocorrelation functions\cite{eckmann_ruelle,Ghil2002}. The numerical results show that the spectrum of $x(t)$ contains two broadband peaks at low frequencies that stand out above an exponentially decaying background. As shown in Fig.~\ref{PSD_chafee}, these two peaks, as well as the exponential background, are strikingly well approximated by a 6D-Galerkin approximation, in both amplitude and frequency, as well as in the rate of decay for high frequencies. The approximation by a truly low-dimensional, 6D-Galerkin model of these key features of the power spectrum of the solutions to the DDE~\eqref{Eq_DDE_chafee} has deep dynamical implications in terms of the Ruelle-Pollicott resonances and mixing properties of the dynamics on the attractor \cite{Chek_al14_RP}. These implications are beyond the scope of this article but {\mg we intend to discuss them} elsewhere. \begin{rem} \label{Rmk_Chafee} Equation.~\eqref{Eq_DDE_chafee} can actually be seen as a highly idealized ENSO model with memory effects; see \cite{BH89,Cane_al90,Dijkstra05,GCStep15,Galanti_al00, GZT08, Sieber2014, Munnich_al91, TSCJ94,Zaliapin_al10,Zivkovic_al13} and references therein. Indeed, for the given parameter values, the solution of Eq.~\eqref{Eq_DDE_chafee} admits two metastable states, as can be seen from the two islands in the attractor given in Figure~\ref{Super_chafee}. These two metastable states {\mg are analogous to the warm, El Ni\~no and the cold, La Ni\~na} states in ENSO dynamics. Moreover, the two broadband peaks at low frequencies in the spectral density of the solution, as shown in Figure~\ref{PSD_chafee}, are also reminiscent of the quasi-quadrennial and the quasi-biennial mode in ENSO \cite{MG_AWR'00, Ghil2002, Jiang95} dynamics. The important role of such low-frequency variability in the understanding and prediction of climate dynamics on various time scales was emphasized in \cite{CKG11,HD_MG'05, Ghil2001, MG_AWR'00, MG_AWR'02} and references therein. \qed \end{rem} \subsection{A highly idealized ENSO model with memory effects} \label{Sect_ENSO} In this section, we consider the following {\mg periodically} forced DDE with two discrete delays \cite{GZT08}: \begin{equation} \label{ENSO_model} \dot{x}= -\alpha \tanh(\kappa x(t-\tau_1)) + \beta \tanh(\kappa x(t-\tau_2)) + \gamma \cos(2\pi t). \end{equation} This equation is a slightly modified version of the model used in \cite{TSCJ94} for the study of the interaction between the seasonal forcing and the intrinsic ENSO variability.\footnote{The nonlinearity used in \cite{TSCJ94} consists of a sigmoid made of two $\tanh$ segments joined continuously by a line segment, cf. \cite[Eq.~(9)]{Munnich_al91}; this sigmoid is simplified here to be just a hyperbolic tangent function.} We also refer to \cite{BH89,Cane_al90,Dijkstra05,GCStep15,Galanti_al00,GZT08, Munnich_al91,TSCJ94,Zaliapin_al10,Zivkovic_al13,GZ15} and references therein for other models with retarded arguments used in this context. The purpose of this subsection is to show that the Galerkin scheme developed in this article can also be easily adapted to deal with DDEs with multiple delays, as well as with non-autonomous, forced DDEs. For this purpose, Eq.~\eqref{ENSO_model} is placed in a quasi-periodic regime {\mg by choosing the parameter values $ \alpha = 2.1, \beta = 1.05, \gamma =3, \kappa = 10, \tau_1 = 0.95,$ and $\tau_2 = 5.13$.} We outline now the necessary modifications to the nonlinear term $G(y)$ in the Galerkin system \eqref{Galerkin_cptForm} for the case of multiple discrete delays. As a direct generalization of the case with a single discrete delay, given in footnote~\ref{footnote_discrete_delay}, the nonlinearity $F$ in \eqref{Eq_DDE} takes the form $F(x(t), x(t-\tau_1), \cdots, x(t-\tau_p))$, where $0 < \tau_1 < \cdots < \tau_p=:\tau$. In this more general situation, the identity \eqref{GalerkinCalc_part2a} becomes \begin{equation} \begin{aligned} \hspace*{-1em}\Bigl \langle \Pi_N \mathcal{F} \Bigl ( \sum_{n=0}^{N-1} y_n(t) \mathcal{K}^\tau_n \Bigr ), \mathcal{K}^\tau_j \Bigr \rangle_{\mathcal{H}} & = F \Biggl( \sum_{n=0}^{N-1} y_n(t), \sum_{n=0}^{N-1} y_n(t) K_n(-\frac{\tau_1}{\tau}), \cdots, \\ & \qquad \sum_{n=0}^{N-1} y_n(t) K_n(-\frac{\tau_{p-1}}{\tau}), \sum_{n=0}^{N-1} y_n(t) K_n(-1) \Biggr), \end{aligned} \end{equation} which leads to the corresponding change in the formula \eqref{eq:G} for the nonlinear vector field in Eq.~\eqref{Galerkin_cptForm}. Note also that the forcing term is dealt with in Remark~\ref{Rmk_Galerkin_forcing}. Again, we compare the DDE's attractor in the left panel of Fig.~\ref{Fig_ENSO_attractors} with the attractor obtained from an associated Galerkin system, in the right panel. Despite the complexity of the DDE's attractor, a $40$-dimensional Galerkin system can already provide a very accurate reconstruction of this attractor. The need for a higher dimensionality of the Galerkin approximation in this case arises from the presence of incommensurable frequencies in the periodically forced model, namely the seasonal cycle and the internal frequencies \cite{Dijkstra05, JNG'94, MG_AWR'00, MG_AWR'02, TSCJ94}. \begin{figure}[hbtp] \centering \includegraphics[height=0.4\textwidth,width=.8\textwidth]{tanh_2delays_attractor.pdf} \caption{{\footnotesize The attractor associated with Eq.~\eqref{ENSO_model} (left panel) and its approximation obtained from a 40D-Galerkin approximation (right panel).}} \label{Fig_ENSO_attractors} \end{figure} \section*{Acknowledgments} This work has been partially supported by the Office of Naval Research (ONR) Multidisciplinary University Research Initiative (MURI) grant N00014-12-1-0911 (MDC, HL and MG) and by National Science Foundation (NSF) grant DMS-1049253 (SW).
{ "timestamp": "2015-09-11T02:00:57", "yymm": "1509", "arxiv_id": "1509.02945", "language": "en", "url": "https://arxiv.org/abs/1509.02945", "abstract": "This article revisits the approximation problem of systems of nonlinear delay differential equations (DDEs) by a set of ordinary differential equations (ODEs). We work in Hilbert spaces endowed with a natural inner product including a point mass, and introduce polynomials orthogonal with respect to such an inner product that live in the domain of the linear operator associated with the underlying DDE. These polynomials are then used to design a general Galerkin scheme for which we derive rigorous convergence results and show that it can be numerically implemented via simple analytic formulas. The scheme so obtained is applied to three nonlinear DDEs, two autonomous and one forced: (i) a simple DDE with distributed delays whose solutions recall Brownian motion; (ii) a DDE with a discrete delay that exhibits bimodal and chaotic dynamics; and (iii) a periodically forced DDE with two discrete delays arising in climate dynamics. In all three cases, the Galerkin scheme introduced in this article provides a good approximation by low-dimensional ODE systems of the DDE's strange attractor, as well as of the statistical features that characterize its nonlinear dynamics.", "subjects": "Chaotic Dynamics (nlin.CD); Classical Analysis and ODEs (math.CA); Dynamical Systems (math.DS)", "title": "Low-Dimensional Galerkin Approximations of Nonlinear Delay Differential Equations", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.983342958526322, "lm_q2_score": 0.8152324938410783, "lm_q1q2_score": 0.8016531323804775 }
https://arxiv.org/abs/0710.0829
Computing the Conditioning of the Components of a Linear Least Squares Solution
In this paper, we address the accuracy of the results for the overdetermined full rank linear least squares problem. We recall theoretical results obtained in Arioli, Baboulin and Gratton, SIMAX 29(2):413--433, 2007, on conditioning of the least squares solution and the components of the solution when the matrix perturbations are measured in Frobenius or spectral norms. Then we define computable estimates for these condition numbers and we interpret them in terms of statistical quantities. In particular, we show that, in the classical linear statistical model, the ratio of the variance of one component of the solution by the variance of the right-hand side is exactly the condition number of this solution component when perturbations on the right-hand side are considered. We also provide fragment codes using LAPACK routines to compute the variance-covariance matrix and the least squares conditioning and we give the corresponding computational cost. Finally we present a small historical numerical example that was used by Laplace in Theorie Analytique des Probabilites, 1820, for computing the mass of Jupiter and experiments from the space industry with real physical data.
\section{Introduction} We consider the linear least squares problem (LLSP) $\min_{x \in \mathbb{R}^n}\|Ax-b\|_2$, where $b \in \mathbb{R}^m$ and $A \in \mathbb{R}^{m\times n}$ is a matrix of full column rank $n$.\\ Our concern comes from the following observation: in many parameter estimation problems, there may be random errors in the observation vector $b$ due to instrumental measurements as well as roundoff errors in the algorithms. The matrix $A$ may be subject to errors in its computation (approximation and/or roundoff errors). In such cases, while the condition number of the matrix $A$ provides some information about the sensitivity of the LLSP to perturbations, a single global conditioning quantity is often not relevant enough since we may have significant disparity between the errors in the solution components. We refer to the last section of the manuscript for illustrative examples.\\ There are several results for analyzing the accuracy of the LLSP by components. For linear systems $Ax=b$ and for LLSP, ~\cite{CHA.IPS.95} defines so called componentwise condition numbers that correspond to amplification factors of the relative errors in solution components due to perturbations in data $A$ or $b$ and explains how to estimate them. For LLSP,~\cite{KEN.LAU.98} proposes to estimate componentwise condition numbers by a statistical method. More recently,~\cite{ABG.07} developed theoretical results on conditioning of linear functionals of LLSP solutions.\\ The main objective of our paper is to provide computable quantities of these theoretical values in order to assess the accuracy of an LLSP solution or some of its components. To achieve this goal, traditional tools for the numerical linear algebra practitioner are condition numbers or backward errors whereas the statistician usually refers to variance or covariance. Our purpose here is to show that these mathematical quantities coming either from numerical analysis or statistics are closely related. In particular, we will show in Equation~(\ref{eq:1}) that, in the classical linear statistical model, the ratio of the variance of one component of the solution by the variance of the right-hand side is exactly the condition number of this component when perturbations on the right-hand side are considered. In that sense, we attempt to clarify, similarly to~\cite{HIGH.STEW.87}, the analogy between quantities handled by the linear algebra and the statistical approaches in linear least squares. Then we define computable estimates for these quantities and explain how they can be computed using the standard libraries LAPACK or ScaLAPACK. This paper is organized as follows. In Section~\ref{sec:theoback}, we recall and exploit some results of practical interest coming from~\cite{ABG.07}. We also define the condition numbers of an LLSP solution or one component of it. In Section~\ref{sec:link}, we recall some definitions and results related to the linear statistical model for LLSP, and we interpret the condition numbers in terms of statistical quantities. In Section~\ref{sec:lapack} we provide practical formulas and FORTRAN code fragments for computing the variance-covariance matrix and LLSP condition numbers using LAPACK. In Section~\ref{sec:numerics}, we propose two numerical examples that show the relevance of the proposed quantities and their practical computation. The first test case is a historical example from Laplace and the second example is related to gravity field computations. Finally some concluding remarks are given in Section~\ref{sec:concl}.\\ Throughout this paper we will use the following notations. We use the Frobenius norm $\nfro{.}$ and the spectral norm $\neuc{.}$ on matrices and the usual Euclidean norm $\neuc{.}$ on vectors. $A^{\dagger}$ denotes the Moore-Penrose pseudo inverse of $A$, the matrix $I$ is the identity matrix and $e_i$ is the $i$-th canonical vector. \section{Theoretical background for linear least squares conditioning}\label{sec:theoback} Following the notations in~\cite{ABG.07}, we consider the function \begin{equation} \label{funct} \begin{array}{c c c c} g\ : & \mathbb{R} ^{m \times n} \times \mathbb{R} ^m & \longrightarrow & \mathbb{R} ^k \\ & A,b & \longmapsto & g(A,b)=L^{T}x(A,b)=L^T(A^TA)^{-1}A^Tb,\\ \end{array} \end{equation} where $L$ is an $n \times k$ matrix, with $k \leq n$. Since $A$ has full rank $n$, $g$ is continuously F-differentiable in a neighbourhood of $(A,b)$ and we denote by $g'$ its F-derivative.\\ Let $\alpha$ and $\beta$ be two positive real numbers. In the present paper we consider the Euclidean norm for the solution space $\mathbb{R}^k$. For the data space $\mathbb{R}^{m\times n}\times \mathbb{R}^m$, we use the product norms defined by $$\norm{(A,b)}_{\rm{F~or~2}}= \sqrt{\alpha^2\norm{A}_{\rm{F~or~2}}^2+\beta^2\neuc{b}^2},~\alpha,\beta>0.$$ Following~\cite{GEURTS}, the absolute condition number of $g$ at the point $(A,b)$ using the product norm defined above is given by: $$\kappa_{g,{\rm{F~or~2}}}(A,b) =\max_{(\Delta A,\Delta b)} \frac{\neuc{g'(A,b).(\Delta A,\Delta b)}}{\norm{(\Delta A,\Delta b)}_{\rm{F~or~2}}}.$$ The corresponding relative condition number of $g$ at $(A,b)$ is expressed by $$\kappa_{g,{\rm{F~or~2}}}^{(rel)}(A,b)= \frac{\kappa_{g,F}(A,b)~\norm{(A,b)}_{\rm{F~or~2}}}{\neuc{g(A,b)}}.$$ To address the special cases where only $A$ (resp. $b$) is perturbed, we also define the quantities $\kappa_{g,{\rm{F~or~2}}}(A) =\max_{\Delta A} \frac{\neuc{\frac{\partial g}{\partial A}(A,b).\Delta A}}{\norm{\Delta A}_{\rm{F~or~2}}}$ (resp. $\kappa_{g,2}(b) =\max_{\Delta b} \frac{\neuc{\frac{\partial g}{\partial b}(A,b).\Delta b}}{\norm{\Delta b}_2}$). \begin{Remark} {\em The product norm for the data space is very flexible; the coefficients $\alpha$ and $\beta$ allow us to monitor the perturbations on $A$ and $b$. For instance, large values of $\alpha$ (resp. $\beta$ ) enable us to obtain condition number problems where mainly $b$ (resp. $A$) are perturbed. In particular, we will address the special cases where only $b$ (resp. $A$) is perturbed by choosing the $\alpha$ and $\beta$ parameters as $\alpha=+\infty~{\rm and}~\beta=1$ (resp. $\alpha=1~{\rm and}~\beta=+\infty$) since we have $$\lim_{\alpha \rightarrow +\infty}\kappa_{g,\rm{F~or~2}}(A,b) =\frac{1}{\beta}\kappa_{g,2}(b) ~{\rm and}~ \lim_{\beta \rightarrow +\infty}\kappa_{g,\rm{F~or~2}}(A,b) =\frac{1}{\alpha}\kappa_{g,\rm{F~or~2}}(A). $$ This can be justified as follows: \begin{eqnarray*} \kappa_{g,\rm{F~or~2}}(A,b) & = & \max_{(\Delta A,\Delta b)} \frac{\neuc{\frac{\partial g}{\partial A}(A,b).\Delta A +\frac{\partial g}{\partial b}(A,b).\Delta b}} {\sqrt{\alpha^2\norm{\Delta A}_{\rm{F~or~2}}^2+\beta^2\neuc{\Delta b}^2}}\\ & = & \max_{(\Delta A,\Delta b)} \frac{\neuc{\frac{\partial g}{\partial A}(A,b).\frac{\Delta A}{\alpha} +\frac{\partial g}{\partial b}(A,b).\frac{\Delta b}{\beta}}} {\sqrt{\norm{\Delta A}_{\rm{F~or~2}}^2+\neuc{\Delta b}^2}}.\\ \end{eqnarray*} The above expression represents the operator norm of a linear functional depending continuously on $\alpha$, and then we get $$\lim_{\alpha \rightarrow +\infty}\kappa_{g,\rm{F~or~2}}(A,b) =\max_{(\Delta A,\Delta b)} \frac{\neuc{\frac{\partial g}{\partial b}(A,b).\frac{\Delta b}{\beta}}} {\sqrt{\norm{\Delta A}_{\rm{F~or~2}}^2+\neuc{\Delta b}^2}} =\max_{\Delta b} \frac{\neuc{\frac{\partial g}{\partial b}(A,b).\frac{\Delta b}{\beta}}} {\neuc{\Delta b}} =\frac{1}{\beta}\kappa_{g,2}(b). $$ The proof is the same for the case where $\beta=+\infty$.\\ } \end{Remark}\\ The condition numbers related to $L^Tx(A,b)$ are referred to as {\bf partial condition numbers} (PCN) of the LLSP with respect to the linear operator $L$ in~\cite{ABG.07}.\\ We are interested in computing the PCN for two special cases. The first case is when $L$ is the identity matrix (conditioning of the solution) and the second case is when $L$ is a canonical vector $e_i$ (conditioning of a solution component). We can extract from~\cite{ABG.07} two theorems that can lead to computable quantities in these two special cases.\\ \begin{theo} \label{theobound} In the general case where~$(L \in \mathbb{R}^{n\times k})$, the absolute condition numbers of $g(A,b)=L^Tx(A,b)$ in the Frobenius and spectral norms can be respectively bounded as follows $$\frac{1}{\sqrt{3}}f(A,b) \leq \kappa_{g,F}(A,b) \leq f(A,b)$$ $$\frac{1}{\sqrt{3}}f(A,b) \leq \kappa_{g,2}(A,b) \leq \sqrt{2} f(A,b)$$ where \begin{equation}\label{eq:equationforf(A,b)} f(A,b)= \left(\neuc{L^T(A^TA)^{-1}}^2 \frac{\neuc{r}^2}{\alpha^2} +\neuc{L^T A^{\dagger}}^2 (\frac{\neuc{x}^2}{\alpha^2} +\frac{1}{\beta^2})\right)^{\frac{1}{2}}. \end{equation} \end{theo} \begin{theo}\label{corocond} In the two particular cases: \begin{enumerate} \item $L$ is a vector ($L \in \mathbb{R}^n$), or \item $L$ is the $n$-by-$n$ identity matrix ($L=I$) \end{enumerate} the absolute condition number of $g(A,b)=L^Tx(A,b)$ in the Frobenius norm is given by the formula: $$ \kappa_{g,F}(A,b)=\left(\neuc{L^T(A^TA)^{-1}}^2 \frac{\neuc{r}^2}{\alpha^2} +\neuc{L^T A^{\dagger}}^2 (\frac{\neuc{x}^2}{\alpha^2} +\frac{1}{\beta^2})\right)^{\frac{1}{2}}. $$ \end{theo} Theorem~\ref{corocond} provides the exact value for the condition number in the Frobenius norm for our two cases of interest ($L=e_i$ and $L=I$). From Theorem~\ref{theobound}, we observe that \begin{equation}\label{eq:condfrob_or_condspec} \frac{1}{\sqrt{3}} \kappa_{g,F}(A,b) \leq \kappa_{g,2}(A,b) \leq \sqrt{6}\kappa_{g,F}(A,b). \end{equation} which states that the partial condition number in spectral norm is of the same order of magnitude as the one in Frobenius norm. In the remainder of the paper, the focus is given to the partial condition number in Frobenius norm only.\\ For the case $L=I$, the result of Theorem~\ref{corocond} is similar to~\cite{GR.96} and~\cite[p. 92]{GEURTS}. The upper bound for $\kappa_{2,F}(A,b)$ that can be derived from Equation~(\ref{eq:condfrob_or_condspec}) is also the one obtained by~\cite{GEURTS} when we consider pertubations in $A$.\\ Let us denote by $\kappa_i(A,b)$ the condition number related to the component $x_i$ in Frobenius norm (i.e $\kappa_i(A,b)=\kappa_{g,F}(A,b)$ where $g(A,b)=e_i^Tx(A,b)=x_i(A,b)$). Then replacing $L$ by $e_i$ in Theorem~\ref{corocond} provides us with an exact expression for computing $\kappa_i(A,b)$, this gives \begin{equation}\label{eq:componentwise_formula} \kappa_i(A,b)=\left(\neuc{e_i^T(A^TA)^{-1}}^2 \frac{\neuc{r}^2}{\alpha^2} +\neuc{e_i^T A^{\dagger}}^2 (\frac{\neuc{x}^2}{\alpha^2} +\frac{1}{\beta^2})\right)^{\frac{1}{2}}. \end{equation} $\kappa_i(A,b)$ will be referred to as {\bf the condition number of the solution component $x_i$}.\\ Let us denote by $\kappa_{LS}(A,b)$ the condition number related to the solution $x$ in Frobenius norm (i.e $\kappa_{LS}(A,b)=\kappa_{g,F}(A,b)$ where $g(A,b)=x(A,b)$). Then Theorem~\ref{corocond} provides us with an exact expression for computing $\kappa_{LS}(A,b)$, that is \begin{equation}\label{eq:normwise_formula} \kappa_{LS}(A,b)= \neuc{(A^TA)^{-1}}^{1/2} \left( \frac{ \neuc{(A^TA)^{-1}} \neuc{r}^2 + \neuc{x}^2 } { \alpha^2 } + \frac{1}{\beta^2} \right)^{\frac{1}{2}}. \end{equation} where we have used the fact that $\neuc{ (A^T A )^{-1}} = \neuc{ A^\dagger }^2$.\\ $\kappa_{LS}(A,b)$ will be referred to as {\bf the condition number of the least squares solution}.\\ Note that~\cite{IRLLS.07} defines condition numbers for both $x$ and $r$ in order to derive error bounds for $x$ and $r$ but uses infinity-norm to measure perturbations.\\ In this paper, we will also be interested in the special case where only $b$ is perturbed ($\alpha=+\infty$ and $\beta=1$). In this case, we will call $\kappa_i(b)$ the condition number of the solution component $x_i$, and $\kappa_{LS}(b)$ the condition number of the least squares solution. When we restrict the perturbations to be on $b$, Equation~(\ref{eq:componentwise_formula}) simplifies to \begin{equation}\label{eq:justb_componentwise} \kappa_i(b)=\neuc{e_i^T A^{\dagger}}, \end{equation} and Equation~(\ref{eq:normwise_formula}) simplifies to \begin{equation}\label{eq:justb_normwise} \kappa_{LS}(b)= \neuc{A^{\dagger}}. \end{equation} This latter formula is standard and is in accordance with ~\cite[p. 29]{BJORCK}. \section{Condition numbers and statistical quantities}\label{sec:link} \subsection{Background for the linear statistical model}\label{sec:cov} We consider here the classical linear statistical model $$ b = Ax +\epsilon, ~A \in \mathbb{R}^{m\times n}, ~b \in \mathbb{R}^m, {\rm rank}(A)=n, $$ where $\epsilon$ is a vector of random errors having expected value $E(\epsilon)=0$ and variance-covariance $V(\epsilon)=\sigma_b^2I$. In statistical language, the matrix $A$ is referred to as the regression matrix and the unknown vector $x$ is called the vector of regression coefficients.\\ Following the Gauss-Markov theorem~\cite{ZELEN}, the least squares estimates $\hat{x}$ is the linear unbiased estimator of $x$ satisfying $$\|A\hat{x}-b\|_2=\min_{x \in \mathbb{R}^n}\|Ax-b\|_2,$$ with minimum variance-covariance equal to \begin{equation}\label{eq;variance-covariance} C=\sigma_b^2 (A^TA)^{-1}. \end{equation} Moreover $\frac{1}{m-n}\neuc{b-A\hat{x}}^2$ is an unbiased estimate of $\sigma_b^2$. This quantity is sometimes called the mean squared error (MSE).\\ The diagonal elements $c_{ii}$ of $C$ give the variance of each component $\hat{x}_i$ of the solution. The off-diagonal elements $c_{ij},~i \neq j$ give the covariance between $\hat{x}_i$ and $\hat{x}_j$.\\ We define $\sigma_{\hat x_i}$ as the standard deviation of the solution component $\hat x_i$ and we have \begin{equation}\label{eq:654} \sigma_{\hat x_i} = \sqrt{c_{ii}}. \end{equation} In the next section, we will prove that the condition numbers $\kappa_{i}(A,b)$ and $\kappa_{LS}(A,b)$ can be related to the statistical quantities $\sigma_{\hat x_i}$ and $\sigma_b$. \subsection{Perturbation on $b$ only} Using Equation~(\ref{eq;variance-covariance}), the variance $c_{ii}$ of the solution component $\hat x_i$ can be expressed as $$ c_{ii} = e_i^T C e_i = \sigma_b^2 e_i^T ( A^{T} A )^{-1} e_i. $$ We note that $ ( A^TA ) ^{-1} = A^{\dagger}A^{\dagger T}$ so that $$ c_{ii} = \sigma_b^2 e_i^T ( A^{\dagger}A^{\dagger T} ) e_i = \sigma_b^2 \neuc{e_i^T A^{\dagger}}^2. $$ Using Equation~(\ref{eq:654}), we get $$ \sigma_{\hat x_i} = \sqrt{c_{ii}} = \sigma_b \neuc{e_i^T A^{\dagger}}. $$ Finally from Equation~(\ref{eq:justb_componentwise}), we get \begin{equation}\label{eq:1} \sigma_{\hat x_i} = \sigma_b \kappa_i(b). \end{equation} Equation~(\ref{eq:1}) shows that the condition number $\kappa_i(b)$ relates linearly the standard deviation of $\sigma_b$ with the standard deviation of $\sigma_{\hat x_i}$.\\ Now if we consider the constant vector $\ell$ of size $n$, we have~(see \cite{ZELEN}) $$ {\rm variance}( \ell^T \hat x) = \ell^T C \ell. $$ Since $C$ is symmetric, we can write $$ \max_{\| \ell \|_2=1} {\rm variance}(\ell^T \hat x) = \neuc{C}. $$ Using the fact that $\| C \|_2 =\sigma_b^2 \neuc{ (A^TA)^{-1}}=\sigma_b^2 \neuc{ A^{\dagger} }^2 $, and Equation~(\ref{eq:justb_normwise}), we get $$ \max_{\| \ell \|_2=1} {\rm variance}(\ell^T \hat x) = \sigma_b^2 \kappa_{LS}(b)^2 $$ or, if we call $\sigma( \ell^T \hat x ) $ the standard deviation of $\ell^T \hat x$, $$ \max_{\| \ell \|_2=1} \sigma(\ell^T \hat x) = \sigma_b \kappa_{LS}(b). $$ Note that $\sigma_b = \max_{\| \ell \|_2=1} \sigma(\ell^T \epsilon) $ since $V(\epsilon)=\sigma_b^2I$. \begin{Remark} {\em Matlab proposes a routine LSCOV that computes the quantities $\sqrt{c_{ii}}$ in a vector STDX and the mean squared error MSE using the syntax [X,STDX,MSE] = LSCOV(A,B).\\ Then the condition numbers $\kappa_i(b)$ can be computed by the matlab expression STDX/sqrt(MSE). } \end{Remark} \subsection{Perturbation on $A$ and $b$} We now provide the expression of the condition number provided in Equation~(\ref{eq:componentwise_formula}) and in Equation~(\ref{eq:normwise_formula}) in term of statistical quantities.\\ Observing the following relations $$ C_i = \sigma_b^2 e_i^T(A^TA)^{-1} \quad {\rm and} \quad c_{ii} = \sigma_b^2 \neuc{e_i^T A^{\dagger}}^2, $$ where $C_i$ is the $i$-th column of the variance-covariance matrix, the condition number of $x_i$ given in Formula~(\ref{eq:componentwise_formula}) can expressed as $$ \kappa_i(A,b)=\frac{1}{\sigma_b}\left( \frac{\neuc{C_i}^2}{\sigma_b^2} \frac{\neuc{r}^2}{\alpha^2} +c_{ii}(\frac{\neuc{x}^2}{\alpha^2} +\frac{1}{\beta^2})\right)^{\frac{1}{2}}. $$ The quantity $\sigma_b^2$ will often be estimated by $\frac{1}{m-n}\neuc{r}^2$ in which case the expression can be simplified \begin{equation} \label{form:cii2} \kappa_i(A,b)=\frac{1}{\sigma_b}\left( \frac{1}{\alpha^2} \frac{\neuc{C_i}^2}{(m-n)} + \frac{c_{ii}\neuc{x}^2}{\alpha^2} + \frac{c_{ii}}{\beta^2})\right)^{\frac{1}{2}}. \end{equation} From Equation~(\ref{eq:normwise_formula}), we obtain $$ \kappa_{LS}(A,b)= \frac{\neuc{C}^{1/2}}{\sigma_b} \left( \frac{ \neuc{C} \neuc{r}^2 } { \alpha^2 \sigma_b^2 } + \frac{\neuc{x}^2}{\alpha^2} + \frac{1}{\beta^2} \right)^{\frac{1}{2}}. $$ The quantity $\sigma_b^2$ will often be estimated by $\frac{1}{m-n}\neuc{r}^2$ in which case the expression can be simplified $$ \kappa_{LS}(A,b)= \frac{\neuc{C}^{1/2}}{\sigma_b} \left( \frac{ 1 } { \alpha^2 } \frac{ \neuc{C} } { (m-n) } + \frac{\neuc{x}^2}{\alpha^2} + \frac{1}{\beta^2} \right)^{\frac{1}{2}}. $$ \section{Computation with LAPACK}\label{sec:lapack} Section~\ref{sec:theoback} provides us with formulas to compute the condition numbers $\kappa_i$ and $\kappa_{LS}$. As explained in Section~\ref{sec:link}, those quantities are intimately interrelated with the entries of the variance-covariance matrix. The goal of this section is to present practical methods and codes to compute those quantities efficiently with LAPACK. The assumption made is that the LLSP has already been solved with either the normal equations method or a QR factorization approach. Therefore the solution vector $\hat x$, the norm of the residual $\| \hat r \|_2$, and the R-factor $R$ of the QR factorization of $A$ are readily available (we recall that the Cholesky factor of the normal equations is the R-factor of the QR factorization up to some signs). In the example codes, we have used the LAPACK routine DGELS that solves the LLSP using QR factorization of A. Note that it is possible to have a more accurate solution using extra-precise iterative refinement~\cite{IRLLS.07}. \subsection{Variance-covariance computation} We will use the fact that $\frac{1}{m-n}\neuc{b-A\hat{x}}^2$ is an unbiased estimate of $\sigma_b^2$. We wish to compute the following quantities related to the variance-covariance matrix $C$ \begin{itemize} \item the $i$-th column $ C_i =\sigma_b^2 e_i^T ( A^T A )^{-1} $ \item the $i$-th diagonal element $c_{ii} =\sigma_b^2 \| e_i ^T A^\dagger \|_2^2 $ \item the whole matrix $ C $ \end{itemize} We note that the quantities $ C_i $, $c_{ii}$, and $C$ are of interest for statisticians. The NAG routine F04YAF~\cite{nag} is indeed an example of tool to compute these three quantities.\\ For the two first quantities of interest, we note that $$ \neuc{e_i^T A^{\dagger}}^2=\neuc{R^{-T}e_i}^2 ~{\rm and}~ \neuc{e_i^T(A^TA)^{-1}}=\neuc{R^{-1}(R^{-T}e_i)}. $$ \subsubsection{Computation of the $i$-th column $ C_i $} $C_i$ can be computed with two $n$--by--$n$ triangular solves \begin{equation}\label{form:trsys} R^Ty=e_i~{\rm and}~ Rz=y. \end{equation} The $i$-th column of $C$ can be computed by the following code fragment.\\ \\ {\bf Code 1:\\ CALL DGELS(~'N',~M,~N,~1,~A,~LDA,~B,~LDB,~WORK,~LWORK,~INFO~)\\ RESNORM = DNRM2(~(M-N),~B(N+1),~1)\\ SIGMA2 = RESNORM**2/DBLE(M-N)\\ E(1:N) = 0.D0\\ E(I) = 1.D0\\ CALL DTRSV(~'U',~'T',~'N',~N-I+1,~A(I,I),~LDA,~E(I),~1)\\ CALL DTRSV(~'U',~'N',~'N',~N,~A,~LDA,~E,~1)\\ CALL DSCAL(~N,~SIGMA2,~E,~1)\\ } \\ This requires about $2n^2$ flops (in addition to the cost of solving the linear least squares problem using DGELS).\\ $c_{ii}$ can be computed by one $n$--by--$n$ triangular solve and taking the square of the norm of the solution which involves about $(n-i+1)^2$ flops. It is important to note that the larger $i$, the less expensive to obtain $c_{ii}$. In particular if $i=n$ then only one operation is needed: $c_{nn} = R_{nn}^{-2}$. This suggests that a correct ordering of the variables can save some computation.\\ \\ \subsubsection{Computation of the $i$-th diagonal element $ c_{ii} $} From $c_{ii} = \sigma_b^2 \neuc{e_i^TR^{-1}}^2$, it comes that each $ c_{ii} $ corresponds to the $i$-th row of $R^{-1}$. Then the diagonal elements of $C$ can be computed by the following code fragment.\\ \\ {\bf Code 2:\\ CALL DGELS(~'N',~M,~N,~1,~A,~LDA,~B,~LDB,~WORK,~LWORK,~INFO~)\\ RESNORM~=~DNRM2((M-N), B(N+1), 1)\\ SIGMA2~=~RESNORM**2/DBLE(M-N)\\ CALL DTRTRI(~'U',~'N',~N,~A,~LDA,~INFO)\\ DO~I=1,N \\ \hspace*{0.5cm} CDIAG(I)~=~DNRM2(~N-I+1,~A(I,I),~LDA)\\ \hspace*{0.5cm} CDIAG(I)~=~SIGMA2~*~CDIAG(I)**2\\ END DO\\ } \\ This requires about $n^3/3$ flops (plus the cost of DGELS).\\ \\ \subsubsection{Computation of the whole matrix $C$} In order to compute explicity all the coefficients of the matrix $C$, one can use the routine DPOTRI which computes the inverse of a matrix from its Cholesky factorization. First the routine computes the inverse of $R$ using DTRTRI and then performs the triangular matrix-matrix multiply $R^{-1}R^{-T}$ by DLAUUM. This requires about $2n^3/3$ flops. We can also compute the variance-covariance matrix without inverting $R$ using for instance the algorithm given in~\cite[p. 119]{BJORCK} but the computational cost remains $2n^3/3$ (plus the cost of DGELS).\\ \\ We can obtain the upper triangular part of $C$ by the following code fragment.\\ \\ {\bf Code 3:\\ CALL DGELS(~'N',~M,~N,~1,~A,~LDA,~B,~LDB,~WORK,~LWORK,~INFO~)\\ RESNORM~=~DNRM2((M-N), B(N+1), 1)\\ SIGMA2~=~RESNORM**2/DBLE(M-N)\\ CALL DPOTRI(~'U',~N,~A,~LDA,~INFO)\\ CALL DLASCL(~'U',~0,~0,~N,~N,~1.D0,~SIGMA2,~N,~N,~A,~LDA,~INFO)\\ } \\ \subsection{Condition numbers computation} For computing $\kappa_i(A,b)$, we need to compute both the $i$-th diagonal element and the norm of the $i$-th column of the variance-covariance matrix and we cannot use direcly Code 1 but the following code fragment\\ \\ {\bf Code 4:\\ ALPHA2 = ALPHA**2\\ BETA2 = BETA**2\\ CALL DGELS(~'N',~M,~N,~1,~A,~LDA,~B,~LDB,~WORK,~LWORK,~INFO~)\\ XNORM = DNRM2(N,~B(1),~1)\\ RESNORM = DNRM2((M-N), B(N+1), 1)\\ CALL DTRSV(~'U',~'T',~'N',~N-I+1,~A(I,I),~LDA,~E(I),~1~)\\ ENORM = DNRM2(N, E, 1)\\ K = (ENORM**2)*(XNORM**2/ALPHA2+1.d0/BETA2)\\ CALL DTRSV(~'U',~'N',~'N',~N,~A,~LDA,~E,~1~)\\ ENORM = DNRM2(N, E, 1)\\ K = SQRT((ENORM*RESNORM)**2/ALPHA2 + K)\\ } \\ For computing all the $\kappa_i(A,b)$, we need to compute the columns $C_i$ and the diagonal elements $c_{ii}$ using Formula~(\ref{form:cii2}) and then we have to compute the whole variance-covariance matrix. This can be performed by a slight modification of Code 3.\\ When only $b$ is perturbed, then we have to invert $R$ and we can use a modification of Code 2 (see numerical example in Section~\ref{sec:cnes}).\\ \\ For estimating $\kappa_{LS}(A,b)$, we need to have an estimate of $\neuc{R^{-1}}$. The computation of $\neuc{R^{-1}}$ requires to compute the minimum singular value of the matrix $A$ (or $R$). One way is to compute the full SVD of $A$ (or $R$) which requires $\mathcal{O} (n^3)$ flops. As an alternative, $\neuc{R^{-1}}$ can be estimated for instance by considering other matrix norms through the following inequalities \begin{eqnarray*} \frac{1}{\sqrt{n}}\nfro{R^{-1}} & \leq \neuc{R^{-1}} \leq & \nfro{R^{-1}}, \\ \frac{1}{\sqrt{n}}\|R^{-1}\|_{\infty} & \leq \neuc{R^{-1}} \leq & \sqrt{n} \|R^{-1}\|_{\infty},\\ \frac{1}{\sqrt{n}}\|R^{-1}\|_1 & \leq \neuc{R^{-1}} \leq & \sqrt{n} \|R^{-1}\|_1.\\ \end{eqnarray*} $\|R^{-1}\|_1$ or $\|R^{-1}\|_{\infty}$ can be estimated using Higham modification~\cite[p. 293]{HIGHAM} of Hager's~\cite{HAGER} method as it is implemented in LAPACK~\cite{LAPACK} DTRCON routine (see Code 5). The cost is $\mathcal{O} ( n^2 )$.\\ \\ {\bf Code 5:\\ CALL DTRCON(~'I',~'U',~'N',~N,~A,~LDA,~RCOND,~WORK,~IWORK,~INFO)\\ RNORM~=~DLANTR(~'I',~'U',~'N',~N,~N,~A,~LDA,~WORK)\\ RINVNORM~=~(1.D0/RNORM)/RCOND\\ } \\ We can also evaluate $\neuc{R^{-1}}$ by considering $\nfro{R^{-1}}$ since we have \begin{eqnarray*} \nfro{R^{-1}}^2 & = & \nfro{R^{-T}}^2\\ & = & {\rm tr}(R^{-1}R^{-T})\\ & = & \frac{1}{\sigma_b^2} {\rm tr}(C),\\ \end{eqnarray*} where tr($C$) denotes the trace of the matrix $C$, i.e $\sum_{i=1}^{n} c_{ii}$. Hence the condition number of the least-squares solution can be approximated by \begin{equation} \label{form:trace} \kappa_{LS}(A,b) \simeq \left( \frac{{\rm tr}(C)}{\sigma_b^2} \left(\frac{{\rm tr}(C) \neuc{r}^2+\sigma_b^2 \neuc{x}^2} {\sigma_b^2 \alpha^2}+\frac{1}{\beta^2}\right) \right)^{\frac{1}{2}}. \end{equation} Then we can estimate $\kappa_{LS}(A,b)$ by computing and summing the diagonal elements of $C$ using Code 2.\\ When only $b$ is perturbed ($\alpha = +\infty~{\rm and}~\beta=1$), then we get $$\kappa_{LS}(b) \simeq \frac{\sqrt{{\rm tr}(C)}}{\sigma_b}.$$ This result relates to~\cite[p. 167]{FAREBROTHER} where ${\rm tr}(C)$ measures the squared effect on the LLSP solution $x$ to small changes in $b$.\\ \\ We give in Table~\ref{TabCompar} the LAPACK routines used for computing the condition numbers of an LLSP solution or its components and the corresponding number of floating-point operations per second. Since the LAPACK routines involved in the covariance and/or LLSP condition numbers have their equivalent in the parallel library ScaLAPACK~\cite{SCALAPACK}, then this table is also available when using ScaLAPACK. This enables us to easily compute these quantities for larger LLSP. \begin{table}[hbtp!] \centering \caption{Computation of least squares conditioning with (Sca)LAPACK} \vspace{0.4cm} \begin{tabular}{|c|c|c|c|} \hline condition number&linear algebra operation&LAPACK routines&flops count\\ &&&\\ \hline $\kappa_i(A,b)$&$R^Ty=e_i~{\rm and}~ Rz=y$& 2 calls to (P)DTRSV&$2n^2$\\ &&&\\ \hline all $\kappa_i(A,b),~i=1,n$&$RY=I~{\rm and~compute}~YY^T$&(P)DPOTRI&$2n^3/3$\\ &&&\\ \hline all $\kappa_i(b),~i=1,n$&invert $R$&(P)DTRTRI&$n^3/3$\\ &&&\\ \hline $\kappa_{LS}(A,b)$&estimate $\|R^{-1}\|_{1~{\rm or}~\infty}$&(P)DTRCON&${\cal O}(n^2)$\\ &compute $\nfro{R^{-1}}$&(P)DTRTRI&$n^3/3$\\ \hline \end{tabular} \label{TabCompar} \end{table} \begin{Remark} {\em The cost for computing all the $\kappa_i(A,b)$ or estimating $\kappa_{LS}(A,b)$ is always ${\cal O}(n^3)$. This seems affordable when we compare it to the cost of the least squares solution using Householder QR factorization ($2mn^2-2n^3/3$) or the normal equations ($mn^2+n^3/3$) because we have in general $m \gg n$.\\ } \end{Remark} \section{Numerical experiments}\label{sec:numerics} \subsection{Laplace's computation of the mass of Jupiter and assessment of the validity of its results} In~\cite{LAPLACE}, Laplace computes the mass of Jupiter, Saturn and Uranus and provides the variances associated with those variables in order to assess the quality of the results. The data comes from the French astronomer Bouvart in the form of the normal equations given in Equation~(\ref{eq:laplace}). \begin{equation}\label{eq:laplace} \begin{array}{rcl} 795938 z_0 - 12729398 z_1 + 6788.2 z_2 - 1959.0 z_3 + 696.13 z_4 + 2602 z_5 & = & 7212.600 \\ -12729398 z_0 + 424865729 z_1 - 153106.5 z_2 - 39749.1 z_3 - 5459 z_4 + 5722 z_5 & = & -738297.800 \\ 6788.2 z_0 - 153106.5 z_1 + 71.8720 z_2 - 3.2252 z_3 + 1.2484 z_4 + 1.3371 z_5 & = & 237.782 \\ -1959.0 z_0 - 39749.1 z_1 - 3.2252 z_2 + 57.1911 z_3 + 3.6213 z_4 + 1.1128 z_5 & = & -40.335 \\ 696.13 z_0 - 5459 z_1 + 1.2484 z_2 + 3.6213 z_3 + 21.543 z_4 + 46.310 z_5 & = & -343.455 \\ 2602 z_0 + 5722 z_1 + 1.3371 z_2 + 1.1128 z_3 + 46.310 z_4 + 129 z_5 & = & -1002.900 \\ \end{array} \end{equation} For computing the mass of Jupiter, we know that Bouvart performed $ m = 129 $ observations and there are $n=6$ variables in the system. The residual of the solution $\| b - A\hat x \|_2^2$ is also given by Bouvart and is $31096$. On the $6$ unknowns, Laplace only seeks one, the second variable $z_1$. The mass of Jupiter in term of the mass of the Sun is given by $z_1$ and the formula: $$ \textmd{mass of Jupiter} = \frac{1+z_1}{1067.09}.$$ It turns out that the first variable $z_0$ represents the mass of Uranus through the formula $$ \textmd{mass of Uranus} = \frac{1+z_0}{19504}.$$ If we solve the system~(\ref{eq:laplace}), we obtain the solution vector\\ \begin{center} \begin{minipage}{10cm} Solution vector\\ 0.08954 -0.00304 -11.53658 -0.51492 5.19460 -11.18638 \end{minipage} \end{center} From $z_1$, we can compute the mass of Jupiter as a fraction of the mass of the Sun and we obtain $1070$. This value is indeed accurate since the correct value according to NASA is $1048$. From $z_0$, we can compute the mass of Uranus as a fraction of the mass of the Sun and we obtain $17918$. This value is inaccurate since the correct value according to NASA is $22992$.\\ Laplace has computed the variance of $z_0$ and $z_1$ to assess the fact that $z_1$ was probably correct and $z_0$ probably inaccurate. To compute those variances, Laplace first performed a Cholesky factorization from right to left of the system~(\ref{eq:laplace}), then, since the variables were correctly ordered the number of operations involved in the computation of the variances of $z_0$ and $z_1$ were minimized. The variance-covariance matrix for Laplace's system is: $$ \left( \begin{array}{cccccc} 0.005245 & -0.000004 & -0.499200 & 0.137212 & 0.235241 & -0.186069 \\ \cdot & 0.000004 & 0.009873 & 0.003302 & 0.002779 & -0.001235 \\ \cdot & \cdot & 71.466023 & -5.441882 & -16.672689 & 14.922752 \\ \cdot & \cdot & \cdot & 10.860492 & 5.418506 & -4.896579 \\ \cdot & \cdot & \cdot & \cdot & 66.088476 & -28.467391 \\ \cdot & \cdot & \cdot & \cdot & \cdot & 15.874809 \\ \end{array} \right) $$ Our computation gives us that the variance for the mass of Jupiter is $4.383233\cdot10^{-6}$. For reference, Laplace in 1820 computed $4.383209\cdot10^{-6}$. (We deduce the variance from Laplace's value 5.0778624. To get what we now call the variance, one needs to compute the quantity: $ 1/(2*10**5.0778624)*m/(m-n)$.) From the variance-covariance matrix, one can assess that the computation of the mass of Jupiter (second variable) is extremely reliable while the computation of the mass of Uranus (first variable) is not. For more details, we recommend to read \cite{langoureview2007}. \subsection{Gravity field computation}\label{sec:cnes} A classical example of parameter estimation problem is the computation of the Earth's gravity field coefficients. More specifically, we estimate the parameters of the gravitational potential that can be expressed in spherical coordinates $(r,\theta,\lambda)$ by~\cite{BALMINO} \begin{equation} \label{potential} V(r,\theta,\lambda)=\frac{GM}{R}\sum_{\ell=0}^{\ell_{max}}\left(\frac{R}{r}\right)^{\ell+1} \sum_{m=0}^{\ell}\overline{P}_{\ell m}(\cos{\theta})\left[\overline{C}_{\ell m} \cos{m\lambda}+\overline{S}_{\ell m}\sin{m\lambda}\right] \end{equation} where $G$ is the gravitational constant, $M$ is the Earth's mass, $R$ is the Earth's reference radius, the $\overline{P}_{\ell m}$ represent the fully normalized Legendre functions of degree $\ell$ and order $m$ and $\overline{C}_{\ell m}$,$\overline{S}_{\ell m}$ are the corresponding normalized harmonic coefficients. The objective here is to compute the harmonic coefficients $\overline{C}_{\ell m}$ and $\overline{S}_{\ell m}$ the most accurately as possible. The number of unknown parameters is expressed by $n=(\ell_{max}+1)^2.$ These coefficients are computed by solving a linear least squares problem that may involve millions of observations and tens of thousands of variables. More details about the physical problem and the resolution methods can be found in \cite{PHD.MB}. The data used in the following experiments were provided by CNES\footnote{Centre National d'Etudes Spatiales, Toulouse, France} and they correspond to 10 days of observations using GRACE\footnote{Gravity Recovery and Climate Experiment, NASA, launched March 2002} measurements (about $166,000$ observations). We compute the spherical harmonic coefficients $\overline{C}_{\ell m}$ and $\overline{S}_{\ell m}$ up to a degree $\ell_{max}=50$; except the coefficients $\overline{C}_{11}, \overline{S}_{11}, \overline{C}_{00}, \overline{C}_{10}$ that are a priori known. Then we have $n=2,597$ unknowns in the corresponding least squares problems (note that the GRACE satellite enables us to compute a gravity field model up to degree 150). The problem is solved using the normal equations method and we have the Cholesky decomposition $A^TA=U^TU$.\\ We compute the relative condition numbers of each coefficient $x_i$ using the formula $$\kappa^{(rel)}_i(b)=\neuc{e_i^TU^{-1}} \neuc{b}/|x_i|,$$ and the following code fragment, derived from Code 2, in which the array $D$ contains the normal equations $A^TA$ and the vector $X$ contains the right-hand side $A^Tb$.\\ \\ {\bf CALL DPOSV(~'U',~N,~1,~D,~LDD,~X,~LDX,~INFO)\\ CALL DTRTRI(~'U',~'N',~N,~D,~LDD,~INFO)\\ DO~I=1,N \\ \hspace*{0.5cm} KAPPA(I)~=~DNRM2(~N-I+1,~D(I,I),~LDD)~*~BNORM/ABS(X(I)) \\ END DO}\\ \\ Figure~\ref{plotcond} represents the relative condition numbers of all the $n$ coefficients. We observe the disparity between the condition numbers (between $10^2$ and $10^8$). To be able to give a physical interpretation, we need first to sort the coefficients by degrees and orders as given in the development of $V(r,\theta,\lambda)$ in Expression~(\ref{potential}).\\ In Figure~\ref{plotcos}, we plot the coefficients $\overline{C}_{\ell m}$ as a function of the degrees and orders (the curve with the $\overline{S}_{\ell m}$ is similar). We notice that for a given order, the condition number increases with the degree and that, for a given degree, the variation of the sensitivity with the order is less significant.\\ We can also study the effect of regularization on the conditioning. The physicists use in general a Kaula~\cite{KAULA} regularization technique that consists of adding to $A^TA$ a diagonal matrix $D=diag(0,\cdots,0,\delta,\cdots,\delta)$ where $\delta$ is a constant that is proportional to $\frac{10^{-5}}{\ell_{max}^2}$ and the nonzero terms in $D$ correspond to the variables that need to be regularized. An example of the effect of Kaula regularization is shown in Figure~\ref{zonaux} where we consider the coefficients of order $0$ also called zonal coefficients. We compute here the absolute condition numbers of these coefficients using the formula $\kappa_i(b)=\neuc{e_i^TU^{-1}}$. Note that the $\kappa_i(b)$ are much lower that 1. This is not surprising because typically in our application $\neuc{b} \sim 10^5$/ and $|x_i| \sim 10^{-12}$ which would make the associated relative condition numbers greater than 1. We observe that the regularization is effective on coefficients of highest degree that are in general more sensitive to perturbations. \begin{figure}[!ht] \begin{center} {\epsfig{file=condpar.eps,width=0.7 \textwidth}} \\ \caption{\label{plotcond} Amplitude of the relative condition numbers for the gravity field coefficients. } \end{center} \end{figure} \begin{figure}[!ht] \begin{center} {\epsfig{file=cosinus.eps,width=0.7 \textwidth}} \\ \caption{\label{plotcos} Conditioning of spherical harmonic coefficients $\overline{C}_{\ell m}~(2 \leq \ell \leq 50~,~1 \leq m\leq 50)$. } \end{center} \end{figure} \begin{figure}[!ht] \begin{center} {\epsfig{file=zonaux.eps,width=0.7 \textwidth}} \\ \caption{\label{zonaux} Effect of regularization on zonal coefficients $\overline{C}_{{\ell} 0}~(2 \leq {\ell} \leq 50)$ } \end{center} \end{figure} \newpage \section{Conclusion}\label{sec:concl} To assess the accuracy of a linear least squares solution, the practitioner of numerical linear algebra uses generally quantities like condition numbers or backward errors when the statistician is more interested in covariance analysis. In this paper we proposed quantities that talk to both communities and that can assess the quality of the solution of a least squares problem or one of its component. We provided pratical ways to compute these quantities using (Sca)LAPACK and we experimented these computations on pratical examples including a real physical application in the area of space geodesy. \bibliographystyle{siam}
{ "timestamp": "2007-10-03T18:29:12", "yymm": "0710", "arxiv_id": "0710.0829", "language": "en", "url": "https://arxiv.org/abs/0710.0829", "abstract": "In this paper, we address the accuracy of the results for the overdetermined full rank linear least squares problem. We recall theoretical results obtained in Arioli, Baboulin and Gratton, SIMAX 29(2):413--433, 2007, on conditioning of the least squares solution and the components of the solution when the matrix perturbations are measured in Frobenius or spectral norms. Then we define computable estimates for these condition numbers and we interpret them in terms of statistical quantities. In particular, we show that, in the classical linear statistical model, the ratio of the variance of one component of the solution by the variance of the right-hand side is exactly the condition number of this solution component when perturbations on the right-hand side are considered. We also provide fragment codes using LAPACK routines to compute the variance-covariance matrix and the least squares conditioning and we give the corresponding computational cost. Finally we present a small historical numerical example that was used by Laplace in Theorie Analytique des Probabilites, 1820, for computing the mass of Jupiter and experiments from the space industry with real physical data.", "subjects": "Numerical Analysis (math.NA); Statistics Theory (math.ST)", "title": "Computing the Conditioning of the Components of a Linear Least Squares Solution", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9833429599907709, "lm_q2_score": 0.8152324915965392, "lm_q1q2_score": 0.8016531313671922 }
https://arxiv.org/abs/2009.04050
Minimal universality criterion sets on the representations of quadratic forms
For a set $S$ of (positive definite and integral) quadratic forms with bounded rank, a quadratic form $f$ is called $S$-universal if it represents all quadratic forms in $S$. A subset $S_0$ of $S$ is called an $S$-universality criterion set if any $S_0$-universal quadratic form is $S$-universal. We say $S_0$ is minimal if there does not exist a proper subset of $S_0$ that is an $S$-universality criterion set. In this article, we study various properties of minimal universality criterion sets. In particular, we show that for `most' binary quadratic forms $f$, minimal $S$-universality criterion sets are unique in the case when $S$ is the set of all subforms of the binary form $f$.
\section{Introduction} Let $S$ be a set of (positive definite integral) quadratic forms with bounded rank. A quadratic form $f$ is called {\it $S$-universal} if it represents all quadratic forms in the set $S$. A subset $S_0$ of $S$ is called {\it an $S$-universality criterion set} if any $S_0$-universal quadratic form is $S$-universal. Conway and Schneeberger's 15-theorem \cite{c} says that the set $\{1,2,3,5,6,7,10,14,15\}$ is an $S$-universality criterion set, where $S$ is the set of all positive integers (see also \cite{b}). Note that any positive integer $a$ corresponds to the unary quadratic form $ax^2$. For an arbitrary set $S$ of quadratic forms with bounded rank, the existence of a finite $S$-universality criterion set was proved by the third author and his collaborators in \cite{kko}. An $S$-universality criterion set $S_0$ is called {\it minimal} if any proper subset of $S_0$ is not an $S$-universality criterion set. In \cite{kko}, the authors proposed the following two questions on the minimal universality criterion sets: Let $\Gamma(S)$ be the set of all $S$-universality criterion sets. \begin{itemize} \item [(i)] For which $S$ is there a unique minimal $S_0 \in \Gamma(S)$? \item [(ii)] Is $\vert S_0\vert=\gamma(S)$ for every minimal $S_0 \in \Gamma(S)$? If not, when? \end{itemize} For the question (i), when $S$ is the set of all quadratic forms of rank $k$, the uniqueness of the minimal $S$-universality criterion set was proved by Bhargava \cite{b} for the case when $k=1$, and by Kominers \cite{kom1}, \cite{kom2} for the cases when $k=2$, $k=8$, respectively (see also \cite{k}, \cite{kko0}, and \cite{o1}). For the question (ii), Elkies, Kane, and Kominers \cite{ekk} answered in the negative for some special set $S$ of quadratic forms. In fact, they considered the set $S_f$ of all subforms of the quadratic form $f(x,y,z)= x^2+y^2+2z^2$. Clearly, the set $\{f\}$ itself is a minimal $S_f$-universality criterion set. They proved that any quadratic form that represents both subforms $x^2+y^2+8z^2$ and $2x^2+2y^2+2z^2$ of $f$ also represents $f$ itself. Therefore the set $\{x^2+y^2+8z^2, 2x^2+2y^2+2z^2\}$ is also a minimal $S_f$-universality criterion set. In this article, we prove that the question (i) is true if $S$ is any set of positive integers, that is, any set of unary quadratic forms. We also prove that minimal $\Phi_n$-universality criterion sets are not unique when $\Phi_n$ is the set of all quadratic forms of rank $n$ for any $n\ge 9$. To analyze the example given by Elkies, Kane, and Kominers more closely, we introduce the notion of a {\it recoverable} quadratic form. A quadratic form $f$ is called recoverable if minimal $S_f$-universality criterion sets are not unique in the case when $S_f$ is the set of all subforms of $f$. The third author and his collaborators proved in \cite{jko} that any unary quadratic form is not recoverable. In this article, we show that `most' binary quadratic forms are not recoverable, and in fact, there are infinitely many recoverable binary quadratic forms up to isometry. The subsequent discussion will be conducted in the better adapted geometric language of quadratic spaces and lattices. Throughout this article, we always assume that every ${\mathbb Z}$-lattice $L={\mathbb Z} x_1+{\mathbb Z} x_2+\dots+{\mathbb Z} x_k$ is {\it positive definite and integral}, that is, the corresponding symmetric matrix $$ M_L=(B(x_i,x_j)) \in M_{k\times k}({\mathbb Z}) $$ is positive definite and the scale $\mathfrak s(L)$ of the ${\mathbb Z}$-lattice $L$ is ${\mathbb Z}$. The corresponding quadratic map $Q: L \to {\mathbb Z}$ will be defined by $Q(x)=B(x,x)$ for any $x\in L$. For any positive integer $a$, the ${\mathbb Z}$-lattice obtained from $L$ by scaling $a$ will be denoted by $L^a$. Hence we have $M_{L^a}=a\cdot M_L$. The discriminant $dL$ of the ${\mathbb Z}$-lattice $L$ will be defined by $dL=\det(M_L)$. We call $L$ is diagonal if $B(x_i,x_j)=0$ for any $i$ and $j$ with $i\ne j$. If $L$ is diagonal, then we simply write $$ L=\langle Q(x_1),\dots,Q(x_k)\rangle. $$ We say $L$ is even if $Q(x)$ is even for any $x \in L$. If an integer $n$ is represented by $L$ over ${\mathbb Z}_p$ for any prime $p$ including infinite prime, then we say that $n$ is represented by the genus of $L$, and we write $n {\ \rightarrow\ } \text{gen}(L)$. Note that $n$ is represented by the genus of $L$ if and only if $n$ is represented by a ${\mathbb Z}$-lattice in the genus of $L$. When $n$ is represented by the ${\mathbb Z}$-lattice $L$ itself, then we write $n{\ \rightarrow\ } L$. We define $$ Q(L)=\{ n \in {\mathbb Z} : n {\ \rightarrow\ } L\}. $$ For any positive integer $n$, The cubic ${\mathbb Z}$-lattice $I_n={\mathbb Z} e_1+{\mathbb Z} e_2+\dots+{\mathbb Z} e_n$ is the ${\mathbb Z}$-lattice satisfying $B(e_i,e_j)=\delta_{ij}$. Any unexplained notation and terminology can be found in \cite{ki} or \cite{om}. \section{Uniqueness of the minimal universality criterion set} In general, minimal $S$-universality criterion sets are not unique for an arbitrary set $S$ of quadratic forms with bounded rank. In this section, we prove that the minimal $S$-universality criterion set is unique if $S$ is any subset of positive integers. Let $\mathbb N$ be the set of positive integers. For a positive integer $k$ and a nonnegative integer $\alpha$, we define the set of arithmetic progressions $$ A_{k,\alpha}=\{kn+\alpha : n\in \mathbb N \cup \{0\}\}. $$ If a ${\mathbb Z}$-lattice $L$ represents all elements in $A_{k,\alpha}$, we simply write $A_{k,\alpha} {\ \rightarrow\ } L$. \begin{prop} \label{keyp} Let $S=\{ s_0,s_1,s_2,\dots\}$ be a subset of ${\mathbb N}$ such that $s_i < s_{i+1}$ for any nonnegative integer $i$, and let $k$ be a positive integer. If there is a ${\mathbb Z}$-lattice $\ell$ such that $$ s_0,s_1,\dots,s_{k-1} \in Q(\ell) \quad \text{and} \quad s_k \not \in Q(\ell), $$ then there is a ${\mathbb Z}$-lattice $L$ such that $Q(L)\cap S=S-\{s_k\}$. \end{prop} \begin{proof} First, we define $$ \mathfrak C=\{ 0 \le u \le s_{k+1}-1 : A_{s_{k+1},u} \cap \{s_{k+1},s_{k+2},\dots\} \ne \emptyset\}=\{ c_1,c_2,\dots,c_v\}, $$ and for each $c \in \mathfrak C$, let $s(c)=\min(A_{s_{k+1},c} \cap \{s_{k+1},s_{k+2},\dots\})$. We define $$ L=\ell \perp s_{k+1} I_4 \perp \langle s(c_1),s(c_2),\cdots,s(c_v) \rangle. $$ Since $s_{k+1} >s_k$ and $s(c_j) > s_k$ for any $j=1,2,\dots,v$, we see that $s_k$ is not represented by $L$. Furthermore, for any integer $a \in \{s_{k+1},s_{k+2},\dots\}$, there is a nonnegative integer $M$ and an integer $j$ with $1\le j \le v$ such that $a=s_{k+1}M+s(c_j)$. Since $M$ is represented by $I_4$, the integer $a$ is represented by $L$. The proposition follows directly from this. \end{proof} \begin{thm} For any set $S=\{s_0,s_1,s_2,\dots\}$ of positive integers, the minimal $S$-universality criterion set is unique. \end{thm} \begin{proof} Without loss of generality, we may assume that $s_i < s_{i+1}$ for any nonnegative integer $i$. A positive integer $s_i \in S$ is called a truant of $S$ if there is a ${\mathbb Z}$-lattice $L$ such that $L$ represents all integers in the set $\{s_0,s_1,\dots,s_{i-1}\}$, whereas $L$ does not represent $s_i$. Clearly, $s_0$ is a truant of $S$. Let $T(S)$ be the set of truants of $S$. Then, by Proposition \ref{keyp}, any $S$-universality criterion set should contain $T(S)$. Hence it suffices to show that $T(S)$ itself is an $S$-universality criterion set. Let $L$ be a ${\mathbb Z}$-lattice that represents all integers in $T(S)$. Suppose that $L$ is not $S$-universal. Let $m$ be the smallest integer such that $s_m$ is not represented by $L$. Then, clearly, $s_m$ is a truant of $S$, and hence $s_m \in T(S)$. This is a contradiction. Therefore, $T(S)$ is the unique minimal $S$-universality criterion set. \end{proof} \begin{rmk} {\rm As pointed out by \cite{h}, Bhargava also proved the above result. However, no proof of this has appeared in the literature to the author's knowledge.} \end{rmk} Let $L$ be a ${\mathbb Z}$-lattice of rank $m$. For any positive integer $j$ less than or equal to $m$, the $j$-th successive minimum of $L$ will be denoted by $\mu_j(L)$ ( for the definition of the successive minimum, see Chapter 12 of \cite{ca}). It is well-known that there is a constant $\gamma_m$, which is called the Hermite constant, such that $$ dL\le \mu_1(L)\mu_2(L)\cdots \mu_m(L) \le \gamma_m^mdL $$ (for the proof, see Proposition 2.3 of \cite{ea}). We define $\min(L)=\min\{ Q(x)\mid x \in L-\{0\}\}$. Note that $\min(L)=\mu_1(L)$. \begin{thm} For any positive integer $k$, there is a subset $S$ of positive integers such that the cardinality of its minimal universality criterion set is exactly $k$. \end{thm} \begin{proof} Let $L= {\mathbb Z} x_{1} + {\mathbb Z} x_{2} + \cdots + {\mathbb Z} x_{k}$ be a ${\mathbb Z}$-lattice such that $Q(x_{i}) = k+i$ for any $i$ with $1 \le i \le k$. If $m$ is the rank of $L$, then we have $m \le k$ and $\mu_{m}(L) \le 2k$. It follows from $\mu_{1}(L) \le \mu_{2}(L) \le \dots \le \mu_{m}(L)$ that $$ dL \le \mu_{1}(L) \mu_{2}(L) \cdots \mu_{m}(L) \le (2k)^{k}. $$ Hence there are only finitely many candidates for $L$ up to isometry since the discriminant and the rank of $L$ are bounded. Let $\{ L_{1}, L_{2}, \dots, L_{t} \}$ be the set of all possible candidates for $L$. We define $$ S = \cap_{i=1}^{t} Q(L_{i}). $$ Then from the definition of $S$, it is obvious that $\{k+1, k+2, \dots, 2k\}$ is an $S$-universality criterion set. Put $M_{1} = \langle k+2 \rangle$. Since $k+1$ is not represented by $M_{1}$, there is a ${\mathbb Z}$-lattice $N_{1}$ such that $Q(N_{1}) \cap S = S-\{k+1\}$ by Proposition \ref{keyp}. For each $i=2,3, \dots, k$, put $$ M_{i} = \langle k+1, \dots, k+i-1 \rangle. $$ Then, one may easily show that $k+j {\ \rightarrow\ } M_{i}$ for any $j = 1, \dots, i-1$, whereas $k+i$ is not represented by $M_{i}$. Then, by Proposition \ref{keyp} again, there is a ${\mathbb Z}$-lattice $N_{i}$ such that $Q(N_{i}) \cap S = S-\{k+i\}$. This implies that $\{k+1, k+2, \dots, 2k\}$ is the minimal $S$-universality criterion set. \end{proof} Let $\Phi_n$ be the set of all ${\mathbb Z}$-lattices of rank $n$. As explained in the introduction, it is known that the minimal $\Phi_n$-universality criterion set is unique if $n=1,2$, or $8$. For all the other positive integers $n$, the explicit minimal $\Phi_n$-universality criterion set is not known yet. \begin{prop} For any integer $n$ with $n \ge 9$, there are infinitely many minimal $\Phi_n$-universality criterion sets. \end{prop} \begin{proof} Let $\Phi_n^0=\{L_1,L_2,\dots,L_s\}$ be a minimal $\Phi_n$-universality criterion set. Assume that $L_i=I_{k_i} \perp \ell_i$, where $\min(\ell_i)\ge 2$. If $n_0=\max\{k_i\}<n$, then $I_{n_0} \perp \ell_1\perp\dots\perp\ell_s$ represents all ${\mathbb Z}$-lattices in $\Phi_n^0$, but it does not represent $I_n$. This is a contradiction. Therefore $n_0=n$, that is, $I_n \in \Phi_n^0$. Similarly, one may easily show that there is an integer $j$ such that $L_j$ represents $D_m[1]$ for some integer $m \equiv 0 \Mod 4$ with $n-4\le m<n$. Note that $L_j=D_m[1]\perp M$ for some ${\mathbb Z}$-lattice $M$ with rank less than or equal to $4$. Without loss of generality, assume that $L_1=I_n$ and $L_2=D_m[1]\perp M$. Note that any ${\mathbb Z}$-lattice that represents both $L_1$ and $L_2$ should represent $I_n\perp D_m[1]$. Furthermore, since $I_n$ is $4$-universal, $L_j$ cannot represent $D_m[1]$ for any $j\ge 3$. Now we show that for any ${\mathbb Z}$-lattice $N$ with rank $n-m$, $$ \Phi_n^0(N)=\{I_n, D_m[1]\perp N,L_3,\dots, L_s\} $$ is also a minimal $\Phi_n$-universality criterion set. Assume that a ${\mathbb Z}$-lattice $\mathcal L$ represents all ${\mathbb Z}$-lattices in $\Phi_n^0(N)$. Since $I_n\perp D_m[1]$ is represented by $\mathcal L$, $L_2=D_m[1]\perp M$ is also represented by $\mathcal L$. Therefore, $\mathcal L$ is $n$-universal from the assumption that $\Phi_n^0$ is a $\Phi_n$-universality criterion set. By using similar argument, one may easily show that $\Phi_n^0(N)$ is, in fact, minimal. \end{proof} \begin{rmk}{\rm We conjecture that $$ \left\lbrace I_4, A_4, A_2\perp A_2, A_2\perp \begin{pmatrix} 2&1\\1&3\end{pmatrix}\right\rbrace $$ is the unique minimal $\Phi_{4}$-universality criterion set. Here, $A_m={\mathbb Z}(e_1-e_2)+{\mathbb Z}(e_2-e_3)+\dots+{\mathbb Z}(e_m-e_{m+1})$ is a root ${\mathbb Z}$-lattice, where $\{e_i\}$ is the standard orthonormal basis of the cubic ${\mathbb Z}$-lattice $I_n$.} \end{rmk} \section{Recoverable ${\mathbb Z}$-lattices} In this section, we introduce the notion of {\it recoverable} $\mathbb{Z}$-lattices and give some properties on those $\mathbb{Z}$-lattices, and we show some necessary conditions and some sufficient conditions for ${\mathbb Z}$-lattices to be recoverable. In \cite{ekk}, Elkies and his collaborators gave an example of a set $S$ of ternary ${\mathbb Z}$-lattices such that the sizes of minimal $S$-universality criterion sets vary. To explain their example more precisely, let $T$ be the set of all ternary sublattices of $\langle 1,1,2\rangle$. Then, clearly, $T_0=\{\langle1,1,2\rangle\}$ is a minimal $T$-universality criterion set. Furthermore, they proved that $$ T_1=\{ \langle 1,1,8\rangle, \langle 2,2,2\rangle\} $$ is also a minimal $T$-universality criterion set. The point is that any ${\mathbb Z}$-lattice that represents both $\langle 1,1,8\rangle$ and $\langle 2,2,2\rangle$, which are sublattices of $\langle1,1,2\rangle$, also represents $\langle 1,1,2\rangle$ itself. From this point of view, the following definition seems to be quite natural: \begin{defn} Let $\ell$ be a ${\mathbb Z}$-lattice and let $S_0=\{\ell_1,\ell_2,\dots,\ell_t\}$ be the set of proper ${\mathbb Z}$-sublattices of $\ell$. We say $\ell$ is {\it recoverable by $S_0$} if any $S_0$-universal ${\mathbb Z}$-lattice represents $\ell$ itself. \end{defn} Note that the ternary ${\mathbb Z}$-lattice $\langle 1,1,2 \rangle$ is recoverable by $T_{1}$. We simply say $\ell$ is {\it recoverable} if there is a finite set of proper sublattices satisfying the above property. Note that if $\ell$ is recoverable, then there is a minimal $S$-universality criterion set whose cardinality is greater than $1$, where $S$ is the set of all sublattices of $\ell$. \begin{lem} \label{not-recover} A ${\mathbb Z}$-lattice $\ell$ is not recoverable if and only if there is a ${\mathbb Z}$-lattice that represents all proper sublattices of $\ell$, but not $\ell$ itself. \end{lem} \begin{proof} Suppose that $\ell$ is not recoverable. Let $S$ be the set of all proper sublattices of $\ell$ and let $S_{0}=\{\ell_1,\ell_2,\dots,\ell_t\}$ be a minimal $S$-universality criterion set. Since we are assuming that $\ell$ is not recoverable, there is a ${\mathbb Z}$-lattice $L$ that represents all ${\mathbb Z}$-lattices in $S_0$, whereas $L$ does not represent $\ell$ itself. Now, since the set $S_0$ is an $S$-universality criterion set, $L$ represents all proper sublattices of $\ell$, but not $\ell$ itself. The converse is trivial. \end{proof} \begin{lem} \label{scaling} Let $\ell$ be a ${\mathbb Z}$-lattice and let $a$ be a positive integer. If $\ell^a$ is recoverable, then so is $\ell$. \end{lem} \begin{proof} Assume that $\ell^a$ is recoverable by $\{\ell^a_1,\ell^a_2,\dots,\ell^a_t\}$, where $\ell_i$ is a proper sublattice of $\ell$ for any $i=1,2,\dots,t$. Let $M$ be any ${\mathbb Z}$-lattice that represents $\ell_i$ for any $i$. Then $\ell_i^a {\ \rightarrow\ } M^a$ for any $i$, and hence $\ell^a {\ \rightarrow\ } M^a$. Therefore $\ell {\ \rightarrow\ } M$ and $\ell$ is recoverable by $\{\ell_1,\ell_2,\dots,\ell_t\}$ . \end{proof} \begin{rmk}{\rm Any unary ${\mathbb Z}$-lattice $\ell$ cannot be recoverable. Let $\ell=\langle 1 \rangle$. Note that $\langle 2,2,5 \rangle$ represents all squares of integers except for $1$ (see \cite{jko}). Hence $\langle 2,2,5 \rangle$ represents all proper sublattices of $\ell$, but not $\ell$ itself. Therefore $\ell$ is not recoverable by Lemma \ref{not-recover}. Moreover, since every unary ${\mathbb Z}$-lattice can be obtained from $\ell$ by a suitable scaling, it is not recoverable by Lemma \ref{scaling}.} \end{rmk} \begin{rmk} {\rm Note that the converse of the above lemma does not hold in general. Let $\ell=\langle1,4\rangle$ be the binary ${\mathbb Z}$-lattice. Let $L$ be any ${\mathbb Z}$-lattice representing both $\ell_1=\langle1,16\rangle$ and $\ell_2=\langle4,4\rangle$. Since $L$ represents $\ell_1$, there is a vector $e_1 \in L$ and a ${\mathbb Z}$-sublattice $L_1$ of $L$ such that $L={\mathbb Z} e_1+L_1$, where $Q(e_1)=1$ and $B(e_1,L_1)=0$. Furthermore, since $L$ represents $\ell_2 = \langle4,4\rangle$, there are nonnegative integers $a,b$ and vectors $x,y \in L_1$ such that $$ Q(ae_1+x)=a^{2}+Q(x)=Q(be_1+y)=4 \quad \text{and} \quad B(ae_1+x,be_1+y)=0. $$ If $a=2$, then $x=0$ and $b=0$. Hence $\langle4\rangle {\ \rightarrow\ } L_1$. If $a=1$, then $$ b=0 \quad \text{and}\quad Q(y)=4 \quad\text{or}\quad b=1, \ Q(x)=Q(y)=3, \ \text{and}\ B(x,y)=-1. $$ For the latter case, we have $Q(x+y)=4$. Finally, if $a=0$, then $Q(x)=4$. Therefore $L_1$ represents $4$ in any case, which implies that $L$ represents $\ell$. Hence $\ell$ is recoverable by $\{\ell_1,\ell_2\}$. Now, we show that $\ell^2=\langle2,8\rangle$ is not recoverable. To show this, let $S$ be the set of all binary ${\mathbb Z}$-lattices with minimum greater than or equal to $9$, and let $S_0=\{\mathfrak m_1,\dots,\mathfrak m_t\}$ be a finite minimal $S$-universality criterion set. Then $\mathfrak m_1\perp \cdots\perp \mathfrak m_t$ represents all binary ${\mathbb Z}$-lattices with minimum greater than or equal to $9$. Now, we define $$ L=K \perp \mathfrak m_1\perp \cdots\perp \mathfrak m_t, \quad \text{where} \quad K=\begin{pmatrix} 2&1&1&0\\1&8&0&0\\1&0&8&4\\0&0&4&10\end{pmatrix}. $$ Clearly, $\ell^2=\langle2,8\rangle$ is not represented by $L$. We show that any proper sublattice of $\ell^2$ is represented by $L$. Let $\ell_{3}$ be any proper sublattice of $\ell^2$. If $\min(\ell_3) \ge 9$, then $\ell_3$ is represented by $\mathfrak m_1\perp\cdots\perp \mathfrak m_t$. Hence we may assume that $\min(\ell_3)=2$ or $8$. For the former case, we have $\ell_3 \simeq \langle2,8m^2\rangle$ for some integer $m\ge2$. Since $\langle 8m^2\rangle {\ \rightarrow\ } \mathfrak m_1\perp \cdots\perp \mathfrak m_t$, $L$ represents $\ell_3$. For the latter case, one may easily check that $\ell_3$ is isometric to one of the binary ${\mathbb Z}$-lattices $$ \langle 8,2m^2\rangle \quad \text{and} \quad \begin{pmatrix} 8&4\\4&2+8n^2\end{pmatrix}, $$ where $m\ge2$ and $n\ge1$. Note that $K$ represents the binary ${\mathbb Z}$-lattice $\ell_3$ for $m=2$ or $n=1$. If $m\ge3$ or $n\ge2$, then one may easily show that $$ \langle 8,2m^2\rangle {\ \rightarrow\ } \langle 8,8, 2m^2-8\rangle {\ \rightarrow\ } L \ \ \text{or} \ \ \begin{pmatrix} 8&4\\4&2+8n^2\end{pmatrix} {\ \rightarrow\ } \begin{pmatrix}8&4\\4&10\end{pmatrix} \perp \langle 8n^2-8\rangle {\ \rightarrow\ } L. $$ Therefore $L$ represents all proper sublattices of $\ell^2$, but not $\ell^2$ itself. Consequently, $\ell^2$ is not recoverable.} \end{rmk} One may easily check that every additively indecomposable ${\mathbb Z}$-lattice is not recoverable. We further prove that every indecomposable ${\mathbb Z}$-lattice $L$ with rank less than $4$ is not recoverable. \begin{prop}\label{bin} Any indecomposable binary ${\mathbb Z}$-lattice is not recoverable. \end{prop} \begin{proof} Let $\ell$ be an indecomposable binary ${\mathbb Z}$-lattice. Let $\{x,y\}$ be a Minkowski-reduced (ordered) basis for $\ell$, that is, $0 \le 2\vert B(x,y)\vert \le Q(x) \le Q(y)$. Let $S$ be the set of all proper sublattices of $\ell$, and let $S_{0}=\{\ell_{1},\ell_{2},\dots,\ell_{t}\}$ be a minimal $S$-universality criterion set. If we define $L=\ell_1\perp\dots\perp \ell_t$, then $L$ represents all proper sublattices of $\ell$. Hence it suffices to show that $L$ does not represent $\ell$ itself. To do this, let $\ell_i={\mathbb Z} x_i+{\mathbb Z} y_i$, where $\{x_i,y_i\}$ is a Minkowski reduced basis for $\ell_i$ for any $i=1,2,\dots,t$. Suppose on the contrary that there is a representation $\phi:\ell \rightarrow L$. Since $\ell_{i}$ is a sublattice of $\ell$ for any $i=1,2,\dots,t$, we may assume, without loss of generality, that $\phi(x)=x_{1}$. Suppose that $\phi(y) = \alpha x_{1} + \beta y_{1} + z$, where $\alpha,\beta \in {\mathbb Z}$ and $z \in ({\mathbb Z} x_{2}+{\mathbb Z} y_{2}) \perp \cdots \perp ({\mathbb Z} x_{t}+{\mathbb Z} y_{t})$. Since $\ell$ is indecomposable, $\beta$ cannot be zero. Then, we have $$ d\ell=d(\phi(\ell)) = d({\mathbb Z} x_{1}+{\mathbb Z} (\alpha x_{1} + \beta y_{1} + z)) \ge d({\mathbb Z} x_{1}+{\mathbb Z} (\alpha x_{1} + \beta y_{1})) \ge d\ell_{1} > d\ell, $$ which is a contradiction. Therefore $L$ does not represent $\ell$ and hence, by lemma \ref{not-recover}, the binary ${\mathbb Z}$-lattice $\ell$ is not recoverable. \end{proof} \begin{prop} Any indecomposable ternary ${\mathbb Z}$-lattice is not recoverable. \end{prop} \begin{proof} Suppose that there is a ternary ${\mathbb Z}$-lattice, say $L$, that is recoverable. Then there are proper sublattices $L_{1}, L_{2}, \dots, L_{t}$ of $L$ such that $L$ is represented by $L_{1} \perp L_{2} \perp \cdots \perp L_{t}$. Without loss of generality, we may assume that all $L_{i}$'s are of rank 3. Let $$ \phi : L {\ \rightarrow\ } L_{1} \perp L_{2} \perp \cdots \perp L_{t} $$ be a representation. Let $\{ u,v,w \}$ be a Minkowski reduced (ordered) basis for $L$. Without loss of generality, we may assume that $\phi(u)=x_{1} \in L_1$. Clearly, there exists a Minkowski reduced basis for $L_{1}$, say $\{ x_{1}, x_{2}, x_{3} \}$, containing $x_{1}$. Assume that $$ \phi(v) = a_{1}x_{1}+x+y, $$ where $a_1\in {\mathbb Z}$, $x \in {\mathbb Z} x_{2}+ {\mathbb Z} x_{3}$ and $y \in L_{2} \perp \cdots \perp L_{t}$. First, assume that $x=0$. Since $$ 2|a_{1}| Q(x_{1}) =2|B(x_{1}, a_{1}x_{1}+y)|=2\vert B(u,v)\vert\le Q(u)=Q(x_{1}), $$ we have $a_{1}=0$. Put $$ \phi(w) = b_{1}x_{1}+b_{2}x_{2}+b_{3}x_{3}+z, $$ where $b_i\in {\mathbb Z}$ for any $i=1,2,3$ and $z \in L_{2} \perp \cdots \perp L_{t}$. Then, we have $$ \phi(L)= {\mathbb Z} x_1 +{\mathbb Z} y +{\mathbb Z} (b_{1}x_{1}+b_{2}x_{2}+b_{3}x_{3}+z). $$ If $b_{3} \ne 0$, then $$ \begin{array}{lll} \mu_{3}(L) = Q(b_{1}x_{1}+b_{2}x_{2}+b_{3}x_{3}+z)\!\!\!&= Q(b_{1}x_{1}+b_{2}x_{2}+b_{3}x_{3})+Q(z)\\[0.2cm] \!\!\!&\ge \mu_{3}(L_{1}) +Q(z) \ge \mu_{3}(L)+Q(z), \end{array} $$ which implies that $z=0$. Hence, $\phi(L)$ is decomposable, which is a contradiction. Therefore, we have $b_{3}=0$. Observe that $$ L_{1} = {\mathbb Z} x_{1} + {\mathbb Z} x_{2} + {\mathbb Z} x_{3} \subseteq L={\mathbb Z} u +{\mathbb Z} v+{\mathbb Z} w \simeq \phi(L) $$ Then, $b_{1}x_{1} + b_{2}x_{2} = \alpha u + \beta v + \gamma w$ for some integers $\alpha, \beta$ and $\gamma$. If $\gamma \ne 0$, then $$ \mu_{3}(L) = Q(b_{1}x_{1}+b_{2}x_{2})+Q(z) \ge Q(w)+Q(z) = \mu_{3}(L)+Q(z). $$ This implies that $z=0$, which is a contradiction. Hence, $b_{1}x_{1} + b_{2}x_{2} = \alpha u + \beta v$. Similarly, we have $x_{1} = \alpha_{1} u + \beta_{1} v$ for some integers $\alpha_1$ and $\beta_1$. Since $Q(x_{1})=Q(u)$ and $B(u,v)=0$, we have $x_{1} = \pm u$ and $b_{1}x_{1} + b_{2}x_{2} = \beta v$ or $x_{1} = \pm v$ and $b_{1}x_{1} + b_{2}x_{2} = \alpha u$. In any case, $\phi(L)$ is decomposable, which is a contradiction. Finally, assume that $x \ne 0$. Since $$ \begin{array}{lll} \mu_{2}(L) = Q(v)=Q(a_{1}x_{1}+x+y) \!\!\!&= Q(a_{1}x_{1}+x)+Q(y)\\[0.2cm] \!\!\!& \ge \mu_{2}(L_{1}) +Q(y) \ge \mu_{2}(L) +Q(y), \end{array} $$ we have $y=0$. Put $$ \phi(v) = a_{1}x_{1}+a_{2}x_{2}+a_{3}x_{3} \quad \mbox{and}\quad \phi(w) = b_{1}x_{1}+b_{2}x_{2}+b_{3}x_{3}+z $$ where $a_i,b_i\in {\mathbb Z}$ for any $i=1,2,3$ and $z \in L_{2} \perp \cdots \perp L_{t}$. If $b_{3} \ne 0$, then $$ \mu_{3}(L) = Q(b_{1}x_{1}+b_{2}x_{2}+b_{3}x_{3})+Q(z) \ge \mu_{3}(L_{1})+Q(z) \ge \mu_{3}(L)+Q(z). $$ This implies that $z=0$. Then $\phi(L) \subseteq L_{1}$, which is a contradiction. Hence, $b_{3}=0$. Suppose that $a_{3} \ne 0$. Since $$ \mu_{2}(L)=Q(a_{1}x_{1}+a_{2}x_{2}+a_{3}x_{3}) \ge \mu_{3}(L_{1}) \ge \mu_{3}(L),$$ we have $\mu_{2}(L)= \mu_{3} (L) = \mu_{2}(L_{1}) = \mu_{3}(L_{1})$. Then, we have $$ \mu_{2}(L) = \mu_{3} (L) = Q(b_{1}x_{1}+b_{2}x_{2}+z) \ge \mu_{2}(L_{1})+Q(z) = \mu_{2}(L)+Q(z). $$ This implies that $z=0$, which is a contradiction. Therefore $a_{3} = 0$. Since $a_{2} \ne 0$, we have $$ \mu_{2}(L) = Q(a_{1}x_{1}+a_{2}x_{2}) \ge \mu_{2}(L_{1}) =Q(x_{2}) \ge \mu_{2}(L). $$ This means that $Q(a_{1}x_{1}+a_{2}x_{2})=Q(x_{2})$. Let $$ {\mathbb Z} x_{1}+{\mathbb Z} x_{2} = \begin{pmatrix} s & r \\ r & t \end{pmatrix}. $$ Then, we have $$ t=a_{1}^{2}s+2a_{1}a_{2}r+a_{2}^{2}t=s \left( a_{1} + \frac{ra_{2}}{s} \right)^{2} +a_{2}^{2} \left( t-\frac{r^{2}}{s} \right) \ge a_{2}^{2} \left( t-\frac{s}{4} \right). $$ If $|a_{2}| \ge 2$, then $t \ge 4t-s>t$, which is a contradiction. Hence, we have $a_{2} = \pm 1$. Then $\phi(L) = {\mathbb Z} x_{1}+{\mathbb Z} x_{2} +{\mathbb Z} z$ is decomposable, which is a contradiction. This completes the proof. \end{proof} \section{Recoverable binary $\mathbb{Z}$-lattices} In this section, we focus on recoverable binary ${\mathbb Z}$-lattices. We find some necessary conditions and some sufficient conditions for binary ${\mathbb Z}$-lattices to be recoverable. Let $n$ be a positive integer and let $S$ be the set of all binary ${\mathbb Z}$-lattices with minimum greater than or equal to $n$. Then there is a finite minimal $S$-universality criterion set $S_{n}=\{m_{1},\dots,m_{t}\}$ by \cite{kko}. If we define $M=m_{1}\perp \cdots\perp m_{t}$, then $M$ represents all binary ${\mathbb Z}$-lattices with minimum greater than or equal to $n$. In this section, $\MM{n}$ stands for a ${\mathbb Z}$-lattice representing all binary ${\mathbb Z}$-lattices with minimum greater than or equal to $n$ and $\min(\MM{n})=n$. From the above argument, such a ${\mathbb Z}$-lattice always exists. \begin{prop}\label{2leab} Let $a$ and $b$ be positive integers such that $2 \le a < b$ and $a$ does not divide $b$. Then the diagonal binary ${\mathbb Z}$-lattice $\ell=\langle a, b \rangle $ is not recoverable. \end{prop} \begin{proof} It suffices to show that there is a ${\mathbb Z}$-lattice such that it represents all proper sublattices of $\ell$, whereas it does not represent $\ell$ itself. Let $h$ be a positive integer such that $h^{2}a < b < (h+1)^{2}a$. For any $h \ge 2$, we define a ${\mathbb Z}$-lattice $$ K(h) =\perp_{\substack{2\le i\le h \\ 1 \le j \le [\frac{i}{2}]}} \begin{pmatrix}i^{2}a&ija\\ija&j^{2}a+b\end{pmatrix}. $$ Now, we define $$ L(h) = \begin{cases} \left( {\mathbb Z} x + {\mathbb Z} y \right) \perp \MM{b+1} & \text{if} \ h=1,\\ \left( {\mathbb Z} x + {\mathbb Z} y \right) \perp K(h) \perp \MM{b+1} & \text{otherwise}, \end{cases} $$ where ${\mathbb Z} x + {\mathbb Z} y = \begin{pmatrix} a&1\\1&b \end{pmatrix}$. We claim that $L(h)$ represents all proper sublattices of $\ell$, whereas it does not represent $\ell$ itself. First, we will prove that $L(h)$ represents all proper sublattices of $\ell$. Let $\ell'$ be a proper sublattice of $\ell$. If $\min(\ell') >b$, then $\MM{b+1}$ represents $\ell'$ and so does $L(h)$. If $\min(\ell') =b$, then $\ell' \simeq \langle b, \alpha^{2}a \rangle$ for some integer $\alpha$ with $\alpha^{2}a > b$. Since $\alpha^{2}a$ is represented by $\MM{b+1}$, $L(h)$ represents $\ell'$. Now, assume that $a<\min(\ell')<b$. Since $4a \le \min(\ell')$, we have $h \ge 2$. Furthermore, there are integers $i,j$, and $\beta$ with $2 \le i \le h$, $0 \le j \le \left[\frac{i}{2}\right]$ and $\beta \ge 1$ such that $$ \ell' \simeq \begin{pmatrix} i^{2}a & ija \\ ija & j^{2}a+\beta^{2}b \end{pmatrix}. $$ If $\beta=1$, then clearly $\ell' {\ \rightarrow\ } L(h)$. Assume that $\beta \ge 2$. Since $(\beta^{2}-1)b>b$, we have $$ \ell' \simeq \begin{pmatrix} i^{2}a & ija \\ ija & j^{2}a+\beta^{2}b \end{pmatrix} {\ \rightarrow\ } \begin{pmatrix} i^{2}a & ija \\ ija & j^{2}a+b \end{pmatrix} \perp \MM{b+1}, $$ which implies that $L(h)$ represents $\ell'$. Finally, if $\min(\ell') = a$, then $\ell' \simeq \langle a, \beta^{2}b \rangle$ for some integer $\beta \ge 2$. Since $\beta^{2}b$ is represented by $\MM{b+1}$, $L(h)$ represents $\ell'$. Next, we will show that $L(h)$ does not represent $\ell$. Clearly, $L(1)$ does not represent $\ell$. Assume $h \ge 2$. For any $i,j$ with $2 \le i \le h$ and $ 1 \le j \le \left[ \frac{i}{2} \right]$, let $$ K_{ij}={\mathbb Z} z + {\mathbb Z} w =\begin{pmatrix} i^{2}a&ija\\ija&j^{2}a+b \end{pmatrix}. $$ Since $$ Q(sz+tw) = (si+tj)^{2}a + t^{2}b >a, $$ for any integers $s$ and $t$, the binary ${\mathbb Z}$-lattice $K_{ij}$ does not represent $a$. Suppose that $Q(sz+tw)=b$ for some integers $s$ and $t$. Since $a$ does not divide $b$, we have $t^{2}=1$ and $si+tj=0$. Furthermore, since $j = |si| \le \left[ \frac{i}{2} \right]$, we have $s=j=0$. This is a contradiction. Hence $K_{ij}$ does not represent $b$ for any possible integers $i$ and $j$. Therefore we have \begin{equation}\label{qvalue} Q(K(h)) \subseteq (\left\{ ma+nb \mid m,n \in {\mathbb N} \cup\{0\} \right\} \setminus \left\{ a, b \right\}). \end{equation} Suppose that $L(h)$ represents $\ell$. Let $u \in L(h)$ be a vector with $Q(u) = b$. Then $$ u=\alpha x + \beta y + z + w, $$ for some integers $\alpha, \beta$ and some vectors $z \in K(h)$ and $w \in \MM{b+1}$. Since $$ Q(u) = Q(\alpha x + \beta y) + Q(z) + Q(w)=b \quad \text{and}\quad Q(w) > b, $$ we have $w=0$. By \eqref{qvalue}, we have $Q(z)=0$ or $Q(z) = \delta a$ for some integer $\delta \ge 2$. If $|\beta| \ge 2$, then $Q(\alpha x + \beta y) \ge \beta^{2}(b-1) >b $. If $\beta=0$, then $Q(\alpha x)$ is a multiple of $a$, and so is $Q(u)$. Hence, we have $|\beta|=1$. This implies that $$ u= \begin{cases} \pm (x-y) \ \text{or} \ \pm y & \text{if} \ a=2,\\ \pm y& \text{if} \ a \ge 3. \end{cases} $$ Let $v \in L(h)$ be a vector with $Q(v) = a$. Then $v=x$ or $v=-x$. Finally, we have $B(u, v) \ne 0$, which is a contradiction. This completes the proof. \end{proof} \begin{prop}\label{odd} For any positive odd integer $m$, the diagonal binary ${\mathbb Z}$-lattice $\ell=\langle 1, m \rangle$ is not recoverable. \end{prop} \begin{proof} For a positive odd integer $m$, let $\ell=\langle 1, m \rangle$ be the diagonal ${\mathbb Z}$-lattice. Since $\langle 1 \rangle \perp \MM{2}$ represents all proper sublattices of $\langle 1,1 \rangle$, but not $\langle 1,1 \rangle$ itself, the binary ${\mathbb Z}$-lattice $\langle 1, 1 \rangle$ is not recoverable. Hence we may assume that $m \ge 3$. Let $N$ be any quinary ${\mathbb Z}$-lattice that represents all binary even ${\mathbb Z}$-lattices. Note that such a ${\mathbb Z}$-lattice exists, for example, the root lattice $D_{5}$ is one of such quinary ${\mathbb Z}$-lattices (see \cite{jko2}). Define a ${\mathbb Z}$-lattice $$ L=\langle 1 \rangle \perp N \perp \MM{m+1}. $$ It is obvious that $\langle 1, m \rangle$ is not represented by $L$. Let $\ell_1$ be any proper ${\mathbb Z}$-sublattice of $\ell$. First, suppose that $\min(\ell_1)=1$. Then $\ell_1 \simeq \langle 1, m\beta^{2} \rangle$ for some integer $\beta \ge 2$. Since $\langle m\beta^{2} \rangle {\ \rightarrow\ } \MM{m+1}$, we have $\ell {\ \rightarrow\ } L$. Now, suppose that $\min(\ell_1)>1$. Then clearly, $\min(\ell_1) \ge 3$. Choose a Minkowski reduced basis for $\ell_1$ so that for some integers $a,b$, and $c$ with $0 \le 2b \le a \le c$ such that $$ \ell_1 \simeq \begin{pmatrix} a&b\\ b&c \end{pmatrix}. $$ Note that we are assuming that $a\ge 3$. If $a \equiv c \equiv 0 \pmod{2}$, then $\ell_1{\ \rightarrow\ } N$ and so $\ell_1 {\ \rightarrow\ } L$. If $a \equiv c \equiv 1 \pmod{2}$, then we define a ${\mathbb Z}$-lattice $$ \ell_1' = \begin{pmatrix} a-1&b-1\\ b-1&c-1 \end{pmatrix}. $$ Since $d\ell_1' \ge \frac{3}{4}ac-c =\frac{3c}{4}(a-\frac43) > 0$, the even ${\mathbb Z}$-lattice $\ell_1'$ is positive definite. Hence $\ell_1' {\ \rightarrow\ } N$ and therefore $\ell_1 {\ \rightarrow\ } \langle 1 \rangle \perp N {\ \rightarrow\ } L$. If $a \equiv 1 \pmod{2}$ and $c\equiv 0 \pmod{2}$, then we define a ${\mathbb Z}$-lattice $$ \ell_1'' = \begin{pmatrix} a-1&b\\ b&c \end{pmatrix}. $$ Since $d\ell_1'' = \left( \frac{ac}{4}-b^{2} \right) +\frac{c}{4} \left( 3a-4 \right) >0$, the even ${\mathbb Z}$-lattice $\ell_1''$ is positive definite. Hence $\ell_1'' {\ \rightarrow\ } N$ and therefore $\ell_1 {\ \rightarrow\ } \langle 1\rangle \perp \ell_1'' {\ \rightarrow\ } \langle 1\rangle \perp N {\ \rightarrow\ } L$. Since the proof of the case when $a \equiv 0 \pmod{2}$ and $c\equiv 1 \pmod{2}$ is quite similar to this, the proof is left to the readers. \end{proof} \begin{prop} \label{2m4} For any positive integer $m$ with $m \equiv 2 \pmod{4}$, the diagonal binary ${\mathbb Z}$-lattice $\ell=\langle 1, m \rangle$ is not recoverable. \end{prop} \begin{proof} For a positive integer $m \equiv 2 \pmod{4}$, let $\ell=\langle 1,m \rangle$ be the diagonal ${\mathbb Z}$-lattice. Since $\langle 1,1 \rangle \perp \MM{3}$ represents all proper sublattices of $\langle 1,2 \rangle$, but it does not represent $\langle 1,2 \rangle$ itself, the binary ${\mathbb Z}$-lattice $\langle 1,2 \rangle$ is not recoverable. If we define $$ L=\langle 1 \rangle \perp \begin{pmatrix} 4 & 0 & 2 \\ 0 & 5 & 1 \\ 2 & 1 & 7 \end{pmatrix} \perp \MM{7}, $$ then one may check that $L$ represents all proper sublattices of $\langle 1, 6 \rangle$, but it does not represent $\langle 1, 6 \rangle$ itself. From now on, we assume that $m \ge 10$. Define $$ L_m' = {\mathbb Z} x + {\mathbb Z} y + {\mathbb Z} z + {\mathbb Z} t = \langle 1,3,5,m-1 \rangle. $$ Let $N$ be an even 2-universal quinary ${\mathbb Z}$-lattice and let $\mathcal{N}$ be the ${\mathbb Z}$-lattice obtained from $N$ by scaling the quadratic space $\mathbb{Q} \otimes N$ by $2$. Hence $\mathcal N$ represents all binary ${\mathbb Z}$-lattices whose norm is contained in $4{\mathbb Z}$. Now, we define $$ L_m = L_m' \perp \mathcal{N} \perp \MM{m+1}. $$ We will show that any proper sublattice of $\ell=\langle 1, m \rangle$ is represented by $L_m$, whereas $\ell$ itself is not represented by $L_m$. Suppose, on the contrary, that $\langle 1, m \rangle$ is represented by $L_m$. Then, one may easily check that $$ \langle m \rangle \longrightarrow \langle 3,5 \rangle \perp \mathcal{N}. $$ Hence we have $m \equiv 3a^2+5b^2 \pmod{4}$ for some integers $a$ and $b$, which is a contradiction to the fact that $m \equiv 2 \pmod{4}$. Let $\ell= {\mathbb Z} u +{\mathbb Z} v = \langle 1, m \rangle$ and let $\ell_1$ be any proper sublattice of $\ell$. If $\min(\ell_1)=1$, then $\ell_1\simeq \langle1,m\beta^2 \rangle$ for some integer $\beta\ge2$. Clearly, $\ell_1 {\ \rightarrow\ } L_m$. Assume that $1< \min(\ell_1)< m$. Then there are integers $a,b$, and $c$ such that $\ell_1= {\mathbb Z} (au) + {\mathbb Z} (bu+ cv)$. If $|c| \ge 2$, then we have $\ell_1 \subseteq {\mathbb Z} u + {\mathbb Z} (cv) = \langle 1, c^{2}m \rangle {\ \rightarrow\ } L_m$. Hence we may assume that $\ell_1 = {\mathbb Z} (au)+{\mathbb Z} (bu+v)$, where the integers $a$ and $b$ satisfy $a \ge 2$ and $0 \le b < a$. Note that $$ \ell_1 = \begin{pmatrix} a^{2} & ab \\ ab & b^{2}+m \end{pmatrix}. $$ First, assume that $a \equiv b \equiv 0 \pmod{2}$. Since $$ \begin{pmatrix} a^{2} & ab \\ ab & b^{2}+m-6 \end{pmatrix} {\ \rightarrow\ } \mathcal{N}, $$ we have $\ell_1{\ \rightarrow\ } \langle 1,5 \rangle \perp \mathcal{N} {\ \rightarrow\ } L_m$. Now, assume that $a \equiv 0 \pmod{2}$ and $b\equiv 1 \pmod{2}$. Since $$ \begin{pmatrix} a^{2} & ab \\ ab & b^{2}+m-3 \end{pmatrix} {\ \rightarrow\ } \mathcal{N}, $$ we have $\ell_1 {\ \rightarrow\ } \langle 3 \rangle \perp \mathcal{N} {\ \rightarrow\ } L_m$. Assume that $a \equiv b \equiv 1 \Mod{2}$. Let $w\in\mathcal{N}$ be a vector with $Q(w)=m-2$. Then, we have $$ {\mathbb Z} (x+y+z)+{\mathbb Z}(y+w)=\begin{pmatrix} 9&3\\ 3&1+m \end{pmatrix} {\ \rightarrow\ } L_m. $$ Hence we may assume that $a \ge 5$. Consider the following ${\mathbb Z}$-lattice $$ \ell_1' = \begin{pmatrix} a^{2}-9 & ab-3 \\ ab-3 & b^{2}+m-3 \end{pmatrix}. $$ Since $m > a^2$, we have $d (\ell_1')>0$. Hence $\ell_1' {\ \rightarrow\ } \mathcal{N}$. Therefore there are vectors $w_{1}, w_{2} \in \mathcal{N}$ such that $$ \ell_1' \simeq {\mathbb Z} w_{1}+{\mathbb Z} w_{2} \subseteq \mathcal{N}. $$ Then, we have $$ {\mathbb Z} (x+y+z+w_{1}) + {\mathbb Z} (y+w_{2}) = \begin{pmatrix} a^{2} & ab \\ ab & b^{2}+m \end{pmatrix}. $$ This implies that $\ell_1$ is represented by $L_m$. Finally, assume that $a \equiv 1 \pmod{2}$ and $b\equiv 0 \pmod{2}$. If $a=3$, then $b=0$ or $2$. If $b=0$, then $$ \ell_1 = \langle 9,m \rangle \longrightarrow \langle 1,5,m-1 \rangle \perp \mathcal{N} {\ \rightarrow\ } L_m. $$ If $b=2$, then we have $$ \ell_1 = \begin{pmatrix} 9 & 6 \\ 6 & m+4 \end{pmatrix} \simeq \begin{pmatrix} 9 & 3 \\ 3 & 1+m \end{pmatrix} {\ \rightarrow\ } L_m. $$ Now, assume that $a \ge 5$. Consider the following ${\mathbb Z}$-lattice $$ \ell_1'' = \begin{pmatrix} a^{2}-9 & ab-4 \\ ab-4 & b^{2}+m-6 \end{pmatrix}. $$ Since $d (\ell_1'')>0$, we have $\ell_1'' {\ \rightarrow\ } \mathcal{N}$. Hence there are vectors $w'_{1}, w'_{2} \in \mathcal{N}$ such that $$ \ell_1'' \simeq {\mathbb Z} w'_{1}+{\mathbb Z} w'_{2} \subseteq \mathcal{N}. $$ Then, we have $$ {\mathbb Z} (x+y+z+w'_{1}) + {\mathbb Z} (-x+z+w'_{2}) = \begin{pmatrix} a^{2} & ab \\ ab & b^{2}+m \end{pmatrix}. $$ Therefore, we have $\ell_1 {\ \rightarrow\ } L_m$. If $\min(\ell_1)=m$, then $\ell_1 \simeq \langle \alpha^2,m \rangle$ for some integer $\alpha$ with $\alpha^2>m$. Hence, we have $\ell_1 {\ \rightarrow\ } \langle 1,m-1\rangle\perp \MM{m+1} {\ \rightarrow\ } L_m$. Finally, if $\min(\ell_1)>m$, then we have $\ell_1 {\ \rightarrow\ } \MM{m+1} {\ \rightarrow\ } L_m$. This completes the proof. \end{proof} \section{Recoverable numbers} From Propositions \ref{bin}, \ref{2leab}, \ref{odd}, and \ref{2m4}, one may conclude that if a binary ${\mathbb Z}$-lattice $\ell$ is recoverable, then $\ell = \langle a, 4ma \rangle$ for some positive integers $a$ and $m$. In this section, we focus on the case when $a=1$. A positive integer $m$ is called \textit{recoverable} if the diagonal binary ${\mathbb Z}$-lattice $\langle 1,4m \rangle$ is recoverable. We prove that any square of an integer is a recoverable number. We also prove that there are infinitely many non square recoverable numbers. \begin{prop} Any square of an integer is recoverable, that is, the diagonal binary ${\mathbb Z}$-lattice $\ell = \langle 1, 4m^{2} \rangle$ is recoverable for any integer $m$. \end{prop} \begin{proof} Let $\ell= \langle 1, 4m^{2} \rangle$ be the diagonal binary ${\mathbb Z}$-lattice. Let $S$ be the set of all proper sublattices of $\ell$. By Lemma \ref{not-recover}, it suffices to show that any $S$-universal ${\mathbb Z}$-lattice represents $\ell$ itself. Let $L$ be an $S$-universal ${\mathbb Z}$-lattice. Since $\langle 1, 16m^{2} \rangle {\ \rightarrow\ } L$, we have $L={\mathbb Z} e_{1} \perp L' =\langle 1 \rangle \perp L'$ for some ${\mathbb Z}$-lattice $L'$. Since $\langle 4, 4m^{2} \rangle {\ \rightarrow\ } L$, one of the following holds: \begin{enumerate} \item there is a vector $y \in L'$ such that ${\mathbb Z} (2e_{1}) + {\mathbb Z} y = \langle 4, 4m^{2} \rangle$; \item there are vectors $x,y \in L'$ and an integer $a$ such that $$ {\mathbb Z} (e_{1}+x) + {\mathbb Z} (ae_{1}+y) = \langle 4, 4m^{2} \rangle; $$ \item there are vectors $x,y \in L'$ and an integer $a$ such that $$ {\mathbb Z} x + {\mathbb Z} (ae_{1}+y) = \langle 4, 4m^{2} \rangle. $$ \end{enumerate} If (1) holds, then $Q(y)=4m^{2}$. Hence $L$ represents $\ell$. If (2) holds, then $$ {\mathbb Z} x+{\mathbb Z} y = \begin{pmatrix} 3&-a\\ -a & 4m^{2}-a^{2}\end{pmatrix}. $$ Hence we have $Q(ax+y)=4m^{2}$. Therefore $L$ represents $\ell$. Finally, if (3) holds, then we have $Q(mx)=4m^{2}$. Therefore $L$ represents $\ell$. This completes the proof. \end{proof} Let $\mathscr{L}$ be the set of all isometry classes of binary ${\mathbb Z}$-lattices and let $\mathscr{L}_{13}$ be the set of all isometry classes of binary ${\mathbb Z}$-lattices whose second successive minimum is greater than or equal to $13$. We define a map $\phi_{9} : \mathscr{L}_{13} {\ \rightarrow\ } \mathscr{L}$ by $$ \phi_{9} \left( \begin{pmatrix} a & b \\ b & c \end{pmatrix} \right) = \begin{pmatrix} a & b \\ b & c-9 \end{pmatrix}, $$ where $\begin{pmatrix} a & b \\ b & c \end{pmatrix}$ is a Minkowski-reduced form in the class so that $0 \le 2b \le a \le c$. Since $$ d(\phi_{9}(K))=ac-b^2-9a=\left(\frac {ac}4-b^2\right)+\frac{3a}4(c-12)>0, $$ the above map $\phi_9$ is well-defined. \begin{lem}\label{phi9} Let $L$ be a ${\mathbb Z}$-lattice and let $K$ be a binary ${\mathbb Z}$-lattice in $\mathscr{L}_{13}$. If $\phi_{9}^{k}(K)$ is represented by $L$ for some nonnegative integer $k$, then $$ K \longrightarrow L \perp 9I_{5}. $$ Here, $9I_{5}$ is the quinary ${\mathbb Z}$-lattice obtained from the cubic lattice $I_{5}$ by scaling the quadratic space $\mathbb{Q} \otimes I_{5}$ by $9$. \end{lem} \begin{proof} Let $L$ be a ${\mathbb Z}$-lattice and let $K$ be a binary ${\mathbb Z}$-lattice in $\mathscr{L}_{13}$. Let $\begin{pmatrix} a & b \\ b & c \end{pmatrix}$ be the Minkowski-reduced form in the isometry class of $K$. Note that $9I_{5}$ represents all binary ${\mathbb Z}$-lattices whose scale is contained in $9{\mathbb Z}$. We use an induction on $k$. Suppose that $\phi_{9}(K)$ is represented by $L$. Then, it is obvious that $$ K= \begin{pmatrix} a & b \\ b & c \end{pmatrix} = \begin{pmatrix} a & b \\ b & c-9 \end{pmatrix} + \begin{pmatrix} 0 & 0 \\ 0 & 9 \end{pmatrix} {\ \rightarrow\ } L \perp 9I_{5}. $$ Suppose that the assertion is true for $k$. Assume that $\phi_{9}^{k+1}(K) {\ \rightarrow\ } L$. Let $K'=\phi_{9}(K)$. Then $\phi_{9}^{k}(K') = \phi_{9}^{k+1}(K) {\ \rightarrow\ } L$. It follows from the induction hypothesis that $K' {\ \rightarrow\ } L \perp 9I_{5}$. This implies that $$ K' =\begin{pmatrix} a & b \\ b & c-9 \end{pmatrix} = \begin{pmatrix} \alpha_{1} & \beta_{1} \\ \beta_{1} & \gamma_{1} \end{pmatrix} + \begin{pmatrix} \alpha_{2} & \beta_{2} \\ \beta_{2} & \gamma_{2} \end{pmatrix}, $$ where $\begin{pmatrix} \alpha_{1} & \beta_{1} \\ \beta_{1} & \gamma_{1} \end{pmatrix} {\ \rightarrow\ } L$ and $\begin{pmatrix} \alpha_{2} & \beta_{2} \\ \beta_{2} & \gamma_{2} \end{pmatrix} {\ \rightarrow\ } 9I_{5}$. Since $ \begin{pmatrix} \alpha_{2} & \beta_{2} \\ \beta_{2} & \gamma_{2}+9 \end{pmatrix} {\ \rightarrow\ } 9I_{5}, $ we have $$ K = \begin{pmatrix} a & b \\ b & c \end{pmatrix} =\begin{pmatrix} \alpha_{1} & \beta_{1} \\ \beta_{1} & \gamma_{1} \end{pmatrix} + \begin{pmatrix} \alpha_{2} & \beta_{2} \\ \beta_{2} & \gamma_{2}+9 \end{pmatrix} {\ \rightarrow\ } L \perp 9I_{5}. $$ This completes the proof. \end{proof} \begin{lem}\label{ord3m} Any proper sublattice of $\langle 1,1 \rangle$ is represented by both $$ L_{1}= \langle 1,2,3 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix} \perp 9I_{5} \ \mbox{ and }\ L_{2}=\langle 1,2,6 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix} \perp 9I_{5}. $$ \end{lem} \begin{proof} Since the proof is quite similar to each other, we only provide the proof of the first case. Let $\ell$ be any proper sublattice of $\langle 1,1 \rangle$. One may directly check that the quinary ${\mathbb Z}$-lattice $\langle 1,2,3 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix}$ represents all binary ${\mathbb Z}$-lattices whose second successive minimum is less than or equal to $12$ except for the following $15$ binary ${\mathbb Z}$-lattices: \begin{equation}\label{15lattices} \begin{array}{ll} & \begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix}, \begin{pmatrix} 1 & 0 \\ 0 & 6 \end{pmatrix}, \begin{pmatrix} 2 & 1 \\ 1 & 2 \end{pmatrix}, \begin{pmatrix} 2 & 1 \\ 1 & 3 \end{pmatrix}, \begin{pmatrix} 2 & 1 \\ 1 & 4 \end{pmatrix},\\[0.3cm] & \begin{pmatrix} 4 & 0 \\ 0 & 6 \end{pmatrix}, \begin{pmatrix} 4 & 1 \\ 1 & 4 \end{pmatrix}, \begin{pmatrix} 4 & 1 \\ 1 & 13 \end{pmatrix},\begin{pmatrix} 4 & 2 \\ 2 & 7 \end{pmatrix}, \begin{pmatrix} 6 & 0 \\ 0 & 7 \end{pmatrix}, \\[0.3cm] & \begin{pmatrix} 6 & 0 \\ 0 & 10 \end{pmatrix}, \begin{pmatrix} 6 & 3 \\ 3 & 7 \end{pmatrix}, \begin{pmatrix} 6 & 3 \\ 3 & 10 \end{pmatrix}, \begin{pmatrix} 7 & 1 \\ 1 & 10 \end{pmatrix}, \begin{pmatrix} 10 & 2 \\ 2 & 10 \end{pmatrix}. \end{array} \end{equation} Note that the above $15$ binary ${\mathbb Z}$-lattices are not proper sublattices of $\langle 1,1 \rangle$. Hence we may assume that $\mu_2(\ell)\ge 13$. Since $\mu_{2}(\phi_{9}(\ell)) \le \max\{\mu_{1}(\ell), \mu_{2}(\ell)-9 \}$, there exists a positive integer $k$ such that $\phi_{9}^{k-1}(\ell) \in \mathscr{L}_{13}$ and $\mu_{2}(\phi_{9}^{k}(\ell)) \le 12$. When $k=1$, then we let $\phi_9^0(\ell)=\ell$. If $$ \phi_{9}^{k}(\ell) {\ \rightarrow\ } \langle 1,2,3 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix}, $$ then, by Lemma \ref{phi9}, we have $$ \ell {\ \rightarrow\ } \langle 1,2,3 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix} \perp 9I_{5}. $$ Hence, we may assume that $\phi_{9}^{k}(\ell)$ is isometric to one of 15 binary ${\mathbb Z}$-lattices listed in \eqref{15lattices}. Since $d\ell$ is a square of an integer and $d(\phi_{9}(\ell)) = d\ell -9\mu_{1}(\ell)$, we see that $\text{ord}_{3}(d(\phi_{9}^{k}(\ell)))$ cannot be one. Hence $\phi_{9}^{k}(\ell)$ is isometric to one of the following ${\mathbb Z}$-lattices: $$ \begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix}, \begin{pmatrix} 2 & 1 \\ 1 & 3 \end{pmatrix}, \quad \text{and} \quad \begin{pmatrix} 2 & 1 \\ 1 & 4 \end{pmatrix}. $$ Since $\phi_{9}^{k-1}(\ell) \in \mathscr{L}_{13}$ and $$ \begin{array}{ll} & \phi_{9}^{-1}\left(\begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix}\right)= \left\{ \begin{pmatrix} 1 & 0 \\ 0 & 10 \end{pmatrix}, \begin{pmatrix} 2 & 1 \\ 1 & 10 \end{pmatrix}, \begin{pmatrix} 5 & 2 \\ 2 & 10 \end{pmatrix}, \begin{pmatrix} 10 & 3 \\ 3 & 10 \end{pmatrix}\right\},\\[0.3cm] & \phi_{9}^{-1}\left(\begin{pmatrix} 2 & 1 \\ 1 & 3 \end{pmatrix}\right)= \left\{ \begin{pmatrix} 2 & 1 \\ 1 & 12 \end{pmatrix}, \begin{pmatrix} 3 & 1 \\ 1 & 11 \end{pmatrix}, \begin{pmatrix} 7 & 3 \\ 3 & 11 \end{pmatrix}, \begin{pmatrix} 10 & 5 \\ 5 & 12 \end{pmatrix} \right\},\\[0.3cm] & \phi_{9}^{-1}\left(\begin{pmatrix} 2 & 1 \\ 1 & 4 \end{pmatrix}\right)= \left\{ \begin{pmatrix} 2 & 1 \\ 1 & 13 \end{pmatrix}, \begin{pmatrix} 4 & 1 \\ 1 & 11 \end{pmatrix}, \begin{pmatrix} 8 & 3 \\ 3 & 11 \end{pmatrix}\right\}, \end{array} $$ we have $$ \phi_{9}^{k}(\ell)\simeq \begin{pmatrix} 2 & 1 \\ 1 & 4 \end{pmatrix}\quad \text{and}\quad \phi_{9}^{k-1}(\ell)\simeq \begin{pmatrix} 2 & 1 \\ 1 & 13 \end{pmatrix}. $$ One may check that $$ \phi_{9}^{k-1}(\ell)\simeq\begin{pmatrix} 2 & 1 \\ 1 & 13 \end{pmatrix} {\ \rightarrow\ } \langle 1,2,3 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix}. $$ Hence, by Lemma \ref{phi9}, we have $$ \ell \longrightarrow \langle 1,2,3 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix} \perp 9I_{5}. $$ This completes the proof. \end{proof} \begin{prop} If $m$ is a positive integer with $\text{ord}_{3}(m)=1$, then $m$ is not a recoverable number. \end{prop} \begin{proof} Let $m$ be a positive integer with $\text{ord}_{3}(m)=1$. Then, we write $m=3m'$ with $m'\equiv 1$ or $2 \Mod 3$. We define $$ L_m = \begin{cases} \langle 1,2,6,4m-1 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix} \perp 9I_{5}\perp \MM{4m+1} & \text{if $m'\equiv 1 \Mod 3$},\\[0.5cm] \langle 1,2,3,4m-1 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix} \perp 9I_{5}\perp \MM{4m+1} & \text{if $m'\equiv 2 \Mod 3$}. \end{cases} $$ Clearly, $L_m$ does not represent $\langle 1, 4m \rangle$. As in the proof of Lemma \ref{2m4}, it is enough to show that $L_m$ represents every proper sublattice of $\langle 1,4m \rangle$, which is of the form $\begin{pmatrix} a^{2} & ab \\ ab & b^{2}+4m \end{pmatrix}$ for some integers $a$ and $b$ with $a \ge 2$ and $0\le 2b \le a$. By Lemma \ref{ord3m}, we have $$ \begin{pmatrix} a^{2} & ab \\ ab & b^{2}+1 \end{pmatrix} {\ \rightarrow\ } \langle 1,2,3 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix} \perp 9I_{5} \ \text{and} \ \langle 1,2,6 \rangle \perp \begin{pmatrix} 2 & 1 \\ 1 & 5 \end{pmatrix} \perp 9I_{5}, $$ which implies that $$ \begin{pmatrix} a^{2} & ab \\ ab & b^{2}+4m \end{pmatrix} {\ \rightarrow\ } L_m. $$ This completes the proof. \end{proof} \begin{prop}\label{12lattices} Any integer $m$ is a recoverable number if $4m$ is represented by all of the following binary ${\mathbb Z}$-lattices $$ \begin{array}{lll} &\begin{pmatrix} 2 & 1 \\ 1 & 4 \end{pmatrix}, \begin{pmatrix} 3 & 1 \\ 1 & 4 \end{pmatrix}, \begin{pmatrix} 4 & 0 \\ 0 & 4 \end{pmatrix}, \begin{pmatrix} 4 & 0 \\ 0 & 5 \end{pmatrix}, \begin{pmatrix} 4 & 1 \\ 1 & 6 \end{pmatrix}, \begin{pmatrix} 4 & 1 \\ 1 & 7 \end{pmatrix},\\[0.3cm] &\begin{pmatrix} 4 & 0 \\ 0 & 8 \end{pmatrix}, \begin{pmatrix} 4 & 1 \\ 1 & 8 \end{pmatrix}, \begin{pmatrix} 4 & 2 \\ 2 & 8 \end{pmatrix}, \begin{pmatrix} 4 & 0 \\ 0 & 9 \end{pmatrix}, \begin{pmatrix} 4 & 1 \\ 1 & 9 \end{pmatrix}, \quad \text{and}\quad \begin{pmatrix} 4 & 2 \\ 2 & 9 \end{pmatrix}. \end{array} $$ \end{prop} \begin{proof} Let $L$ be any ${\mathbb Z}$-lattice that represents all proper sublattices of $\langle 1,4m \rangle$. Since $\langle 1, 16m \rangle {\ \rightarrow\ } L$, we have $L={\mathbb Z} e_1 \perp L'= \langle 1 \rangle \perp L'$ for some ${\mathbb Z}$-lattice $L'$. To prove the proposition, it suffices to show that $\langle 1,4m\rangle$ is represented by $L$, that is, $4m$ is represented by $L'$. Since $\langle 4, 4m \rangle {\ \rightarrow\ } L$, one of the followings holds: \begin{enumerate} \item there is a vector $y \in L'$ such that ${\mathbb Z} (2e_{1}) + {\mathbb Z} y = \langle 4, 4m \rangle$; \item there are vectors $x,y \in L'$ and an integer $a$ such that $$ {\mathbb Z} (e_{1}+x) + {\mathbb Z} (ae_{1}+y) = \langle 4, 4m \rangle; $$ \item there are vectors $x,y \in L'$ and an integer $a$ such that $$ {\mathbb Z} x + {\mathbb Z} (ae_{1}+y) = \langle 4, 4m \rangle. $$ \end{enumerate} If (1) or (2) holds, then $4m$ is represented by $L'$. Therefore we may assume that (3) holds. Hence $4$ is represented by $L'$. Now, note that $\langle 9,4m \rangle$ is also represented by $L$. Similarly to the above, one may easily show that $4m$ is represented by $L'$ or one of binary ${\mathbb Z}$-lattices $\begin{pmatrix} 8 & -a \\ -a & 4m-a^2 \end{pmatrix}$ and $\langle 9, 4m-a^2\rangle$ is represented by $L'$. Hence $8$ or $9$ is represented by $L'$. Suppose that $L'$ represents 4 and 8. Then $L'$ represents at least one of the following binary ${\mathbb Z}$-lattices: $$ \begin{pmatrix} 4 & 0 \\ 0 & 8 \end{pmatrix}, \begin{pmatrix} 4 & 1 \\ 1 & 8 \end{pmatrix}, \begin{pmatrix} 4 & 2 \\ 2 & 8 \end{pmatrix}, \begin{pmatrix} 4 & 3 \\ 3 & 8 \end{pmatrix}, \begin{pmatrix} 4 & 4 \\ 4 & 8 \end{pmatrix}, \quad \text{and}\quad \begin{pmatrix} 4 & 5 \\ 5 & 8 \end{pmatrix}. $$ Here, we have $$ \begin{pmatrix} 4 & 3 \\ 3 & 8 \end{pmatrix} \simeq \begin{pmatrix} 4 & 1 \\ 1 & 6 \end{pmatrix}, \begin{pmatrix} 4 & 4 \\ 4 & 8 \end{pmatrix} \simeq \begin{pmatrix} 4 & 0 \\ 0 & 4 \end{pmatrix},\quad \text{and}\quad \begin{pmatrix} 4 & 5 \\ 5 & 8 \end{pmatrix} \simeq \begin{pmatrix} 2 & 1 \\ 1 & 4 \end{pmatrix}. $$ Finally, suppose that $L'$ represents 4 and 9. Then $L'$ represents at least one of the following binary ${\mathbb Z}$-lattices: $$ \begin{pmatrix} 4 & 0 \\ 0 & 9 \end{pmatrix}, \begin{pmatrix} 4 & 1 \\ 1 & 9 \end{pmatrix}, \begin{pmatrix} 4 & 2 \\ 2 & 9 \end{pmatrix}, \begin{pmatrix} 4 & 3 \\ 3 & 9 \end{pmatrix}, \begin{pmatrix} 4 & 4 \\ 4 & 9 \end{pmatrix}, \quad\text{and}\quad \begin{pmatrix} 4 & 5 \\ 5 & 9 \end{pmatrix}, $$ Here, we have $$ \begin{pmatrix} 4 & 3 \\ 3 & 9 \end{pmatrix} \simeq \begin{pmatrix} 4 & 1 \\ 1 & 7 \end{pmatrix}, \begin{pmatrix} 4 & 4 \\ 4 & 9 \end{pmatrix} \simeq \begin{pmatrix} 4 & 0 \\ 0 & 5 \end{pmatrix}, \quad\text{and}\quad \begin{pmatrix} 4 & 5 \\ 5 & 9 \end{pmatrix} \simeq \begin{pmatrix} 3 & 1 \\ 1 & 4 \end{pmatrix}. $$ Therefore, if $4m$ is represented by all of the above $12$ binary ${\mathbb Z}$-lattices, then $4m$ is represented by $L'$. This completes the proof. \end{proof} \begin{cor} Let $m \equiv 1 \pmod 8$ be a prime. If $m$ is a quadratic residue modulo $q$ for any prime $q\in\{3,5,7,11,23,31\}$, then $m$ is a recoverable number. In particular, any prime $m \equiv 5569 \pmod {3\cdot 5\cdot 7\cdot 11\cdot 23\cdot 31}$ is a recoverable number. Therefore there are infinitely many non square recoverable numbers. \end{cor} \begin{proof} Note that $4m$ is represented by all of $12$ binary ${\mathbb Z}$-lattices in Proposition \ref{12lattices} if and only if $m$ is represented by all of the following binary ${\mathbb Z}$-lattices \begin{equation}\label{9lattices} \begin{array}{lll} &\hspace{0.7cm}\begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix}, \begin{pmatrix} 1 & 0 \\ 0 & 2 \end{pmatrix}, \begin{pmatrix} 1 & \frac12 \\ \frac12 & 2 \end{pmatrix}, \begin{pmatrix} 1 & \frac12 \\ \frac12 & 3 \end{pmatrix},\\[0.3cm] & \begin{pmatrix} 1 & 0 \\ 0 & 5 \end{pmatrix}, \begin{pmatrix} 1 & \frac12 \\ \frac12 & 7 \end{pmatrix}, \begin{pmatrix} 1 & 0 \\ 0 & 8 \end{pmatrix}, \begin{pmatrix} 1 & \frac12 \\ \frac12 & 9 \end{pmatrix}, \quad\text{and}\quad \begin{pmatrix} 1 & 0 \\ 0 & 9 \end{pmatrix}, \end{array} \end{equation} and furthermore, $m$ is represented by both \begin{equation}\label{2genera} \text{gen}\left( \begin{pmatrix} 1 & \frac12 \\ \frac12 & 6 \end{pmatrix}\right) \quad \text{and} \quad \text{gen}\left(\begin{pmatrix} 1 & \frac12 \\ \frac12 & 8 \end{pmatrix}\right). \end{equation} As a sample, assume that $4m$ is represented by $\begin{pmatrix} 4 & 1 \\ 1 & 6 \end{pmatrix}$. Then $2m$ is represented by $\begin{pmatrix} 2 & \frac12 \\ \frac12 & 3 \end{pmatrix}$. Assume that $2x^2+xy+3y^2=2m$ for some integers $x$ and $y$. Then either $x+y$ or $y$ is even. If $x+y=2z$ for some integer $z$, then $2x^2-5xz+6z^2=m$. Hence $m$ is represented by $\begin{pmatrix} 2 & \frac12 \\ \frac12 & 3 \end{pmatrix}$. If $y=2z$, then $x^2+xz+6z^2=m$. Hence $m$ is represented by $\begin{pmatrix} 1 & \frac12 \\ \frac12 & 6 \end{pmatrix}$. Note that $$ \text{gen}\left(\begin{pmatrix} 1 & \frac12 \\ \frac12 & 6 \end{pmatrix}\right)\big/\sim=\left\{ \begin{pmatrix} 1 & \frac12 \\ \frac12 & 6 \end{pmatrix}, \begin{pmatrix} 2 & \frac12 \\ \frac12 & 3 \end{pmatrix} \right\}. $$ Conversely, assume that $m$ is represented by the genus of $\begin{pmatrix} 1 & \frac12 \\ \frac12 & 6 \end{pmatrix}$. If $m$ is represented by $\begin{pmatrix} 1 & \frac12 \\ \frac12 & 6 \end{pmatrix}$, then we have $$ \langle 4m \rangle {\ \rightarrow\ } \begin{pmatrix} 4 &2 \\2 & 24 \end{pmatrix} {\ \rightarrow\ } \begin{pmatrix} 4 &1 \\1 & 6 \end{pmatrix}. $$ If $m$ is represented by $\begin{pmatrix} 2 & \frac12 \\ \frac12 & 3 \end{pmatrix}$, then we have $$ \langle 4m \rangle {\ \rightarrow\ } \begin{pmatrix} 8 & 2 \\ 2 & 12 \end{pmatrix} \simeq \begin{pmatrix} 8 & 6 \\ 6 & 16 \end{pmatrix} {\ \rightarrow\ } \begin{pmatrix} 8 & 3 \\ 3 & 4 \end{pmatrix} \simeq \begin{pmatrix} 4 &1 \\1 & 6 \end{pmatrix}. $$ Hence $4m$ is represented by \!$\begin{pmatrix} 4 & 1 \\ 1 & 6 \end{pmatrix}$ if and only if $m$ is represented by \!$\text{gen}\left(\!\begin{pmatrix} 1 & \frac12 \\ \frac12 & 6 \end{pmatrix}\!\right)$. Note that 9 binary ${\mathbb Z}$-lattices in \eqref{9lattices} have class number $1$. Therefore if $m\equiv 1 \pmod 8$ is prime, and for any prime $q\in \{3,5,7,11,23,31\}$, $m$ is a quadratic residue modulo $q$, then one may easily check that $m$ is represented by $9$ binary ${\mathbb Z}$-lattices in \eqref{9lattices} and by both genera in \eqref{2genera}. This implies that $m$ is a recoverable number by Proposition \ref{12lattices}. This completes the proof. \end{proof} \begin{rmk}{\rm We checked that any non square integer less than or equal to $35$ is not a recoverable number. } \end{rmk}
{ "timestamp": "2020-09-10T02:06:33", "yymm": "2009", "arxiv_id": "2009.04050", "language": "en", "url": "https://arxiv.org/abs/2009.04050", "abstract": "For a set $S$ of (positive definite and integral) quadratic forms with bounded rank, a quadratic form $f$ is called $S$-universal if it represents all quadratic forms in $S$. A subset $S_0$ of $S$ is called an $S$-universality criterion set if any $S_0$-universal quadratic form is $S$-universal. We say $S_0$ is minimal if there does not exist a proper subset of $S_0$ that is an $S$-universality criterion set. In this article, we study various properties of minimal universality criterion sets. In particular, we show that for `most' binary quadratic forms $f$, minimal $S$-universality criterion sets are unique in the case when $S$ is the set of all subforms of the binary form $f$.", "subjects": "Number Theory (math.NT)", "title": "Minimal universality criterion sets on the representations of quadratic forms", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9920620070720565, "lm_q2_score": 0.8080672204860316, "lm_q1q2_score": 0.8016527886045105 }
https://arxiv.org/abs/1806.09222
Analysis of Krylov Subspace Solutions of Regularized Nonconvex Quadratic Problems
We provide convergence rates for Krylov subspace solutions to the trust-region and cubic-regularized (nonconvex) quadratic problems. Such solutions may be efficiently computed by the Lanczos method and have long been used in practice. We prove error bounds of the form $1/t^2$ and $e^{-4t/\sqrt{\kappa}}$, where $\kappa$ is a condition number for the problem, and $t$ is the Krylov subspace order (number of Lanczos iterations). We also provide lower bounds showing that our analysis is sharp.
\section{The cubic-regularized problem}\label{sec:upper-cr} We now consider the cubic-regularized problem \begin{equation*}% \minimize_{x\in\R^{d}}~\fcu\left(x\right)\defeq \f(x) +\frac{\rho}{3} \norm{x}^3 = \half x^{\T}Ax+b^{\T}x+\frac{\rho}{3} \norm{x}^3. \end{equation*} Any global minimizer of $\fcu$, denoted $\scu$, admits the characterization~\cite[Theorem 3.1]{CartisGoTo11} \begin{equation} \grad \fcu(\scu) = (A+\rs I)\, \scu+b =0 ~~ \mbox{and} ~~ \rho \norm{\scu} \ge -\lambda_{\min}. \label{eq:cr-optimality} \end{equation} Comparing this characterization to its counterpart~\eqref{eq:tr-optimality} for the trust-region problem, we see that any instance $(A,b,\rho)$ of cubic regularization has an \emph{equivalent trust-region instance} $(A,b,R)$, with $R=\norm{\scu}$. Theses instances are equivalent in that they have the same set of global minimizers. Evidently, the equivalent trust-region instance has optimal Lagrange multiplier $\ltr = \rs$. Moreover, at any trust-region feasible point $x$ (satisfying $\norm{x}\le R=\norm{\scu}=\norm{\str}$), the cubic-regularization optimality gap is smaller than its trust-region equivalent, \begin{equation*} \fcu(x)-\fcu(\scu) = \f(x) - \f(\str) + \frac{\rho}{3}\big( \norm{x}^3 - \norms{\str}^3 \big) \le \f(x) - \f(\str). \end{equation*} Letting $\itercu_t$ denote the minimizer of $\fcu$ in $\Krylov$ and letting $\itertr_t$ denote the Krylov subspace solution of the equivalent trust-region problem, we conclude that \begin{equation}\label{eq:cr-tr-opt-gap-bound} \fcu(\itercu_t) - \fcu(\scu) \le \fcu(\itertr_t) - \fcu(\scu) \le \f(\itertr_t) - \f(\str); \end{equation} cubic regularization Krylov subspace solutions always have a \emph{smaller optimality gap} than their trust-region equivalents. The guarantees of Theorem~\ref{thm:tr} therefore apply to $\fcu(\itercu_t) - \fcu(\scu)$ as well, and we arrive at the following \begin{corollary}\label{cor:cu} For every $t>0$, \begin{equation}\label{eq:cr-lin-time} % % \fcu(\itercu_t) - \fcu(\scu) \le 36\left[\fcu(0) - \fcu(\scu) \right] \exp\left\{ -4t\sqrt{\frac{\lambda_{\min} + \rs}{\lambda_{\max} +\rs}} \right\}, \end{equation} and \begin{equation}\label{eq:cr-sublin-time} \fcu(\itercu_t) - \fcu(\scu) \le \frac{(\lambda_{\max} - \lambda_{\min})\norm{\scu}^2}{(t-\half)^2} \left[ 4 + \frac{\I_{\{\lambda_{\min} < 0\}}}{8} \log^2\left(\frac{4\norm{b}^2}{(u_{\min}^T b)^2}\right) \right]. \end{equation} \end{corollary} \begin{proof} Use the slightly stronger bound~\eqref{eq:tr-lin-time-bound-stronger} derived in the proof of Theorem~\ref{thm:tr} with the inequality $18 \str^T \Atr \str + 4\ltr\norm{\str}^2 \le 36[ \half\scu^T A \scu + \frac{1}{6}\rho \norm{\scu}^3] = 36[\fcu(0)-\fcu(\scu)]$. \end{proof} Here too it is possible to randomly perturb $b$ and obtain a guarantee for cubic regularization that applies in the hard case. In~\cite{CarmonDu16} we carry out such analysis for gradient descent, and show that perturbations to $b$ with norm $\sigma$ can increase $\norms{\scu}^2$ by at most $2\sigma/\rho$ \cite[Lemma 4.6]{CarmonDu16}. Thus the cubic-regularization equivalent of Corollary~\ref{cor:tr-rand} amounts to replacing $R^2$ with $\norm{\scu}^2+2\sigma/\rho$ in~\eqref{eq:tr-pert-sublin-time}. We note briefly---without giving a full analysis---that Corollary~\ref{cor:cu} shows that the practically successful Adaptive Regularization using Cubics (ARC) method~\cite{CartisGoTo11} can find $\epsilon$-stationary points in roughly $\epsilon^{-7/4}$ Hessian-vector product operations (with proper randomization and subproblem stopping criteria). Researchers have given such guarantees for a number of algorithms that are mainly theoretical~\cite{AgarwalAlBuHaMa17,CarmonDuHiSi18}, as well as variants of accelerated gradient descent~\cite{CarmonDuHiSi17,JinNeJo17}, which while more practical still require careful parameter tuning. In contrast, ARC requires very little tuning and it is encouraging that it may also exhibit the enhanced Hessian-vector product complexity $\epsilon^{-7/4}$, which is at least near-optimal~\cite{CarmonDuHiSi17lii}. \section{Numerical experiment details}\label{sec:exp-details} \paragraph{Random problem generation, $\kappa < \infty$} We generate random cubic regularization instances $(A, b, \rho)$ as follows. We take $\lambda_{\max} = 1$ and draw $\lambda_{\min}\sim U[-1,-0.1]$, where $U[a,b]$ denotes the uniform distribution on $[a,b]$. We then fix two eigenvalues of $A$ to be $\lambda_{\min},\lambda_{\max}$ and draw the other $d-2$ eigenvalues independently from $U[\lambda_{\min}, \lambda_{\max}]$. We then take $A$ to be diagonal with said eigenvalues. This is without much loss of generality (as the Krylov subspace method is rotationally invariant), and it allows us to quickly compute matrix-vector products, whose computation nevertheless accounts for much of the experiment running time when using $d=10^{6}$. For a desired condition number $\kappa$, we let \begin{equation*} \ltr \defeq \frac{\lambda_{\max} - \kappa\lambda_{\min}}{\kappa - 1} \end{equation*} and as usual denote $\Atr = A + \ltr I$. To generate $b$, $\rho$, we draw a standard normal $d$-dimensional vector $v\sim \mc{N}(0; I)$ and let \begin{equation*} b = \sqrt{\frac{2}{v^T \Atr^{-1} v + \frac{\ltr}{3} v^T\Atr^{-2} v} }\cdot v ~,~ \rho = \frac{\ltr}{\norms{\Atr^{-1}b}}, \end{equation*} The above choice of $b$ and $\rho$ guarantees that $\rho \norm{\Atr^{-1}b} = \ltr$ and therefore $\scu = -\Atr^{-1}b$ is the unique solution and the problem condition number satisfies \begin{equation*} \frac{\lambda_{\max}+\rs}{\lambda_{\min}+\rs} = \frac{\lambda_{\max} + \ltr}{\lambda_{\min} + \ltr} = \kappa \end{equation*} as desired. Moreover, our scaling of $b$ guarantees that \begin{equation*} \fcu(0)-\fcu(\scu) = \half (\scu)^T \Atr \scu + \frac{\rho}{6}\norm{\scu}^3 = \half \left(b^T \Atr^{-1} b + \frac{\ltr}{3}b^T \Atr^{-2} b\right) =1. \end{equation*} Our technique for generating $(A,b,\rho)$ is similar to the one we used in~\cite{CarmonDu16} to test gradient descent for cubic regularization. The main difference is that in~\cite{CarmonDu16} the value of $\rho$ is fixed and consequently there is no control over the initial optimality gap. For every value of $\kappa$, we generate 5,000 problem instances independently as described above. \paragraph{Random problem generation, $\kappa = \infty$} We let $A=\diag(\lambda)$ where $\lambda_1 = \lambda_{\min} = -0.5$, $\lambda_d = \lambda_{\max} = 0.5$ and $\lambda_2, \ldots, \lambda_{d-1}$ are drawn i.i.d. from $U[\lambda_{\min} + \gamma, \lambda_{\max}]$ where we take the eigen-gap $\gamma=10^{-4}$ and $d=10^6$. As $\kappa=\infty$, we let \begin{equation*} \ltr = -\lambda_{\min} \end{equation*} and denote $\hat{A}_{\ltr} \defeq \diag(\lambda_2+\ltr, \ldots, \lambda_{\max}+\ltr)$. We generate $b$ and $\rho$ by drawing a standard normal $(d-1)$-dimensional vector $v$, and letting \begin{equation*} b_1 = 0 ~,~ b_{2:d} = \sqrt{\frac{2}{v^T \hat{A}_{\ltr}^{-1} v + (1+\tau^2)\frac{\ltr}{3} v^T\hat{A}_{\ltr}^{-2} v} } v ~,~ \rho = \frac{\ltr}{\norms{\hat{A}_{\ltr}^{-1}b_{2:d}}\sqrt{1+\tau^2}}, \end{equation*} where $\tau$ is a parameter that determines the weight of the eigenvector corresponding to $\lambda_{\min}$ in the solution (when $\tau = \infty$ we have a pure eigenvector instance); we take $\tau=10$. A global minimizer $\scu$ of the problem instance $(A,b,\rho)$ generated above has the form, \begin{equation*} [\scu]_1 = \pm \tau \norms{\hat{A}_{\ltr}^{-1}b_{2:d}} ~,~ [\scu]_{2:d} = -\hat{A}_{\ltr}^{-1}b_{2:d}. \end{equation*} As in the case $\kappa < \infty$, it is easy to verify that the scaling of $b$ guarantees $\fcu(0)-\fcu(\scu) = 1$. When $\kappa=\infty$, the choice of eigen-gap $\gamma$ strongly affects optimization performance. We explore this in Figure~\ref{fig:exp-gap}, which repeats the experiment described above with different values of $\gamma$ (and $d=10^5$). As seen in the figure, the non-randomized Krylov subspace solution becomes more suboptimal as $\gamma$ increases. Moreover, randomization ``kicks-in'' after roughly $\log d / \sqrt{\gamma}$ iterations, when eigen-gap-dependent linear convergence begins. To create each plot, we draw 10 independent problem instances from the distribution described above, and for each problem instance run each randomization approach with 50 different random seeds; we observe that sampling problem instances and sampling randomization seeds contribute similar amount of variation to the final ensemble of results. \begin{figure} \centering \includegraphics[width=\nips{\columnwidth}\arxiv{0.975\columnwidth}] {code/figures/pert_experiment_many.png} \vspace{-0.5cm} \caption{\label{fig:exp-gap}Optimality gap of Krylov subspace solutions on random cubic-regularization problems, versus subspace dimension $t$. Each plot shows result for problem instances with a different eigen-gap $\gamma = (\lambda_{\max} - \lambda_{\min})/(\lambda_2 - \lambda_{\min})$, where $\lambda_2$ is the smallest eigenvalue larger than $\lambda_{\min}$. Each line represents median suboptimality, and shaded regions represent inter-quartile range. Different lines correspond to different randomization settings. } \end{figure} \paragraph{Hardness of generated problems} It is well known that the performance of subspace methods improves dramatically when the eigenvalues of $A$ are clustered~\cite{TrefethenBa97}. Taking the eigenvalues of $A$ to be uniformly distributed produces very little clustering, making the instances we draw somewhat hard. However, examining the proof of the lower bound~\eqref{eq:cr-lb-lin} we see that the worst case eigenvalues are of the form $\lambda_k = \lambda_{\min} + (\lambda_{\max}-\lambda_{\min})\sin^2 \theta_k$ where $\theta_1, \ldots \theta_d$ are equally spaced in $[0, \pi/2]$. This is fairly different from a uniform distribution (asymptotically as $d\to\infty$ it becomes an arcsine distribution), and consequently we think that uniformly distributing the eigenvalues makes for a challenging but not quite adversarial test case. \paragraph{Computing Krylov subspace solutions} We use the Lanczos process to obtain a tridiagonal representation of $A$ as described in Section~\ref{sec:lanczos}. To obtain full optimization traces we solve equation~\eqref{eq:cr-lambda-search} after every Lanczos iteration, warm-starting $\lambda$ with the solution from the previous step and the minimum eigenvalue of the current tridiagonal matrix. We use the Newton method described by~\citet[Algorithm 6.1]{CartisGoTo11} to solve the equation~\eqref{eq:cr-lambda-search} in the Krylov subspace. For the $\kappa < \infty$ experiment, we stop the process when $|\norm{A_{\lambda}^{-1}b}-\lambda/\rho| < 10^{-12}$ or after 25 tridiagonal system solves are computed. For the $\kappa = \infty$ experiment we allow up to 100 system solves. \section{Numerical experiments}\label{sec:exp} \arxiv{\vspace{-3pt}} To see whether our analysis applies to non-worst case problem instances, we generate 5,000 random cubic-regularization problems with $d=10^{6}$ and controlled condition number $\kappa=(\lambda_{\max}+\rs)/(\lambda_{\min}+\rs)$ (see Section~\ref{sec:exp-details} in the supplement for more details). We repeat the experiment three times with different values of $\kappa$ and summarize the results in Figure~\ref{fig:exp}a. As seen in the figure, about 20 Lanczos iterations suffice to solve even the worst-conditioned instances to about $10\%$ accuracy, and 100 iterations give accuracy better than $1\%$. Moreover, for $t \gtrapprox \sqrt{\kappa}$, the approximation error decays exponentially with precisely the rate $4/\sqrt{\kappa}$ predicted by our analysis, for almost all the generated problems. For $t \ll \sqrt{\kappa}$, the error decays approximately as $t^{-2}$. We conclude that the rates characterized by Theorem~\ref{thm:tr} are relevant beyond the worst case. We conduct an additional experiment to test the effect of randomization for ``hard case'' instances, where $\kappa = \infty$. We generate such problem instances (see details in Section~\ref{sec:exp-details}), and compare the joint subspace randomization scheme (Corollary~\ref{cor:tr-rand-joint}) to the perturbation scheme (Corollary~\ref{cor:tr-rand}) with different perturbation magnitudes $\sigma$; the results are shown in Figure~\ref{fig:exp}b. For any fixed target accuracy, some choices of $\sigma$ yield faster convergence than the joint subspace scheme. However, for any fixed $\sigma$ optimization eventually hits a noise floor due to the perturbation, while the joint subspace scheme continues to improve. Choosing $\sigma$ requires striking a balance: if too large the noise floor is high and might even be worse than no perturbation at all; if too small, escaping the unperturbed error level will take too long, and the method might falsely declare convergence. A practical heuristic for safely choosing $\sigma$ is an interesting topic for future research. \begin{figure} \centering \hspace{-0.75cm} \begin{minipage}[t]{0.72\textwidth} \centering \includegraphics[height=5.025cm]{code/figures/experiment.png} \\ \footnotesize\textbf{(a)} \end{minipage} \begin{minipage}[t]{0.27\textwidth} \centering \includegraphics[height=5.1cm]{code/figures/pert_experiment.png} \\ \nips{~~~~~~~~~~~~~~~~} \arxiv{~~~~~} \footnotesize\textbf{(b)} \end{minipage} \caption{\label{fig:exp}Optimality gap of Krylov subspace solutions on random cubic-regularization problems, versus subspace dimension $t$. \textbf{(a)} Columns show ensembles with different condition numbers $\kappa$, and rows differ by scaling of $t$. Thin lines indicate results for individual instances, and bold lines indicate ensemble median and maximum suboptimality. \textbf{(b)} Each line represents median suboptimality, and shaded regions represent inter-quartile range. Different lines correspond to different randomization settings. } \end{figure} \section{Introduction} Consider the potentially nonconvex quadratic function \begin{equation*} \f(x) \defeq \half x^T A x + b^T x, \end{equation*} where $A \in \R^{d \times d}$ and $b \in \R^d$. We wish to solve regularized minimization problems of the form \begin{equation}\label{eq:problems} \minimize_{x} \f(x) ~ \subjectto \norm{x} \le R ~~~\mbox{and}~~~ \minimize_{x} \f(x) + \frac{\rho}{3} \norm{x}^3, \end{equation} where $R$ and $\rho \ge 0$ are regularization parameters. These problems arise primarily in the family of trust-region and cubic-regularized Newton methods for general nonlinear optimization problems~\cite{ConnGoTo00,NesterovPo06, Griewank81,CartisGoTo11}, which optimize a smooth function $g$ by sequentially minimizing local models of the form \begin{equation*} g(x_i + \Delta) \approx g(x_i) + \grad g(x_i)^T \Delta + \half \Delta^T \hess g(x_i) \Delta = g(x_i) + \f[\hess g(x_i), \grad g(x_i)](\Delta), \end{equation*} where $x_i$ is the current iterate and $\Delta \in \R^d$ is the search direction. Such models tend to be unreliable for large $\norm{\Delta}$, particularly when $\hess g(x_i)\nsucc 0$. Trust-region and cubic regularization methods address this by constraining and regularizing the direction $\Delta$, respectively. Both classes of methods and their associated subproblems are the subject of substantial ongoing research~\cite{HazanKo16,Ho-NguyenKi16,CarmonDu16,AgarwalAlBuHaMa17,LendersKiPo18}. In the machine learning community, there is growing interest in using these methods for minimizing (often nonconvex) training losses, handling the large finite-sum structure of learning problems by means of sub-sampling~\cite{ReigerJoMcAu17,KohlerLu17,BlanchetCaMeSc16, YaoXuRoMa18,TripuraneniStJiReJo17}. The problems~\eqref{eq:problems} are challenging to solve in high-dimensional settings, where direct decomposition (or even storage) of the matrix $A$ is infeasible. In some scenarios, however, computing matrix-vector products $v\mapsto Av$ is feasible. Such is the case when $A$ is the Hessian of a neural network, where $d$ may be in the millions and $A$ is dense, and yet we can compute Hessian-vector products efficiently on batches of training data~\cite{Pearlmutter94,Schraudolph02}. In this paper we consider a scalable approach for approximately solving~\eqref{eq:problems}, which consists of minimizing the objective in the \emph{Krylov subspace} of order $t$, \begin{equation}\label{eq:krylov-def} \mc{K}_t(A,b) \defeq \mathrm{span}\{b, Ab, \ldots, A^{t-1}b\}. \end{equation} This requires only $t$ matrix-vector products, and the Lanczos method allows one to efficiently find the solution to problems~\eqref{eq:problems} over $\mc{K}_t(A,b)$ (see, e.g.~\cite[Sec.~2]{GouldLuRoTo99,CartisGoTo11}). Krylov subspace methods are familiar in numerous large-scale numerical problems, including conjugate gradient methods, eigenvector problems, or solving linear systems~\cite{HestenesSt52, Nemirovski94,TrefethenBa97,GolubVa89}. It is well-known that, with exact arithmetic, the order $d$ subspace $\Krylov[d]$ generically contains the global solutions to~\eqref{eq:problems}. However, until recently the literature contained no guarantees on the rate at which the suboptimality of the solution approaches zero as the subspace dimension $t$ grows. This is in contrast to the two predominant Krylov subspace method use-cases---convex quadratic optimization~\cite{GolubVa89,NemirovskiYu83,Nesterov04} and eigenvector finding~\cite{KuczynskiWo92}---where such rates of convergence have been known for decades. \citet{ZhangShLi17} make substantial progress on this gap, establishing bounds implying a linear rate of convergence for the trust-region variant of problem~\eqref{eq:problems}. In this work we complete the picture, proving that the optimality gap of the order $t$ Krylov subspace solution to either of the problems~\eqref{eq:problems} is bounded by both $e^{-4t/\sqrt{\kappa}}$ and $t^{-2}\log^2(\norms{b}/|u_{\min}^T b|)$. Here $\kappa$ is a condition number for the problem that naturally generalizes the classical condition number of the matrix $A$, and $u_{\min}$ is an eigenvector of $A$ corresponding to its smallest eigenvalue. Using randomization, we may replace $|u_{\min}^T b|$ with a term proportional to $1/\sqrt{d}$, circumventing the well-known ``hard case'' of the problem~\eqref{eq:problems} (see Section~\ref{sec:upper-rand}). Our analysis both leverages and unifies the known results for convex quadratic and eigenvector problems, which constitute special cases of~\eqref{eq:problems}. \paragraph{Related work} \citet{ZhangShLi17} show that the error of certain polynomial approximation problems bounds the suboptimality of Krylov subspace solutions to the trust region-variant of the problems~\eqref{eq:problems}, implying convergence at a rate exponential in $-t/\sqrt{\kappa}$. Based on these bounds, the authors propose novel stopping criteria for subproblem solutions in the trust-region optimization method, showing good empirical results. However, the bounds of~\cite{ZhangShLi17} become weak for large $\kappa$ and vacuous in the hard case where $\kappa=\infty$. Prior works develop algorithms for solving~\eqref{eq:problems} with convergence guarantees that hold in the hard case. \citet{HazanKo16}, \citet{Ho-NguyenKi16}, and \citet{AgarwalAlBuHaMa17} propose algorithms that obtain error roughly $t^{-2}$ after computing $t$ matrix-vector products. The different algorithms these papers propose all essentially reduce the problems~\eqref{eq:problems} to a sequence of eigenvector and convex quadratic problems to which standard algorithms apply. In previous work~\cite{CarmonDu16}, we analyze gradient descent---a direct, local method---for the cubic-regularized problem. There, we show a rate of convergence roughly $t^{-1}$, reflecting the well-known complexity gap between gradient descent (respectively, the power method) and conjugate gradient (respectively, Lanczos) methods~\cite{TrefethenBa97,GolubVa89}. Our development differs from this prior work in the following ways. \begin{enumerate}[leftmargin=*] \item We analyze a practical approach, implemented in efficient optimization libraries~\cite{GouldOrTo03,LendersKiPo18}, with essentially no tuning parameters. Previous algorithms~\cite{HazanKo16,Ho-NguyenKi16, AgarwalAlBuHaMa17} are convenient for theoretical analysis but less conducive to efficient implementation; each has several parameters that require tuning, and we are unaware of numerical experiments with any of the approaches. \item We provide both linear ($e^{-4t/\sqrt{\kappa}})$ and sublinear ($t^{-2}$) convergence guarantees. In contrast, the papers~\cite{HazanKo16,Ho-NguyenKi16,AgarwalAlBuHaMa17} provide only a sublinear rate; \citet{ZhangShLi17} provide only the linear rate. \item Our analysis applies to both the trust-region and cubic regularization variants in~\eqref{eq:problems}, while~\cite{HazanKo16,Ho-NguyenKi16,ZhangShLi17} consider only the trust-region problem, and \cite{ZhangShLi17,CarmonDu16} consider only cubic regularization. \item We provide lower bounds---for adversarially constructed problem instances---showing our convergence guarantees are tight to within numerical constants. By a resisting oracle argument~\cite{NemirovskiYu83}, these bounds apply to any deterministic algorithm that accesses $A$ via matrix-vector products. \item Our arguments are simple and transparent, and we leverage established results on convex optimization and the eigenvector problem to give short proofs of our main results. \end{enumerate} \paragraph{Paper organization} In Section~\ref{sec:upper} we state and prove our convergence rate guarantees for the trust-region problem. Then, in Section~\ref{sec:upper-cr} we quickly transfer those results to the cubic-regularized problem by showing that it always has a smaller optimality gap. Section~\ref{sec:lower} gives our lower bounds, stated for cubic regularization but immediately applicable to the trust-region problem by the same optimality gap bound. Finally, in Section~\ref{sec:exp} we illustrate our analysis with some numerical experiments. \newcommand{\lambda_{\min}}{\lambda_{\min}} \newcommand{\lambda_{\max}}{\lambda_{\max}} \paragraph{Notation} For a symmetric matrix $A\in\R^{d\times d}$ and vector $b$ we let $\f(x) \defeq \half x^T A x + b^T x.$ We let $\lambda_{\min}(A)$ and $\lambda_{\max}(A)$ denote the minimum and maximum eigenvalues of $A$, and let $u_{\min}(A), u_{\max}(A)$ denote their corresponding (unit) eigenvectors, dropping the argument $A$ when clear from context. For integer $t\ge1$ we let $\Polys[t] \defeq \left\{ c_0 + c_1 x + \cdots + c_{t-1}x^{t-1}\mid c_i\in \R \right\}$ be the polynomials of degree at most $t-1$, so that the Krylov subspace~\eqref{eq:krylov-def} is $\Krylov = \left\{ p(A)b \mid p\in\Polys[t] \right\}$. We use $\norm{\cdot}$ to denote Euclidean norm on $\R^d$ and $\ell_2$-operator norm on $\R^{d\times d}$. Finally, we denote $(z)_+\defeq\max\{z,0\}$ and $(z)_- \defeq \min\{z,0\}$. \section*{Acknowledgments} We thank the anonymous reviewers for several helpful questions and suggestions. Both authors were supported by NSF-CAREER Award 1553086 and the Sloan Foundation. YC was partially supported by the Stanford Graduate Fellowship. \setlength{\bibsep}{6pt} \bibliographystyle{abbrvnat} \section{Computing Krylov subspace solutions}\label{sec:lanczos} Generic instances of the trust-region and cubic-regularized problems can be globally optimized by solving the one-dimensional equations \begin{equation}\label{eq:tr-lambda-search} \norm{A_{\lambda}^{-1}b} = R~,~ \lambda > \max\{-\lambda_{\min}, 0\}. \end{equation} and \begin{equation}\label{eq:cr-lambda-search} \norm{A_\lambda^{-1}b} = \lambda/\rho ~~,~~ \lambda \ge -\lambda_{\min}, \end{equation} respectively. However, when $d$ is very large, even a single exact evaluation of $\norms{A_\lambda^{-1}b}$ (which requires a direct linear system solution) can become prohibitively expensive. In this case, a general approach to obtaining approximate solutions is to constrain the domain to a linear subspace $\mc{Q}_t \subset \R^d$ of dimension $t \ll d$. Let $Q_t \in \R^{d\times t}$ be an orthogonal basis for $\mc{Q}_t$ ($Q_t^T Q_t = I$). Finding the global minimizer in $\mc{Q}_t$ is equivalent to re-parameterizing $x$ as $x=Q_t \tilde{x}$ and solving for $\tilde{x}\in\R^t$, which is also equivalent to solving a $t$-dimensional problem instance with $\tilde{A} = Q_t^T A Q_t$ and $\tilde{b} = Q_t^T b$. For sufficiently large $d$, the time to solve such problems will be dominated by the $t$ matrix-vector products required to construct $\tilde{A}$. In this paper we focus on the choice $\mc{Q}_t = \Krylov$ the Krylov subspace of order $t$. This choice offers a significant efficiency boost: we can efficiently construct a basis $Q_t$ for which $Q_t^T A Q_t$ is tridiagonal, using the Lanczos process, which consists of the following recursion, starting with $q_1 = b/\norm{b}, q_0 = 0$, \begin{equation*} \alpha_t = q_t^T A q_t ~,~ q'_{t+1} = A q_t - \alpha_t q_t - \beta_t q_{t-1} ~,~ \beta_{t+1} = \norm{q'_{t+1}} ~,~ q_{t+1} = q'_{t+1}/\norm{q'_{t+1}}. \end{equation*} The vectors $q_1, \ldots, q_t$ give the columns of $Q_t$ while $\alpha_1, \ldots, \alpha_t$ and $\beta_2, \ldots, \beta_t$ respectively give the diagonal and off-diagonal elements of the symmetric tridiagonal matrix $\tilde{A} = Q_t^T A Q_t$; this makes solving equations~\eqref{eq:tr-lambda-search} and~\eqref{eq:cr-lambda-search} easy. One straightforward approach is to directly compute the factorization $\tilde{A}$, which for a symmetric tridiagonal matrix of size $t$ takes $O(t\log t)$ time~\cite{CoakleyRo13}. A more efficient approach---and the one used in practice---is to iteratively solve systems of the form $\tilde{A}_{\lambda}x = z$ and update $\lambda$ using Newton steps~\cite{ConnGoTo00,CartisGoTo11}. Every tridiagonal system solution can be done in time $O(t)$, and the Newton steps are shown in~\cite{ConnGoTo00,CartisGoTo11} to be globally linearly convergent, with local quadratic convergence. In our experience less than 20 Newton steps generally suffice to reach machine precision, and so the computational cost is essentially linear in $t$. It is also possible to avoid keeping $Q_t$ in memory (when $t\cdot d$ storage is too demanding) by running the Lanczos process twice, once for evaluating $\tilde{x}$ and again to obtain $x=Q_t \tilde{x}$. The Lanczos process produces the same result as Gram-Schmidt orthonormalization of the vectors $\left[b, Ab, \ldots, A^{t-1}b\right]$ but uses the special structure of that matrix to avoid computing inner products that are known in advance to be zero. When run for many iterations, the Lanczos process has well-documented numerical stability issues~\cite{TrefethenBa97}. However, in our setting we usually seek low to moderate accuracy solutions and will usually stop at $t < 100$, for which Lanczos is reasonably stable with floating point arithmetic even when $d$ is quite large. The application of the Lanczos process---which is typically used for eigenvector computation---in the context of regularized quadratic optimization is sometimes referred to as the generalized Lanczos process~\cite{GouldLuRoTo99}. \subsection{Computing joint Krylov subspace solutions}\label{sub:block-lanczos} To solve equations~\eqref{eq:tr-lambda-search} and~\eqref{eq:cr-lambda-search} in subspaces of the form \begin{equation*} \Krylov[mt][A, \{v_1, \ldots, v_m\}] \defeq \mathrm{span}\{A^j v_i\}_{i\in\{1,\ldots, m\},j\in\{0,\ldots,t-1\}} \end{equation*} we may use the block Lanczos method~\cite{CullumDo74,Golub77}, a natural generalization of the Lanczos method that creates an orthonormal basis for the subspace $\Krylov[mt][A, \{v_1, \ldots, v_m\}]$ in which $A$ has a block tridiagonal form. Overloading the notation defined above so that now $q_t \in \R^{d\times m}$ and $\alpha_t, \beta_t \in \R^{m\times m}$, the block Lanczos recursion is given by, \begin{equation*} \alpha_t = q_t^T A q_t ~,~ q'_{t+1} = A q_t - q_t \alpha_t - q_{t-1} \beta_t^T ~,~ (q_{t+1}, \beta_{t+1}) = \mathrm{QR}(q'_{t+1}). \end{equation*} where $\mathrm{QR}$ stands for the QR decomposition (i.e.\ if $(q, \beta) = \mathrm{QR}(a)$ then $q$ is orthogonal, $\beta$ is upper diagonal and $a = q\cdot \beta$), and the initial conditions are that $q_1$ is an orthonormalized version of $[v_1, \ldots, v_m]$ and $q_0=0$. The matrix $\tilde{A} = Q_t^T A Q_t$ is now block tridiagonal, with the diagonal and sub-diagonal blocks given by $\{\alpha_i\}_{i\in\{1,\ldots,t\}}$ and $\{\beta_i\}_{i\in\{2,\ldots,t\}}$ respectively. Since the $\beta$ matrices are upper diagonal, $\tilde{A}$ is a symmetric banded matrix with $m$ non-zeros sub-diagonal bands. Such matrix admits fast Cholesky decomposition (in time linear in $m^2 t$), and consequently the Newton method described above is still efficient. \section{Proof of lower bounds}\label{sec:lb-proof} In what follows, we break Theorem~\ref{thm:lb} into two parts, one for the linear convergence lower bound~\eqref{eq:cr-lb-lin} and one for the sublinear lower bounds~\eqref{eq:cr-lb-sub-ncvx} and~\eqref{eq:cr-lb-sub-cvx}. We restate each sub-theorem in a way that clearly shows our control over problem-dependent parameters when constructing the hard problem instances. In our proofs we will make use of the following expression for the optimality gap in the cubic-regularization problem, \begin{equation}\label{eq:fx-s-expression} \fcu\left(x\right)-\fcu\left(\scu\right)= \frac{1}{2}(x-\scu)^T\Acu (x-\scu) +\frac{\rho}{6}\left(\| \scu\| -\| x\| \right)^{2}\left(\| \scu\| +2\| x\| \right), \end{equation} where $\Acu = A + \rs I$. \subsection{Proof of linear convergence lower bound} \begin{customthm}{\ref{thm:lb}, part I} Let $\lambda_{\min}, \lambda_{\max}, \ltr, \Delta \in \R$ such that $\lambda_{\min} \le \lambda_{\max}$, $\ltr > \max\{-\min,0\}$ and $R, \Delta>0$. For every $t \ge 1$ and every $d>t$ there exists $A\in\R^{d\times d}$, $b\in\R^d$ and $\rho > 0$ such that \begin{itemize} \item all eigenvalues of $A$ are in $[\lambda_{\min}, \lambda_{\max}]$, \item the solution $\scu = \argmin_{x\in\R^d}\fcu(x)$ satisfies % $\rs = \ltr$, \item $\fcu(0) - \fcu(\scu) = \Delta$, and \end{itemize} \begin{equation*} \fcu(s) - \fcu(\scu) > % \left(1+\frac{\rs}{3(\rs + \lambda_{\min})} \right)^{-1} \left[\fcu(0) - \fcu(\scu) \right] \exp\left\{ -\frac{4t}{\sqrt{\frac{\rs + \lambda_{\max}}{\rs +\lambda_{\min}}}-1} \right\}. \end{equation*} for every $s\in\Krylov$. \end{customthm} \begin{proof} From Lemma~\ref{lem:cheby-first} with $\alpha = \ltr + \lambda_{\min}$, $\beta = \ltr + \lambda_{\max}$ and $n=t$, we have that there exist $\xi_0, \ldots, \xi_t \in [\alpha, \beta]$ and probability distribution $\pi_0, \ldots, \pi_{t}$ such that \begin{equation*} \min_{p \in \Polys[t]} \sum_{k=0}^{t} \pi_k (1-\xi_k p(\xi_k))^2 \ge e^{-4t/(\sqrt{\kappa}-1)}, \end{equation*} where $\kappa = \beta/\alpha = (\lambda_{\max}+\ltr)/(\lambda_{\min}+\ltr)$. We let $\xi$ and $\sqrt{\pi}$ denote vectors with entries $\xi_0, \ldots, \xi_{t}$ and $\sqrt{\pi_0}, \ldots, \sqrt{\pi_t}$ respectively. To construct the problem instance $(A,b,\rho)$ we assume without loss of generality $d=t+1$, as higher dimensional instances can be obtained by zero-padding a $(t+1)$-dimensional construction. We set \begin{equation*} A = \diag( \xi - \ltr ) ,~ b = \mu \Atr^{1/2}\sqrt{\pi} ~\mbox{and}~ \rho =\ltr/ \norm{\Atr^{-1}b}, \end{equation*} where we will choose $\mu > 0$ to set the value of $\fcu(0)-\fcu(\scu)$. First, we note that for any value of $\mu$ our choice of $\rho$ guarantees that $\norm{\Atr^{-1}b} = \ltr/\rho$, making $\scu = -\Atr^{-1}b$ the unique global minimizer of $\fcu$. We therefore have by equation~\eqref{eq:fx-s-expression} \begin{equation*} \fcu(0)-\fcu(\scu) = \half \scu^T \Atr \scu + \frac{\rs}{6}\norm{\scu}^2 = \frac{\mu^2}{2}\left( 1 + \frac{\ltr}{3} \sqrt{\pi}^T \Atr^{-1} \sqrt{\pi} \right), \end{equation*} so for every $\Delta > 0$ there is $\mu$ for which $\fcu(0)-\fcu(\scu) = \Delta$. Noting that $\sqrt{\pi}^T \Atr^{-1} \sqrt{\pi} \le (\ltr + \lambda_{\min})^{-1}\norm{\sqrt{\pi}}^2 = (\ltr + \lambda_{\min})^{-1}$, we also have \begin{equation*} \frac{\mu^2}{2} \ge \Delta \left(1+\frac{\ltr}{3(\ltr + \lambda_{\min})} \right)^{-1} . \end{equation*} Now, every $s\in\Krylov$ is of the form $s=-p(\Atr)b$ for $p\in\Polys[t]$, and using equation~\eqref{eq:fx-s-expression} again we have \begin{align*} \fcu(s)-\fcu(\scu) &\ge \half \norm{\Atr^{1/2}(s-\scu)}^2 \stackrel{(a)}{=} \half\norm{(I-\Atr p(\Atr))\Atr^{-1/2}b}^2 \\ & \stackrel{(b)}{=} \frac{\mu^2}{2}\sum_{k=0}^{n} \pi_k (1-\xi_k p(\xi_k))^2 \stackrel{(c)}{\ge} \frac{\mu^2}{2} e^{-4t/(\sqrt{\kappa}-1)}, \end{align*} where in $(a)$ we substituted $s=-p(\Atr)b$ and $\scu=-\Atr^{-1}b$, in $(b)$ we used our construction of $A$ and $b$, and in $(c)$ we used the guarantee from Lemma~\ref{lem:cheby-first}. The result follows from substituting our lower bound on $\mu^2$ and recalling that $\ltr = \rs$. \end{proof} \subsection{A lower bound for finding eigenvectors} The ``non-convex'' lower bound is in its heart a statement about the difficulty of approximating an extremal eigenvector in a Krylov subspace, which we state explicitly here. The proof of the lemma consists of applying ``in reverse'' the same polynomial approximation result (Lemma~\ref{lem:cheby-second}) that~\citet{KuczynskiWo92} use for proving upper bounds on finding eigenvector with the Lanczos method (which we state as Lemma~\ref{lem:eigenvec-tight}). \begin{restatable}[Finding eigenvectors: lower bound]{lemma}{lemEigenLB}\label{lem:eigenvec-lb} For every $d>0$, vector $v\in \R^d$, unit vector $u\in\R^d$ and $t<d$, there exists matrix $M\in\R^{d\times d}$ such that $M \succeq 0$, $M u = 0$, and for every $z \in \Krylov[t][M,v]$, \begin{equation*} \frac{z_t^T M z_t}{\opnorm{M}\norm{z_t}^2} \ge \min\left\{\frac{1}{4}, \frac{1}{64(t-\half)^2} \log^2\left(-3+4\frac{\norm{v}^2}{(u^T v)^2}\right)\right\}. \end{equation*} \end{restatable} \begin{proof} We take $\opnorm{M}=1$ without loss of generality; results for arbitrary norms of $M$ follow by scaling the construction below. Define \begin{equation}\label{eq:eigenvec-lb-err-def} \err \defeq \min\left\{\frac{1}{4}, \frac{1}{64(t-\half)^2} \log^2\left(-3+4\frac{\norm{v}^2}{(u^T v)^2}\right)\right\}. \end{equation} % % % % % % % % % % % % % % % % % % We apply Lemma~\ref{lem:cheby-second} with $n=t-1$, $\alpha = \err$ and $\beta=1$, to obtain $\xi_1, \ldots, \xi_t \in [\err, 1]$ and probability distribution $\pi_1, \ldots, \pi_t$ such that \begin{equation}\label{eq:cheby-second-lb-applied} \min_{p \in \Polys[t-1]} \sum_{k=1}^{t} \pi_k (\xi_k - \err)(1-\xi_k p(\xi_k))^2 \ge \frac{4\err}{e^{2(2t-1)/(\frac{1}{\sqrt{\err}}-1)}-1}. \end{equation} We assume without loss of generality that $d=t+1$ (otherwise we zero-pad), and construct $M$ as follows. First, we take the eigenvalues of $M$ to be $0, \xi_1, \ldots, \xi_t$, satisfying $0\preceq M \preceq I$. Next, we let $u$ be the eigenvector of $M$ corresponding to eigenvalue $0$, satisfying $Mu=0$. Finally, for $i=1,\ldots, t$ we choose the eigenvector $u_i$ corresponding to eigenvalue $\xi_i$ such that $(u_i^T v)^2 = \pi_i (\norm{v}^2 - (u^T v)^2)$. Assume by contradiction \begin{equation}\label{eq:eigenvec-lb-contra-assumption} \min_{z\in\Krylov[t][M,v]} \frac{ z^T M z }{\norm{z}^2 } < \err, \end{equation} and let $q\in \Polys[t]$ be be such that \begin{equation*} \frac{ \sum_{i=1}^t \xi_i q^2(\xi_i) (u_i^T v)^2} { q^2(0)(u^T v)^2 + \sum_{i=1}^t q^2(\xi_i) (u_i^T v)^2} = \frac{ (q(M)v)^T M q(M)v }{\norm{q(M)v}^2 } % % = \min_{z\in\Krylov[t][M,v]} \frac{ z^T M z }{\norm{z}^2 } < \err. \end{equation*} Rearranging, using $(u_i^T v)^2 = \pi_i (\norm{v}^2 - (u^T v)^2)$, and letting $\tilde{q}(x) = q(x)/q(0)$, we have that \begin{equation*} \err > \left( \frac{\norm{v}^2}{(u^T v)^2} - 1\right) \sum_{i=1}^{t} \pi_i (\xi_i - \err) \tilde{q}^2(\xi_i) \ge \left( \frac{\norm{v}^2}{(u^T v)^2} - 1\right) \frac{4 \err}{e^{2(2t-1)/(\frac{1}{\sqrt{\err}}-1)}-1}. \end{equation*} where in the last transition we used that $\tilde{q}(0)=1$ and therefore it is of the form $1-xp(x)$ for some $p\in\Polys[t-1]$, so the lower bound~\eqref{eq:cheby-second-lb-applied} applies. Rearranging gives \begin{equation*} \err > h\left(\frac{1}{16(t-\half)^2} \log^2\left(-3+4\frac{\norm{v}^2}{(u^T v)^2}\right)\right) ,~~ h(x) = \frac{x}{(1+\sqrt{x})^2}. \end{equation*} Using $h(x) \ge \frac{1}{4}\min\{1, x\}$ and the definition~\eqref{eq:eigenvec-lb-err-def} of $\err$, we see that the above bound gives the contradiction $\err > \err$ and therefore assumption~\eqref{eq:eigenvec-lb-contra-assumption} must be false and we have the desired result $\min_{z\in\Krylov[t][M,v]} \frac{ z^T M z }{\norm{z}^2 } \ge \err$. \end{proof} \subsection{Proof of sublinear convergence lower bound} \begin{customthm}{\ref{thm:lb}, part II} Let $\lambda_{\min}, \lambda_{\max}, R, \tau \in \R$ such that $\lambda_{\min}\le\lambda_{\max}$, $\tau \ge 1$ and $R>0$. For every $t \ge 1$ and every $d>t$ there exists $A\in\R^{d\times d}$, $b\in\R^d$ and $\rho > 0$ such that \begin{itemize} \item all eigenvalues of $A$ are in $[\lambda_{\min}, \lambda_{\max}]$, \item the solution $\scu = \argmin_{x\in\R^d}\fcu(x)$ satisfies $\norm{\scu} = R$, % \item there exists unit eigenvector $u_{\min}$ such that $u_{\min}^T A u_{\min} = \lambda_{\min}$ and $\frac{\norm{b}}{|u_{\min}^T b|} = \tau$, and \end{itemize} \begin{equation*} \fcu(s) - \fcu(\scu) > \min\left\{ \lambda_{\max}^{-}-\lambda_{\min}~,~ \frac{\lambda_{\max}-\lambda_{\min}}{16(t-\half)^2} \log^2\left(\frac{\norm{b}^2}{( u_{\min}^T b)^2}\right) \right\}\frac{\norm{\scu}^2}{32}, \end{equation*} where $\lambda_{\max}^{-}=\min\{\lambda_{\max},0\}$, and \begin{equation*} \fcu(s) - \fcu(\scu) > \frac{(\lambda_{\max} - \lambda_{\min})\norm{\scu}^2}{16(t+\half)^2}. \end{equation*} for every $s\in\Krylov$. \end{customthm} \renewcommand{\err}[1][t]{\epsilon_{#1}} \begin{proof} We begin with the first, ``non-convex'' bound, which is essentially a reduction to the eigenvector problem. Here we assume $\lambda_{\min} \le 0$ as otherwise the lower bound is vacuous. We use Lemma~\ref{lem:eigenvec-lb} to construct $M\in\R^{d\times d}$ and unit vectors $u_{\min}, v \in R^d$ such that $M\succeq 0$, $\opnorm{M}=\lambda_{\max} - \lambda_{\min}$, $M u_{\min} = 0$, $\norm{v}/|u_{\min}^T v| = \tau$ and for every $z\in \Krylov[t][M,v]$ \begin{align}\label{eq:cr-sub-lb-err-def} \frac{z^T M z}{\norm{z}^2} & \ge \frac{\lambda_{\max} - \lambda_{\min}}{4} \min\left\{1 \,,\, \frac{1}{16(t-\half)^2} \log^2\left(-3+4\frac{\norm{v}^2}{( u_{\min}^T v)^2}\right)\right\} \nonumber \\ & \ge \frac{1}{4}\min\left\{\lambda_{\max}^{-} -\lambda_{\min} \,,\, \frac{\lambda_{\max} - \lambda_{\min}}{16(t-\half)^2} \log^2\left(\frac{\norm{v}^2}{( u_{\min}^T v)^2}\right)\right\} \defeq \err, \end{align} where $\lambda_{\max}^{-}=\min\{\lambda_{\max},0\}$. We let $\eps > 0$ be a parameter to be specified later. We let \begin{equation*} \ltr = -\lambda_{\min} + \eps \end{equation*} and construct the cubic regularization instance as follows \begin{equation*} A = M + \lambda_{\min} I ~,~ b = \frac{R}{\norm{\Atr^{-1}v}}v ~,~ \rho = \ltr / R. \end{equation*} The solution for this instance is unique and satisfies $\scu = -\Atr^{-1}b = - R \Atr^{-1}v / \norm{\Atr^{-1}v}$ so that $\norm{\scu} = R$, and moreover we note that $\norm{b} \to 0$ as $\eps\to 0$. For every $s\in\Krylov[t][M,v]=\Krylov[t][A,b]$, \begin{equation*} \fcu(s) = \half s^T A s + b^T s + \frac{\rho}{3}\norm{s}^3 \ge -\norm{b}\norm{s} + \half(\lambda_{\min} + \err)\norm{s}^2 + \frac{\rho}{3}\norm{s}^3. \end{equation*} The RHS above is minimal for \begin{equation*} \norm{s} = \tilde{R} \defeq -\frac{\lambda_{\min} + \err}{2\rho} + \sqrt{\left( \frac{\lambda_{\min} + \err}{2\rho} \right)^2 + \frac{\norm{b}}{\rho}} \le \frac{-\lambda_{\min} - \err}{\rho}+ \sqrt{\frac{\norm{b}}{\rho}}, \end{equation*} where the bound holds since our definition of $\err$ implies $\err \le -\lambda_{\min}$ and so $-\lambda_{\min}-\err \ge 0$. The minimum value of the RHS satisfies \begin{equation}\label{eq:cr-sub-lb-kryl-lb} \fcu(s) \ge -\frac{2}{3}\norm{b}\tilde{R} - \frac{1}{6}(-\lambda_{\min} - \err) \tilde{R}^2. \end{equation} Taking without loss of generality $u_{\min}^T b \le 0$ and using $\rho = \ltr / R$, and $\ltr = -\lambda_{\min} + \eps$, we have \begin{equation}\label{eq:cr-sub-lb-opt-lb} \fcu(\scu) \le \fcu(R \cdot u_{\min}) \le \half \lambda_{\min} R^2 + \frac{1}{3}\ltr R^2 = \frac{1}{6} \lambda_{\min} R^2 + \frac{\eps}{3} R^2. \end{equation} Recall that $\norm{b}\to 0$ as $\eps\to0$, and take $\eps>0$ sufficiently small so that \begin{equation*} \eps < \err/24 ~~ \mbox{and} ~~ \norm{b} \le \min\{\err R/24 , \err^2/\rho \}, \end{equation*} which implies also \begin{equation*} \tilde{R} \le \frac{-\lambda_{\min} - \err}{\rho} + \frac{\err}{\rho} = \frac{-\lambda_{\min}}{\rho} \le \frac{\ltr}{\rho} = R. \end{equation*} Using $\tilde{R}\le R$, we may replace $\tilde R$ with $R$ in the bound~\eqref{eq:cr-sub-lb-kryl-lb}, and combining this with~\eqref{eq:cr-sub-lb-opt-lb} and the bounds on $\norm{b}$ and $\eps$ we obtain \begin{equation*} \fcu(s)-\fcu(\scu) \ge \frac{\err}{6}R^2 - \frac{2\norm{b}}{3}R - \frac{\eps}{3}R^2 \ge \frac{\err}{8} R^2. \end{equation*} Recalling $\norm{\scu} = R$ and the definition~\eqref{eq:cr-sub-lb-err-def} of $\err$, we get the desired ``non-convex'' lower bound. To derive the alternative, ``convex'' lower bound, we again let $0<\eps<\lambda_{\max}-\lambda_{\min}$ be a parameter to be determined, and we apply Lemma~\ref{lem:cheby-second} with $n=t$, $\alpha=\eps$, $\beta=\lambda_{\max}-\lambda_{\min}$ to obtain points $\xi_0, \ldots, \xi_t \in [0, \lambda_{\max}-\lambda_{\min}]$ and probability masses $\pi_0, \ldots, \pi_t$ such that \begin{equation*} \min_{p \in \Polys} \sum_{k=0}^{n} \pi_k (\xi_k - \eps)(1-\xi_k p(\xi_k))^2 = \left[\minmaxU[t][\frac{\lambda_{\max}-\lambda_{\min}}{\eps}] \right]^2. \end{equation*} To construct the hard instance we again set \begin{equation*} \ltr = -\lambda_{\min} + \eps. \end{equation*} Letting $\xi$ and $\sqrt{\pi}$ denote vectors with entries $\xi_i$ and $\sqrt{\pi}_i$, we set \begin{equation*} A = \diag(\xi - \ltr) ~,~ b = R\cdot\Atr\sqrt{\pi} ~,~ \rho = \ltr / R. \end{equation*} Again we have that $\scu = -\Atr^{-1}b$ is the unique solution and $\norm{\scu} = R\norm{\sqrt{\pi}}=R$. Let $s\in\Krylov$, then \begin{equation*} s = -p(\Atr)b = p(\Atr)\Atr \scu = -R p(\Atr)\Atr \sqrt{\pi} \end{equation*} for some $p\in\Polys[t]$. By equality~\eqref{eq:fx-s-expression} we have \begin{align*} \fcu(s) - \fcu(\scu) & \ge \frac{1}{2}\norm{\Atr^{1/2}(s-\scu)}^2 = \frac{R^2}{2} \sum_{k=0}^t \pi_k \xi_k (1-\xi_k p(\xi_k) )^2 \\ & \ge \frac{R^2}{2} \sum_{k=0}^t \pi_k (\xi_k -\eps) (1-\xi_k p(\xi_k) )^2 = \frac{R^2}{2} \left[\minmaxU[t][\frac{\lambda_{\max}-\lambda_{\min}}{\eps}] \right]^2. \end{align*} Note that \begin{equation*} \lim_{\eps\to 0} \minmaxU[t][\frac{\lambda_{\max}-\lambda_{\min}}{\eps}] = \frac{\sqrt{\lambda_{\max}-\lambda_{\min}}}{2t+1}. \end{equation*} Therefore, we can choose $\eps$ sufficiently small so that \begin{equation*} \left[\minmaxU[t][\frac{\lambda_{\max}-\lambda_{\min}}{\eps}] \right]^2 \ge \frac{\lambda_{\max}-\lambda_{\min}}{2(2t+1)^2}, \end{equation*} which gives the proof for the ``convex'' lower bound, as $\norm{\scu}=R$. \end{proof} \section{Lower bounds}\label{sec:lower} \arxiv{\vspace{-6pt}} \newcommand{\mathfrak{L}}{\mathfrak{L}} We now show that the guarantees in Theorem~\ref{thm:tr} and Corollary~\ref{cor:cu} are tight up to numerical constants for adversarially constructed problems. We state the result for the cubic-regularization problem; corresponding lower bounds for the trust-region problem are immediate from the optimality gap relation~\eqref{eq:cr-tr-opt-gap-bound}.\footnote{To obtain the correct prefactor in the trust-region equivalent of lower bound~\eqref{eq:cr-lb-lin} we may use the fact that $\fcu(0) - \fcu(\scu) = \frac{1}{2}b^T \Atr^{-1} b + \frac{\rho}{6}\norm{\scu}^3 \ge \frac{1}{3}(\frac{1}{2} b^T \Atr^{-1} b + \frac{\ltr}{2}R^2)= \frac{1}{3}(\f(0)-\f(\str))$.} To state the result, we require a bit more notation. Let $\mathfrak{L}$ map cubic-regularization problem instances of the form $(A,b,\rho)$ to the quadruple $(\lambda_{\min}, \lambda_{\max}, \ltr, \Delta)=\mathfrak{L}(A,b,\rho)$ such that $\lambda_{\min},\lambda_{\max}$ are the extremal eigenvalues of $A$ and the solution $\scu = \argmin_x \fcu(x)$ satisfies $\rho\norm{\scu} = \ltr$, and $\fcu(0) - \fcu(\scu) = \Delta$. Similarly let $\mathfrak{L}'$ map an instance $(A,b,\rho)$ to the quadruple $(\lambda_{\min}, \lambda_{\max}, \tau, R)$ where now $\norm{\scu}=R$ and $\norm{b}/|u_{\min}^T b|=\tau$, with $u_{\min}$ an eigenvector of $A$ corresponding to eigenvalue $\lambda_{\min}$. With this notation in hand, we state our lower bounds. (See supplemental section~\ref{sec:lb-proof} for a proof.) \begin{theorem}\label{thm:lb} Let $d, t \in \N$ with $t < d$ and $\lambda_{\min},\lambda_{\max},\ltr,\Delta$ be such that $\lambda_{\min} \le \lambda_{\max}$, $\ltr > (-\lambda_{\min})_+$, and $\Delta > 0$. There exists $(A, b, \rho)$ such that $\mathfrak{L}(A,b,\rho) = (\lambda_{\min}, \lambda_{\max}, \ltr, \Delta)$ and for all $s\in\Krylov$, \begin{equation}\label{eq:cr-lb-lin} \fcu(s) - \fcu(\scu) > % \frac{1}{K} \left[\fcu(0) - \fcu(\scu) \right] \exp\left\{ -\frac{4t}{\sqrt{\kappa}-1} \right\}, \end{equation} where $K = 1+\frac{\ltr}{3(\ltr + \lambda_{\min})} $ and $\kappa = \frac{\ltr + \lambda_{\max}}{\ltr +\lambda_{\min}}$. Alternatively, for any $\tau\ge 1$ and $R>0$, there exists $(A, b, \rho)$ such that $\mathfrak{L}'(A,b,\rho) = (\lambda_{\min}, \lambda_{\max}, \tau, R)$ and for $s\in\Krylov$, \begin{equation}\label{eq:cr-lb-sub-ncvx} \fcu(s) - \fcu(\scu) > \min\left\{ (\lambda_{\max})_- - \lambda_{\min}, \frac{\lambda_{\max}-\lambda_{\min}}{16(t-\half)^2} \log^2\left(\frac{\norm{b}^2}{( u_{\min}^T b)^2}\right) \right\}\frac{\norm{\scu}^2}{32}, \end{equation} and % \begin{equation}\label{eq:cr-lb-sub-cvx} \fcu(s) - \fcu(\scu) > \frac{(\lambda_{\max} - \lambda_{\min})\norm{\scu}^2}{16(t+\half)^2}. \end{equation} \end{theorem} The lower bounds~\eqref{eq:cr-lb-lin} matches the linear convergence guarantee~\eqref{eq:cr-lin-time} to within a numerical constant, as we may choose $\lambda_{\max},\lambda_{\min}$ and $\ltr$ so that $\kappa$ is arbitrary and $K < 2$. Similarly, lower bounds~\eqref{eq:cr-lb-sub-ncvx} and~\eqref{eq:cr-lb-sub-cvx} match the sublinear convergence rate~\eqref{eq:cr-sublin-time} for $\lambda_{\min} < 0$ and $\lambda_{\min} \ge 0$ respectively. Our proof flows naturally from minimax characterizations of uniform polynomial approximations (Lemmas~\ref{lem:cheby-first} and~\ref{lem:cheby-second} in the supplement), which also play a crucial role in proving our upper bounds. One consequence of the lower bound~\eqref{eq:cr-lb-lin} is the existence of extremely badly conditioned instances, say with $\kappa = (100d)^2$ and $K=3/2$, such that in the first $d-1$ iterations it is impossible to decrease the initial error by more than a factor of 2 (the initial error may be chosen arbitrarily large as well). However, since these instances have finite condition number we have $\scu\in\Krylov[d]$, and so the error supposedly drops to 0 at the $d$th iteration. This seeming discontinuity stems from the fact that in this case $\scu$ depends on the Lanczos basis of $\Krylov[d]$ through a very badly conditioned linear system and cannot be recovered with finite-precision arithmetic. Indeed, running Krylov subspace methods for $d$ iterations with inexact arithmetic often results in solutions that are very far from exact, while guarantees of the form~\eqref{eq:cr-lin-time} are more robust to roundoff errors~\cite{MuscoMuSi17,DruskinKn91,TrefethenBa97}. While we state the lower bounds in Theorem~\ref{thm:lb} for points in the Krylov subspace $\Krylov$, a classical ``resisting oracle'' construction due to~\citet[Chapter 7.2]{NemirovskiYu83} (see also~\cite[\S10.2.3]{Nemirovski94}) shows that (for $d>2t$) these lower bounds hold also for \emph{any deterministic method} that accesses $A$ only through matrix-vector products, and computes a single matrix-vector product per iteration. The randomization we employ in Corollaries~\ref{cor:tr-rand-joint} and~\ref{cor:tr-rand} breaks the lower bound~\eqref{eq:cr-lb-sub-ncvx} when $\lambda_{\min} < 0$ and $\norm{b}/|u_{\min}^T b|$ is very large, so there is some substantial power from randomization in this case. However, \citet{Simchowitz18} recently showed that randomization cannot break the lower bounds for convex quadratics ($\lambda_{\min}\ge0$ and $\rho=0)$. \section{Polynomial approximation results}\label{sec:polys} In this section we state (and prove for ease of reference) two classical results on uniform polynomial approximation (cf.~\cite{KuczynskiWo92,Nemirovski94}) that stand at the core of the technical development in this work. \begin{lemma}\label{lem:cheby-first} Let $n \ge 1$ and $0< \alpha \le \beta$, and let $\kappa = \beta/\alpha$. Then \begin{equation*} \min_{p \in \Polys} \max_{x\in [\alpha,\beta]} | 1 - x p(x) | = \minmaxT \defeq 2 \left( \left(\frac{\sqrt{\kappa}+1}{\sqrt{\kappa}-1}\right)^n + \left(\frac{\sqrt{\kappa}-1}{\sqrt{\kappa}+1}\right)^n \right)^{-1} \end{equation*} and \begin{equation*} 2\left(e^{2n/(\sqrt{\kappa}-1)}+1\right)^{-1} \le \minmaxT \le 2e^{-2n/\sqrt{\kappa}}. \end{equation*} Moreover, there exist $x_0, x_1, \ldots, x_{n} \in [\alpha,\beta]$ and probability distribution $\pi_0, \pi_1, \ldots \pi_{n}$ such that \begin{equation*} \min_{p \in \Polys} \sum_{k=0}^{n} \pi_k (1-x_k p(x_k))^2 = [\minmaxT] ^2. \end{equation*} \end{lemma} \begin{proof} Let \begin{equation*} T_n(x) = \begin{cases} \cos(n\arccos(x)) & |x| \le 1 \\ \half \left( (x + \sqrt{x^2-1})^n + (x - \sqrt{x^2-1})^n \right) & |x| \ge 1 \end{cases} \end{equation*} denote the order $n$ Chebyshev polynomial of the first kind. We claim that $p^\star\in\Polys$ that solves the minimax problem $\min_{p \in \Polys} \max_{x\in [\alpha,\beta]} | 1 - x p(x) |$ is given by \begin{equation*} 1-xp^\star(x) = \minmaxT\cdot T_n \left( \frac{\kappa+1-2x/\alpha} {\kappa-1} \right), \end{equation*} where $\minmaxT = \left[ T_n \left( \frac{\kappa+1}{\kappa-1} \right)\right]^{-1}$ guarantees that the RHS has value 1 at $x=0$ and therefore $p^\star$ is well defined. Since clearly $|T_n(y)|\le 1$ for every $y\in[-1,1]$, we have that $\max_{x\in [\alpha,\beta]} | 1 - x p^\star(x) | = \minmaxT$. We argue that $p^\star$ is optimal using the classical alternating signs argument, sometimes also referred to as Chebyshev's theorem. First, note that $T_n(y)$ has $n+1$ extrema in $[-1,1]$ (at $y_k=\cos(k\pi/n)$ for $k=0,\ldots,n$) and that their values alternate between $-1$ and $1$ (\ie $T_n(y_k)=(-1)^k$). Therefore, there exist $n+1$ distinct points $x_0, x_1, \ldots, x_{n} \in [\alpha,\beta]$ for which $1-x_i p^\star(x_k) = (-1)^{k}\minmaxT$. Let $q\in\Polys$ satisfy $\max_{x\in [\alpha,\beta]} | 1 - x q(x) | \le \minmaxT$. Then, \[p^\star(x_k) - q(x_k) = \frac{[1-x_k q(x_k)] -[1-x_k p^\star(x_k)]}{x_k} \] must be non-positive for even $k$ and non-negative for odd $k$, and therefore $p^\star- q$ must have at least $n$ roots in $[\alpha,\beta]$. However, $p^\star- q$ is a polynomial of degree at most $n-1$ and can have $n$ roots only if it is identically 0, so we have that $q=p^\star$, proving that $p^\star$ is the unique solution of the minimax problem. To see the upper and lower bounds on $\minmaxT$, note that $\minmaxT = 1/\cosh(n \log( 1 + \frac{2}{\sqrt{\kappa}-1}))$, that $\half e^{|y|} \le \cosh(y) \le \half(e^{|y|} +1)$, and that \begin{equation*} \frac{2}{z} \le \log\left( 1 + \frac{2}{z-1}\right) \le \frac{2}{z-1} \end{equation*} for all $z>1$, where the lower bound above can seen by comparing derivatives. To see the final part of the lemma, let $x_0, x_1, \ldots, x_{n} \in [\alpha,\beta]$ be the points constructed in the optimality argument above, and note that this argument continues to hold if the inner maximization is restricted to these points. Therefore, \begin{equation*} \min_{p \in \Polys}\max_{0\le k \le n} (1-x_k p(x_k))^2 = \left[ \min_{p \in \Polys}\max_{0\le k \le n} |1-x_k p(x_k)| \right]^2 = [\minmaxT]^2. \end{equation*} Letting $\Delta_{n+1}$ denote the probability simplex with $n+1$ variables, we may write \begin{equation*} \max_{0\le k \le n} (1-x_k p(x_k))^2 = \max_{\mu\in\Delta_{n+1}} \sum_{k=0}^{n} \mu_k (1-x_k p(x_k))^2. \end{equation*} Finally, noting that the objective $\sum_{k=0}^{n} \mu_k (1-x_k p(x_k))^2$ is linear (and hence concave) in $\mu$ and convex in (the coefficients of) $p$, we may use Von-Neumann's lemma and swap the $\min$ and $\max$ above, writing \begin{equation*} \max_{\mu\in\Delta_{n+1}} \min_{p \in \Polys} \sum_{k=0}^{n} \mu_k (1-x_k p(x_k))^2 = \min_{p \in \Polys}\max_{\mu\in\Delta_{n+1}} \sum_{k=0}^{n} \mu_k (1-x_k p(x_k))^2 = [\minmaxT]^2. \end{equation*} Letting $\pi$ denote the distribution attaining the outer maximum, we get the desired result. We remark in passing that $\pi$ may be constructed explicitly using the orthogonality principle of least squares estimation and orthogonality relations of Chebyshev polynomials. \end{proof} \begin{lemma}\label{lem:cheby-second} Let $n \ge 1$ and $0< \alpha \le \beta$, let $\kappa = \beta/\alpha$ and define $w(x)\defeq \sqrt{x-\alpha}$. Then \begin{equation*} \min_{p \in \Polys} \max_{x\in [\alpha,\beta]} {w(x)}| 1 - x p(x) | = \minmaxU \defeq 2\sqrt{\alpha} \left( \left(\frac{\sqrt{\kappa}+1}{\sqrt{\kappa}-1}\right)^{n + \half} - \left(\frac{\sqrt{\kappa}-1}{\sqrt{\kappa}+1}\right)^{n + \half} \right)^{-1} \end{equation*} and \begin{equation*} 2\sqrt{\alpha} \left(e^{2(2n+1)/(\sqrt{\kappa}-1)} - 1\right)^{-\half} \le \minmaxU \le 2\sqrt{\alpha} \left(e^{2(2n+1)/\sqrt{\kappa}} - 2\right)^{-\half}. \end{equation*} Moreover, there exist $x_0, x_1, \ldots, x_{n} \in [\alpha,\beta]$ and probability distribution $\pi_0, \pi_1, \ldots \pi_{n}$ such that \begin{equation*} \min_{p \in \Polys} \sum_{k=0}^{n} \pi_k w^2(x_k)(1-x_k p(x_k))^2 = [\minmaxU] ^2. \end{equation*} \end{lemma} \begin{proof} Let \begin{equation*} U_n(x) = \begin{cases} \frac{1}{\sqrt{1-x^2}}\sin((n+1)\arccos(x)) & |x| \le 1 \\ \frac{1}{2\sqrt{x^2-1}}\left( (x + \sqrt{x^2-1})^{n+1} - (x - \sqrt{x^2-1})^{n+1}\right) & |x| \ge 1 \end{cases} \end{equation*} denote the order $n$ Chebyshev polynomial of the second kind. We claim that $p^\star\in\Polys$ that solves the minimax problem $\min_{p \in \Polys} \max_{x\in [\alpha,\beta]} (x-\alpha)^{1/2}| 1 - x p(x) |$ is given by \begin{equation*} 1-xp^\star(x) = \frac{\minmaxU}{w(\beta)}\cdot U_{2n} \left( \sqrt{ \frac{\kappa-x/\alpha} {\kappa-1} }\right), \end{equation*} where $\minmaxU = w(\beta)\left[ U_{2n} \left( \sqrt{\frac{\kappa}{\kappa-1}} \right)\right]^{-1}$ guarantees that the RHS has value 1 at $x=0$ and therefore $p^\star$ is well defined (note that $U_{2n}(\cdot)$ is an even polynomial and therefore $U_{2n}(\sqrt\cdot)$ is a polynomial of degree $n$). For $x\in[\alpha,\beta]$, we have by the definition of $p^*$ and the expression for $U_{2n}$, \begin{equation*} w(x)(1-xp^\star(x)) = \minmaxU \cdot \sin\left( (2n+1)\arccos\left( \sqrt{\frac{\kappa-x/\alpha}{\kappa-1} }\right)\right). \end{equation*} Therefore, we have that $w(x)|1-xp^\star(x)| \le \minmaxU$ for every $x\in[\alpha,\beta]$, and moreover we have that $w(x_k)(1-x_k p^\star(x_k)) = (-1)^k \cdot \minmaxU$, for the points $x_0, \ldots x_n \in [\alpha, \beta]$ satisfying \begin{equation*} \sqrt{\frac{\kappa-x_k/\alpha}{\kappa-1} } = \cos \left(\frac{\pi}{2}\cdot \frac{2k+1}{2n+1}\right). \end{equation*} Hence, the alternating signs argument from the proof of Lemma~\ref{lem:cheby-first} holds here as well and we have that $p^\star$ is optimal and that $\min_{p \in \Polys} \max_{x\in [\alpha,\beta]} {w(x)}| 1 - x p(x) | = \minmaxU$. To see the upper and lower bounds on $\minmaxU$, note that $\minmaxU = \sqrt{\alpha}/\sinh((n+\half) \log( 1 + \frac{2}{\sqrt{\kappa}-1}))$, that for $y\ge 0$, $\sinh(y) =\frac{1}{\sqrt{2}}\sqrt{\cosh(2y)-1}$ gives $\half \sqrt{e^{2y}-2} \le \sinh(y) \le \half \sqrt{e^{2y}-1}$, and that (as in Lemma~\ref{lem:cheby-first}) $\frac{2}{z} \le \log\left( 1 + \frac{2}{z-1}\right) \le \frac{2}{z-1}$. The final part of the lemma follows exactly as in Lemma~\ref{lem:cheby-first}. \end{proof} \section{The trust-region problem}\label{sec:upper} Fixing a symmetric matrix $A\in \R^{d\times d}$, vector $b\in\R^d$ and trust-region radius $R>0$, we let \begin{equation*} \str \in \argmin_{x\in\R^d,~\norm{x}\le R} \f(x) = \half x^T A x + b^T x \end{equation*} denote a solution (global minimizer) of the trust region problem. Letting $\lambda_{\min},\lambda_{\max}$ denote the extremal eigenvalues of $A$, $\str$ admits the following characterization~\cite[Ch.~7]{ConnGoTo00}: $\str$ solves problem~\eqref{eq:problems} if and only if there exists $\ltr$ such that \begin{equation} (A + \ltr I) \str = -b, ~~~ \ltr \ge (-\lambda_{\min})_+, ~~~ \mbox{and} ~~~ \ltr(R - \norm{\str}) = 0. % % % \label{eq:tr-optimality} \end{equation} The optimal Lagrange multiplier $\ltr$ always exists and is unique, and if $\ltr > -\lambda_{\min}$ the solution $\str$ is unique and satisfies $\str = -(A+\ltr I )^{-1}b$. Letting $u_{\min}$ denote the eigenvector of $A$ corresponding to $\lambda_{\min}$, the characterization~\eqref{eq:tr-optimality} shows that $u_{\min}^T b \ne 0$ implies $\ltr > -\lambda_{\min}$. Now, consider the Krylov subspace solutions, and for $t>0$, let \begin{equation*} \itertr_t \in \argmin_{x\in\Krylov,~\norm{x}\le R} \f(x) = \half x^T A x + b^T x \end{equation*} denote a minimizer of the trust-region problem in the Krylov subspace of order $t$ . \citet{GouldLuRoTo99} show how to compute the Krylov subspace solution $\itertr_t$ in time dominated by the cost of computing $t$ matrix-vector products using the Lanczos method (see also Section~\ref{sec:lanczos} of the supplement). \subsection{Main result} With the notation established above, our main result follows. \begin{restatable}{theorem}{thmMain}\label{thm:tr} % % For every $t > 0$, \begin{equation}\label{eq:tr-lin-time} % % \f(\itertr_t) - \f(\str) \le 36\left[\f(0) - \f(\str) \right] \exp\left\{ -4t\sqrt{\frac{\lambda_{\min} + \ltr}{\lambda_{\max} +\ltr}} \right\}, \end{equation} and \begin{equation}\label{eq:tr-sublin-time} \f(\itertr_t) - \f(\str) \le \frac{(\lambda_{\max} - \lambda_{\min})\norm{\str}^2}{(t-\half)^2} \left[ 4 + \frac{\I_{\{\lambda_{\min} < 0\}}}{8} \log^2\left(\frac{4\norm{b}^2}{(u_{\min}^T b)^2}\right) \right]. \end{equation} \end{restatable} Theorem~\ref{thm:tr} characterizes two convergence regimes: linear~\eqref{eq:tr-lin-time} and sublinear~\eqref{eq:tr-sublin-time}. Linear convergence occurs when $t \gtrsim \sqrt{k}$, where $\kappa = \frac{\lambda_{\max} + \ltr}{\lambda_{\min} + \ltr} \ge 1$ is the condition number for the problem. There, the error decays exponentially and falls beneath $\epsilon$ in roughly $\sqrt{\kappa}\log{\frac{1}{\epsilon}}$ Lanczos iteration. Sublinear convergence occurs when $t \lesssim \sqrt{k}$, and there the error decays polynomially and falls beneath $\epsilon$ in roughly $\frac{1}{\sqrt{\epsilon}}$ iterations. For worst-case problem instances this characterization is tight to constant factors, as we show in Section~\ref{sec:lower}. The guarantees of Theorem~\ref{thm:tr} closely resemble the well-known guarantees for the conjugate gradient method~\cite{TrefethenBa97}, including them as the special case $R = \infty$ and $\lambda_{\min} \ge 0$. For convex problems, the radius constraint $\norm{x}\le R$ always improves the conditioning of the problem, as $\frac{\lambda_{\max}}{\lambda_{\min}} \ge \frac{\lambda_{\max}+\ltr}{\lambda_{\min}+\ltr}$; the smaller $R$ is, the better conditioned the problem becomes. For non-convex problems, the sublinear rate features an additional logarithmic term that captures the role of the eigenvector $u_{\min}$. The first rate~\eqref{eq:tr-lin-time} is similar to those of \citet[Thm.~4.11]{ZhangShLi17}, though with somewhat more explicit dependence on $t$. In the ``hard case,'' which corresponds to $u_{\min}^T b = 0$ and $\lambda_{\min} + \ltr = 0$ (cf.~\cite[Ch.~7]{ConnGoTo00}), both the bounds in Theorem~\ref{thm:tr} become vacuous, and indeed $\itertr_t$ may not converge to the global minimizer in this case. However, as the bound~\eqref{eq:tr-sublin-time} depends only logarithmically on $u_{\min}^T b$, it remains valid even extremely close to the hard case. In Section~\ref{sec:upper-rand} we describe two simple randomization techniques with convergence guarantees that are valid in the hard case as well. \subsection{Proof sketch}\label{sec:upper-proof-outline} Our analysis reposes on two elementary observations. First, we note that Krylov subspaces are invariant to shifts by scalar matrices, \ie $\mc{K}_t(A,b) = \mc{K}_t(A_\lambda, b)$ for any $A,b,t$ where $\lambda\in\R$, and \begin{equation*} A_\lambda \defeq A + \lambda I. \end{equation*} Second, we observe that for every point $x$ and $\lambda\in\R$ \begin{align}\label{eq:tr-gamma-pivot-outline} \f(x) - \f(\str) & = \f[A_{\lambda},b](x) - \f[A_{\lambda},b](\str) + \frac{\lambda}{2}(\norm{\str}^2 - \norm{x}^2) \end{align} Our strategy then is to choose $\lambda$ such that $A_\lambda \succeq 0$, and then use known results to find $y_t \in \Krylov[t][A_\lambda,b] = \Krylov[t][A,b]$ that rapidly reduces the ``convex error'' term $\f[A_{\lambda},b](y_t) - \f[A_{\lambda},b](\str)$. We then adjust $y_t$ to obtain a feasible point $x_t$ such that the ``norm error'' term $\frac{\lambda}{2}(\norm{\str}^2 - \norm{x_t}^2)$ is small. To establish linear convergence, we take $\lambda=\ltr$ and adjust the norm of $y_t$ by taking $x_t=(1-\alpha)y_t$ for some small $\alpha$ that guarantees $x_t$ is feasible and that the ``norm error'' term is small. To establish sublinear convergence we set $\lambda=-\lambda_{\min}$ and take $x_t = y_t + \alpha \cdot z_t$, where $z_t$ is an approximation for $u_{\min}$ within $\Krylov$, and $\alpha$ is chosen to make $\norm{x_t}=\norm{\str}$. This means the ``norm error'' vanishes, while the ``convex error'' cannot increase too much, as $A_{-\lambda_{\min}}z_t \approx A_{-\lambda_{\min}}u_{\min}=0$. Our approach for proving the sublinear rate of convergence is inspired by~\citet{Ho-NguyenKi16}, who also rely on Nesterov's method in conjunction with Lanczos-based eigenvector approximation. The analysis in~\cite{Ho-NguyenKi16} uses an algorithmic reduction, proposing to apply the Lanczos method (with a random vector instead of $b$) to approximate $u_{\min}$ and $\lambda_{\min}$, then run Nesterov's method on an approximate version of the ``convex error'' term, and then use the approximated eigenvector to adjust the norm of the result. We instead argue that all the ingredients for this reduction already exist in the Krylov subspace $\Krylov$, obviating the need for explicit eigenvector estimation or actual application of accelerated gradient descent. \subsection{Building blocks} Our proof uses the following classical results. \begin{restatable}[Approximate matrix inverse]{lemma}{lemLinCheby}\label{lem:lin-cheby} % Let $\alpha,\beta$ satisfy $0 < \alpha \le \beta$, and let $\kappa = \beta/\alpha$. For any $t\ge1$ there exists a polynomial $p$ of degree at most $t-1$, such that for every $M$ satisfying $\alpha I \preceq M \preceq \beta I$, \begin{equation*} \opnorm{I - Mp(M)} \le 2e^{-2t/\sqrt{\kappa}}. \end{equation*} % \end{restatable} \begin{restatable}[Convex trust-region problem]{lemma}{lemAGD}\label{lem:agd} Let $t\ge1$, $M\succeq 0$, $v\in \R^d$ and $r\ge 0$, and let $\f[M,v](x) =\half x^T M x + v^T x$. There exists $x_t \in \Krylov[t][M,v]$ such that \begin{equation*} \norm{x_t} \le r ~~\mbox{and}~~ \f[M,v](x_t) - \min_{\norm{x}\le r} \f[M,v](x) \le \frac{4\lambda_{\max}(M) \cdot r^2}{(t+1)^2}. \end{equation*} \end{restatable} \begin{restatable}[{Finding eigenvectors, \cite[Theorem 4.2]{KuczynskiWo92}}]{lemma}{lemEigen}\label{lem:eigenvec-tight} Let $M\succeq 0$ be such that $u^T M u = 0$ for some unit vector $u\in\R^d$, and let $v\in\R^d$. For every $t\ge1$ there exists $z_t \in \Krylov[t][M,v]$ such that \begin{equation*} \norm{z_t} = 1 ~~\mbox{and}~~ z_t^T M z_t \le \frac{\opnorm{M}}{16(t-\half)^2} \log^2\left(-2+4\frac{\norm{v}^2}{(u^T v)^2}\right). \end{equation*} \end{restatable} While these lemmas are standard, their explicit forms are useful, and we prove them in Section~\ref{sec:prelim-proofs} in the supplement. Lemmas~\ref{lem:lin-cheby} and~\ref{lem:eigenvec-tight} are consequences of uniform polynomial approximation results (cf.\ supplement, Sec.~\ref{sec:polys}). To prove Lemma~\ref{lem:agd} we invoke Tseng's results on a variant of Nesterov's accelerated gradient method~\cite{Tseng08}, arguing that its iterates lie in the Krylov subspace. \subsection{Proof of Theorem~\ref{thm:tr}} \paragraph{Linear convergence}\label{sec:upper-lin-proof} Recalling the notation $\Atr = A + \ltr I$, let $y_t = -p(\Atr)b = p(\Atr)\Atr \str$, for the $p\in\mc{P}_t$ which Lemma~\ref{lem:lin-cheby} guarantees to satisfy $\opnorm{p(\Atr)\Atr - I}\le 2e^{-2t/\sqrt{\kappa(\Atr)}}$. Let \begin{equation*} x_t = (1-\alpha) y_t, ~\mbox{where}~ \alpha = \frac{\norm{y_t} - \norm{\str}}{\max\{\norm{\str},\norm{y_t}\}}, \end{equation*} so that we are guaranteed $\norm{x_t}\le\norm{\str}$ for any value of $\norm{y_t}$. Moreover \begin{equation*} |\alpha| = \frac{|\norm{y_t}-\norm{\str}|}{\max\{\norm{\str},\norm{y_t}\}} \le \frac{\norm{y_t - \str}}{\norm{\str}} = \frac{\norm{(p(\Atr)\Atr-I)\str}}{\norm{\str}} \le 2e^{-2t/\sqrt{\kappa(\Atr)}}, \end{equation*} where the last transition used $\opnorm{p(\Atr)\Atr - I}\le 2e^{-2t/\sqrt{\kappa(\Atr)}}$. Since $b = -\Atr\str$, we have $\f[\Atr,b](x) = \f[\Atr,b](\str) + \half\norms{\Atr^{1/2}(x-\str)}^2$. The equality~\eqref{eq:tr-gamma-pivot-outline} with $\lambda = \ltr$ and $\norm{x_t}\le \norm{\str}$ therefore implies \begin{equation}\label{eq:tr-lin-subopt-bound} \f(x_t) - \f(\str) \le \half\norm{\Atr^{1/2}(x_t - \str)}^2 + \ltr\norm{\str}(\norm{\str} - \norm{x_t}). \end{equation} When $\norm{y_t} \ge \norm{\str}$ we have $\norm{x_t} = \norm{\str}$ and the second term vanishes. When $\norm{y_t} < \norm{\str}$, \begin{equation}\label{eq:tr-lin-norm-diff} \norm{\str} - \norm{x_t} = \norm{\str} - \norm{y_t} - \frac{\norm{y_t}}{\norm{\str}}\cdot (\norm{\str} - \norm{y_t} ) = \norm{\str}\alpha^2 \le 4e^{-4t/\sqrt{\kappa(\Atr)}}\norm{\str}. \end{equation} We also have, \begin{align}\label{eq:tr-linear-quad} \lefteqn{\norm{\Atr^{1/2}(x_t - \str)} = \norm{\left([1-\alpha]p(\Atr)\Atr-I\right)\Atr^{1/2}\str}} \nonumber \\ & ~~ \le (1+|\alpha|)\norm{\left(p(\Atr)\Atr-I\right)\Atr^{1/2}\str} + |\alpha| \norm{\Atr^{1/2}\str} % \le 6 \norm{\Atr^{1/2}\str} e^{-2t/\sqrt{\kappa(\Atr)}}, \end{align} where in the final transition we used our upper bounds on $\alpha$ and $\opnorm{p(\Atr)\Atr - I}$, as well as $|\alpha|\le 1$. Substituting the bounds~\eqref{eq:tr-lin-norm-diff} and~\eqref{eq:tr-linear-quad} into inequality~\eqref{eq:tr-lin-subopt-bound}, we have \begin{equation}\label{eq:tr-lin-time-bound-stronger} \f(x_t) - \f(\str) \le \left(18 \str^T \Atr \str + 4\ltr\norm{\str}^2\right) e^{-4t/\sqrt{\kappa(\Atr)}}, \end{equation} and the final bound follows from recalling that $\f(0)-\f(\str) = \half \str^T \Atr \str + \frac{\ltr}{2}\norm{\str}^2$ and substituting $\kappa(\Atr) = (\lambda_{\max} + \ltr)/(\lambda_{\min} + \ltr)$. To conclude the proof we note that $(1-\alpha)p(\Atr) = (1-\alpha)p(A + \ltr I) = \tilde{p}(A)$ for some $\tilde{p} \in \mc{P}_t$, so that $x_t \in \mc{K}_t(A, b)$ and $\norm{x_t}\le R$, and therefore $\f(\itertr_t) \le \f(x_t)$. \paragraph{Sublinear convergence}\label{sec:upper-sub-proof} \newcommand{A_{0}}{A_{0}} \newcommand{\tilde{x}^\star_0}{\tilde{x}^\star_0} Let $A_{0} \defeq A - \lambda_{\min} I \succeq 0$ and apply Lemma~\ref{lem:agd} with $M=A_{0}$, $v=b$ and $r=\norm{\str}$ to obtain $y_t \in \Krylov[t][A_{0}, b] = \Krylov$ such that \begin{equation}\label{eq:tr-sublin-yt} \norm{y_t}\le \norm{\str} ~\mbox{and}~ \f[A_{0},b](y_t) - \f[A_{0},b](\str) \le \f[A_{0},b](y_t) - \min_{\norm{x}\le\norm{\str}}\f[A_{0},b](x) \le \frac{4\opnorm{A_{0}}\norm{\str}^2}{(t+1)^2}. \end{equation} If $\lambda_{\min} \ge 0$, equality~\eqref{eq:tr-gamma-pivot-outline} with $\lambda=-\lambda_{\min}$ along with~\eqref{eq:tr-sublin-yt} means we are done, recalling that $\opnorm{A_{0}} = \lambda_{\max}-\lambda_{\min}$. For $\lambda_{\min}< 0$, apply Lemma~\ref{lem:eigenvec-tight} with $M=A_{0}$ and $v=b$ to obtain $z_t\in\Krylov$ such that \begin{equation}\label{eq:tr-sublin-zt} \norm{z_t} = 1 ~~\mbox{and}~~ z_t^T A_{0} z_t \le \frac{\opnorm{A_{0}}}{16(t-\half)^2} \log^2\left(4\frac{\norm{b}^2}{(u_{\min}^T b)^2}\right). \end{equation} We form the vector \begin{equation*} x_t = y_t + \alpha \cdot z_t\in\mc{K}_t(A,b), \end{equation*} and choose $\alpha$ to satisfy \begin{equation*} \norm{x_t} = \norm{\str} ~~\mbox{and}~~ \alpha \cdot z_t ^T (A_{0} y_t + b) = \alpha \cdot z_t ^T \grad \f[A_{0},b](y_t) \le 0. \end{equation*} We may always choose such $\alpha$ because $\norm{y_t}\le\norm{\str}$ and therefore $\norm{y_t + \alpha z_t} = \norm{\str}$ has both a non-positive and a non-negative solution in $\alpha$. Moreover because $\norm{z_t}=1$ we have that $|\alpha| \le 2\norm{\str}$. The property $\alpha \cdot z_t ^T \grad \f[A_{0},b](y_t) \le 0$ of our construction of $\alpha$ along with $\hess \f[A_{0},b] = A_{0}$, gives us, \begin{equation*} \f[A_{0},b](x_t) = \f[A_{0},b](y_t) + \alpha \cdot z_t ^T \grad \f[A_{0},b](y_t) + \frac{\alpha^2}{2} z_t^T A_{0} z_t \le \f[A_{0},b](y_t) + \frac{\alpha^2}{2} z_t^T A_{0} z_t. \end{equation*} Substituting this bound along with $\norm{x_t}=\norm{\str}$ and $\alpha^2 \le 4\norm{\str}^2$ into~\eqref{eq:tr-gamma-pivot-outline} with $\lambda=-\lambda_{\min}$ gives \begin{equation*} \f(x_t) - \f(\str) \le \f[A_{0},b](y_t) - \f[A_{0},b](\str) + 2\norm{\str}^2 z_t^T A_{0} z_t. \end{equation*} Substituting in the bounds~\eqref{eq:tr-sublin-yt} and~\eqref{eq:tr-sublin-zt} concludes the proof for the case $\lambda_{\min} < 0$. \subsection{Randomizing away the hard case}\label{sec:upper-rand} \newcommand{\hat{s}^{\mathsf{tr}}}{\hat{s}^{\mathsf{tr}}} \newcommand{\tilde{s}^{\mathsf{tr}}}{\tilde{s}^{\mathsf{tr}}} \newcommand{\tilde{b}}{\tilde{b}} \newcommand{\tilde{x}^{\star}_{\mathsf{tr}}}{\tilde{x}^{\star}_{\mathsf{tr}}} \newcommand{\tilde{\lambda}^{\star}}{\tilde{\lambda}^{\star}} \newcommand{A_{\pltr}}{A_{\tilde{\lambda}^{\star}}} Krylov subspace solutions may fail to converge to global solution when both $\ltr = -\lambda_{\min}$ and $u_{\min}^T b = 0$, the so-called hard case~\cite{ConnGoTo00,NocedalWr06}. Yet as with eigenvector methods~\cite{KuczynskiWo92,GolubVa89}, simple randomization approaches allow us to handle the hard case with high probability, at the modest cost of introducing to the error bounds a logarithmic dependence on $d$. Here we describe two such approaches. In the first approach, we draw a spherically symmetric random vector $v$, and consider the \emph{joint Krylov subspace} \begin{equation*} \Krylov[2t][A, \{b, v\}] \defeq \mathrm{span}\{b, Ab, \ldots, A^{t-1}b, v, Av, \ldots, A^{t-1}v\}. \end{equation*} The trust-region and cubic-regularized problems~\eqref{eq:problems} can be solved efficiently in $\Krylov[2t][A, \{b, v\}]$ using the \emph{block Lanczos} method~\cite{CullumDo74,Golub77}; we survey this technique in Section~\ref{sub:block-lanczos} in the supplement. The analysis in the previous section immediately implies the following convergence guarantee. \begin{restatable}{corollary}{corRandJoin}\label{cor:tr-rand-joint} Let $v$ be uniformly distributed on the unit sphere in $\R^d$, and $$\hat{s}^{\mathsf{tr}}_t \in \argmin_{x\in\Krylov[\floor{t/2}][A,\{b,v\}], \norm{x}\le R} \f[A,b](x).$$ For any $\delta > 0$, \begin{equation}\label{eq:tr-joint-sublin-time} \f(\hat{s}^{\mathsf{tr}}_t) - \f(\str) \le \frac{(\lambda_{\max} - \lambda_{\min})R^2}{(t-1)^2} \left[ 16 + 2\cdot\I_{\{\lambda_{\min} < 0\}} \log^2\left(\frac{2\sqrt{d}}{ \delta}\right) \right] \end{equation} with probability at least $1-\delta$ with respect to the random choice of $v$. % % \end{restatable} \begin{proof} In the preceding proof of sublinear convergence, apply Lemma~\ref{lem:agd} on $\Krylov[\floor{t/2}][A,b]$ and Lemma~\ref{lem:eigenvec-tight} on $\Krylov[\floor{t/2}][A,v]$; the constructed solution is in $\Krylov[\floor{t/2}][A,\{b,v\}]$. To bound $|u_{\min}^T v|^2/\|v\|^2$, note that its distribution is $\textrm{Beta}(\frac{1}{2}, \frac{d-1}{2})$ and therefore $|u_{\min}^T v|^2/\|v\|^2 \ge \delta^2 / d$ with probability greater than $1-\delta$ (cf.~\cite[Lemma 4.6]{CarmonDu16}). \end{proof} Corollary~\ref{cor:tr-rand-joint} implies we can solve the trust-region problem to $\epsilon$ accuracy in roughly $\epsilon^{-1/2}\log d$ matrix-vector products, even in the hard case. The main drawback of this randomization approach is that half the matrix-vector products are expended on the random vector; when the problem is well-conditioned or when $|u_{\min}^T b|/\norms{b}$ is not extremely small, using the standard subspace solution is nearly twice as fast. The second approach follows the proposal~\cite{CarmonDu16} to construct a perturbed version of the linear term $b$, denoted $\tilde{b}$, and solve the problem instance $(A,\tilde{b}, R)$ in the Krylov subspace $\Krylov[t][A,\tilde{b}]$. \begin{restatable}{corollary}{corRand}\label{cor:tr-rand} Let $v$ be uniformly distributed on the unit sphere in $\R^d$, let $\sigma > 0$ and let \begin{equation*} \tilde{b} = b + \sigma \cdot v. \end{equation*} Let $\tilde{s}^{\mathsf{tr}}_t \in \argmin_{x\in\Krylov[t][A,\tilde{b}], \norm{x}\le R} \f[A,\tilde{b}](x) \defeq \half x^T A x + \tilde{b}^T x$. For any $\delta > 0$, \begin{equation}\label{eq:tr-pert-sublin-time} \f(\tilde{s}^{\mathsf{tr}}_t) - \f(\str) \le \frac{(\lambda_{\max} - \lambda_{\min})R^2}{(t-\half)^2} \left[ 4 + \frac{\I_{\{\lambda_{\min} < 0\}}}{2} \log^2\left(\frac{2\norms{\tilde{b}}\sqrt{d}}{\sigma \delta}\right) \right] + 2\sigma R \end{equation} with probability at least $1-\delta$ with respect to the random choice of $v$. \end{restatable} \noindent See section~\ref{sec:tr-rand-proof} in the supplement for a short proof, which consists of arguing that $\f$ and $\f[A,\tilde{b}]$ deviate by at most $\sigma R$ at any feasible point, and applying a probabilistic lower bound on $|u_{\min}^T \tilde{b}|$. For any desired accuracy $\epsilon$, using Corollary~\ref{cor:tr-rand} with $\sigma = \epsilon/(4R)$ shows we can achieve this accuracy, with constant probability, in a number of Lanczos iterations that scales as $\epsilon^{-1/2}\log(d/\epsilon^2)$. Compared to the first approach, this rate of convergence is asymptotically slightly slower (by a factor of $\log{\frac{1}{\epsilon}}$), and moreover requires us to decide on a desired level of accuracy in advance. However, the second approach avoids the 2x slowdown that the first approach exhibits on easier problem instances. In Section~\ref{sec:exp} we compare the two approaches empirically. We remark that the linear convergence guarantee~\eqref{eq:tr-lin-time} continues to hold for both randomization approaches. For the second approach, this is due to the fact that small perturbations to $b$ do not drastically change the condition number, as shown in~\cite{CarmonDu16}. However, this also means that we cannot expect a good condition number when perturbing $b$ in the hard case. Nevertheless, we believe it is possible to show that, with randomization, Krylov subspace methods exhibit linear convergence even in the hard case, where the condition number is replaced by the normalized eigen-gap $(\lambda_{\max}-\lambda_{\min})/(\lambda_2 - \lambda_{\min})$, with $\lambda_2$ the smallest eigenvalue of $A$ larger than $\lambda_{\min}$. \section{Proofs from Section~\ref{sec:upper}}\label{sec:ub-proofs} \subsection{Proof of auxiliary lemmas}\label{sec:prelim-proofs} \lemLinCheby* \begin{proof} This is an immediate consequence of Lemma~\ref{lem:cheby-first}, as \begin{equation*} \min_{p\in\Polys[t]}\max_{\alpha I \preceq M \preceq \beta I}\opnorm{I - Mp(M)} = \min_{p\in\Polys[t]}\max_{\lambda\in [\alpha,\beta]}\left|1-\lambda \cdot p(\lambda)\right| = \minmaxT[t]. \end{equation*} \end{proof} \lemAGD* \begin{proof} Let $g:\R^d \to \R$ be convex with $L$-Lipschitz gradient and let $Q\subseteq\R^d$ be a convex set containing the point $0$. Consider Nesterov's accelerated gradient method for minimization of $g$, which comprises the following recursion \cite[Scheme (2.2.17)]{Nesterov04}, \begin{gather*} x_{k+1} = \min_{x\in Q}\left\{ x^T \grad g(y_k) + \frac{L}{2}\norm{x -y_k}^2 \right\} = \Pi_Q\left( y_k - \frac{1}{L}\grad g(y_k) \right) % % \\ \alpha_{k+1}^2/(1-\alpha_{k+1}) = \alpha_k^2 \Rightarrow \alpha_{k+1} = -\frac{\alpha_k^2}{2} + \frac{\alpha_k^2}{2}\sqrt{ 1+\frac{4}{\alpha_k^2}} \\ y_{k+1} = x_{k+1} + \alpha_{k+1}(\alpha_k^{-1}-1) (x_{k+1} - x_k), \end{gather*} where $\Pi_Q(\cdot)$ is the Euclidean projection to $Q$. Letting $\alpha_0=1$ and $y_0 = x_0=0$, and letting $x^\star$ denote any minimizer of $g$ in $Q$, the analysis of~\citet[Corollary 2(b)]{Tseng08} gives\footnote{ translating to the notation of~\cite{Tseng08}, take $\phi(x,v) = g(x)$ and $P(x)$ to be the indicator of $Q$, so that $q^P(\cdot) = g(x^\star)$, note that $\theta_k$ ($\alpha_k$ in our notation) satisfies $\theta_k \le 2/(2+k)$. We discuss alternative references for this result after the proof.}, \begin{equation}\label{eq:tseng-bound} g(x_t) - g(x^\star) \le \frac{4L\max_{z\in Q}\norm{z}^2}{(t+1)^2}. \end{equation} Taking $g=\f[M,v]$ and $Q=B_r=\{x\mid\norm{x}\le r\}$, we note that $\f[M,v]$ is convex with $L\defeq\lambda_{\max}(M)$-Lipschitz gradient, and that the projection step guarantees that $\norm{x_t}\le r$ for every $t$. Therefore, to establish the lemma it remains only to argue that $x_t$ as defined above is in $\Krylov[t][M,v]$; we shall see this by simple induction, whose basis is $y_0, x_0 \in \Krylov[0][M,v] = \{0\}$. Assume now that $y_k, x_k \in \Krylov[k][M,v]$ for some $k\ge0$. This implies \begin{equation*} y_k - \frac{1}{L}\grad g(y_k) = y_k - \frac{1}{L}A y_k - \frac{1}{L} v \in \Krylov[k+1][M,v]. \end{equation*} Further, note that projection to the Euclidean ball $B_r$ is simply scaling: \begin{equation*} \Pi_{Q}(z) = \Pi_{B_r}(z) = \frac{r}{\max\{r, \norm{z}\}} \cdot z, \end{equation*} and therefore $x_{k+1}\in \Krylov[k+1][M,v]$. Finally, $y_{k+1}$ is simply a linear combination of $x_{k+1}$ and $x_k$ and therefore is also in $\Krylov[k+1][M,v]$, concluding the induction and the proof. \end{proof} A bound similar to~\eqref{eq:tseng-bound} appears in Nesterov's earlier analysis~\cite[Theorem~2.2.3]{Nesterov04}, but with an the additional factor proportional to $g(0)-g(x\opt)$ which is not immediately upper bounded by $\half L \max_{z\in Q}\norms{z}^2$ due to the constraint $z\in Q$. The bound~\eqref{eq:tseng-bound} also appears in later work of~\citet{AllenOr17}. \lemEigen* \newcommand{\err}[1][t]{\mathsf{err}_{#1}} \begin{proof} Let $\lambda\parind{1} \le \lambda\parind{2} \le \cdots \le \lambda\parind{d}$ denote the eigenvalues of $M$ and let $u_1, u_2, \ldots, u_d$ denote their corresponding (orthonormal) eigenvectors. By our assumption $\lambda\parind{1} = 0$ and we have also $\lambda\parind{d} = \opnorm{M}$. We let \begin{equation*} v\parind{i} \defeq u_i^T v \end{equation*} denote the component of $v$ in the eigenbasis of $M$. Define \begin{equation*} \err \defeq \min_{p\in\Polys[t]} \frac{ (p(M)v)^T M p(M)v }{\norm{p(M)v}^2 } = \min_{p\in\Polys[t]} \frac{ \sum_{i=1}^d v\parind{i}^2 p^2(\lambda\parind{i}) \lambda\parind{i}} {\sum_{i=1}^d v\parind{i}^2 p^2(\lambda\parind{i})}, \end{equation*} and let $q\in\Polys[t]$ attain the minimum above. Setting $z_t = q(M)v/\norm{q(M)v}$, we see that \begin{equation*} \err = z_t^T M z_t = \frac{ \sum_{i=1}^d v\parind{i}^2 q^2(\lambda\parind{i}) \lambda\parind{i}} {\sum_{i=1}^d v\parind{i}^2 q^2(\lambda\parind{i})}, \end{equation*} and so our proof comprises of bounding $\err$ from above. We invoke Lemma~\ref{lem:cheby-second} with $n=t-1$, $\alpha=\err$ and $\beta=\lambda\parind{d}=\opnorm{M}$; let $\tilde{q}(x) = 1-xp^\star(x)\in\Polys[t]$ be the polynomial for which the Lemma guarantees \begin{equation*} \max_{x\in [\err, \lambda\parind{d}]} (x-\err)^{1/2}|\tilde{q}(x)| = \minmaxU[t-1]. \end{equation*} By the optimality of $q$, we have that \begin{equation*} \err \le \frac{ \sum_{i=1}^d v\parind{i}^2 \tilde{q}^2(\lambda\parind{i}) \lambda\parind{i}} {\sum_{i=1}^d v\parind{i}^2 \tilde{q}^2(\lambda\parind{i})}. \end{equation*} Rearranging and noting that $\tilde{q}(\lambda\parind{1}) = \tilde{q}(0) = 1$, we obtain \begin{equation*} \err \le \sum_{i=2}^d \frac{v\parind{i}^2}{v\parind{1}^2} (\lambda\parind{i} - \err)\tilde{q}^2(\lambda\parind{i}) \le \frac{\norm{v}^2-v\parind{1}^2}{v\parind{1}^2} \max_{\lambda\in [\err, \lambda\parind{d}]} (\lambda - \err) \tilde{q}^2(\lambda) = \left(\frac{\norm{v}^2}{v\parind{1}^2} -1\right)[\minmaxU[t-1]]^2. \end{equation*} Lemma~\ref{lem:cheby-second} provides the bound \begin{equation*} [\minmaxU[t-1]]^2 \le \frac{4\err}{e^{2(2t-1)\sqrt{\err/\opnorm{M}}}-2}. \end{equation*} Substituting the upper bound into $\err \le \big(\frac{\norm{v}^2}{v\parind{1}^2}-1\big)[\minmaxU[t-1]]^2$ and rearranging gives the result. \end{proof} \subsection{Proof of Corollary~\ref{cor:tr-rand}}\label{sec:tr-rand-proof} \corRand* \begin{proof} Let $\tilde{x}^{\star}_{\mathsf{tr}} \in \argmin_{x\in\Krylov[t][A,\tilde{b}], \norm{x}\le R} \f[A,\tilde{b}](x)$ be a solution to the perturbed problem. Since $v$ is a unit vector, for any feasible $x$ we have \begin{align}\label{eq:tr-rand-pointwise-bound} \f(x) - \f(\str) & = \f[A,\tilde{b}](x) - \f[A,\tilde{b}](\str) + \sigma\cdot v^T (\str - x) \le \f[A,\tilde{b}](x) - \f[A,\tilde{b}](\str) + 2\sigma R \nonumber \\ & \le \f[A,\tilde{b}](x) - \f[A,\tilde{b}](\tilde{x}^{\star}_{\mathsf{tr}}) + 2\sigma R, \end{align} and so it suffices to argue about the perturbed optimality gap $\f[A,\tilde{b}](\tilde{s}^{\mathsf{tr}}_t) - \f[A,\tilde{b}](\str)$. Applying the bound~\eqref{eq:tr-sublin-time} on the perturbed problem gives us \begin{equation}\label{eq:tr-rand-pert-sublin-bound} \f[A,\tilde{b}](\tilde{s}^{\mathsf{tr}}_t) - \f[A,\tilde{b}](\tilde{x}^{\star}_{\mathsf{tr}}) \le \frac{(\lambda_{\max} - \lambda_{\min})R^2}{(t-\half)^2} \left[ 4 + \frac{\I_{\{\lambda_{\min} < 0\}}}{2} \log^2\left(2\frac{\norms{\tilde{b}}}{|u_{\min}^T \tilde{b}|}\right) \right], \end{equation} and a simple argument on the density of $u_{\min}^T \tilde{b}$ (cf.~\cite[Lemma 4.6]{CarmonDu16}) shows that \begin{equation}\label{eq:tr-rand-b1-bound} |u_{\min}^T \tilde{b}| \ge \frac{\sigma \cdot \delta}{\sqrt{d}} ~~\mbox{with probability at least $1-\delta$}. \end{equation} Combining the bounds~\eqref{eq:tr-rand-pointwise-bound}, \eqref{eq:tr-rand-pert-sublin-bound} and~\eqref{eq:tr-rand-b1-bound} gives the result~\eqref{eq:tr-pert-sublin-time}. % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % % \end{proof}
{ "timestamp": "2019-01-03T02:14:20", "yymm": "1806", "arxiv_id": "1806.09222", "language": "en", "url": "https://arxiv.org/abs/1806.09222", "abstract": "We provide convergence rates for Krylov subspace solutions to the trust-region and cubic-regularized (nonconvex) quadratic problems. Such solutions may be efficiently computed by the Lanczos method and have long been used in practice. We prove error bounds of the form $1/t^2$ and $e^{-4t/\\sqrt{\\kappa}}$, where $\\kappa$ is a condition number for the problem, and $t$ is the Krylov subspace order (number of Lanczos iterations). We also provide lower bounds showing that our analysis is sharp.", "subjects": "Optimization and Control (math.OC)", "title": "Analysis of Krylov Subspace Solutions of Regularized Nonconvex Quadratic Problems", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9890130563978902, "lm_q2_score": 0.8104789155369047, "lm_q1q2_score": 0.8015742294012015 }
https://arxiv.org/abs/2101.06818
Approximating monomials using Chebyshev polynomials
This paper considers the approximation of a monomial $x^n$ over the interval $[-1,1]$ by a lower-degree polynomial. This polynomial approximation can be easily computed analytically and is obtained by truncating the analytical Chebyshev series expansion of $x^n$. The error in the polynomial approximation in the supremum norm has an exact expression with an interesting probabilistic interpretation. We use this interpretation along with concentration inequalities to develop a useful upper bound for the error.
\section{Motivation and Introduction}\label{sec:intro} We are interested in approximating the monomial $x^n$ by a polynomial of degree $0 \leq k < n$ over the interval $[-1,1]$. The monomials $1,x,x^2,\dots$ form a basis for $C[-1,1]$, so it seems unlikely that we can represent a monomial in terms of lower degree polynomials. In Figure~\ref{fig:monomialbasis}, we plot a few functions from the monomial basis over $[0,1]$; the basis function look increasingly alike as we take higher and higher powers, i.e., they appear to ``lose independence.'' Numerical analysts often avoid the monomial basis in polynomial interpolation since they result in ill-conditioned Vandermonde matrices, leading to poor numerical performance in finite precision arithmetic. This loss of independence means that it is reasonable to approximate the monomial $x^n$ as a linear combination of lower order monomials, i.e., a lower order polynomial approximation. The natural question to ask, therefore, is: how small can $k$ be so that a well-chosen polynomial of degree $k$ can accurately approximate $x^n$? \begin{figure}[!ht] \centering \includegraphics[scale=0.4]{monomial_basis} \caption{Visualization of a few monomials in the interval $[0,1]$.} \label{fig:monomialbasis} \end{figure} The surprising answer to this question is that we can approximate the monomial $x^n$ over $[-1,1]$ by a polynomial of small degree, which we will make precise. Let $\|f\|_\infty = \max_{x \in [-1,1]} | f(x)|$ denote the supremum norm on $C[-1,1]$ and let $\pi_k^*(\cdot)$ be the best polynomial approximation to $x^n$ in this norm; that is \[ E_{n,k} := \min_{\pi \in \mathcal{P}_k} \|x^n- \pi(x) \|_\infty= \|x^n - \pi_k^*(x)\|_\infty, \] where $\mathcal{P}_k$ is a vector space of polynomials with real coefficients of degree at most $k$. The minimizer $\pi_k^*(\cdot)$ exists and is unique~\cite[Chapter 10]{trefethen2013approximation}, but does not have a closed form expression. Newman and Rivlin~\cite[Theorem 2]{newman1976approximation} showed that\footnote{We briefly mention that the notation our manuscript differs from~\cite{newman1976approximation} in that we reverse the roles of $n$ and $k$.} \begin{equation}\label{eqn:newmanrivlin} \frac{p_{n,k}}{4e} \leq \|x^n - \pi_k^*(x)\|_\infty \leq p_{n,k}, \end{equation} where the term $p_{n,k}$ is given by the formula \[ p_{n,k} = \frac{1}{2^{n-1}} \sum_{j = \lfloor (n+k)/2 \rfloor +1}^n \binom{n}{j}. \] Since $p_{n,k}$ involves the sum of binomial coefficients, it has a probabilistic interpretation which we explore in Section~\ref{sec:prob}. To see why a small $k$ is sufficient, consider the upper bound $p_{n,k}$. In Section~\ref{sec:prob} we use the probabilistic interpretation to obtain the following bound $p_{n,k} \leq 2\exp\left(-{k^2}/{2n}\right)$. Suppose we are given a user-defined tolerance $\epsilon > 0$. To ensure \[ \|x^n - \pi_k^*(x)\|_\infty \leq \epsilon,\] we need to choose $k \geq \sqrt{2n\log(2/\epsilon)}$. The accuracy of the polynomial approximation is visualized in Figure~\ref{fig:approx}, where in the left panel we plot the monomial $x^{n}$ for $n=75$ and the best polynomial approximation $\pi_{k}^*$ for $k=5,15,25$. The polynomial $\pi_k^*$ is computed using the Remez algorithm, implemented in chebfun~\cite{Driscoll2014}. We see that for $k=25$, the polynomial approximation looks very accurate. In the right panel, we display $p_{n,k}$, which is the upper bound of the best polynomial approximation, as well as the upper bound for $p_{n,k}$. We see that $p_{n,k}$ and its upper bound both have sharp decay with increasing $k$. Numerical evidence in~\cite{nakatsukasa2018rational} further confirms this analysis; the authors show that the error $E_{n,k}$ behaves approximately like $\frac12 \text{erfc}(k/\sqrt{n})$, where erfc is the complementary error function. . \begin{figure}[!ht] \centering \includegraphics[scale=0.5]{monomial_approx.eps} \caption{(left) Approximation of the monomial $x^{n}$ for $n=75$ by $\pi_k^*$, (right) $p_{n,k}$ and its upper bound $2\exp(-k^2/2n)$. The visualization is restricted to the interval $[0,1]$.} \label{fig:approx} \end{figure} Polynomial and rational approximations to the monomial has received considerable attention in the approximation theory community, and surveys of various results can be found in~\cite{reddy1987approximations,nakatsukasa2018rational}. Polynomial approximations to high order monomials have many applications in numerical analysis. This key insight was exploited by Cornelius Lanczos~\cite{lanczos1988applied} in his ``$\tau$-method'' for the numerical solution of differential equations. For a simulating discussion on this topic, please see~\cite{nakatsukasa2018rational}. In numerical linear algebra, this has been exploited to efficiently compute matrix powers and the Schatten p-norm of a matrix~\cite{avron2011randomized,dudley2020monte}. In this short note, we show to construct a polynomial approximation $x^n \approx \phi_k(x)$ using a truncated Chebyshev polynomial expansion. The error in the truncated representation equals the sum of the discarded coefficients and is precisely $p_{n,k}$. The polynomial $\phi_k$ and the resulting error can both be computed analytically and, therefore, is of great practical use. We briefly review Chebyshev polynomials in Section~\ref{sec:cheby} and state and prove the main result in Section~\ref{sec:main}. In Section~\ref{sec:prob}, we explore probabilistic interpretations of $p_{n,k}$ and obtain bounds for partial sums of binomial coefficients. \section{Chebyshev polynomials}\label{sec:cheby} The Chebyshev polynomials of the first kind $T_n(x)$ for $n=0,1,2,\dots$ can be represented as \[T_n(x) = \cos(n\arccos{x}) \qquad x \in [-1,1]. \] Starting with $T_0(x) = 1$, $T_1(x) = x$, the Chebyshev polynomials satisfy a recurrence relationship of the form $T_{n+1}(x) = 2xT_n(x) - T_{n-1}(x)$ for $n\geq 1$. The Chebyshev polynomials are orthogonal with respect to the weighted inner product \[\langle u,v \rangle = \int_{-1}^1 w(x) u(x) v(x) dx \] where the weight function takes the form $w(x) = (1-x^2)^{-1/2}$. Any function $f \in C[-1,1]$ that is Lipschitz continuous can be represented in terms of a Chebyshev polynomial expansion of the form \[ f(x) = \frac12 c_0 + \sum_{j=1}^\infty c_jT_j(x) \qquad x \in [-1,1], \] where the coefficients $c_j$ are obtained as $c_j = \frac{2}{{\pi}}\langle f(x), T_j(x)\rangle$ and the series is uniformly convergent. The monomial $x^n$ is rather special since it has the following exact representation in terms of the Chebyshev polynomials~\cite[Section 4]{cody1970survey} \begin{equation}\label{eqn:xncheby} x^n = \sum_{j=0}^{n}{}^{'} c_jT_j(x),\end{equation} where ${}^{'}$ means the summand corresponding to $j=0$ is halved (if it appears) and the coefficients $c_j$ for $j=0,\dots,n$ are \begin{equation}\label{eqn:cj} c_j = \left\{ \begin{array}{ll} 2^{1-n}\binom{n}{(n-j)/2} & n-j \text{ even} \\ 0 & \text{otherwise}.\end{array} \right.\end{equation} Equation~\eqref{eqn:xncheby} takes a more familiar form, when we consider the trigonometric perspective of Chebyshev polynomials. For example, the well-known trigonometric identity $\cos(3\theta) = 4\cos^3 \theta - 3\cos \theta$, can be arranged as \[ \cos^3 \theta = \frac{3}{4}\cos\theta + \frac{1}{4} \cos(3\theta) = \frac{1}{2^2} \left( \binom{3}{1} \cos\theta + \binom{3}{0} \cos(3\theta)\right). \] With $x=\cos\theta$, we get $x^3 = 2^{-2} \binom{3}{1} T_1(x) + 2^{-2} \binom{3}{0} T_3(x)$. For completeness, we provide a derivation of~\eqref{eqn:xncheby} in Appendix~\ref{app:der}. It is important to note here that the series in~\eqref{eqn:xncheby} is finite, but can be truncated to obtain an accurate approximation; see Section~\ref{sec:main}. Chebyshev polynomials have many applications in approximation theory and numerical analysis~\cite{trefethen2013approximation} but we limit ourselves to two such examples here. First, if the function is differentiable $r$ times or analytic, the Chebyshev coefficients exhibit decay (algebraic or geometric respectively). Therefore, the Chebyshev series can be truncated to obtain an polynomial approximation of the function and the accuracy of the approximation depends on the rate of decay of the coefficients. Another application of Chebyshev polynomials is in the theory and practice of polynomial interpolation. The polynomial ${q}_{n-1}^*(x) = x^{n} - 2^{1-n}T_{n}(x)$ solves the minimax problem \begin{equation}\label{eqn:minmax} \min_{q \in \mathcal{P}_{n-1}} \|x^{n} - q(x)\|_\infty = 2^{1-n}. \end{equation} Based on the minimax characterization, to interpolate a function over $[-1,1]$ by a polynomial of degree $n-1$, the function to be interpolated should be evaluated at the roots of the Chebyshev polynomial $T_{n}$ given by the points $x_j = \cos\left(\frac{2j+1}{2n }\pi\right)$ for $j=0,\dots,n-1$. \section{Main result}\label{sec:main} We construct the polynomial approximation $ x^n \approx \phi_k(x)$ by truncating the Chebyshev polynomial expansion in~\eqref{eqn:xncheby} beyond the term $j=k$. That is \[\phi_k(x) := \sum_{j=0}^k{}^{'}c_jT_j(x).\] Our main result is the following theorem, which quantifies the error in the polynomial approximation. The proof of this theorem is based on the expression in~\eqref{eqn:xncheby}. We believe this result is new. \begin{theorem}\label{thm:main} The error in the polynomial approximation $\phi_k(x)$ satisfies \[ \|x^n - \phi_k(x)\|_\infty = p_{n,k}. \] \end{theorem} \begin{proof} From~\eqref{eqn:xncheby}, $x^{n} - \phi_k(x) = \sum_{j=k+1}^nc_jT_j(x)$. Using triangle inequality, we find that $\|x^n-\phi_k(x)\|_\infty \leq \sum_{j=k+1}^n c_j$ since the coefficients are nonnegative and the Chebyshev polynomials are bounded as $|T_j(x)| \leq 1$. Substituting the coefficients $c_j$ from~\eqref{eqn:cj}, to get \begin{equation} \label{eqn:inter} \|x^n-\phi_k(x)\|_\infty \leq \frac{1}{2^{n-1}}\sum_{\substack{j=k+1 \\ n-j \text{ even}}}^n \binom{n }{(n-j)/2}. \end{equation} Using the properties of the binomial coefficients, the summation simplifies as \[ \sum_{\substack{j=k+1 \\ n-j \text{ even}}}^n \binom{n }{(n-j)/2} = \sum_{\substack{j=k+1 \\ n+j \text{ even}}}^n \binom{n }{(n+j)/2} =\sum_{j=\lfloor (n+k)/2\rfloor+1}^n \binom{n }{j}. \] Plug this identity into~\eqref{eqn:inter} to get $\|x^n - \phi_k(x)\|_\infty \leq p_{n,k}$. The bound is clearly achieved at $x= 1$, where all the Chebyshev polynomials take the value $1$. \end{proof} This theorem shows that the polynomial approximation $\phi_k$ is nearly optimal, and the error due to this approximation is $p_{n,k}$. However, it is the optimal polynomial for the special case $k=n-1$. It is easy to see that $x^n - \phi_{n-1}(x) = 2^{1-n}T_n(x)$ and so $\phi_{n-1}$ is the same as the best polynomial approximation $q^*_{n-1}$ in~\eqref{eqn:minmax}. For $k < n-1$, from~\eqref{eqn:newmanrivlin} and Theorem~\ref{thm:main} \[ \frac{ \|x^n-\phi_k(x)\|_\infty}{4e} \leq \|x^n-\pi_k^*(x)\|_\infty \leq \|x^n-\phi_k(x)\|_\infty, \] so that the error in the Chebyshev polynomial approximation is suboptimal by at most the factor $4e \approx 10.87$. Therefore, by using $\phi_k$ we lose only one significant digit of accuracy compared to $\pi^*_k$. \section{A probabilistic digression}\label{sec:prob} In Section~\ref{sec:intro}, we saw that the error in the monomial approximation depends on $p_{n,k}$. Since $p_{n,k}$ depends on the sum of binomial coefficients, it has a probabilistic interpretation. Newman and Rivlin~\cite{newman1976approximation} observed that if a fair coin is tossed $n$ times, $p_{n,k}$ is the probability that the magnitude of the difference between the number of heads and the number of tails exceeds $k$. They used this insight along with the de Moivre-Laplace theorem~\cite[Section 1.3]{vershynin2018high} (which is a special case of the Central Limit Theorem) to obtain the approximation $p_{n,k} \approx 2 \text{erfc}(k/\sqrt{n})$. To convert this into a rigorous inequality for $p_{n,k}$ we use a different tool from probability, namely, concentration inequalities. The inequalities are useful in quantifying how much a random variable deviates from its mean. We start with the following alternative interpretation for $p_{n,k}$: it is twice the probability that greater than $\lfloor (n+k)/2 \rfloor$ coin tosses result in heads (or equivalently tails). We associate each coin toss with an independent Bernoulli random variable $X_i$ with parameter $p=1/2$ since the coin is fair. The random variable $X = \sum_{i=1}^nX_i$ has the Binomial distribution with parameters $n$ and $p$. Then, \[ p_{n,k} = 2\mathbb{P}\left( \lfloor (n+k)/2 \rfloor +1 \leq X \leq n\right) \leq 2\mathbb{P}\left(X \geq (n+k)/2 \right). \] Since $X$ has the Binomial distribution, we can once again use the de Moivre-Laplace theorem, to say that as $n\rightarrow \infty$, \[ \frac{X - np}{\sqrt{np(1-p)}} \longrightarrow \mathcal{N}(0,1), \qquad \text{in distribution}.\] Roughly speaking, this theorem says that $X$ behaves as a normal random variable with mean $n/2$ and variance $n/4$. Since the tails of normal distributions decay exponentially, we expect that $X$ lies in the range $\frac{n}{2} \pm 1.96 \sqrt{\frac{n}{4}}$ with nearly $95\%$ probability; alternatively, the probability that it is outside this range is very small. To make this more precise, we apply Hoeffding's concentration inequality~\cite[Theorem 2.2.6]{vershynin2018high}, to obtain \[ \mathbb{P}\left(X \geq (n+k)/2 \right) = \mathbb{P}\left(X - \mathbb{E}[X] \geq k/2 \right) \leq \exp\left(-\frac{k^2}{2n}\right). \] This gives our desired bound $p_{n,k} \leq 2\exp(-{k^2}/{2n})$. We can use a similar technique to prove the following result which may be of independent interest. If $0 \leq k \leq n/2$, then \[ \sum_{j=0}^k \binom{n}{j} \leq 2^n \exp\left( - \frac{(n-2k)^2}{2n}\right).\] Other concentration inequalities such as Chernoff and Bernstein (see~\cite[Chapter 2]{vershynin2018high}) also give equally interesting bounds. We invite the reader to explore such results. \section{Acknowledgements} The author would like to thank Alen Alexanderian, Ethan Dudley, Ivy Huang, Ilse Ipsen, and Nathan Reading for comments and feedback. The work was supported by the National Science Foundation through the grants DMS-1745654 and DMS-1845406. \section{Declaration of Interest} The author has no relevant financial or non-financial competing interests to report.
{ "timestamp": "2021-01-19T02:27:42", "yymm": "2101", "arxiv_id": "2101.06818", "language": "en", "url": "https://arxiv.org/abs/2101.06818", "abstract": "This paper considers the approximation of a monomial $x^n$ over the interval $[-1,1]$ by a lower-degree polynomial. This polynomial approximation can be easily computed analytically and is obtained by truncating the analytical Chebyshev series expansion of $x^n$. The error in the polynomial approximation in the supremum norm has an exact expression with an interesting probabilistic interpretation. We use this interpretation along with concentration inequalities to develop a useful upper bound for the error.", "subjects": "Numerical Analysis (math.NA)", "title": "Approximating monomials using Chebyshev polynomials", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9890130583409233, "lm_q2_score": 0.8104789063814616, "lm_q1q2_score": 0.8015742219211363 }
https://arxiv.org/abs/1712.00744
The quaternionic Gauss-Lucas Theorem
The classic Gauss-Lucas Theorem for complex polynomials of degree $d\ge2$ has a natural reformulation over quaternions, obtained via rotation around the real axis. We prove that such a reformulation is true only for $d=2$. We present a new quaternionic version of the Gauss-Lucas Theorem valid for all $d\geq2$, together with some consequences.
\section{Introduction} Let $p$ be a complex polynomial of degree $d\geq2$ and let $p'$ be its derivative. The Gauss-Lucas Theorem asserts that the zero set of $p'$ is contained in the convex hull $\mathcal K(p)$ of the zero set of $p$. The classic proof uses the logarithmic derivative of $p$ and it strongly depends on the commutativity of ${\mathbb{C}}$. This note deals with a quaternionic version of such a classic result. We refer the reader to \cite{GeStoSt2013} for the notions and properties concerning the algebra ${\mathbb{H}}$ of quaternions we need here. The ring ${\mathbb{H}}[X]$ of quaternionic polynomials is defined by fixing the position of the coefficients w.r.t.\ the indeterminate $X$ (e.g.\ on the right) and by imposing commutativity of $X$ with the coefficients when two polynomials are multiplied together (see e.g.\ \cite[\S 16]{Lam}). Given two polynomials $P,Q\in{\mathbb{H}}[X]$, let $P\cdot Q$ denote the product obtained in this way. If $P$ has real coefficients, then $(P\cdot Q)(x)=P(x)Q(x)$. In general, a direct computation (see \cite[\S 16.3]{Lam}) shows that if $P(x)\ne0$, then \begin{equation}\label{product} (P\cdot Q)(x)=P(x)Q(P(x)^{-1}xP(x)), \end{equation} while $(P\cdot Q)(x)=0$ if $P(x)=0$. In this setting, a {(left) root} of a polynomial $P(X)=\sum_{h=0}^dX^h a_h$ is an element $x\in{\mathbb{H}}$ such that $P(x)=\textstyle\sum_{h=0}^dx^h a_h=0$. Given $P(X)=\sum_{k=0}^dX^ka_k\in {\mathbb{H}}[X]$, consider the polynomial $P^c(X)=\sum_{k=0}^dX^k\bar a_k$ and the \emph{normal polynomial} $N(P)=P \cdot P^c=P^c\cdot P$. Since $N(P)$ has real coefficients, it can be identified with a polynomial in ${\mathbb{R}}[X]\subset{\mathbb{C}}[X]$. We recall that a subset $A$ of ${\mathbb{H}}$ is called \emph{circular} if, for each $x\in A$, $A$ contains the whole set (a 2-sphere if $x\not\in{\mathbb{R}}$, a point if $x\in{\mathbb{R}}$) \begin{equation}\label{sx} {\mathbb{S}}_x=\{pxp^{-1}\in{\mathbb{H}}\;|\;p\in{\mathbb{H}}^*\}, \end{equation} where ${\mathbb{H}}^*:={\mathbb{H}}\setminus\{0\}$. In particular, for any imaginary unit $I\in{\mathbb{H}}$, ${\mathbb{S}}_I={\mathbb{S}}$ is the 2-sphere of all imaginary units in ${\mathbb{H}}$. For every subset $B$ of ${\mathbb{H}}$ we define its \emph{circularization} as the set $\bigcup_{x\in B}{\mathbb{S}}_x$. It is well-known (\cite[\S3.3]{GeStoSt2013}) that the zero set $V(N(P))\subset{\mathbb{H}}$ of the normal polynomial is the circularization of the zero set $V(P)$, which consists of isolated points or isolated 2-spheres of the form \eqref{sx} if $P\neq0$. Let the degree $d$ of $P$ be at least 2 and let $P'(X)=\sum_{k=1}^dX^{k-1}ka_k$ be the derivative of $P$. It is known (see e.g.\ \cite{GGSarxiv}) that the Gauss-Lucas Theorem does not hold directly for quaternionic polynomials. For example, the polynomial $P(X)=(X-i)\cdot (X-j)=X^2-X(i+j)+k$ has zero set $V(P)=\{i\}$, while $P'$ vanishes at $x=(i+j)/2$. Since the zero set $V(P $ of $P$ is contained in the set $V(N(P) $, a natural reformulation in ${\mathbb{H}}[X]$ of the classic Gauss-Lucas Theorem is the following: $V(N(P'))\subset\mathcal K(N(P))$ or equivalently \begin{equation}\label{eq:g-l} V(P')\subset\mathcal K(N(P)), \end{equation} where $\mathcal K(N(P))$ denotes the convex hull of $V(N(P))$ in ${\mathbb{H}}$. This set is equal to the circularization of the convex hull of the zero set of $N(P)$ viewed as a polynomial in ${\mathbb{C}}[X]\subset{\mathbb{H}}[X]$. Recently two proofs of the above inclusion \eqref{eq:g-l} were presented in \cite{VlacciGL,GGSarxiv}. Our next two propositions prove that inclusion \eqref{eq:g-l} is correct in its full generality only when $d=2$. \section{Gauss-Lucas polynomials} \begin{definition} Given a polynomial $P\in{\mathbb{H}}[X]$ of degree $d\geq 2$ we say that $P$ is a \emph{Gauss-Lucas polynomial} if $P$ satisfies \eqref{eq:g-l}. \end{definition} \begin{proposition} If $P$ is a polynomial in ${\mathbb{H}}[X]$ of degree $2$, then $V(N(P))={\mathbb{S}}_{x_1}\cup{\mathbb{S}}_{x_2}$ for some $x_1,x_2\in{\mathbb{H}}$ (possibly with ${\mathbb{S}}_{x_1}={\mathbb{S}}_{x_2}$) and \[ V(P')\subset\bigcup_{y_1\in{\mathbb{S}}_{x_1},y_2\in{\mathbb{S}}_{x_2}}\left\{\frac{y_1+y_2}{2}\right\}. \] In particular every polynomial $P\in{\mathbb{H}}[X]$ of degree $2$ is a Gauss-Lucas polynomial. \end{proposition} \begin{proof} Let $P(X)=X^2a_2+Xa_1+a_0\in{\mathbb{H}}[X]$ with $a_2\neq0$. Since $P\cdot a_2^{-1}=Pa_2^{-1}$, $(P\cdot a_2^{-1})'=P'\cdot a_2^{-1}=P'a_2^{-1}$, we can assume $a_2=1$. Consequently, $P(X)=(X-x_1)\cdot(X-x_2)=X^2-X(x_1+x_2)+x_1x_2$ for some $x_1,x_2\in{\mathbb{H}}$. Then $x_1\in V(P)$ and $\bar x_2\in V(P^c)$, since $P^c(X)=(X-\bar x_2)\cdot(X-\bar x_1)$. Therefore $x_1,x_2\in V(N(P))$. On the other hand $V(P')=\{(x_1+x_2)/2\}$ as desired. \end{proof} \begin{remark} Let $P(X)=\sum_{k=0}^dX^ka_k\in{\mathbb{H}}[X]$ of degree $d\geq2$ and let $Q:=P\cdot a_d^{-1}$ be the corresponding monic polynomial. Since $V(P)=V(Q)$ and $V(P')=V(Q')$, $P$ is a Gauss-Lucas polynomial if and only if $Q$ is. \end{remark} \begin{proposition} \label{prop:-4} Let $P\in{\mathbb{H}}[X]$ of degree $d\geq3$. Suppose that $N(P)(X)=X^{2e}\cdot(X^2+1)^{d-e}$ for some $e<d$ and that $N(P')(X)=\sum_{k=0}^{2d-2}X^kb_k$ contains a unique monomial of odd degree, that is, $b_k\neq0$ for a unique odd $k$. Then $P$ is not a Gauss-Lucas polynomial. \end{proposition} \begin{proof} Since $V(N(P))\subset\{0\}\cup{\mathbb{S}}$, $\mathcal K(N(P))\subset\im({\mathbb{H}})=\{x\in{\mathbb{H}}\,|\,\re(x)=0\}$. Then it suffices to show that $N(P')$ has at least one root in ${\mathbb{H}}\setminus\im({\mathbb{H}})$. Let $G,L\in{\mathbb{R}}[X]$ be the unique real polynomials such that $G(t)+iL(t)=N(P')(it)$ for every $t\in{\mathbb{R}}$. Since $N(P')$ contains a unique monomial of odd degree, say $X^{2\ell+1}b_{2\ell+1}$, then $L(t)=(-1)^\ell b_{2\ell+1}t^{2\ell+1}$ and hence $V(N(P'))\cap i{\mathbb{R}}\subset\{0\}$. Being $V(N(P'))$ a circular set, it holds $V(N(P'))\cap\im({\mathbb{H}})\subset\{0\}$. Since $N(P')$ contains at least two monomials, namely those of degrees $2\ell+1$ and $2d-2$, we infer that it must have a nonzero root. Therefore $V(N(P'))\not\subset\{0\}$ and hence $V(N(P'))\not\subset\im({\mathbb{H}})$, as desired. \end{proof} \begin{corollary}\label{counterexample} Let $d\geq3$ and let \[ P(X)=X^{d-3}\cdot(X-i)\cdot(X-j)\cdot(X-k). \] Then $N(P)(X)=X^{2d-6}\cdot(X^2+1)^3$ and $N(P')$ contains a unique monomial of odd degree, namely $-4X^{2d-5}$. In particular $P$ is not a Gauss-Lucas polynomial. \end{corollary} \begin{proof} By a direct computation we obtain: \begin{align} P(X)&=X^d-X^{d-1}(i+j+k)+X^{d-2}(i-j+k)+X^{d-3},\\ P'(X)&=dX^{d-1}-(d-1)X^{d-2}(i+j+k)+(d-2)X^{d-3}(i-j+k)+(d-3)X^{d-4},\label{P1} \\ N(P')(X)&=d^2X^{2d-2}+3(d-1)^2X^{2d-4}-4X^{2d-5}+3(d-2)^2X^{2d-6}+(d-3)^2X^{2d-8}. \end{align} Proposition \ref{prop:-4} implies the thesis. \end{proof} Let $I\in{\mathbb{S}}$ and let ${\mathbb{C}}_I\subset{\mathbb{H}}$ be the complex plane generated by 1 and $I$. Given a polynomial $P\in{\mathbb{H}}[X]$, we will denote by $P_I:{\mathbb{C}}_I\to{\mathbb{H}}$ the restriction of $P$ to ${\mathbb{C}}_I$. {\it If $P_I$ is not constant, we will denote by $\mathcal K_{{\mathbb{C}}_I}(P)$ the convex hull in the complex plane ${\mathbb{C}}_I$ of the zero set $V(P_I)=V(P)\cap{\mathbb{C}}_I$. If $P_I$ is constant, we set $\mathcal K_{{\mathbb{C}}_I}(P)={\mathbb{C}}_I$.} If ${\mathbb{C}}_I$ contains every coefficient of $P\in{\mathbb{H}}[X]$, then we say that $P$ is a \emph{${\mathbb{C}}_I$-polynomial}. \begin{proposition} The following holds: \begin{enumerate} \item[(1)] Every ${\mathbb{C}}_I$-polynomial of degree $\geq2$ is a Gauss-Lucas polynomial. \item[(2)] Let $d\geq3$, let ${\mathbb{H}}_d[X]=\{P\in{\mathbb{H}}[X]\,|\,\deg(P)=d\}$ and let $E_d[X]$ be the set of all elements of ${\mathbb{H}}_d[X]$ that are not Gauss-Lucas polynomials. Identify each $P(X)=\sum_{k=0}^dX^ka_k$ in ${\mathbb{H}}_d[X]$ with $(a_0,\ldots,a_d)\in{\mathbb{H}}^d\times{\mathbb{H}}^*\subset{\mathbb{R}}^{4d+4}$ and endow ${\mathbb{H}}_d[X]$ with the relative Euclidean topology. Then $E_d[X]$ is a nonempty open subset of ${\mathbb{H}}_d[X]$. Moreover $E_d[X]$ is not dense in ${\mathbb{H}}_d[X]$, being $X^d-1$ an interior point of its complement. \end{enumerate} \end{proposition} \begin{proof} If $P$ is a ${\mathbb{C}}_I$-polynomial, then $P_I$ can be identified with an element of ${\mathbb{C}}_I[X]$. Consequently, the classic Gauss-Lucas Theorem gives $V(P')\cap{\mathbb{C}}_I=V(P_I')\subset\mathcal K_{{\mathbb{C}}_I}(P)$. The zero set of the ${\mathbb{C}}_I$-polynomial $P'$ has a particular structure (see \cite[Lemma~3.2]{GeStoSt2013}): $V(P')$ is the union of $V(P')\cap{\mathbb{C}}_I$ with the set of spheres ${\mathbb{S}}_x$ such that $x,\bar x\in V(P'_I)$. It follows that \[ V(P')\subset\mathcal K(N(P)). \] This proves (1). Now we prove (2). By Corollary \ref{counterexample} we know that $E_d[X]\neq\emptyset$. If $P\in E_d[X]$, then $V(N(P'))\not\subset\mathcal K(N(P))$. $N(P)$ and $N(P')$ are polynomials with real coefficients. Since the roots of $N(P)$ and of $N(P')$ depend continuously on the coefficients of $P$ and $\mathcal K(N(P))$ is closed in ${\mathbb{H}}$, for every $Q\in{\mathbb{H}}_d[X]$ sufficiently close to $P$, $V(N(Q'))$ is not contained in $\mathcal K(N(Q))$, that is $Q\in E_d[X]$. To prove the last statement, observe that $P(X)=X^d-1$ is not in $E_d[X]$ from part (1). Since $V(P')=V(N(P'))=\{0\}$ is contained in the interior of the set $\mathcal K(N(P))$, for every $Q\in{\mathbb{H}}_d[X]$ sufficiently close to $P$, $V(Q')$ is still contained in $\mathcal K(N(Q))$. \end{proof} \section{A quaternionic Gauss-Lucas Theorem} Let $P(X)=\sum_{k=0}^dX^ka_k\in{\mathbb{H}}[X]$ of degree $d\geq2$. For every $I\in{\mathbb{S}}$, let $\pi_I:{\mathbb{H}}\to{\mathbb{H}}$ be the orthogonal projection onto ${\mathbb{C}}_I$ and $\pi_I^\bot=id-\pi_I$. Let $P^I(X):=\sum_{k=1}^dX^ka_{k,I}$ be the ${\mathbb{C}}_I$-polynomial with coefficients $a_{k,I}:=\pi_I(a_k)$. \begin{definition} We define the \emph{Gauss-Lucas snail of $P$} as the following subset $\mk{sn}(P)$ of ${\mathbb{H}}$: \[ \mk{sn}(P):=\bigcup_{I\in{\mathbb{S}}}\mathcal K_{{\mathbb{C}}_I}(P^I). \] \end{definition} Our quaternionic version of the Gauss-Lucas Theorem reads as follows. \begin{theorem}\label{thm} For every polynomial $P\in{\mathbb{H}}[X]$ of degree $\geq2$, \begin{equation}\label{snail} V(P')\subset\mk{sn}(P). \end{equation} \end{theorem} \begin{proof} Let $P(X)=\sum_{k=0}^dX^ka_k$ in ${\mathbb{H}}_d[X]$ with $d\geq2$. We can decompose the restriction of $P$ to ${\mathbb{C}}_I$ as $P_I=\pi_I\circ P_I+\pi_I^\bot\circ P_I={P^I}_{|{\mathbb{C}}_I}+\pi_I^\bot\circ P_I$. If $x\in{\mathbb{C}}_I$, then $P^I(x)\in{\mathbb{C}}_I$ while $(\pi_I^\bot\circ P_I)(x)\in{\mathbb{C}}_I^\bot$. The same decomposition holds for $P'$. This implies that $V(P')\cap{\mathbb{C}}_I\subset V((P^I)')\cap{\mathbb{C}}_I$. The classic Gauss-Lucas Theorem applied to $P^I$ on ${\mathbb{C}}_I$ gives $V(P')\cap{\mathbb{C}}_I\subset \mathcal K_{{\mathbb{C}}_I}(P^I)$. Since $V(P')=\bigcup_{I\in{\mathbb{S}}}(V(P')\cap{\mathbb{C}}_I)$, the inclusion \eqref{snail} is proved. \end{proof} If $P$ is monic Theorem \ref{thm} has the following equivalent formulation: {\it For every monic polynomial $P\in{\mathbb{H}}[X]$ of degree $\geq2$, it holds} \begin{equation} \mk{sn}(P')\subset\mk{sn}(P). \end{equation} \begin{remark}\label{rem:monic} If $P$ is a nonconstant monic polynomial in ${\mathbb{H}}[X]$, then two properties hold: \begin{itemize} \item[(a)] $\mathcal K_{{\mathbb{C}}_I}(P^I)$ is a compact subset of ${\mathbb{C}}_I$ for every $I\in{\mathbb{S}}$. \item[(b)] $\mathcal K_{{\mathbb{C}}_I}(P^I)$ depends continuously on $I$. \end{itemize} Let $I\in{\mathbb{S}}$. Since $P$ is monic, also $P^I$ is a monic, nonconstant polynomial and then $\mathcal K_{{\mathbb{C}}_I}(P^I)$ is a compact subset of ${\mathbb{C}}_I$. This prove property (a). To see that (b) holds, one can apply the Continuity theorem for monic polynomials (see e.g.~\cite[Theorem~1.3.1]{RahmanSchmeisser}). The roots of $P^I$ depend continuously on the coefficients of $P^I$, which in turn depend continuously on $I$. Therefore the convex hull $\mathcal K_{{\mathbb{C}}_I}(P^I)$ depends continuously on $I$. Observe that (a) and (b) can not hold for polynomials that are not monic. For example, let $P(X)=X^2i$. Then, given $I=\alpha_1i+\alpha_2j+\alpha_3k\in{\mathbb{S}}$, $\mathcal K_{{\mathbb{C}}_I}(P^I)=\{0\}$ if $\alpha_1\ne0$ and $\mathcal K_{{\mathbb{C}}_I}(P^I)={\mathbb{C}}_I$ if $\alpha_1=0$, since in this case $P^I$ is constant. \end{remark} \begin{remark}\label{rem:closed} If $P$ is a monic polynomial in ${\mathbb{H}}[X]$ of degree $\geq2$, then its Gauss-Lucas snail is a closed subset of ${\mathbb{H}}$. To prove this fact, consider $q\in{\mathbb{H}}\setminus\mk{sn}(P)$ and choose $I\in{\mathbb{S}}$ such that $q\in{\mathbb{C}}_I$. Write $q=\alpha+I\beta\in{\mathbb{C}}_I$ for some $\alpha,\beta\in{\mathbb{R}}$ and define $z:=\alpha+i\beta\in{\mathbb{C}}$. Since $P$ is monic, $\mk{sn}(P)\cap{\mathbb{C}}_I=\mathcal K_{{\mathbb{C}}_I}(P^I)$ is a compact subset of ${\mathbb{C}}_I$. Moreover, $\mathcal K_{{\mathbb{C}}_I}(P^I)$ depends continuously on $I$, and then there exist an open neighborhood $U_I$ of $z$ in ${\mathbb{C}}$ and an open neighborhood $W_I$ of $I$ in ${\mathbb{S}}$ such that the set \[ \textstyle [U_I,W_I]:=\bigcup_{J\in W_I}\{a+Jb\in{\mathbb{C}}_J\;|\; a+ib\in U_I\} \] is an open neighborhood of $q$ in $\bigcup_{J\in W_I}{\mathbb{C}}_J$, and it is disjoint from $\mk{sn}(P)$. If $q\not\in{\mathbb{R}}$ then $q$ is an interior point of ${\mathbb{H}}\setminus\mk{sn}(P)$, because $[U_I,W_I]$ is a neighborhood of $q$ in ${\mathbb{H}}$ as well. Now assume that $q\in{\mathbb{R}}$. Since ${\mathbb{S}}$ is compact there exist $I_1,\ldots,I_n\in{\mathbb{S}}$ such that $\bigcup_{\ell=1}^nW_{I_\ell}={\mathbb{S}}$. It follows that $[\bigcap_{\ell=1}^nU_\ell,{\mathbb{S}}]$ is a neighborhood of $q$ in ${\mathbb{H}}$, which is disjoint from $\mk{sn}(P)$. Consequently $q$ is an interior point of ${\mathbb{H}}\setminus\mk{sn}(P)$ also in this case. This proves that $\mk{sn}(P)$ is closed in ${\mathbb{H}}$. In Proposition \ref{pro:sn-compact} below we will show that the Gauss-Lucas snail of a monic polynomial in ${\mathbb{H}}[X]$ of degree $\geq2$ is also a compact subset of ${\mathbb{H}}$. \end{remark} If all the coefficients of $P$ are real, then $\mk{sn}(P)$ is a circular set. In general, $\mk{sn}(P)$ is neither closed nor bounded nor circular, as shown in the next example. \begin{example} Let $P(X)=X^2i+X$. Given $I=\alpha_1i+\alpha_2j+\alpha_3k\in{\mathbb{S}}$, $P^I(x)=X^2I\alpha_1+X$ and then $\mathcal K_{{\mathbb{C}}_I}(P^I)=\{0\}$ if $\alpha_1=0$ while $\mathcal K_{{\mathbb{C}}_I}(P^I)$ is the segment from 0 to $I{\alpha_1}^{-1}$ if $\alpha_1\ne0$. It follows that $\mk{sn}(P)=\{x_1i+x_2j+x_3k\in\im({\mathbb{H}})\,|\,0<x_1\le1\}\cup\{0\}$. Finally, observe that the monic polynomial $Q(X)=-P(X)\cdot~i=X^2-Xi$ corresponding to $P$ has compact Gauss-Lucas snail $\mk{sn}(Q)=\{x_1i+x_2j+x_3k\in\im({\mathbb{H}})\,|\,(x_1-1/2)^2+x_2^2+x_3^2\le 1/4\}$. \end{example} \begin{remark} Even for ${\mathbb{C}}_I$-polynomials, the Gauss-Lucas snail of $P$ can be strictly smaller than the circular convex hull $\mathcal K(N(P))$. For example, consider the ${\mathbb{C}}_i$-polynomial $P(X)=X^3+3X+2i$, with zero sets $V(P)=\{-i,2i\}$ and $V(P')={\mathbb{S}}$. The set $\mathcal K(N(P))$ is the closed three-dimensional disc in $\im({\mathbb{H}})$, with center at the origin and radius 2. The Gauss-Lucas snail $\mk{sn}(P)$ is the subset of $\im({\mathbb{H}})$ obtained by rotating around the $i$-axis the following subset of the coordinate plane $L=\{x=x_1i+x_2j\in\im({\mathbb{H}})\,|\,x_1,x_2\in{\mathbb{R}}\}$: \[\{x=\rho\cos(\theta)i+\rho\sin(\theta)j\in L\;|\;0\le\theta\le\pi,\,0\le\rho\le2\cos(\theta/3)\}.\] Therefore $\mk{sn}(P)$ is a proper subset of $\mathcal K(N(P))$ (the boundaries of the two sets intersect only at the point $2i$). Its boundary is obtained by rotating a curve that is part of the \emph{lima\c con trisectrix} (see Figure \ref{fig:limacon}). \end{remark} \begin{figure} \begin{center} \includegraphics[width=6cm]{GaussLucas.png}\\ \caption{Cross-sections of $\mk{sn}(P)$ (gray), of $V(P')$ and $\mathcal K(N(P))$ (dashed). \label{fig:limacon} \end{center} \end{figure} \subsection{Estimates on the norm of the critical points} Let $p(z)=\sum_{k=0}^da_kz^k$ be a complex polynomial of degree $d\ge1$. The norm of the roots of $p$ can be estimated making use of the norm of the coefficients $\{a_k\}_{k=0}^d$ of $p$. There are several classic results in this direction (see e.g.\ \cite[\S8.1]{RahmanSchmeisser}). For instance the estimate \cite[(8.1.2)]{RahmanSchmeisser} (with $\lambda=1,p=2$) asserts that \begin{equation}\label{eq:cauchy} \textstyle \max_{z\in V(p)}|z|\leq|a_d|^{-1}\sqrt{\sum_{k=0}^d|a_k|^2}\;. \end{equation} \begin{proposition}\label{pro:sn-compact} For every monic polynomial $P\in{\mathbb{H}}[X]$ of degree $d\geq2$, the Gauss-Lucas snail $\mk{sn}(P)$ is a compact subset of ${\mathbb{H}}$. \end{proposition} \begin{proof} Since $P=\sum_{k=0}^dX^ka_k$ is monic, every polynomial $P^I$ is monic. From \eqref{eq:cauchy} it follows that $\max_{x\in V(P^I)}|x|^2\le\sum_{k=0}^d|\pi_I(a_k)|^2\le\sum_{k=0}^d|a_k|^2$ and hence $\mk{sn}(P)\subset \{x\in{\mathbb{H}}\,|\, |x|^2\le \sum_{k=0}^d|a_k|^2\}$ is bounded. Since $\mk{sn}(P)$ is closed in ${\mathbb{H}}$, as seen in Remark \ref{rem:closed}, it is also a compact subset of ${\mathbb{H}}$. \end{proof} Define a function $C:{\mathbb{H}}[X]\to{\mathbb{R}}\cup\{+\infty\}$ as follows: $C(a):=+\infty$ if $a$ is a quaternionic constant and \[ \textstyle C(P):=|a_d|^{-1}\sqrt{\sum_{k=0}^d|a_k|^2} \qquad \text{if $P(X)=\sum_{k=0}^dX^ka_k$ with $d\ge1$ and $a_d\neq0$.} \] \begin{proposition}\label{pro:estimate} For every polynomial $P\in{\mathbb{H}}[X]$ of degree $d\geq1$, it holds \begin{equation}\label{eq:C} \max_{x\in V(P)}|x|\leq C(P). \end{equation} \end{proposition} \begin{proof} We follow the lines of the proof of estimate \eqref{eq:cauchy} for complex polynomials given in \cite{RahmanSchmeisser}. Let $P(X)=\sum_{k=0}^dX^ka_k$ with $d\ge1$ and $a_d\neq0$. We can assume that $P(X)$ is not the monomial $X^d a_d$, since in this case the thesis is immediate. Let $b_k=|a_ka_d^{-1}|$ for every $k=0,\ldots,d-1$. The real polynomial $h(z)=z^d-\sum_{k=0}^{d-1}b_kz^k$ has exactly one positive root $\rho$ and is positive for real $z>\rho$ (see \cite[Lemma~8.1.1]{RahmanSchmeisser}). Let $S:=\sum_{k=0}^{d-1}b_k^2=C(P)^2-1$. From the Cauchy-Schwartz inequality, it follows that \[\left(\sum_{k=0}^{d-1}b_kC(P)^k\right)^2\le S\sum_{k=0}^{d-1}C(P)^{2k}= (C(P)^2-1)\frac{C(P)^{2d}-1}{C(P)^2-1}<C(P)^{2d}. \] Therefore $h(C(P))>0$ and then $C(P)>\rho$. Let $x\in V(P)$. It remains to prove that $|x|\le\rho$. Since $x^d=-\sum_{k=0}^{d-1}x^ka_ka_d^{-1}$, it holds \[|x|^d\le\sum_{k=0}^{d-1}|x|^k\left|a_ka_d^{-1}\right|=\sum_{k=0}^{d-1}|x|^kb_k.\] This means that $h(|x|)\le0$, which implies $|x|\le\rho$. \end{proof} From Proposition~\ref{pro:estimate} it follows that for every polynomial $P\in{\mathbb{H}}[X]$ of degree $d\geq2$, it holds \begin{equation}\label{eq:C} \max_{x\in V(P')}|x|\leq C(P'). \end{equation} Theorem \ref{thm} allows to obtain a new estimate. \begin{proposition} Given any polynomial $P\in{\mathbb{H}}[X]$ of degree $d\geq2$, it holds: \begin{equation}\label{eq:ijk'} \max_{x\in V(P')}|x|\leq\sup_{I\in{\mathbb{S}}}\{C(P^I)\}. \end{equation} \end{proposition} \begin{proof} If $x\in V(P')\cap{\mathbb{C}}_I$, Theorem \ref{thm} implies that $x\in\mathcal K_{{\mathbb{C}}_I}(P^I)$. Therefore \begin{equation*}\label{eq:ijk} \max_{x\in V(P')\cap{\mathbb{C}}_I}|x|\leq C(P^I) \quad \text{ for every $I\in{\mathbb{S}}$ with $V(P')\cap{\mathbb{C}}_I\ne\emptyset$}, \end{equation*} from which inequality \eqref{eq:ijk'} follows. \end{proof} Our estimate \eqref{eq:ijk'} can be strictly better than classic estimate \eqref{eq:C}, as explained below. \begin{remark} Let $d\geq3$ and let $P(X)=X^{d-3}\cdot(X-i)\cdot(X-j)\cdot(X-k)$. Using \eqref{P1}, by a direct computation we obtain: \[ C(P')=d^{-1}\sqrt{8d^2-24d+24}. \] Moreover, given $I=\alpha_1i+\alpha_2j+\alpha_3k\in{\mathbb{S}}$ for some $\alpha_1,\alpha_2,\alpha_3\in{\mathbb{R}}$ with $\alpha_1^2+\alpha_2^2+\alpha_3^2=1$, we have $ \pi_I(i+j+k)=\langle I,i+j+k\rangle I=(\alpha_1+\alpha_2+\alpha_3)I$ and $\pi_I(i-j+k)=\langle I,i-j+k\rangle I=(\alpha_1-\alpha_2+\alpha_3)I$ and hence \[ C(P^I)=\sqrt{4+4\alpha_1\alpha_3}\leq\sqrt{4+2(\alpha_1^2+\alpha_3^2)}\leq\sqrt{6}. \] This implies that \[ \sup_{I\in{\mathbb{S}}}\{C(P^I)\}\leq\sqrt{6}. \] For every $d\geq11$ it is easy to verify that $\sqrt{6}<C(P')$ so \[ \sup_{I\in{\mathbb{S}}}\{C(P^I)\}<C(P'), \] as announced. \end{remark} \begin{remark} Some of the results presented here can be generalized to real alternative *-algebras, a setting in which polynomials can be defined and share many of the properties valid on the quaternions (see \cite{GhPe_AIM}). The polynomials given in Corollary \ref{counterexample} can be defined every time the algebra contains an Hamiltonian triple $i,j,k$. This property is equivalent to say that the algebra contains ${\mathbb{H}}$ as a subalgebra (see \cite[\S8.1]{Numbers}). For example, this is true for the algebra of octonions and for the Clifford algebras with signature $(0,n)$, with $n\ge2$. Therefore in all such algebras there exist polynomials for which the zero set $V(P')$ (as a subset of the \emph{quadratic cone}) is not included in the circularization of the convex hull of $V(N(P))$ viewed as a complex polynomial. \end{remark}
{ "timestamp": "2018-03-05T02:12:34", "yymm": "1712", "arxiv_id": "1712.00744", "language": "en", "url": "https://arxiv.org/abs/1712.00744", "abstract": "The classic Gauss-Lucas Theorem for complex polynomials of degree $d\\ge2$ has a natural reformulation over quaternions, obtained via rotation around the real axis. We prove that such a reformulation is true only for $d=2$. We present a new quaternionic version of the Gauss-Lucas Theorem valid for all $d\\geq2$, together with some consequences.", "subjects": "Complex Variables (math.CV)", "title": "The quaternionic Gauss-Lucas Theorem", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9830850906978876, "lm_q2_score": 0.8152324848629214, "lm_q1q2_score": 0.8014429013213294 }
https://arxiv.org/abs/1812.11519
Two-scale methods for convex envelopes
We develop two-scale methods for computing the convex envelope of a continuous function over a convex domain in any dimension.This hinges on a fully nonlinear obstacle formulation [A. M. Oberman, "The convex envelope is the solution of a nonlinear obstacle problem", Proc. Amer. Math. Soc. 135(6):1689--1694, 2007]. We prove convergence and error estimates in the max norm. The proof utilizes a discrete comparison principle, a discrete barrier argument to deal with Dirichlet boundary values, and the property of flatness in one direction within the non-contact set. Our error analysis extends to a modified version of the finite difference wide stencil method of [A. M. Oberman, "Computing the convex envelope using a nonlinear partial differential equation", Math. Models Meth. Appl. Sci, 18(05):759--780, 2008].
\section{Introduction} Given an open set $\Omega \subset \mathbb{R}^d$ and a continuous function $f: \overline{\Omega} \rightarrow \mathbb{R}$, its convex envelop in $\Omega$ is defined as \begin{equation}\label{E:def-CE} u(x) = \sup\left\{ l(x): l \leq f \text{ in } \overline{\Omega}, \; l \text{ is affine} \right\}, \end{equation} which in fact is the largest convex function majorized by $f$ in $\overline{\Omega}$. This function $u$ can also be viewed as the viscosity solution of the following fully nonlinear, degenerate elliptic PDE introduced by Oberman \cite{Ob1} \begin{equation}\label{E:pde-int-CE} T[u;f](x) := \min\left\{ f(x) - u(x), \lambda_1[D^2u](x)\right\} = 0, \end{equation} where $\lambda_1[D^2u]$ denotes the smallest eigenvalue of the Hessian $D^2u$. This is the complementarity form of the fully nonlinear obstacle problem at hand. \begin{figure}[!htb] \vspace{-10pt} \includegraphics[width=0.8\textwidth]{figures/pde_CE.jpg} \vspace{-20pt} \caption{\small Illustration of the equation \eqref{E:pde-int-CE}. In the non-contact set $\{u < f\}$, the function $u$ must be flat in one direction, i.e. $\lambda_1[D^2u] = 0$.} \label{F:pde_CE} \end{figure} \Cref{F:pde_CE} illustrates the pde formulation \eqref{E:pde-int-CE}. Roughly speaking, in the contact set \[ \mathcal{C}(f) := \left\{ x \in \overline{\Omega}: u(x) = f(x) \right\}, \] we have the equality $u = f$ and the inequality $\lambda_1[D^2u] \geq 0$ given by the convexity of $u$. Outside the contact set, we have $u < f$ and that $u$ is flat in at least one direction which implies $\lambda_1[D^2u] = 0$. In this paper, we consider the case $\Omega$ bounded and strictly convex, which guarantees the Dirichlet boundary condition $u = f$ on $\partial\Omega$ is attained. Therefore the convex envelope $u$ of $f$ is the viscosity solution of the following problem: \begin{equation} \label{E:pde-CE} \left\{ \begin{aligned} T[u;f](x) = \min\left\{ f(x) - u(x), \lambda_1[D^2u](x)\right\} = 0 & \quad {\rm in} \;\; \Omega, \\ u =f & \quad {\rm on} \;\; \partial\Omega. \end{aligned} \right. \end{equation} The regularity study of convex envelopes dates back to \cite{TrudUrbas1984, CaNiSp}, thus before the PDE formulation \eqref{E:pde-CE} of \cite{Ob1}. However, the problem considered in \cite{TrudUrbas1984, CaNiSp} is a Dirichlet problem for the degenerate Monge-Amp\`{e}re equation, $\det(D^2u)=0$, which corresponds to the convex envelope of function $f$ given on the boundary $\partial\Omega$ as a Dirichlet condition. For the convex envelope $u$ in \eqref{E:def-CE}, De Philippis and Figalli \cite{DeFi} obtained recently the optimal regularity $u \in C^{1,1}(\overline{\Omega})$ under the assumption that $\Omega$ is a uniformly convex domain of class $C^{3,1}$ and $f \in C^{3,1}(\overline{\Omega})$. There are a handful of papers regarding the numerical approximation of convex envelopes. Oberman \cite{Ob2} proposed a wide stencil method to approximate \eqref{E:pde-int-CE}. Dolzmann \cite{Dolzmann} developed a method to compute rank-one convex envelopes, a related notion of critical importance in materials science. Dolzmann and Walkington \cite{DolzWalk} proved an $O(h^{1/3})$ rate of convergence. Finally, Bartels \cite{Bartels} improved the error estimate of \cite{DolzWalk} to $O(h)$ upon increasing the number of directions and function evaluations within elements, thus at the expense of extra computational cost. In this paper, we construct and study a two-scale method for \eqref{E:pde-CE}, which is somewhat related to the wide stencil method of \cite{Ob2}. Two-scale methods are developed in \cite{NoNtZh1}, whereas suboptimal pointwise error estimates are derived in \cite{NoNtZh2} and optimal ones in \cite{LiNo}. We prove existence, uniqueness, and uniform convergence, as well as pointwise error estimates under realistic regularity assumptions on $u$. Our proof hinges on a discrete comparison principle and discrete barrier functions, and is thus classical. However, we exploit that $u$ is flat in at least one direction outside the contact set $\mathcal{C}(f)$ \cite{CaNiSp, ObSi}, a crucial property that plays an essential role in dealing with low regularity of $u$. Our techniques extend to a modified wide stencil method obtained from that in \cite{Ob2} upon adding a two-scale structure. The remainder of this paper is organized as follows. In section 2, we introduce the two-scale method for convex envelope problem \eqref{E:pde-CE} and prove several properties of it. In section 3, we prove our main error estimate in the $L^{\infty}$ norm after reviewing geometric properties of $u$ and studying the consistency error. We next extend our analysis to a modified wide stencil method in section 4. We conclude in section 5 with numerical experiments which illustrate the performance of the two-scale methods and compare with theory. \section{Two-Scale Method} \label{S:TwoSc} In this section, we extend the two-scale method developed in \cite{NoNtZh1} to solve \eqref{E:pde-CE}, and prove several important properties including convergence. \subsection{Definition of the Two-Scale Method} \label{S:IntroTwoSc} Let $\{\mathcal{T}_h\}$ be a sequence of meshes made of closed simplices $T$. Let $\mathcal{T}_h$ be shape-regular and quasi-uniform with mesh size $h$ and shape-regular constant $\sigma$, i.e. \begin{equation}\label{E:shape-regularity} \max_{h} \ \max_{T \in \mathcal{T}_h} \ \frac{h_T}{\rho_T} \le \sigma, \end{equation} where $h_T$ denotes the diameter of $T$ and $\rho_T$ the diameter of the largest ball inscribed in $T$. Let $\Omega_h$ be the interior of the union of elements $T\in\mathcal{T}_h$, $\mathcal{N}_h$ be the nodes of $\mathcal{T}_h$, $\mathcal{N}_h^b := \{x_i \in \mathcal{N}_h: x_i \in \partial \Omega\}$ be the boundary nodes and $\mathcal{N}_h^0 := \mathcal{N}_h \setminus \mathcal{N}_h^b$ be the interior nodes; since we require that $\mathcal{N}_h^b \subset \partial \Omega$ we deduce that $\Omega_h\subset\Omega$ is also convex. Let $\mathbb{V}_h$ be the space of continuous piecewise linear functions over $\mathcal{T}_h$. Before introducing the two-scale method we need additional notation. Let $\mathbb{S}$ be the unit sphere in $\mathbb{R}^d$. We consider a finite discretization $\mathbb S_{\theta} \subset \mathbb{S}$ of $\mathbb{S}$ governed by the parameter $\theta$: given any $v \in \mathbb{S}$, there exists $v^\theta \in \mathbb S_{\theta}$ such that \begin{equation*} |v - v^{\theta} | \leq \theta. \end{equation*} Let the meshsize $h$ be the fine scale and $\delta \geq h$ (to be chosen later) be the coarse scale. For every $x_i\in\mathcal{N}_h^0$, let \begin{equation}\label{E:deltai} \delta_i := \min\big\{\delta,\textrm{dist}(x_i,\partial\Omega_h)\big\}, \end{equation} and observe that $\delta_i\ge C(\sigma) h$ and the open ball $B(x_i,\delta_i)$ centered at $x_i$ with radius $\delta_i$ is contained in $\Omega_h$. For any function $w \in C(\overline{\Omega}_h)$, in particular for $w \in \mathbb{V}_h$, let the centered second difference operator be \begin{equation} \label{E:2Sc2Dif} \sdd w(x_i;v) := \frac{ w(x_i+ \delta_i v) -2 w(x_i) + w(x_i- \delta_i v) }{ \delta_i^2} \end{equation} and note that it is well defined for all $x_i\in\mathcal{N}_h^0$ and $v \in \mathbb{S}$. Since \begin{equation}\label{E:lam1} \lambda_1[D^2 w](x) = \min_{v \in \mathbb{S}} \partial_{vv}^2 w(x), \end{equation} we consider the following approximation of $\lambda_1[D^2 w]$ at $x=x_i \in \mathcal{N}_h^0$ \begin{equation*} \lambda_1[D^2 w](x_i) \approx \min_{v \in \mathbb S_{\theta}} \sdd w(x_i;v). \end{equation*} If $\varepsilon := (h,\delta,\theta)$ encodes the discretetization parameters, our two-scale operator $T_{\varepsilon}$ for the convex envelope problem \eqref{E:pde-int-CE} is finally given by \begin{equation}\label{E:disc-oper} T_{\varepsilon}[w_h;f](x_i) = \min\left\{f(x_i) - w_h(x_i), \; \min_{v \in \mathbb S_{\theta}} \sdd w_h(x_i;v) \right\} \quad\forall \, x_i \in \mathcal{N}_h^0 \end{equation} for any $w_h\in\mathbb{V}_h$. The corresponding two-scale method reads: seek $u_{\varepsilon} \in \mathbb{V}_h$ \begin{equation} \label{E:2ScOp} T_{\varepsilon}[u_{\varepsilon};f](x_i) = 0 \quad \; \forall x_i \in \mathcal{N}_h^0, \end{equation} and $u_{\varepsilon}(x_i) = f(x_i)$ for all $x_i \in \mathcal{N}_h^b$. We say that $w_h \in \mathbb{V}_h$ is a discrete subsolution (supersolution) of \eqref{E:2ScOp} if \[ T_{\varepsilon}[w_h; f](x_i) \ge 0 \ (\le 0) \quad \forall x_i \in \mathcal{N}_h^0 \ ; \quad w_h(x_i) \le (\ge) f(x_i) \quad \forall x_i \in \mathcal{N}_h^b . \] Therefore, a discrete solution of \eqref{E:2ScOp} is both a discrete sub and supersolution. Although this discrete solution $u_{\varepsilon}$ fails to be convex in general, it is still discretely convex, which is a notion of approximate convexity introduced in \cite{NoNtZh1}. We say that $w_h \in \mathbb{V}_h$ is \textit{discretely convex} \cite{NoNtZh1} if \begin{equation*} \sdd w_h(x_i;v) \geq 0 \qquad \forall x_i \in \mathcal{N}_h^0, \quad \forall v \in \mathbb S_{\theta}. \end{equation*} \subsection{Discrete Comparison Principle}\label{S:DCP} One important feature of the definition \eqref{E:disc-oper} of the discrete operator $T_{\varepsilon}$ is its monotonicity. This is similar to the two-scale method for Monge-Amp\`{e}re equation in \cite[Lemma 2.3]{NoNtZh1}. \begin{Lemma}[monotonicity] \label{L:Monotonicity} Let $x_i \in \mathcal{N}_h^0$ be an interior node and $u_h,w_h \in \mathbb{V}_h$. If $u_h(x_i) \geq w_h(x_i)$ and \begin{equation*} \sdd u_h(x_i;v) \leq \sdd w_h(x_i;v) \end{equation*} for any $v \in \mathbb S_{\theta}$, then \begin{equation*} T_{\varepsilon}[u_h;f](x_i) \leq T_{\varepsilon}[w_h;f](x_i). \end{equation*} In particular, if $u_h - w_h$ attains a non-negative maximum at $x_i$, then \begin{equation*} T_{\varepsilon}[u_h;f](x_i) \leq T_{\varepsilon}[w_h;f](x_i). \end{equation*} \end{Lemma} \begin{proof} If $\sdd u_h(x_i;v) \leq \sdd w_h(x_i;v)$ for any $v \in \mathbb S_{\theta}$, then \begin{equation*} \min_{v \in \mathbb S_{\theta}} \sdd u_h(x_i;v) \le \min_{v \in \mathbb S_{\theta}} \sdd w_h(x_i;v). \end{equation*} Recalling the definition \eqref{E:disc-oper} of $T_{\varepsilon}$ and combining with the fact $u_h(x_i) \geq w_h(x_i)$, this implies \begin{equation*} T_{\varepsilon}[u_h;f](x_i) \le T_{\varepsilon}[w_h;f](x_i). \end{equation*} On the other hand, if $u_h - w_h$ attains a non-negative maximum at $x_i \in \mathcal{N}_h^0$, then we have $u_h(x_i) \geq w_h(x_i)$ and \begin{equation*} u_h(x_i) - w_h(x_i) \geq u_h(z) - w_h(z) \quad \forall z \in \overline{\Omega}_h. \end{equation*} By definition \eqref{E:2Sc2Dif} of operator $\sdd$, we obtain \begin{equation*} \sdd u_h(x_i;v) \leq \sdd w_h(x_i;v) \quad \forall v \in \mathbb S_{\theta}, \end{equation*} and thus use the previous result to conclude the proof. \end{proof} Monotonicity leads to the following discrete comparison principle. \begin{Lemma}[discrete comparison principle] \label{L:DCP} Let $u_h,w_h \in \mathbb{V}_h$ with $u_h(x_i) \leq w_h(x_i)$ for all $x_i \in \mathcal{N}_h^b$ and \begin{equation}\label{E:comparison} T_{\varepsilon}[u_h;f](x_i) \geq T_{\varepsilon}[w_h;f](x_i) \quad \forall x_i \in \mathcal{N}_h^0. \end{equation} Then, $u_h \leq w_h$ in $\Omega_h$. \end{Lemma} \begin{proof} The proof splits into two steps. \textbf{Step 1.} We first consider the case with strict inequality \begin{equation}\label{E:strict} T_{\varepsilon}[u_h;f](x_i) > T_{\varepsilon}[w_h;f](x_i) \quad \forall x_i \in \mathcal{N}_h^0. \end{equation} We assume by contradiction that there exists an interior node $x_k \in \mathcal{N}_h^0$ such that $u_h - w_h$ attains a maximum at $x_k$, and $u_h(x_k) > w_h(x_k)$. Then, by \Cref{L:Monotonicity} (monotonicity) we obtain the contradiction \begin{equation} \label{E:Contr} T_{\varepsilon}[u_h;f](x_k) \leq T_{\varepsilon}[w_h;f](x_k). \end{equation} \textbf{Step 2.} Now we deal with \eqref{E:comparison} without the strict inequality. We introduce the auxiliary strictly convex function $q(x) = \frac{1}{2}|x-x_0|^2- \frac{1}{2}R^2$, which satisfies $q \leq 0$ on $\overline{\Omega}$, and in particular $q \leq 0$ on $\partial\Omega_h$ provided $R = \textrm{diam} (\Omega)$ and $x_0 \in \Omega$. Its Lagrange interpolant $q_h=\mathcal I_h q$ is discretely convex and satisfies \begin{equation*} \sdd q_h(x_i;v) \geq \sdd q(x_i;v) = \partial^2_{vv} q(x_i) = 1 \quad\forall x_i \in \mathcal{N}_h^0, \quad\forall v \in \mathbb S_{\theta}, \end{equation*} because $q$ is quadratic. For arbitrary $\alpha>0$, consider the function $u_{\alpha} = u_h + \alpha q_h - \alpha$, which satisfies $u_{\alpha} < u_h \leq w_h$ on $\partial\Omega_h$ and \begin{equation*} \begin{aligned} T_{\varepsilon}[u_{\alpha};f](x_i) & = \min\left\{f(x_i) - u_{\alpha}(x_i), \; \min_{v \in \mathbb S_{\theta}} \sdd u_{\alpha}(x_i)(x_i;v) \right\} \\ & \geq \min\left\{f(x_i) - (u_h(x_i) - \alpha), \; \min_{v \in \mathbb S_{\theta}} \left(\sdd u_h(x_i;v) + \alpha \right) \right\} \\ &= T_{\varepsilon}[u_h;f](x_i) + \alpha > T_{\varepsilon}[w_h;f](x_i) \quad \forall x_i \in \mathcal{N}_h^0. \end{aligned} \end{equation*} Applying Step 1 we deduce \begin{equation*} u_h + \alpha q_h - \alpha \leq w_h \quad \forall \alpha>0. \end{equation*} Finally, let $\alpha\to0$ to obtain the asserted inequality. \end{proof} \subsection{Existence, Uniqueness and Stability}\label{S:Exist-Uniq} We now prove several properties of our discrete system \eqref{E:2ScOp} which are useful for the proof of convergence. \begin{Lemma} [existence, uniqueness and stability]\label{L:Exist-Uniq-Stab} There exists a unique $u_{\varepsilon}\in\mathbb{V}_h$ that solves the discrete equation \eqref{E:2ScOp}. The solution $u_{\varepsilon}$ is stable in the sense that $\|u_{\varepsilon}\|_{L^{\infty}(\Omega_h)} \leq \|f\|_{L^{\infty}(\Omega)}$ regardless of the parameters $\varepsilon=(h,\delta,\theta)$ of the method. \end{Lemma} \begin{proof} Since uniqueness is a trivial consequence of Lemma \ref{L:DCP} (discrete comparison principle), we just have to prove existence and stability. \textbf{Step 1 - Stability:} We first show that $u_h^- = \mathcal I_h u$ is a discrete subsolution where $u$ is the exact convex envelope and $u_h^+ = \mathcal I_h f$ is a discrete supersolution, where again $\mathcal I_h$ stands for the Lagrange interpolation operator. Since $u$ is the exact convex envelope, for any $x_i \in \mathcal{N}_h^0$, we have $u_h^-(x_i) \leq f(x_i)$ and $\sdd u_h^-(x_i;v) \geq 0$ because $u$ is convex. By definition \eqref{E:disc-oper} of $T_{\varepsilon}$, this gives us $T_{\varepsilon}[u_h^-;f](x_i) \geq 0$ for all $x_i \in \mathcal{N}_h^0$. It is also clear that we have $T_{\varepsilon}[u_h^+;f](x_i) \leq f(x_i) - u_h^+(x_i) = 0$ for all $x_i \in \mathcal{N}_h^0$. Therefore combining with the fact that $u_h^+(x_i) = u_h^-(x_i) = f(x_i)$ for $x_i \in \mathcal{N}_h^b$, we see that $u_h^-$ and $u_h^+$ are discrete subsolution and supersolution respectively. By Lemma \ref{L:DCP} (discrete comparison principle), this implies \begin{equation}\label{E:disc-sol-bounds} u_h^- \leq u_{\varepsilon} \leq u_h^+, \end{equation} and we thus obtain the stability of $u_{\varepsilon}$ because both $\|u_h^- \|_{L^{\infty}(\Omega_h)}$ and $\|u_h^+ \|_{L^{\infty}(\Omega_h)}$ are bounded by $\|f\|_{L^{\infty}(\Omega)}$. \textbf{Step 2 - Discrete Perron Method:} It remains to prove the existence of $u_{\varepsilon}$. We proceed as in \cite{NoNtZh1, NoZh2} and use the discrete Perron's method to construct a monotone increasing sequence of functions $\left\{ u_h^k \right\}_{k=0}^\infty$. The initial iterate $u_h^0$ is chosen to be $u_h^-$, and thus satisfies the boundary condition $u_h^0(x_i) = f(x_i)$ for all $x_i \in \mathcal{N}_h^b$ and \begin{equation}\label{E:discrete-sub} T_{\varepsilon}[u_h^0;f](x_i) \geq 0 \quad\forall x_i\in\mathcal{N}_h^0. \end{equation} We construct $\left\{ u_h^k \right\}$ by induction. Suppose that we have already built $u_h^k\in\mathbb{V}_h$ satisfying both the boundary condition and \eqref{E:discrete-sub}. To construct $u_h^{k+1}\in\mathbb{V}_h$ such that $u_h^{k+1}\geq u_h^k$ and also satisfies both the boundary condition and \eqref{E:discrete-sub}, we consider all interior nodes in order and construct auxiliary functions $u_h^{k,i-1}\in\mathbb{V}_h$ using the first $i-1$ nodes and starting from $u_h^{k,0}:=u_h^k$ as follows. At $x_i\in\mathcal{N}_h^0$ we check whether or not $T_{\varepsilon}[u_h^{k,i-1};f](x_i) > 0$. If so, we increase the value of $u_h^{k,i-1}(x_i)$ and denote the resulting function by $u_h^{k,i}$, until \begin{equation*} T_{\varepsilon}[u_h^{k,i};f](x_i) = 0. \end{equation*} This is possible because $T_{\varepsilon}[u_h^{k,i};f](x_i)$ is strictly decreasing with respect to $u_h^{k,i}(x_i)$. Expression \eqref{E:disc-oper} also shows that this process does not decrease $T_{\varepsilon}[u_h^{k,i};f](x_j)$ for any $x_j \ne x_i$, whence \begin{equation*} T_{\varepsilon}[u_h^{k,i};f](x_j) \geq T_{\varepsilon}[u_h^{k,i-1};f](x_j) \geq 0 \quad\forall x_j\ne x_i. \end{equation*} We repeat this process with the remaining nodes $x_j$ for $i<j\leq N$ where $N$ is the number of all interior points, and set $u_h^{k+1} := u_h^{k,N}$ to be the last intermediate function. By construction, we clearly obtain \begin{equation*} T_{\varepsilon}[u_h^{k+1};f](x_i) \geq 0, \quad u_h^{k+1}(x_i) \geq u_h^k(x_i) \quad\forall x_i\in\mathcal{N}_h^0, \end{equation*} and $u_h^k(x_i) = f(x_i)$ for all $x_i \in \mathcal{N}_h^b$. \textbf{Step 3 - Convergence of $u_h^k$:} By construction we have $u_h^k \geq u_h^0 = u_h^-$ and by \Cref{L:DCP} (discrete comparison principle), $u_h^k \leq u_h^+$ and thus $u_h^k(x_i)$ is uniformly bounded. Since the sequence $\{u_h^k\}_k$ is monotone, it must converge to a limit \begin{equation*} u_{\varepsilon}(x_i) = \lim_{k\to\infty} u_h^k(x_i) = \lim_{k\to\infty} u_h^{k,i}(x_i) \quad\forall x_i\in\mathcal{N}_h. \end{equation*} Due to continuity of $T_{\varepsilon}[w_h;f]$ with respect to $w_h(x_j)$, we have $T_{\varepsilon}[u_{\varepsilon};f](x_i) = \lim_{k\to\infty}T_{\varepsilon}[u_h^{k,i};f](x_i) = 0 $ for any $x_i \in \mathcal{N}_h^0$. This implies that the limit $u_{\varepsilon}$ is the solution of discrete equation \eqref{E:2ScOp} and finishes the proof. \end{proof} Another way to prove existence and uniqueness is to take advantage of the existing results for Bellman equation and Howard's algorithm as we can see in section \ref{S:Exp}. \looseness=-1 We define for $x \in \overline{\Omega}$ \begin{equation}\label{E:def-uo-uu} \overline{u}(x) := \limsup_{\varepsilon,\frac{h}{\delta} \to 0, \;y \to x} u_{\varepsilon}(y), \quad \underline{u}(x) := \liminf_{\varepsilon,\frac{h}{\delta} \to 0, \;y \to x} u_{\varepsilon}(y), \end{equation} where the limits are taken for $y \in \Omega_h$. From equation \eqref{E:disc-sol-bounds} and the continuity of both $u$ and $f$, we immediately obtain the following lemma characterizing the behavior of $\overline{u}$ and $\underline{u}$ on the boundary $\partial \Omega$. \begin{Lemma}[boundary behavior] \label{L:disc-sol-boundary} Let $\Omega$ be a strictly convex bounded domain, let $u_{\varepsilon}$ be the discrete solution of \eqref{E:2ScOp}, and let $\overline{u}(x)$ and $\underline{u}(x)$ be defined in \eqref{E:def-uo-uu}. Then we have $\overline{u}(x) = \underline{u}(x) = f(x)$ for all $x \in \partial \Omega$. \end{Lemma} \begin{proof} Since $\Omega$ is strictly convex, the Dirichlet boundary condition $u=f$ on $\partial\Omega$ is attained as a direct consequence of \cite[Corollary 17.1.5]{Rockafellar2015convex}, or can be proved in the same way as \cite[Theorem 1.5.2]{Gutierrez}. Next use \begin{equation*} \mathcal I_h u(x) = u_h^-(x) \leq u_{\varepsilon}(x) \leq u_h^+(x) = \mathcal I_h f(x) \quad x \in \Omega_h \end{equation*} with equality on $\partial\Omega$ to deduce the assertion. \end{proof} \subsection{Consistency} \label{S:Consistency} We now quantify the consistency error of our discrete operator $T_{\varepsilon}[\mathcal I_h u;f]$ for a smooth function $u$, which is enough for the proof of convergence. In \Cref{S:RoC} we will carry out a more delicate analysis of the consistency error which enables us to prove error estimates for solutions with weaker but realistic regularity. In the meantime, we stress that the convex envelope $u$ is generically never better than of class $C^{1,1}(\overline{\Omega})$ \cite{DeFi}. Given a node $x_i \in \mathcal{N}_h^0$ we denote \begin{equation}\label{E:Bi} B_i := \cup \{T: T\in\mathcal{T}_h, \, \textrm{dist}(x_i,T) \leq \delta_i \}, \end{equation} where $\delta_i$ is defined in \eqref{E:deltai}. We also denote by $\Omega_{h,s}$ the following $s$-interior region of $\Omega_h$ for any parameter $s > 0$ \begin{equation*} \Omega_{h,s} = \left\{ x \in \Omega_h \; : \; \textrm{dist}(x,\partial \Omega_h ) \geq s \right\}. \end{equation*} Hereafter, we use the symbols $C(d,\sigma)$, $C(d)$ and $C$ to denote constants that depend only on the dimension $d$ and the shape-regularity constant $\sigma$, but are independent of the two scales $h$ and $\delta$, the parameter $\theta$ and the function $u$. \Cref{L:Consistency-smooth} below establishes a consistency error estimate for the two-scale method similar to \cite[Lemma 4.1]{NoNtZh1} and \cite[Lemma 4.2]{NoNtZh1}. The proof follows along the lines of \cite{NoNtZh1}. \begin{Lemma} [consistency for smooth functions] \label{L:Consistency-smooth} Let $u\in C^{2+k,\alpha}(B_i)$ for $k=0,1$ and $\alpha \in (0,1]$, $\mathcal I_h u$ be its Lagrange interpolant, and $B_i$ be defined in \eqref{E:Bi}. The following estimates are then valid: \begin{enumerate}[(i)] \item For all $x_i \in \mathcal{N}_h^0$ and all $v \in \mathbb{S}$, we have \begin{equation}\label{E:sddIhuB} \left|\sdd \mathcal I_h u (x_i;v) \right| \leq C(d,\sigma) \; |u|_{W^2_{\infty}(B_i) }, \end{equation} \item For all $x_i \in \mathcal{N}_h^0 \cap \Omega_{h,\delta}$ and all $v \in \mathbb{S}$, we have \begin{equation}\label{E:sdd-error} \left|\sdd \mathcal I_h u(x_i;v) - \sd{u}{v}(x_i) \right| \leq C(d,\sigma) \left( |u|_{C^{2+k,\alpha}(B_i)} \delta^{k+\alpha}+ |u|_{W^2_{\infty}(B_i)}\frac{h^2}{\delta^2} \right), \end{equation} \item For all $x_i \in \mathcal{N}_h^0 \cap \Omega_{h,\delta}$ and all $v \in \mathbb{S}$, we have \begin{equation}\label{E:Op-error-smooth} \small \bigg| T_{\varepsilon}[\mathcal I_h u;f](x_i) - T[u;f](x_i) \bigg| \leq C(d,\sigma) \left[ |u|_{C^{2+k,\alpha}(B_i)} \delta^{k+\alpha} + |u|_{W^2_{\infty}(B_i)} \left( \frac{h^2}{\delta^2} + \theta^2 \right) \right]. \end{equation} \end{enumerate} \end{Lemma} \begin{proof} For the proof of \eqref{E:sddIhuB} and \eqref{E:sdd-error}, the readers may refer to \cite[Lemma 4.1]{NoNtZh1}. Here we only prove \eqref{E:Op-error-smooth}. Recalling the definitions of $T$ in \eqref{E:pde-int-CE} and $T_{\varepsilon}$ in \eqref{E:disc-oper} we only need to prove {\small\begin{equation* \left| \lambda_1[D^2u](x_i) - \min_{v \in \mathbb S_{\theta}} \sdd \mathcal I_h u(x_i;v) \right| \leq C(d,\sigma) \left[ |u|_{C^{2+k,\alpha}(B_i)} \delta^{k+\alpha} + |u|_{W^2_{\infty}(B_i)} \left( \frac{h^2}{\delta^2} + \theta^2 \right) \right]. \end{equation*}} To this end, first let $v_{\theta}$ be the direction such that \begin{equation*} \sdd \mathcal I_h u (x_i;v_{\theta}) = \min_{v \in \mathbb S_{\theta}} \sdd \mathcal I_h u (x_i;v). \end{equation*} We use \eqref{E:lam1} and \eqref{E:sdd-error} to get \begin{equation*} \begin{aligned} \lambda_1[D^2u](x_i) - \min_{v \in \mathbb S_{\theta}} \sdd \mathcal I_h u(x_i;v) \leq & \; \sd{u}{v_{\theta}}(x_i) - \sdd \mathcal I_h u(x_i;v_{\theta}) \\ \leq & \; C(d,\sigma) \left( |u|_{C^{2+k,\alpha}(B_i)} \delta^{k+\alpha}+ |u|_{W^2_{\infty}(B_i)} \frac{h^2}{\delta^2} \right), \end{aligned} \end{equation*} which proves one inequality of \eqref{E:Op-error-smooth}. To show the reverse inequality we let $v$ be the direction that realizes the minimum in \eqref{E:lam1}, which means \begin{equation*} \partial_{vv}^2 u(x_i) = \lambda_1[D^2 u](x_i), \end{equation*} and we also know that $v$ is the eigenvector of $D^2 u(x_i)$ corresponding to the smallest eigenvalue $\lambda_1$. By definition of $\mathbb S_{\theta}$, there exists $v_{\theta} \in \mathbb S_{\theta}$ such that $|v - v_{\theta}| \leq \theta$, and we can thus write \begin{equation*} \min_{v \in \mathbb S_{\theta}} \sdd \mathcal I_h u(x_i;v) - \lambda_1[D^2u](x_i) \leq \sdd \mathcal I_h u(x_i;v_{\theta}) - \partial_{vv}^2 u(x_i) = I_1 + I_2, \end{equation*} where \begin{equation*} I_1 = \sdd \mathcal I_h u(x_i;v_{\theta}) - \partial_{v_{\theta}\vt}^2 u(x_i), \qquad I_2 = \partial_{v_{\theta}\vt}^2 u(x_i) - \partial_{vv}^2 u(x_i). \end{equation*} It is clear that $I_1$ can be bounded by \eqref{E:sdd-error}. For $I_2$, write $v_{\theta} = v + w$, then \begin{equation*} \begin{aligned} \partial_{v_{\theta}\vt}^2 u(x_i) &= v_{\theta}^T D^2u(x_i) v_{\theta} = \partial_{vv}^2 u(x_i) + 2 w^T D^2u(x_i) v + w^T D^2u(x_i) w \\ &= \partial_{vv}^2 u(x_i) + 2 \lambda_1 v \cdot w + w^T D^2u(x_i) w. \end{aligned} \end{equation*} Since \begin{equation*} 1= |v_{\theta} |^2 = |v|^2 + 2 v \cdot w + |w|^2, \end{equation*} and $|v| = 1$, we observe that \begin{equation*} | v \cdot w | = \frac{1}{2} |w|^2 \leq \frac{1}{2}\theta^2, \end{equation*} whence we obtain \begin{equation*} I_2 \leq C |u|_{W^2_{\infty}(B_i)} \theta^2. \end{equation*} Combining the bounds for both $I_1$ and $I_2$ we have {\small \begin{equation*} \min_{v \in \mathbb S_{\theta}} \sdd \mathcal I_h u(x_i;v) - \lambda_1[D^2u](x_i) \leq C(d,\sigma) \left[ |u|_{C^{2+k,\alpha}(B_i)} \delta^{k+\alpha} + |u|_{W^2_{\infty}(B_i)} \left( \frac{h^2}{\delta^2} + \theta^2 \right) \right] . \end{equation*}} This finishes the proof of \eqref{E:Op-error-smooth}. \end{proof} \subsection{Convergence}\label{S:Convergence} We are now ready to prove the convergence result. \begin{Theorem}[convergence] \label{T:Convergence} If $\Omega$ is a bounded and strictly convex domain and $f \in C(\overline{\Omega})$, then the discrete solution $u_{\varepsilon}$ of \eqref{E:2ScOp} converges uniformly to the convex envelope $u$ of $f$ as $\varepsilon = (h, \delta, \theta) \to 0$ and $\frac{h}{\delta} \to 0$. \end{Theorem} \begin{proof} Our approximation scheme \eqref{E:2ScOp} satisfies monotonicity (\Cref{L:DCP}), stability (\Cref{L:Exist-Uniq-Stab}), and consistency (\Cref{L:Consistency-smooth}). Moreover, the PDE \eqref{E:pde-CE} for the convex envelope problem admits a comparison principle \cite[Proposition 2.7]{ObRu} for Dirichlet boundary conditions in the classical sense. Similarly to \cite[Section 4]{JeSm}, \cite[Theorem 17]{FeJe} and \cite[Section 5]{NoNtZh1}, in order to use the convergence theorem of Barles and Souganidis \cite{BaSoug}, we still need the additional fact that $\overline{u}(x) = \underline{u}(x) = f(x)$ on $\partial\Omega$. Since this is proved in \Cref{L:disc-sol-boundary} (boundary behavior), \cite{BaSoug} yields uniform convergence of the discrete solution $u_{\varepsilon}$ to the viscosity solution $u$ of \eqref{E:pde-CE}. \end{proof} \section{Rates of Convergence} \label {S:RoC} In this section, we prove convergence rates for solutions of class $C^{k,\alpha}(\overline{\Omega})$ for $k=0,1$ and $0<\alpha \leq 1$. Since in general we could only expect $u \in C^{1,1}(\overline{\Omega})$ even for smooth $f$ and $\Omega$, our estimate of consistency error in \Cref{S:Consistency} fails. The challenge is thus to estimate the consistency error for solutions with less regularity. We first show a key geometric lemma about convex envelopes which enables us to give an estimate of the consistency error for $u \in C^{k,\alpha}(\overline{\Omega})$. On the basis on this result, we next prove the convergence rate using \Cref{L:DCP} (discrete comparison principle). \subsection{Flatness}\label{S:Flatness} The heuristic behind the governing PDE \eqref{E:pde-int-CE} is that the convex envelope $u$ must be flat at least in one direction within the non-contact set, i.e. $\lambda_1[D^2 u](x) = 0$ for all $x \notin \mathcal{C}(f)$. The question whether there is a line segment containing $x$, on which $u$ is flat, is studied in \cite[Section 3]{ObSi} for the Dirichlet convex envelope problem in which $f$ is only defined on $\partial\Omega$. For $f \in C(\overline{\Omega})$ defined in the entire $\Omega$, and corresponding definition \eqref{E:def-CE} of convex envelope $u$, we have a similar property. \begin{Lemma}[flatness in one direction]\label{L:Flatness} Let $f \in C(\overline{\Omega})$ and $x \in \Omega$ be such that $\textrm{dist}(x,\mathcal{C}(f)) \geq d \delta$. Then for any slope $p \in \partial u(x)$, there exists a direction $v \in \mathbb{S}$ such that \begin{equation*} x_{\pm} = x \pm \delta v, \quad u(x_{\pm}) = u(x) \pm \delta (p \cdot v), \quad \sdd u(x;v) = 0. \end{equation*} Moreover, $p$ belongs also to the subdifferential sets $\partial u(x_{\pm})$. \end{Lemma} This lemma says that if $x$ is away from the contact set $\mathcal{C}(f)$ at least at distance $d \delta$, then there exists a line segment centered at $x_i$ with length at least $2\delta$ such that the convex envelope $u$ is flat on this segment. The flattness means the second difference of $u$ in this direction is $0$, which plays an important role in obtaining consistency error for $x$ far away from $\mathcal{C}(f)$. To prove \Cref{L:Flatness}, we need the following definition and subsequent result: given $x \in \Omega \setminus \mathcal{C}(f)$ and $p \in \partial u(x)$, let \begin{equation*} \mathcal{C}(f;x,p) := \left\{ y \in \overline{\Omega}: f(y) = u(x) + p \cdot (y-x) \right\}, \end{equation*} and note that $\mathcal{C}(f;x,p) \subset \mathcal{C}(f)$ because $u$ is convex and $u(y) \ge u(x) + p \cdot (y-x)$ whence $u(y)=f(y)$. The following auxiliary result is exactly the same as \cite[Lemma 3.3]{DeFi} and similar to \cite[Lemma 2]{CaNiSp} and \cite[Theorem 3.2]{ObSi}. We still give a proof here for completeness. \begin{Lemma}[structure of non-contact set]\label{L:noncontact-set} Let $f \in C(\overline{\Omega})$ and $x \in \Omega \setminus \mathcal{C}(f)$. Then for any slope $p \in \partial u(x)$, there exist points $x_1,\ldots,x_k \in \mathcal{C}(f)$ with $2 \leq k \leq d+1$ such that \begin{equation*} x \in \textrm{conv} \;(x_1,\ldots,x_k), \end{equation*} and $u$ is affine in the convex hull $\textrm{conv} \;(x_1,\ldots,x_k)$ of $(x_i)_{i=1}^k$. Moreover, $p$ is also in the subdifferential set $\partial u(y)$ for any $y \in \textrm{conv} \;(x_1,\ldots,x_k)$. \end{Lemma} \begin{proof} For any $p \in \partial u(x)$, define $P(y) := u(x) + p \cdot (y-x)$ and observe that \begin{equation*} \mathcal{C} := \mathcal{C}(f;x,p) = \left\{ y \in \overline{\Omega}: f(y) = P(y) \right\}. \end{equation*} We claim that $x \in \textrm{conv}(\mathcal{C})$. Argue by contradiction, suppose $x \notin \textrm{conv}(\mathcal{C})$, and use the hyperplane separation theorem to find an affine function $L$ such that $L(x) > 0$ and $L(y) < 0$ for every $y \in \mathcal{C}$. By the definition of $\mathcal{C}$ and the fact that $P \leq u \leq f$, it is clear that $f - P$ is strictly positive in the compact set $\overline{\Omega} \cap \{ L \geq 0\}$: in fact, if $f(y) \le P(y)$ then $f(y) = P(y) = u(y)$ and $y \in \mathcal{C}$, whence $L[y] < 0$. Therefore it is easy to see that for some small $\alpha > 0$, we have \begin{equation*} \wt{L}(y) := P(y) + \alpha L(y) \leq f(y) \quad \forall y \in \overline{\Omega}, \end{equation*} but $\wt{L}(x) > P(x) = u(x)$. This contradicts the definition of convex envelope $u$ and thus proves the claim $x \in \textrm{conv}(\mathcal{C})$. Now we use Carath\'{e}odory's theorem to obtain the existence of $x_1,\ldots,x_k \in \mathcal{C}$ with $k \leq d+1$ such that $x \in \textrm{conv}(x_1,\ldots,x_k)$. To prove that $p \in \partial u(y)$ for any $y \in \textrm{conv}(x_1,\ldots,x_k)$, we define \begin{equation*} \mathcal{K} := \left\{ y \in \overline{\Omega}: u(y) = u(x) + p \cdot (y - x) \right\} = \left\{ y \in \overline{\Omega}: u(y) = P(y) \right\}, \end{equation*} whence $u$ is affine in $\mathcal{K}$. We claim that $\mathcal{K}$ is convex. Let $y_1,y_2 \in \mathcal{K}, \lambda \in (0,1)$ and $z = \lambda y_1 + (1- \lambda) y_2$. Since $u$ is convex, we have \begin{equation*} u(z) \leq \lambda u(y_1) + (1- \lambda) u(y_2) = \lambda P(y_1) + (1- \lambda) P(y_2) = P(z). \end{equation*} On the other hand, since $p \in \partial u(x)$, the supporting plane $P$ must be below $u$, and in particular \begin{equation*} u(z) \geq P(z). \end{equation*} Therefore $u(z) = P(z)$, and thus $z \in \mathcal{K}$, which implies the convexity of $\mathcal{K}$. Since $P \leq u \leq f$, we have $\{x_1, \ldots, x_k \} \subset \mathcal{C} \subset \mathcal{K}$ and $\textrm{conv} \;(x_1, \ldots, x_k) \subset \mathcal{K}$. It is clear that for any $y \in \mathcal{K}$, we have $u(y) = u(x) + p \cdot (y - x)$ and \begin{equation*} P(z) = u(x) + p \cdot (z - x) = u(y) + p \cdot (z - y) \leq u(z) \quad \forall z \in \overline{\Omega}. \end{equation*} By definition of $\partial u(y)$ this implies $p \in \partial u(y)$ for any $y \in \textrm{conv}(x_1,\ldots,x_k)$. In addition, $u$ is affine in $\textrm{conv} \;(x_1, \ldots, x_k)$. \end{proof} \begin{proof}[Proof of \Cref{L:Flatness}] For any $p \in \partial u(x)$, by \Cref{L:noncontact-set} (structure of non-contact set), there exist $k$ ($2\le k \le d+1$) points $x_i \in \mathcal{C}(f;x,p)$ such that \begin{equation*} x = \sum_{i=1}^{k} \lambda_i x_i, \quad \lambda_i \geq 0, \quad \sum_{i=1}^{k} \lambda_i = 1, \end{equation*} and $p$ belongs to the subdifferential set $\partial u(y)$ for any $y \in \textrm{conv}(x_1,\ldots,x_k)$. If $j$ is such that $\lambda_j = \max_{1 \leq i \leq k} \lambda_i$, then we have \begin{equation*} \lambda_j \geq \frac{1}{k} \sum_{i=1}^{k} \lambda_i = \frac{1}{k} \geq \frac{1}{d+1}. \end{equation*} Now let $x_0 = \sum_{i \neq j} \frac{\lambda_i}{1-\lambda_j} x_i \in \textrm{conv} \;(x_1, \ldots, x_k)$ to get \begin{equation*} x = \sum_{i=1}^{k} \lambda_i x_i = \lambda_j x_j + \sum_{i \neq j} \lambda_i x_i = \lambda_j x_j + (1-\lambda_j) x_0. \end{equation*} Since both $x_0,x_j \in \textrm{conv} \;(x_1,\ldots,x_k)$, the segment $\overline{x_0x_j}$ is also in $\textrm{conv} \;(x_1,\ldots,x_k)$. Due to the fact $\textrm{dist}(x,\mathcal{C}(f)) \geq d \delta$, we have $|x_j - x| \geq d \delta$, and \begin{equation*} |x_0 - x| = \frac{\lambda_j}{1 - \lambda_j} |x_j - x| \geq \frac{1/(d+1)}{1-1/(d+1)} \; d \delta = \delta. \end{equation*} Therefore, if $v = \frac{x_j - x}{|x_j - x|}$ and $x_{\pm} = x \pm \delta v$, clearly $x_{\pm}$ lie in the segment $\overline{x_0x_j}$, and thus also inside $\textrm{conv}(x_1,\ldots,x_k)$. Finally, \Cref{L:noncontact-set} (structure of non-contact set) shows $p \in \partial u(x_{\pm})$ and $u(x_{\pm}) = u(x) \pm \delta (p \cdot v)$, which immediately leads to $\sdd u(x;v) = 0$. \end{proof} \subsection{Consistency for Solutions with H\"older Regularity}\label{S:Consistency-lowreg} In this section, we take advantage of results in \Cref{S:Flatness} to derive a consistency error for solutions with realistic H\"older regularity $u \in C^{k,\alpha}(\overline{\Omega})$ for $k=0,1$ and $0<\alpha \leq 1$, which improves upon the consistency error estimates in \Cref{S:Consistency}. The Lagrange interpolant $\mathcal I_h u \in \mathbb{V}_h$ of $u$ satisfies for all interior nodes $x_i \in \mathcal{N}_h^0$ \begin{equation*} \mathcal I_h u(x_i) = u(x_i) \leq f(x_i), \quad \sdd \mathcal I_h u(x_i;v) \geq \sdd u(x_i;v) \geq 0 \quad \forall v \in \mathbb{S} \end{equation*} because of the convexity of $u$. In view of definition \eqref{E:disc-oper} of $T_{\varepsilon}$, this in turn implies $T_{\varepsilon}[\mathcal I_h u;f](x_i) \geq 0$ for all $x_i \in \mathcal{N}_h^0$. The following proposition yields upper bounds for $T_{\varepsilon}[\mathcal I_h u;f](x_i)$ depending on the location of $x_i$ relative to $\mathcal{C}(f)$ and $\partial\Omega$. \begin{Proposition}[consistency for $u$ with H\"older regularity]\label{Prop:Consistency-lowreg} Let $\Omega$ be a bounded strictly convex domain, $u \in C^{k,\alpha}(\overline{\Omega})$ for $k=0,1$ and $0<\alpha \leq 1$ be the exact solution of the convex envelope problem \eqref{E:pde-CE}. In addition, let $B_i$ be defined in \eqref{E:Bi} and set \begin{equation}\label{E:wtBi} \wt{B_i} := \{x \in \overline{\Omega}: |x - x_i| \leq d\delta \}. \end{equation} For $x_i \in \mathcal{N}_h^0$, the following estimates are then valid: \begin{enumerate}[(i)] \item If $\textrm{dist}(x_i,\mathcal{C}(f)) \geq d \delta$, we have \begin{equation}\label{E:error-noncontact} \min_{v_{\theta} \in \mathbb S_{\theta}} \sdd \mathcal I_h u(x_i;v_{\theta}) \leq C(d,\sigma) \frac{(\delta \theta)^{k+\alpha} + h^{k+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(B_i)}. \end{equation} \item If $\textrm{dist}(x_i,\mathcal{C}(f)) < d \delta, \; \textrm{dist}(x_i,\partial\Omega) \geq d\delta$, and $f \in C^{k,\alpha}(\overline{\Omega})$, then for $k = 0$ we have \begin{equation}\label{E:error-contact-0} f(x_i) - u(x_i) \leq C(d,\sigma) \delta^{\alpha} \left( |u|_{C^{0,\alpha}(\wt{B_i})} + |f|_{C^{0,\alpha}(\wt{B_i})} \right), \end{equation} whereas for $k = 1$ we have \begin{equation}\label{E:error-contact-1} f(x_i) - u(x_i) \leq C(d,\sigma) \delta^{1+\alpha} |f|_{C^{1,\alpha}(\wt{B_i})}. \end{equation} \item If $0 < \textrm{dist}(x_i,\partial\Omega) < d\delta$, then for all $v \in \mathbb{S}$, we have \begin{equation}\label{E:error-boundary} \sdd \mathcal I_h u(x_i;v) \leq C(d,\sigma) \delta_i^{k+\alpha-2} |u|_{C^{k,\alpha}(B_i)}, \end{equation} and \eqref{E:error-contact-0} also holds provided $k=0$. \end{enumerate} \end{Proposition} \begin{proof} Since $\Omega$ is strictly convex, we have $\partial\Omega \subset \mathcal{C}(f)$. This implies that $x_i \in \mathcal{N}_h^0$ must fall within one of the following three mutually exclusive cases. \textbf{Case 1: } $\textrm{dist}(x_i,\mathcal{C}(f)) \geq d \delta$. By \Cref{L:Flatness} (flatness in one direction), for any $p \in \partial u(x_i)$, there exists $v \in \mathbb{S}$ such that \begin{equation*} x_{\pm} = x_i \pm \delta v, \quad u(x_{\pm}) = u(x_i) \pm \delta (p \cdot v), \quad \sdd u(x_i;v) = 0. \end{equation*} By the definition of $\mathbb S_{\theta}$, there exists $v_{\theta} \in \mathbb S_{\theta}$ such that $|v - v_{\theta}| \leq \theta$. We claim that \begin{equation*}\sdd \mathcal I_h u(x_i;v_{\theta}) \leq C(d,\sigma) \frac{(\delta \theta)^{k+\alpha} + h^{k+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(B_i)}, \end{equation*} which implies \eqref{E:error-noncontact}. Using $\; \textrm{dist}(x_i,\mathcal{C}(f)) \geq d \delta$, we have $\delta_i = \delta$ in definition \eqref{E:2Sc2Dif}. Let $x^{\theta}_{\pm} = x_i \pm \delta v_{\theta}$, then $x^{\theta}_{\pm} \in B_i$ and $|x^{\theta}_{\pm} - x_{\pm} | \leq \delta \theta$. Since the interpolation error satisfies \begin{equation}\label{E:interp-error} |u - \mathcal I_h u|_{L^{\infty}(B_i)} \leq C(d,\sigma) h^{k+\alpha}|u|_{C^{k,\alpha}(B_i)}, \end{equation} we infer that \begin{equation}\label{E:sdd-u-uh} \left| \sdd \mathcal I_h u(x_i;v_{\theta}) - \sdd u(x_i;v_{\theta}) \right| \leq C(d,\sigma) \frac{h^{k+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(B_i)}, \end{equation} whence it remains to prove \begin{equation*} \sdd u(x_i;v_{\theta}) \leq C(d,\sigma) \frac{(\delta \theta)^{k+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(B_i)}. \end{equation*} For $k = 0$, by definition of $|u|_{C^{k,\alpha}(B_i)}$ seminorm, we see that \begin{align*} |u(x_{\pm}) - u(x^{\theta}_{\pm})| \leq |x^{\theta}_{\pm} - x_{\pm} |^{\alpha} \; |u|_{C^{k,\alpha}(B_i)} \leq (\delta \theta)^{\alpha} \; |u|_{C^{k,\alpha}(B_i)}. \end{align*} Using this inequality, along with $\sdd u(x_i;v) = 0$, yields the desired bound \begin{equation*} \begin{aligned} \sdd u(x_i;v_{\theta}) & \leq \sdd u(x_i;v) + \frac{|u(x_+) - u(x^{\theta}_+)| + |u(x_+) - u(x^{\theta}_+)|}{\delta^2} \\ & \leq \frac{2(\delta \theta)^{k+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(B_i)}. \end{aligned} \end{equation*} For $k=1$, we know $p = \nabla u(x_i) = \nabla u(x_{\pm})$. If $w = v_{\theta} - v$, we then have \begin{equation*} \begin{aligned} u(x^{\theta}_{\pm}) &= u(x_{\pm}) \pm \int_0^1 \delta \; \nabla u\left(x_{\pm} \pm t\delta w \right) \cdot w \; dt \\ &= u(x_{\pm}) \pm \delta \nabla u(x_{\pm}) \cdot w \pm \int_0^1 \delta \; \left[ \nabla u\left(x_{\pm} \pm t\delta w \right) - \nabla u(x_{\pm}) \right] \cdot w \; dt, \end{aligned} \end{equation*} whence \begin{equation}\label{E:C1a-error} \begin{aligned} u(x^{\theta}_{\pm}) &\leq u(x_{\pm}) \pm \delta \nabla u(x_{\pm}) \cdot w + \int_0^1 \delta \; |t \delta w|^{\alpha}|u|_{C^{k,\alpha}(B_i)} \; |w| \; dt \\ &\leq u(x_{\pm}) \pm \delta p \cdot w + C (\delta \theta)^{1+\alpha} |u|_{C^{k,\alpha}(B_i)} . \end{aligned} \end{equation} Therefore plugging the above inequalities into the expression of $\sdd u(x_i;v_{\theta})$ we obtain \begin{equation*} \begin{aligned} \sdd u(x_i;v_{\theta}) &\leq \sdd u(x_i;v) + \frac{1}{\delta^2} \left( \delta p \cdot w - \delta p \cdot w + 2 C (\delta \theta)^{1+\alpha} |u|_{C^{k,\alpha}(B_i)} \right) \\ &\leq C \frac{(\delta \theta)^{1+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(B_i)}, \end{aligned} \end{equation*} and finish the proof of our claim. \textbf{Case 2: } $\textrm{dist}(x_i,\mathcal{C}(f)) < d \delta$ and $\; \textrm{dist}(x_i,\partial\Omega) \geq d\delta$. By the assumptions, there exists $y \in \mathcal{C}(f) \setminus \partial\Omega$ such that $|x_i - y| < d \delta$. We claim that if $k = 0$, \begin{equation*} f(x_i) - \mathcal I_h u(x_i) \leq C(d,\sigma) \delta^{\alpha} \left( |u|_{C^{0,\alpha}(\wt{B_i})} + |f|_{C^{0,\alpha}(\wt{B_i})} \right), \end{equation*} which is \eqref{E:error-contact-0}. This claim is a consequence of $\mathcal I_h u(x_i) = u(x_i), u(y) = f(y)$ and \begin{equation*} \begin{aligned} &\left| u(x_i) - u(y) \right| \leq |x_i - y|^{\alpha} |u|_{C^{0,\alpha}(\wt{B_i})} \leq d^{\alpha} \delta^{\alpha} |u|_{C^{0,\alpha}(\wt{B_i})}, \\ &\left| f(x_i) - f(y) \right| \leq |x_i - y|^{\alpha} |f|_{C^{0,\alpha}(\wt{B_i})} \leq d^{\alpha} \delta^{\alpha} |f|_{C^{0,\alpha}(\wt{B_i})}. \end{aligned} \end{equation*} If $k = 1$, we claim that \begin{equation*} f(x_i) - \mathcal I_h u(x_i) \leq C(d,\sigma) \delta^{1+\alpha} |f|_{C^{1,\alpha}(\wt{B_i})}, \end{equation*} which is \eqref{E:error-contact-1}. To prove this claim, we let $p = \nabla u(y)$, then consider the supporting hyperplane $P(x) := u(y) + (x - y) \cdot p$. Since $f$ is differentiable, $f(y) = P(y)$ and $f(x) \geq u(x) \geq P(x)$, we know $p = \nabla f(y)$. Proceeding similarly to \eqref{E:C1a-error}, we end up with \begin{equation*} \left| f(x_i) - P(x_i) \right| = \left| f(x_i) - f(y) - (x_i - y) \cdot p \right| \leq C(d,\sigma) \delta^{1+\alpha} |f|_{C^{1,\alpha}(\wt{B_i})}. \end{equation*} Therefore our claim holds because \begin{equation*} f(x_i) - \mathcal I_h u(x_i) = f(x_i) - u(x_i) \leq f(x_i) - P(x_i) \leq C(d,\sigma) \delta^{1+\alpha} |f|_{C^{1,\alpha}(\wt{B_i})}. \end{equation*} \textbf{Case 3:} $0 < \textrm{dist}(x_i,\partial\Omega) < d\delta$. We point out that, unlike the first two cases, the upper bound given in \eqref{E:error-boundary} does not converge to zero as $\delta_i \to 0$. However, this result is still useful in our proof of error estimates. We claim that for all $v \in \mathbb{S}$, \begin{equation*} \sdd \mathcal I_h u(x_i;v) \leq C(d,\sigma) \delta_i^{k+\alpha-2} |u|_{C^{k,\alpha}(B_i)}, \end{equation*} which is \eqref{E:error-boundary}. Using \eqref{E:interp-error} and the fact $ \delta_i/h \geq C(d,\sigma)$ due to the shape-regularity assumption on the mesh $\mathcal{T}_h$, we have \begin{equation*} \begin{aligned} \left| \sdd u(x_i;v) - \sdd \mathcal I_h u(x_i;v) \right| &\leq C(d,\sigma) \frac{h^{k+\alpha}}{\delta_i^2} |u|_{C^{k,\alpha}(B_i)} \\ &\leq C(d,\sigma) \delta_i^{k+\alpha - 2} |u|_{C^{k,\alpha}(B_i)}. \end{aligned} \end{equation*} Consequently, it just suffices to prove \begin{equation*} \sdd u(x_i;v) \leq C(d,\sigma) \delta_i^{k+\alpha-2} |u|_{C^{k,\alpha}(B_i)}. \end{equation*} If $k = 0$, this is obtained from \begin{equation*} \left| u(x_i \pm \delta_i v) - u(x_i) \right| \leq \delta_i^{\alpha} |u|_{C^{0,\alpha}(B_i)}. \end{equation*} If $k = 1$, let $p = \nabla u(x_i)$ and $P(x) = u(x_i) + (x-x_i) \cdot p$, we have similarly to \eqref{E:C1a-error} \begin{equation*} \left| (u-P)(x_i \pm \delta_i v) \right| \leq C \delta_i^{1+\alpha} |u|_{C^{1,\alpha}(B_i)}. \end{equation*} Therefore since $\sdd P(x_i;v) = 0$, our claim is a consequence of \begin{equation*} \sdd u(x_i;v) \leq \sdd P(x_i;v) + \frac{C \delta_i^{1+\alpha} |u|_{C^{1,\alpha}(B_i)}}{\delta_i^2} = C \delta_i^{\alpha-1} |u|_{C^{1,\alpha}(B_i)}. \end{equation*} This concludes the proof. \end{proof} \begin{comment} \begin{remark} When $x_i \in \mathcal{N}_h^0$ falls into the third case in our proof, we could also give an upper bound of $T_{\varepsilon}[\mathcal I_h u;f](x_i)$ through estimating $f(x_i) - \mathcal I_h u(x_i) = f(x_i) - u(x_i)$. To illustrate this idea, we need to further assume $f \in C^{k,\alpha}(\overline{\Omega})$. Since $\textrm{dist}(x_i,\partial\Omega) < d\delta$, there exists $y \in \partial\Omega$ such that $|x_i - y| < d\delta$, and thus we have $u(y) = f(y)$. If $k + \alpha \geq 1$, then by our assumptions both $u$ and $f$ are Lipschitz, so \begin{equation*} \begin{aligned} &\left| u(x_i) - u(y) \right| \leq |x_i - y| \; |u|_{C^{0,1}(B_i)} < d\delta |u|_{C^{0,1}(B_i)}, \\ &\left| f(x_i) - f(y) \right| \leq |x_i - y| \; |f|_{C^{0,1}(B_i)} < d\delta |f|_{C^{0,1}(B_i)}, \end{aligned} \end{equation*} which combining with $u(y) = f(y)$ implies that \begin{equation*} \left| \mathcal I_h u(x_i) - f(x_i) \right| = \left| u(x_i) - f(x_i) \right| \leq d\delta \left(|u|_{C^{0,1}(B_i)} + |f|_{C^{0,1}(B_i)} \right). \end{equation*} If $k+ \alpha < 1$, similarly we have \begin{equation*} \left| \mathcal I_h u(x_i) - f(x_i) \right| = \left| u(x_i) - f(x_i) \right| \leq (d\delta)^{\alpha} \left(|u|_{C^{0,\alpha}(B_i)} + |f|_{C^{0,\alpha}(B_i)} \right). \end{equation*} One drawback of this estimate is when $k = 1$, the estimate is $O(\delta)$ instead of $O(\delta^{k+\alpha})$. This is due to the fact that when $y \in \partial\Omega$, even though $u(y) = f(y)$, we do not have $\nabla u(y) = \nabla f(y)$, which must hold if $y$ is in the interior of $\Omega$. \end{remark} \end{comment} \subsection{Discrete Barrier Functions}\label{S:DBarrier} In \Cref{Prop:Consistency-lowreg} (consistency for $u$ with H\"older regularity) we estimate the consistency error for the convex envelope $u \in C^{k,\alpha}(\overline{\Omega})$ for $k=0,1$ and $0<\alpha \leq 1$. In order to take advantage of this result for error analysis, we now introduce two discrete barrier functions. The first one is used to handle those $x_i \in \mathcal{N}_h^0$ far from the contact set $\mathcal{C}(f)$, which satisfy the condition in \Cref{Prop:Consistency-lowreg}(i). The second discrete barrier function is used to handle those $x_i \in \mathcal{N}_h^0$ close to the boundary of $\Omega$, which satisfy the condition in \Cref{Prop:Consistency-lowreg}(iii). First we collect properties of the discrete barrier function $q_h$ introduced in the proof of \Cref{L:DCP} (discrete comparison principle); see also \cite[Lemma 4.1]{LiNo}. \begin{Lemma}[discrete barrier $q_h$]\label{L:Barrier_qh} Let $x_0 \in \Omega$ and $R = \textrm{diam} (\Omega)$. The interpolant $q_h = \mathcal I_h q \in \mathbb{V}_h$ of the function $q(x) = \frac{1}{2}|x-x_0|^2 - \frac{1}{2}R^2$ satisfies \begin{subequations} \begin{align} \label{E:Barrier_qh-1} \sdd q_h(x_i;v_j) \geq 1 \quad &\forall \; x_i \in \mathcal{N}_h^0, \; v_j \in \mathbb{S}, \\ \label{E:Barrier_qh-2} -C \leq \; q_h(x) \; \leq 0 \quad &\forall \; x \in \Omega_h, \end{align} \end{subequations} where constant $C$ only depends on $\Omega$. \end{Lemma} Now we construct our second discrete barrier function $p_h(x)$. For $k=0,1$ and $0<\alpha \leq 1$, $p_h$ is to satisfy the property \begin{equation*} \max_{v_{\theta} \in \mathbb S_{\theta}} \sdd p_h(x_i;v_{\theta}) \geq \delta_i^{k+\alpha-2}, \quad \forall \; x_i \in \mathcal{N}_h^0 \setminus \Omega_{h,d\delta}. \end{equation*} We consider a convex function $\eta: [0,\infty) \rightarrow (-\infty,0]$ satisfying \begin{equation}\label{E:eta-prop-1} \eta''(t) = 2^{4-k-\alpha} \ t^{k+\alpha-2} \quad t \in (0,2d\delta); \quad \eta(0) = 0; \quad \eta'(t) = 0 \quad t \geq 2d\delta. \end{equation} Simple calculations reveal that for $k + \alpha \neq 1$, \begin{equation*} \eta(t) = \left\{ \begin{array}{ll} \frac{2^{4-k-\alpha}}{k+\alpha-1} \left( \frac{1}{k+\alpha}t^{k+\alpha} - (2d\delta)^{k+\alpha-1}t \right) \quad & \quad 0 \leq t \leq 2d\delta\\ -\frac{16}{k+\alpha} (d\delta)^{k+\alpha} \quad & \quad t > 2d\delta, \end{array} \right. \end{equation*} and for $k + \alpha = 1$, \begin{equation*} \eta(t) = \left\{ \begin{array}{ll} 8t \left( \ln{t} - \ln(2d\delta) - 1 \right) \quad & \quad 0 \leq t \leq 2d\delta \\ -16d\delta \quad & \quad t > 2d\delta. \end{array} \right. \end{equation*} It can be seen immediately that $\eta$ is monotonically non-increasing, and satisfies \begin{equation}\label{E:eta-prop-2} -C \delta^{k+\alpha} \leq \eta(t) \leq 0 \qquad \forall t \geq 0. \end{equation} Then we define the barrier function $p_h$ as \begin{equation}\label{E:def_p} p(x) := \eta(\textrm{dist}(x, \partial \Omega_h)) \quad x \in \Omega_h, \end{equation} and denote by $p_h = \mathcal I_h p \in \mathbb{V}_h$ its Lagrange interpolant. The following lemma is similar to \cite[Section 6.2]{NoZh1} and \cite[Lemma 4.2]{LiNo}. \begin{Lemma}[discrete barrier $p_h$]\label{L:Barrier_p} If $\Omega$ is strictly convex and $\theta \leq 1$, then the discrete barrier function $p_h$ defined in \eqref{E:def_p} satisfies \begin{subequations} \begin{gather}\label{E:Barrier_ph-1} \max_{v_{\theta} \in \mathbb S_{\theta}} \sdd p_h(x_i;v_{\theta}) \geq \; C\delta_i^{k+\alpha-2} \quad \forall \; x_i \in \mathcal{N}_h^0 \setminus \Omega_{h,d\delta}, \\ \label{E:Barrier_ph-2} \sdd p_h(x_i;v) \geq \; 0 \quad \forall \; x_i \in \mathcal{N}_h^0, \; v \in \mathbb{S} , \\ \label{E:Barrier_ph-3} -C\delta^{k+\alpha} \leq \; p_h(x) \; \leq 0 \quad \forall \; x \in \Omega_h. \end{gather} \end{subequations} Moreover, for $x_i \in \mathcal{N}_h^0 \setminus \Omega_{h,d\delta}$, we could choose $v_{\theta} \in \mathbb S_{\theta}$ only depending on $x_i, \mathbb S_{\theta}$ to satisfy $\sdd p_h(x_i;v_{\theta}) \geq \; \delta_i^{k+\alpha-2}$. \end{Lemma} \begin{proof} We proceed as in \cite[Lemma 4.2]{LiNo}. We first study the function $p$ defined on the convex domain $\Omega_h \subset \Omega$; the properties of $p_h$ will be simple consequences of those of $p$. Define $d(x) := \textrm{dist}(x,\Omega_h)$ for any $x \in \Omega_h$. Given any $x_0 \in \Omega_h$, let $y \in \partial \Omega_h$ be a (closest) point so that \begin{equation*} |y - x_0| = d(x_0). \end{equation*} Since $\Omega_h$ is convex, there exists a supporting hyperplane $P$ of $\Omega_h$ touching $\Omega_h$ at $y$ and perpendicular to $\nu := \frac{x_0 - y}{|x_0 - y|}$. Consider any two points $x_+, x_- \in\Omega_h$ so that $x_0 = (x_+ + x_-)/2$. Then there exists a vector $v$ such that $x_{\pm} = x_0 \pm v$ and, without loss of generality, $\langle v, \nu \rangle \geq 0$; hence \begin{equation}\label{E:proof-ph} d(x_{\pm}) \leq \textrm{dist}(x_{\pm},P) = d(x_0) \pm \langle v, \nu \rangle. \end{equation} We now show that $p(x)$ is convex. We exploit that $\eta$ is a nonincreasing convex function, and $d(x_0) - \langle v, \nu \rangle \geq 0$, to write \begin{equation*} \begin{aligned} p(x_+) + p(x_-) \geq \; \eta \left( d(x_0) + \langle v, \nu \rangle \right) + \eta \left( d(x_0) - \langle v, \nu \rangle \right) \geq \; 2 \eta \left(d(x_0) \right) = \; 2 p(x_0). \end{aligned} \end{equation*} Since this holds for any $x_{\pm}, x_0$ satisfying $x_0 = (x_+ + x_-)/2$, we deduce that $p(x)$ is convex in $\Omega_h$. This immediately implies \eqref{E:Barrier_ph-2}: \[ \sdd p_h(x_i;v) \geq \; \sdd p(x_i;v) \ge \; 0 \quad \forall \; x_i \in \mathcal{N}_h^0, \; v \in \mathbb{S}. \] We next prove \eqref{E:Barrier_ph-1}. If $x_i \in \mathcal{N}_h^0 \setminus \Omega_{h,d\delta}$, then $\delta_i \leq d(x_i) \leq d\delta_i \le d\delta$ and $d(x_i) \pm \delta_i \in [0,2d(x_i)] \subset [0,2d\delta]$, where $\delta_i \leq \delta$ is defined in \eqref{E:deltai}. It follows from the definition \eqref{E:def_p} of $p$, inequality \eqref{E:proof-ph} and the monotonicity of $\eta$ that \begin{equation*} \begin{aligned} \sdd p_h(x_i;v) \geq & \sdd p(x_i; v) = \frac{p(x_i+ \delta_i v)+p(x_i- \delta_i v)-2p(x_i)}{\delta_i^2} \\ \geq & \frac{\eta \left(d(x_i) + \delta_i \langle v,\nu \rangle \right) + \eta \left(d(x_i) - \delta_i \langle v,\nu \rangle \right) -2\eta \left(d(x_i)\right)}{\delta_i^2}, \end{aligned} \end{equation*} for all $v \in \mathbb{S}$. Using the fact that for $t \in [0,2d(x_i)]$, \begin{equation*} \eta''(t) \geq 2^{4-k-\alpha} \left(2d(x_i)\right)^{k+\alpha-2} = 4d(x_i)^{k+\alpha-2} \geq 4 (d\delta_i)^{k+\alpha-2}, \end{equation*} Taylor expansion gives \begin{equation*} \begin{aligned} \sdd p_h(x_i;v) \ge & \frac{\eta''(\xi) \left(\delta_i \langle v,\nu \rangle \right)^2}{\delta_i^2} \ge \frac{4 (d\delta_i)^{k+\alpha-2} \; \delta_i^2 \langle v,\nu \rangle^2}{\delta_i^2} = 4 \langle v,\nu \rangle^2 (d\delta_i)^{k+\alpha-2}, \end{aligned} \end{equation*} where $\xi \in (0, 2d(x_i))$. By definition of $\mathbb S_{\theta}$, there exists $v_{\theta} \in \mathbb S_{\theta}$ such that $|v_{\theta} - \nu| \leq \theta \leq 1$, whence \begin{equation*} \langle v_{\theta}, \nu \rangle = \frac{|v_{\theta}|^2+|\nu|^2 - |v_{\theta} - \nu|^2}{2} \geq \frac{1}{2}, \end{equation*} which yields $\sdd p_h(x_i;v_{\theta}) \geq 4 \langle v_{\theta},\nu \rangle^2 (d\delta_i)^{k+\alpha-2} \geq C\delta_i^{k+\alpha-2} $. This proves \eqref{E:Barrier_ph-1}, whereas \eqref{E:Barrier_ph-3} is a direct consequence of \eqref{E:eta-prop-2}. \end{proof} \begin{remark}[boundary resolution] Notice that we only assume $\theta \leq 1$ here. Our two-scale method can actually be generalized in such a way that each $x_i \in \mathcal{N}_h^0$ has a different choice of $\mathbb S_{\theta}(x_i)$. In fact, in our derivation of error estimate later, for those $x_i$ with $\textrm{dist}(x_i,\partial\Omega) < d\delta$, we only require the $\mathbb S_{\theta}(x_i)$ to satisfy requirements of discretization for $\theta \leq 1$. This means in practice, for nodes near the boundary $\partial\Omega$, we do not need as many directions as for the nodes in the interior region. \end{remark} \subsection{Error Estimates for Solutions with H\"older Regularity}\label{S:RatesHolder} In this subsection we deal with solutions $u$ of \eqref{E:pde-CE} of class $C^{k,\alpha}(\overline{\Omega})$ for $k=0,1$ and $0<\alpha \leq 1$, and derive convergence rates in the $L^{\infty}$ norm. Our main analytic tool is \Cref{L:DCP} (discrete comparison principle), along with the results of Sections \ref{S:Consistency-lowreg} and \ref{S:DBarrier}. \begin{Theorem}[error estimate]\label{T:error-estimate} Let $\Omega$ be strictly convex. Let $u$ be the viscosity solution of \eqref{E:pde-CE} and $u_{\varepsilon}$ be the discrete solution of \eqref{E:2ScOp}. If $u \in C^{k,\alpha}(\overline{\Omega})$ for $k=0,1$ and $0<\alpha \leq 1$, and $\theta \leq 1$, there exists a constant $C = C(\Omega,d,\sigma)$ such that {\small \begin{equation}\label{E:error-estimate} \Vert \mathcal I_h u - u_{\varepsilon} \Vert_{L^{\infty}(\Omega_h)} \leq C \left[ |u|_{C^{k,\alpha}(\overline{\Omega})} \frac{(\delta \theta)^{k+\alpha} + h^{k+\alpha}+ \delta^{2+k+\alpha}}{\delta^2} + |f|_{C^{k,\alpha}(\overline{\Omega})} \delta^{k+\alpha} \right]. \end{equation}} \end{Theorem} \begin{proof} We find lower and upper bounds of $u_{\varepsilon}$ in terms of $\mathcal I_h u$. For the lower bound, we recall that $u_h^- = \mathcal I_h u$ is a discrete subsolution of \eqref{E:2ScOp} and satisfies $u_h^- \leq u_{\varepsilon}$ from \eqref{E:disc-sol-bounds} in the proof of \Cref{L:Exist-Uniq-Stab} (existence, uniqueness and stability), thereby yielding a lower bound of $u_{\varepsilon}$. For the upper bound, we construct a discrete supersolution $u_h^+ \in \mathbb{V}_h$ such that \begin{equation*} \left\{ \begin{aligned} T_{\varepsilon}[u_h^+;f](x_i) &\leq 0 \quad \forall x_i \in \mathcal{N}_h^0 \\ u_h^+(x_i) &\geq f(x_i) \quad \forall x_i \in \mathcal{N}_h^b, \end{aligned} \right. \end{equation*} upon suitably modifying $\mathcal I_h u$. We let $u_h^+ \in \mathbb{V}_h$ be of the form \begin{equation*} u_h^+ = \mathcal I_h u - K_1 q_h + K_2 - K_3 p_h, \end{equation*} where $q_h, p_h \leq 0$ in $\Omega_h$ according to \eqref{E:Barrier_qh-2} and \eqref{E:Barrier_ph-3}, and the positive constants $K_1, K_2, K_3$ are to be chosen properly. Since \begin{equation*} u_h^+(x_i) \geq \mathcal I_h u(x_i) = f(x_i) \quad\forall \; x_i \in \mathcal{N}_h^b, \end{equation*} to guarantee that $u_h^+$ is a discrete supersolution, it remains to show $T_{\varepsilon}[u_h^+;f](x_i) \leq 0$ for all $x_i \in \mathcal{N}_h^0$. We divide the subsequent discussion into three cases based on the position of $x_i$ relative to $\mathcal{C}(f)$ and $\partial\Omega$, exactly as in \Cref{Prop:Consistency-lowreg}. If $\textrm{dist}(x_i,\mathcal{C}(f)) \geq d \delta$, using the estimate \eqref{E:error-noncontact} of \Cref{Prop:Consistency-lowreg} (consistency for $u$ with H\"older regularity) and the properties \eqref{E:Barrier_qh-1} of $q_h$ and \eqref{E:Barrier_ph-2} of $p_h$, we have \begin{equation*} \begin{aligned} \min_{v \in \mathbb S_{\theta}} \sdd u_h^+(x_i;v) &\leq \min_{v_{\theta} \in \mathbb S_{\theta}} \sdd [\mathcal I_h u - K_1 q_h](x_i;v) \leq \min_{v_{\theta} \in \mathbb S_{\theta}} \sdd \mathcal I_h u(x_i;v) - K_1 \\ &\leq C(d,\sigma) \frac{(\delta \theta)^{k+\alpha} + h^{k+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(B_i)} - K_1 \leq 0, \end{aligned} \end{equation*} provided that $K_1 = C(d,\sigma) \frac{(\delta \theta)^{k+\alpha} + h^{k+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(\overline{\Omega})}$. Consequently, \begin{equation*} T_{\varepsilon}[u_h^+;f](x_i) \leq \min_{v \in \mathbb S_{\theta}} \sdd u_h^+(x_i;v) \leq 0. \end{equation*} If $\; \textrm{dist}(x_i,\mathcal{C}(f)) < d \delta, \; \textrm{dist}(x_i,\partial\Omega) \geq d\delta$, from \eqref{E:error-contact-0} and \eqref{E:error-contact-1} in \Cref{Prop:Consistency-lowreg}, we have \begin{equation*} \begin{aligned} f(x_i) - u_h^+(x_i) &\leq f(x_i) - \mathcal I_h u(x_i) - K_2 \\ &\leq C(d,\sigma) \delta^{k+\alpha} \left( |u|_{C^{k,\alpha}(\wt{B_i})} + |f|_{C^{k,\alpha}(\wt{B_i})} \right) - K_2 \leq 0, \end{aligned} \end{equation*} with $K_2 = C(d,\sigma) \delta^{k+\alpha} \left( |u|_{C^{k,\alpha}(\overline{\Omega})} + |f|_{C^{k,\alpha}(\overline{\Omega})} \right)$. This implies $T_{\varepsilon}[u_h^+;f](x_i) \leq f(x_i) - u_h^+(x_i) \leq 0$. If $\; \textrm{dist}(x_i,\partial\Omega) < d\delta$, we have $x_i \in \mathcal{N}_h^0 \setminus \Omega_{h,d\delta}$. Choosing $K_3 = C(d,\sigma)|u|_{C^{k,\alpha}(\overline{\Omega})}$ and invoking \eqref{E:error-boundary} in \Cref{Prop:Consistency-lowreg} and the property \eqref{E:Barrier_ph-1} of $p_h$, we have \begin{equation*} \begin{aligned} \small \min_{v \in \mathbb S_{\theta}} \; \sdd u_h^+(x_i;v) &\leq \min_{v \in \mathbb S_{\theta}} \; \sdd [\mathcal I_h u - K_3 p_h](x_i;v) \\ &\leq C(d,\sigma) \delta_i^{k+\alpha-2} |u|_{C^{k,\alpha}(B_i)} - K_3 \max_{v \in \mathbb S_{\theta}} \sdd p_h(x_i;v) \\ & \leq C(d,\sigma) \delta_i^{k+\alpha-2} |u|_{C^{k,\alpha}(B_i)} - C(d,\sigma)|u|_{C^{k,\alpha}(\overline{\Omega})} \; \delta_i^{k+\alpha-2} \le 0. \end{aligned} \end{equation*} Therefore $T_{\varepsilon}[u_h^+;f](x_i) \leq \min_{v \in \mathbb S_{\theta}} \sdd u_h^+(x_i;v) \leq 0$. The three cases show that $u_h^+$ is a discrete supersolution, and thus by \Cref{L:DCP} (discrete comparison principle), \begin{equation*} \begin{aligned} u_{\varepsilon} \leq \; & \mathcal I_h u - K_1 q_h + K_2 - K_3 p_h \\ = \; & \mathcal I_h u + C(d,\sigma, \Omega) \frac{(\delta \theta)^{k+\alpha} + h^{k+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(\overline{\Omega})} \\ &+ C(d,\sigma) \delta^{k+\alpha} \left( |u|_{C^{k,\alpha}(\overline{\Omega})} + |f|_{C^{k,\alpha}(\overline{\Omega})} \right) + C(d,\sigma)|u|_{C^{k,\alpha}(\overline{\Omega})} \delta^{k+\alpha}. \end{aligned} \end{equation*} This, conjunction with the lower bound of $u_{\varepsilon}$, completes the proof. \end{proof} \begin{Corollary}[convergence rate]\label{C:convergence-rate} Let $\Omega$ be strictly convex. Let $u$ be the viscosity solution of \eqref{E:pde-CE} and $u_{\varepsilon}$ be the discrete solution of \eqref{E:2ScOp}. If $u \in C^{k,\alpha}(\overline{\Omega})$ for $k=0,1$ and $0<\alpha \leq 1$, and $\theta \leq 1$, we have \begin{equation}\label{E:convergence-rate} \|u-u_{\varepsilon}\|_{L^\infty(\Omega_h)} \le C(\Omega,d,\sigma) \Big( |u|_{C^{k,\alpha}(\overline{\Omega})} + |f|_{C^{k,\alpha}(\overline{\Omega})} \Big) \; h^{\frac{(k+\alpha)^2}{2+k+\alpha}}, \end{equation} provided $R_\alpha(u) := |u|_{C^{k,\alpha}(\overline{\Omega})}^{\frac{1}{2+k+\alpha}} \Big(|u|_{C^{k,\alpha}(\overline{\Omega})} + |f|_{C^{k,\alpha}(\overline{\Omega})} \Big)^{-\frac{1}{2+k+\alpha}}$ and \begin{equation*} \delta = R_\alpha(u) h^{\frac{k+\alpha}{2+k+\alpha}}, \quad \theta = R_\alpha(u)^{-1} h^{\frac{2}{2+k+\alpha}}. \end{equation*} \end{Corollary} \begin{proof} Since the pointwise interpolation error satisfies \cite{BrennerScott} \begin{equation*} \|u - \mathcal I_h u\|_{L^\infty(\Omega_h)} \le C h^{k+\alpha} |u|_{C^{k,\alpha}(\overline{\Omega})} \le C \frac{h^{k+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(\overline{\Omega})}, \end{equation*} and $h \le \delta$, we end up with the error estimate \begin{equation*} \|u - u_{\varepsilon}\|_{L^\infty(\Omega_h)} \le C \left[ |u|_{C^{k,\alpha}(\overline{\Omega})} \frac{h^{k+\alpha} + (\delta \theta)^{k+\alpha}}{\delta^2} + \Big(|u|_{C^{k,\alpha}(\overline{\Omega})} + |u|_{C^{k,\alpha}(\overline{\Omega})}\Big) \delta^{k+\alpha} \right]. \end{equation*} In order to balance all contributions, we first choose $\theta=\frac{h}{\delta}$ and next equate the two terms on the right-hand side to obtain the asserted relations between $\delta,\theta$ and $h$. This completes the proof. \end{proof} \begin{remark}[two important scenarios]\label{R:two-scenarios} We want to point out two important scenarios based on the regularity of $u$ for \Cref{C:convergence-rate} (convergence rate). \begin{enumerate}[$\bullet$] \item Full regularity $u \in C^{1,1}(\overline{\Omega})$, i.e. $k = \alpha = 1$. The optimal choice of parameters $\delta \sim O(h^{1/2}), \theta \sim O(h^{1/2})$ in \Cref{C:convergence-rate} yields either a linear decay rate $O(h)$ or a quadratic rate $O(\delta^2)$ in terms of the fine scale $h$ or the coarse scale $\delta$. \item Lipschitz regularity $u \in C^{0,1}(\overline{\Omega})$, i.e. $k = 0, \ \alpha = 1$. Choosing optimal parameters $\delta \sim O(h^{1/3}), \theta \sim O(h^{2/3})$ in \Cref{C:convergence-rate} gives us either a rate $O(h^{1/3})$ in terms of the fine scale $h$ or a linear rate $O(\delta)$ in terms of the coarse scale $\delta$. \end{enumerate} We point out that, since $|u|_{C^{0,1}(\overline{\Omega})} \lesssim |f|_{C^{1,1}(\overline{\Omega})}$ and $|u|_{C^{1,1}(\overline{\Omega})} \lesssim |f|_{C^{3,1}(\overline{\Omega})}$ under proper assumptions of $\Omega$ \cite{DeFi}, the right hand side of \eqref{E:convergence-rate} can be bounded with only norms of $f$. Our error estimates are thus realistic in terms of regularity. \end{remark} \begin{remark}[fine scale vs regularity] It is instructive to realize that the coarse scale $\delta$ gets finer with increasing regularity $k+\alpha$ of $u$, whereas the angular scale $\theta$ gets coarser. This behavior is opposite to the error estimates in \cite[Remark 5.4]{LiNo}. \end{remark} \begin{remark}[alternate proof]\label{R:alternate-proof-k0} When $k=0$, the proof of \Cref{T:error-estimate} (error estimate) can be simplified a little bit. To be more specific, we can construct a discrete supersolution $u_h^+ \in \mathbb{V}_h$ of the form \begin{equation*} u_h^+ = \mathcal I_h u - K_1 q_h + K_2 \end{equation*} provided that \begin{equation*} K_1 = C(d,\sigma) \frac{(\delta \theta)^{\alpha} + h^{\alpha}}{\delta^2} |u|_{C^{0,\alpha}(\overline{\Omega})}, \quad K_2 = C(d,\sigma) \delta^{\alpha} \left( |u|_{C^{0,\alpha}(\overline{\Omega})} + |f|_{C^{0,\alpha}(\overline{\Omega})} \right). \end{equation*} This is due to the fact that if $0 < \textrm{dist}(x_i,\partial\Omega) < d\delta$, then invoking \eqref{E:error-contact-0} with our choice of $K_2$ implies $T_{\varepsilon}[u_h^+;f](x_i) \leq 0$. \end{remark} \subsection{Non-attainment of Dirichlet condition}\label{S:non-attainment} Although we mainly focus on the case that the domain $\Omega$ is strictly convex, it is also possible to modify and extend our two-scale method to compute the convex envelope over {\it convex polytopes} $\Omega$, thus domains with piecewise linear boundary. For simplicity, we only explain the ideas in ${\mathbb{R}}^2$, but higher dimensions $d>2$ can be dealt with in a similar manner. We need additional notation. A convex polytope $\Omega$ can be described by a set $\mathcal{N}^v$ of vertices on its boundary; thus $\Omega = \textrm{conv} (\mathcal{N}^v)$. We then let $\mathcal{N}^e = \partial\Omega \setminus \mathcal{N}^v$ be the set of boundary edges of $\Omega$ excluding vertices. While $u = f$ is no longer true on $\partial\Omega$ if $\Omega$ is not strictly convex, it can be shown using \cite[Corollary 17.1.5]{Rockafellar2015convex} that $u = f$ at vertices of $\mathcal{N}^v$, and on each edge of $\mathcal{N}^e$, the function $u$ is the convex envelope of $f$ restricted to that edge. One can thus show that $u$ is the viscosity solution of the following fully nonlinear obstacle problem: \begin{equation}\label{E:pde-CE-polytope} \left\{ \begin{array}{ll} T[u;f](x) = 0 \quad \; & \forall x \in \Omega, \\ \min\left\{f(x) - u(x), e^T(x) D^2u(x) e(x) \right\} = 0\quad \; & \forall x \in \mathcal{N}^e, \\ u(x) = f(x) \quad \; & \forall x \in \mathcal{N}^v, \end{array} \right. \end{equation} where $e(x)$ is a unit vector parallel to the edge of $\Omega$ containing $x \in \mathcal{N}^e$; note that \eqref{E:pde-CE-polytope} is a modification of \eqref{E:pde-CE} on $\partial\Omega$. To discretize this system, let $\mathcal{N}_h^v := \mathcal{N}^v \subset \mathcal{N}_h^b$ and $\mathcal{N}_h^e := \mathcal{N}_h^b \cap \mathcal{N}^e$, then our discrete problem is to find $u_{\varepsilon} \in \mathbb{V}_h$ satisfying \begin{equation}\label{E:2ScOp-Ex4} \left\{ \begin{array}{ll} T_{\varepsilon}[u_{\varepsilon};f](x_i) = 0 \quad \; & \forall x_i \in \mathcal{N}_h^0,\\ \min\left\{f(x_i) - u_{\varepsilon}(x_i), \sdd u_{\varepsilon}(x_i, e(x_i)) \right\} = 0\quad \; & \forall x_i \in \mathcal{N}_h^e, \\ u_{\varepsilon}(x_i) = f(x_i) \quad \; & \forall x_i \in \mathcal{N}_h^v, \end{array} \right. \end{equation} where the step size of $\sdd u_{\varepsilon}(x_i, e(x_i))$ should be defined as the maximum number $\delta_i$ in $(0,\delta]$ such that $x_i \pm \delta_i e(x_i)$ are both inside $\overline{\Omega}$. The convergence of $u_{\varepsilon}$ can be derived in a similar way to \Cref{S:TwoSc}. We now prove an error estimate. \begin{Proposition}[convergence rate for polytopes]\label{L:conv-rate-polytopes} Let $\Omega$ be a convex polytope and $u \in C^{k,\alpha}(\overline{\Omega})$ with $k=0,1, \ 0<\alpha \leq 1$, and $\theta \leq 1$. Let $u_{\varepsilon}\in\mathbb{V}_h$ be the discrete solution of \eqref{E:2ScOp-Ex4}. If the discretization parameters $\varepsilon = (h,\delta,\theta)$ obey relations similar to those in \Cref{C:convergence-rate} (convergence rate), then \[ \| u - u_{\varepsilon} \|_{L^\infty(\Omega)} \le C(u,\Omega,d,\sigma) \, h^{\frac{(k+\alpha)^2}{2+k+\alpha}}. \] \end{Proposition} \begin{proof} We first notice that $\Omega_h=\Omega$ and that \Cref{L:DCP} (discrete comparison principle) implies the following stability result: if $u_h, w_h \in \mathbb{V}_h$ satisfy $T_{\varepsilon}[u_h;f](x_i) = T_{\varepsilon}[w_h;f](x_i)$ for all $x_i \in \mathcal{N}_h^0$, then \begin{equation}\label{E:stability} \max_{x_i \in \mathcal{N}_h} \left| u_h(x_i) - w_h(x_i) \right| \leq \max_{x_i \in \mathcal{N}_h^b} \left| u_h(x_i) - w_h(x_i) \right|. \end{equation} We consider an auxiliary discrete problem: seek $\wt{u}_{\varepsilon} \in \mathbb{V}_h$ that solves \begin{equation*} \left\{ \begin{array}{ll} T_{\varepsilon}[\wt{u}_{\varepsilon};f](x_i) = 0 \quad \; & \forall x_i \in \mathcal{N}_h^0,\\ \wt{u}_{\varepsilon}(x_i) = u(x_i) \quad \; & \forall x_i \in \mathcal{N}_h^b. \end{array} \right. \end{equation*} We observe that \Cref{C:convergence-rate} still holds for $\wt{u}_{\varepsilon}$, without the strict convexity assumption on $\Omega$, because the Dirichlet boundary is attained. Therefore, choosing $\delta$ and $\theta$ as in \Cref{C:convergence-rate}, we obtain \begin{equation*} \Vert u - \wt{u}_{\varepsilon} \Vert_{L^{\infty}(\Omega_h)} \leq C(u,\Omega,d,\sigma) \, h^{\frac{(k+\alpha)^2}{2+k+\alpha}}. \end{equation*} It remains to estimate $\Vert \wt{u}_{\varepsilon} - u_{\varepsilon} \Vert_{L^{\infty}(\Omega_h)}$, for which we resort to \eqref{E:stability} because both $\wt{u}_{\varepsilon},u_{\varepsilon}\in\mathbb{V}_h$. Since the boundary subsystem \begin{equation*} \left\{ \begin{array}{ll} \min\left\{f(x_i) - u_{\varepsilon}(x_i), \sdd u_{\varepsilon}(x_i, e(x_i)) \right\} = 0\quad \; & \forall x_i \in \mathcal{N}_h^e, \\ u_{\varepsilon}(x_i) = f(x_i) \quad \; & \forall x_i \in \mathcal{N}_h^v, \end{array} \right. \end{equation*} can be viewed as several one dimensional two-scale discretizations of the convex envelope problem, \Cref{C:convergence-rate} again implies \begin{equation*} \max_{x_i \in \mathcal{N}_h^b} \left| \wt{u}_{\varepsilon}(x_i) - u_{\varepsilon}(x_i) \right| = \max_{x_i \in \mathcal{N}_h^b} \left| u(x_i) - u_{\varepsilon}(x_i) \right| \leq C(u,\Omega,d,\sigma) \, h^{\frac{(k+\alpha)^2}{2+k+\alpha}}. \end{equation*} This concludes the proof. \end{proof} It is worth pointing out that we may not need a two-scale structure on the boundary since it reduces to a one dimensional problem on the edge of a polytope in 2D. However, notice that this procedure extends to dimensions $d>2$, and in such case boundary subproblems possess dimension higher than one and require a two-scale structure. \section{Modified Wide Stencil Method}\label{S:modified-wide-stencil} Our numerical analysis of the previous sections could be applied to derive error estimates for a modified wide stencil method obtained upon adding a two-scale structure into that of \cite{Ob2}. Since key ideas and techniques are identical to those for the two-scale method, we present them without proofs. First let us briefly introduce the wide stencil method in a way convenient to our analysis; we refer the readers to \cite{Ob2} and \cite{ObRu} for more details. For a strictly convex domain $\Omega \subset \mathbb{R}^d$, with abuse of notations, let $\mathcal{N}_h^0 := \Omega \cap h\mathbb{Z}^d$ be a Cartesian grid in $\Omega$, and $\mathbb{V}_h$ be the space consisting of all maps $u_h: \mathcal{N}_h^0 \cup \partial\Omega \rightarrow \mathbb{R}$. Let a coarse scale $\delta \ge \sqrt{d}{h}$ be used to define the set of discrete directions \[ D_{\varepsilon} := \left\{ x \in h\mathbb{Z}^d: \textrm{dist}\big(x, \partial B(0,\delta) \big) \le \frac{\sqrt{d}}{2}h \right\}, \] where $\varepsilon := (h,\delta)$ and $B(0,\delta)$ is the ball centered at the origin with radius $\delta$. It is worth pointing out that $D_{\varepsilon}$ is just a few layers of grid points, and thus its cardinality satisfies $\#D_{\varepsilon} \lesssim \left(\frac{\delta}{h}\right)^{d-1}$. The following lemma is similar to \cite[Lemma 4.4]{DolzWalk} and characterizes the consistency error due to using $D_{\varepsilon}$ instead of $\partial B(0,\delta)$. \begin{Lemma}[properties of $D_{\varepsilon}$]\label{L:consist-Dve} For any $v \in \partial B(0,\delta)$, there exists $v_{\varepsilon} \in D_{\varepsilon}$ such that the angle between the vectors $v$ and $v_{\varepsilon}$ is bounded by $\frac{\sqrt{d}\pi h}{4\delta}$. Moreover, $\frac{\delta}{2}\le |v| \le \frac{3\delta}{2}$ for all $v\in D_\varepsilon$. \end{Lemma} \begin{proof} Choose a Cartesian grid point in $v_{\varepsilon} \in h\mathbb{Z}^d$ closest to $v$, which in turn must satisfy $|v - v_{\varepsilon}| \le \frac{\sqrt{d}h}{2}$, whence $v_{\varepsilon} \in D_{\varepsilon}$. The angle $\theta$ between $v$ and $v_{\varepsilon}$ is dictated by \looseness=-1 \[ \sin\theta \le \frac{|v - v_{\varepsilon}|}{\delta} \le \frac{\sqrt{d}}{2}\frac{h}{\delta}. \] This implies $\theta \leq \frac{\pi}{2}\sin\theta \le \frac{\sqrt{d}\pi h}{4\delta}$. Moreover, by definition of $D_\varepsilon$ we see that $\frac{\delta}{2} \le \delta - \frac{\sqrt{d}}{2}h \le |v| \le \delta + \frac{\sqrt{d}}{2}h \le \frac{3\delta}{2}$ for all $v\in D_\varepsilon$. \end{proof} For any function $w \in \mathbb{V}_h$ and any vector $v \in D_{\varepsilon}$, let the centered second difference operator at any $x_i \in \mathcal{N}_h^0$ in the direction $v$ be \begin{equation*} \nabla^2_\varepsilon w(x_i;v) := \frac{2}{\left(\rho_{+} + \rho_{-}\right)|v|^2} \left( \frac{w(x_i + \rho_{+}v) - w(x_i)}{\rho_{+}} + \frac{w(x_i - \rho_{-}v) - w(x_i)}{\rho_{-}} \right), \end{equation*} where $\rho_{\pm}$ are the biggest numbers in $(0, 1]$ such that $x_i \pm \rho_{\pm} v \in \overline{\Omega}$. Notice that this is well-defined for any $w \in \mathbb{V}_h$ because $x_i \pm \rho_{\pm} v$ are either in $\mathcal{N}_h^0$ or on the boundary $\partial\Omega$. Since for any $v \in D_{\varepsilon}$ we have $\frac{\delta}{2} \le |v| \le \frac{3\delta}{2}$, the parameter $\delta$ plays a role similar to the coarse scale $\delta$ for second differences in our two-scale method. The cardinalities $\#D_{\varepsilon} \approx (\delta/h)^{d-1}$ and $\#\mathbb S_{\theta} \approx \theta^{-(d-1)}$ are consistent provided $\theta \approx h/\delta$. We define the discrete operator for the modified wide stencil method to be \begin{equation*} T_{\varepsilon}[w;f](x_i) := \min\left\{f(x_i) - w(x_i), \min_{v \in D_{\varepsilon}} \nabla^2_\varepsilon w(x_i;v) \right\} \quad\forall \, x_i \in \mathcal{N}_h^0 \end{equation*} for any $w \in \mathbb{V}_h$. Finally, the discrete problem reads: find $u_{\varepsilon} \in \mathbb{V}_h$ such that \begin{equation}\label{E:2ScOp-WD} T_{\varepsilon}[u_{\varepsilon};f](x_i) = 0 \quad\forall \, x_i \in \mathcal{N}_h^0, \end{equation} and $u_{\varepsilon}(x) = f(x)$ for any $x \in \partial\Omega$. It is now easy to check that \Cref{L:DCP} (discrete comparison principle) and \Cref{Prop:Consistency-lowreg} (consistency for $u$ with H\"older regularity) are valid verbatim in the present context, except that instead of \eqref{E:error-noncontact} we now have \[ \min_{v \in D_\varepsilon} \nabla^2_\varepsilon w(x_i;v) \le C(d,\sigma) \frac{h^{k+\alpha}}{\delta^2} |u|_{C^{k,\alpha}(B_i)}. \] In fact, the modified wide stencil method can be viewed as a modified version of two-scale method without interpolation error and $\theta \approx h/\delta$. The following error estimate mimics that in \Cref{S:RatesHolder}. It is a consequence of the discrete comparison principle and consistency for the wide stencil method together with the discrete barrier functions of \Cref{S:DBarrier}. We omit its proof. \begin{Theorem}[error estimate for the wide stencil method]\label{T:error-estimate-WD} Let $\Omega$ be strictly convex. Let $u$ be the viscosity solution of \eqref{E:pde-CE} and $u_{\varepsilon}$ be the discrete solution of \eqref{E:2ScOp-WD}. If $u \in C^{k,\alpha}(\overline{\Omega})$ for $k=0,1$ and $0<\alpha \leq 1$, then the following error estimate holds \begin{equation*}\label{E:error-estimate-WD} \left| u(x_i) - u_{\varepsilon}(x_i) \right| \leq C \left( |u|_{C^{k,\alpha}(\overline{\Omega})} \frac{h^{k+\alpha}+ \delta^{2+k+\alpha}}{\delta^2} + |f|_{C^{k,\alpha}(\overline{\Omega})} \delta^{k+\alpha} \right) \quad \forall x_i \in \mathcal{N}_h^0, \end{equation*} with $C = C(\Omega,d,\sigma)$. If $\delta := |u|_{C^{k,\alpha}(\overline{\Omega})}^{\frac{1}{2+k+\alpha}} \Big(|u|_{C^{k,\alpha}(\overline{\Omega})} + |f|_{C^{k,\alpha}(\overline{\Omega})} \Big)^{-\frac{1}{2+k+\alpha}} h^{\frac{k+\alpha}{2+k+\alpha}}$, we thus obtain the convergence rate \begin{equation*}\label{E:convergence-rate-WD} \left| u(x_i) - u_{\varepsilon}(x_i) \right| \leq C(\Omega,d,\sigma) \Big( |u|_{C^{k,\alpha}(\overline{\Omega})} + |f|_{C^{k,\alpha}(\overline{\Omega})} \Big) \; h^{\frac{(k+\alpha)^2}{2+k+\alpha}} \quad \forall x_i \in \mathcal{N}_h^0. \end{equation*} \end{Theorem} We point out that \Cref{R:two-scenarios} (two important scenarios) applies in this context. In particular, the convergence rate is of order $O(h)$ provided $\delta = O(h^{1/2})$ for functions $u \in C^{1,1}(\overline{\Omega})$. \section{Numerical Experiments}\label{S:Exp} To solve the discrete system \eqref{E:2ScOp}, we use Howard's algorithm which converges superlinearly. We implemented the 2-scale method within MATLAB, using some of the routines provided by the software FELICITY \cite{Walker1, Walker2}. \subsection{Howard's Algorithm}\label{S:Howard-Algorithm} For convenience, let us order the nodes in $\mathcal{N}_h = \{ x_1, \ldots, x_N\}$ with $x_i \in \mathcal{N}_h^0$ for $1 \leq i \leq N_0$ and $x_i \in \mathcal{N}_h^b$ for $N_0+1 \leq i \leq N$; thus $N, N_0$ and $N_b := N-N_0$ are the cardinality of $\mathcal{N}_h, \mathcal{N}_h^0$ and $\mathcal{N}_h^b$ respectively. In addition, let $\bm{u} := (u_h(x_i))_{i=1}^N \in \mathbb{R}^N$ stand for the vector of nodal values of a generic $u_h \in \mathbb{V}_h$, and $\mathbb S_{\theta} = \left\{v_1,\ldots,v_{S} \right\}$, where $S$ is the cardinality of $\mathbb S_{\theta}$. In view of the expression \eqref{E:disc-oper} for the discrete operator $T_{\varepsilon}$, the discrete system \eqref{E:2ScOp} reads \begin{equation}\label{E:Howard-disc-system} \sup_{\bm{\alpha} \in \mathcal{A}} \left( B^{\bm{\alpha}} \bm{u} - F^{\bm{\alpha}} \right) = \bm{0}, \end{equation} where $\mathcal{A} = \left\{(\alpha_1,\ldots,\alpha_{N_0}): \alpha_i \in \{j\}_{j=0}^S \right\}$, matrix $B^{\bm{\alpha}}\in\mathbb{R}^{N\times N}$ satisfies \begin{equation*} \left( B^{\bm{\alpha}} \bm{u} \right)_i = \left\{ \begin{array}{ll} u_h(x_i) \quad & i \geq N_0+1,\; 0 \le \alpha_i \le S \\ u_h(x_i) \quad & 1 \leq i \leq N_0,\;\alpha_i = 0, \\ -\sdd u_h(x_i;v_{\alpha_i}) \quad & 1 \leq i \leq N_0,\;1 \le \alpha_i \le S, \end{array} \right. \end{equation*} and $F^{\bm{\alpha}}$ is given by \begin{equation*} \left( F^{\bm{\alpha}} \right)_i = \left\{ \begin{array}{ll} f(x_i) \quad & i \geq N_0+1, \;0 \le \alpha_i \le S \\ f(x_i) \quad & 1 \leq i \leq N_0,\;\alpha_i = 0, \\ 0 \quad & 1 \leq i \leq N_0,\; 1 \le \alpha_i \le S. \end{array} \right. \end{equation*} We solve \eqref{E:Howard-disc-system} via the Howard's algorithm \cite{BoMaZi}, which is a semi-smooth Newton method \cite{BoMaZi,HIK:2002,SmearsSuli2014,Ulbrich:2011} also known as policy iteration in the financial literature \cite{PutermanBrumelle1979}: \begin{algorithm} \caption{(Howard's Algorithm) \label{alg:Howard}} \begin{algorithmic}[1] \State Select an arbitrary initial $\bm{\alpha}_0 \in \mathcal{A}$, and let $n=0$. \While{} \State Let $\bm{u}_{n}$ be the solution of the linear equations $B^{\bm{\alpha}_n} \bm{u}_{n} - F^{\bm{\alpha}_n} = \bm{0}$. \State Let $\bm{\alpha}_{n+1} = \textrm{arg\,max}_{\bm{\alpha} \in \mathcal{A}} \left( B^{\bm{\alpha}} \bm{u}_{n} - F^{\bm{\alpha}} \right)$. \State If $\bm{\alpha}_{n+1} = \bm{\alpha}_n$, stop; else $n = n+1$. \EndWhile \end{algorithmic} \end{algorithm} \noindent Hereafter, the vector equality in \eqref{E:Howard-disc-system} and inequalities $\ge$ later are understood componentwise. We could immediately see from the above that for any $\bm{\alpha} \in \mathcal{A}$, we have $\left(B^{\alpha}\right)_{ii} > 0$ and $\left(B^{\alpha}\right)_{ij} \leq 0$ for $i \neq j$. In fact, we prove that $B^{\bm{\alpha}}$ is an M-matrix. \begin{Lemma}[M-matrix property]\label{L:M-Matrix} For any $\bm{\alpha} \in \mathcal{A}$, $B^{\bm{\alpha}}$ is an M-matrix. \end{Lemma} \begin{proof} We only need to prove $B^{\bm{\alpha}} \bm{u} \geq \bm{0}$ implies $\bm{u} \geq \bm{0}$. Given two vectors $\bm{u},\bm{w} \in \mathbb{R}^{N}$ so that $B^{\bm{\alpha}} \bm{u} \ge B^{\bm{\alpha}} \bm{w}$ for all $\bm{\alpha} \in \mathcal{A}$, we deduce $u_h \geq w_h$ for the corresponding functions $u_h,w_h\in\mathbb{V}_h$ in view of \Cref{L:DCP} (discrete comparison principle). This immediately implies $\bm{u}\ge\bm{w}$, and, upon taking $\bm{w}=\bm{0}$, that $\bm{u}\ge\bm{0}$ as desired. \end{proof} Invoking the fact that $B^{\bm{\alpha}}$ is an M-matrix and applying \cite[Theorem 2.1]{BoMaZi}, we deduce that the $n$-th iterate $\bm{u}_n$ of Howard's algorithm converges monotonically and superlinearly to $u_{\varepsilon}$ as $n \to \infty$. The latter follows from the semi-smooth Newton structure of Algorithm \ref{alg:Howard}. The former is a consequence of its step 4 because \[ B^{\bm{\alpha}_{n+1}} \bm{u}_{n} - F^{\bm{\alpha}_{n+1}} \ge B^{\bm{\alpha}_{n}} \bm{u}_{n} - F^{\bm{\alpha}_{n}} = \bm{0} = B^{\bm{\alpha}_{n+1}} \bm{u}_{n+1} - F^{\bm{\alpha}_{n+1}} , \] whence $\bm{u}_{n+1} \le \bm{u}_n$. Moreover, \cite[Theorem 2.1]{BoMaZi} automatically gives existence and uniqueness of our discrete system \eqref{E:2ScOp}, which we also proved in \Cref{L:Exist-Uniq-Stab} (existence, uniqueness and stability). In practice, when $\Vert \sup_{\bm{\alpha} \in \mathcal{A}} \left( B^{\bm{\alpha}} \bm{u}_n - F^{\bm{\alpha}} \right) \Vert_2$ is sufficiently small we can stop Algorithm \ref{alg:Howard}; we thus use the criterion \begin{equation*} \Vert T_{\varepsilon}[u_n;f] \Vert_{L^2(\Omega)} \leq 10^{-10} \Vert T_{\varepsilon}[f;f] \Vert_{L^2(\Omega)} \end{equation*} in all numerical experiments below. \subsection{Accuracy}\label{S:Accuracy} We now present several examples to examine the performance of the two-scale method \eqref{E:2ScOp} for the convex envelope problem. We choose $\delta = C_{\delta} h^{\alpha}$ and $\theta = C_{\theta} h^{\beta}$ for different $C_{\delta}, \alpha, C_{\theta}, \beta > 0$ in our experiments, and compare the computational rates with our theoretical rate of \Cref{C:convergence-rate} (convergence rate). \begin{example}[full regularity $u \in C^{1,1}(\overline{\Omega})$] \label{Ex:ex1} Let $\Omega = \{x\in \mathbb{R}^2: |x|<1 \}$ be the unit circle and $f(\bm{x}) = \cos(2 \pi |\bm{x}|)$. Then the convex envelope $u$ is given by \begin{equation*} u(x) = \left\{ \begin{array}{ll} 0 \;, & \text{if} \quad |x| \leq 0.5 \\ \cos \left(2 \pi |x| \right) \;, & \text{if} \quad 0.5 < |x| \leq \alpha_{*} \\ \cos \left(2 \pi \alpha_* \right) - 2\pi \sin \left(2 \pi \alpha_* \right) \left( |x| - \alpha_* \right) \;, & \text{if} \quad \alpha_{*} < |x| \leq 1, \end{array} \right. \end{equation*} where the constant $\alpha_{*} \approx 0.6290$ satisfies the equation \begin{equation*} \cos \left(2 \pi \alpha_* \right) - 2\pi \sin \left(2 \pi \alpha_* \right) \left( 1 - \alpha_* \right) = 1. \end{equation*} The contact set $\mathcal{C}(f)$ consists of two disjoint sets $\{\frac12\le|x|\le\alpha_*\}$ and $\partial\Omega$. In this example we have $f$ smooth and $u \in C^{1,1}(\overline{\Omega})$ (full regularity). Upon choosing $\delta = 0.5 h^{1/2}$ and $\theta\approx 0.25 h^{1/2}$ we obtain computationally a linear convergence rate with respect to $h$, thus consistent with \Cref{C:convergence-rate} (convergence rate), and report it in \Cref{Table:Ex1} and Figure \ref{F:Ex1}. Plots of $u_{\varepsilon}$ and $f$ are shown in \Cref{F:Ex1-function-plot} and slices of these functions on $\{(x,0): x \ge 0\}$ are depicted in \Cref{F:Ex1} (left). In \Cref{F:Ex1} (right), we also display the $L^{\infty}$ error vs meshsize $h$ for several choices $\delta = O(h^{\alpha})$ with different values of $\alpha$ together with $\theta \approx 0.25h^{1/2}$. The convergence rate for $\delta = O(h^{2/3})$ is better than the one predicted in \Cref{C:convergence-rate}, but other rates are consistent with our theory. We choose $\theta$ to be small enough to make the error induced by $\theta$ small relative to those of $\delta$ and $h$. In fact, we can see from \Cref{F:Ex1} (right) that the effect of changing from $\theta \approx 0.25 h^{1/2}$ to $\theta \approx h^{1/2}$ is relatively small, and thus conclude that $\theta$ is not a sensitive parameter. \begin{table}[h!] \begin{center} \begin{tabular}[t]{ | l | l | c | c |} \hline Degrees of freedom & Number of directions & $L^{\infty}-$error & Iteration steps \\ \hline\hline $N = 1557$, $h=2^{-4}$ & \qquad\quad $S = 26$ & $3.769 \times 10^{-2}$ & 6 \\ \hline $N = 6317$, $h=2^{-5}$ & \qquad\quad $S = 36$ & $1.887 \times 10^{-2}$ & 10 \\ \hline $N = 25469$, $h=2^{-6}$ & \qquad\quad $S = 51$ & $9.617 \times 10^{-3}$ & 11 \\ \hline $N = 102445$, $h=2^{-7}$ & \qquad\quad $S = 72$ & $4.801 \times 10^{-3}$ & 11 \\ \hline $N = 410793$, $h=2^{-8} $ & \qquad\quad $S = 101$ & $2.400 \times 10^{-3}$ & 11 \\ \hline \end{tabular} \end{center} \vskip0.2cm \caption{\small \Cref{Ex:ex1}: $\delta = 0.5h^{1/2},\theta \approx 0.25 h^{1/2}$. The convergence rate is about linear (see \Cref{F:Ex1}), thus consistent with \Cref{C:convergence-rate}. The number of search directions $S$ scales like $S\approx\theta^{-1}\approx h^{-1/2}$, whereas the number of Howard's steps is relatively uniform.} \label{Table:Ex1} \end{table} \begin{figure}[!htb] \includegraphics[width=0.48\linewidth]{figures/pic1_f-eps-converted-to.pdf} \includegraphics[width=0.48\linewidth]{figures/pic1_uve-eps-converted-to.pdf} \caption{\small \Cref{Ex:ex1}, left: plot of $f$; right: plot of $u_{\varepsilon}$ for $h = 2^{-6}$. } \label{F:Ex1-function-plot} \end{figure} \begin{figure}[!htb] \includegraphics[width=0.48\linewidth]{figures/pic1_slice-eps-converted-to.pdf} \includegraphics[width=0.48\linewidth]{figures/pic1_rate-eps-converted-to.pdf} \caption{\small \Cref{Ex:ex1}. Left: slice of numerical solution $u_{\varepsilon}$ on $\{(x,0): x \ge 0\}$ with $h = 2^{-6}, \delta = 0.25h^{1/2}, \theta \approx 0.25h^{1/2}$. Right: experimental rates of convergence upon choosing $\theta \approx 0.25h^{1/2}$ and $\delta = O(h^{\alpha})$ with $\alpha = 1/3, 1/2, 2/3, 1$. A least square regression is performed for $h^{-k}$ with $k=6,7,8$ and the case $\delta = O(h)$. The orders are about $0.67, 0.99, 1.30, 0.07$. We also plot the errors for $\theta \approx h^{1/2}, \delta =h^{2/3}$, and the errors are very close to choosing $\theta \approx 0.25h^{1/2}, \delta =h^{2/3}$.} \label{F:Ex1} \end{figure} \end{example} \begin{example}[Lipschitz regularity $u \in C^{0,1}(\overline{\Omega})$] \label{Ex:ex2} Let $\Omega = \{x\in\mathbb{R}^2: |x|<1 \}$ and \begin{equation*} f(x) = \left\{ \begin{array}{ll} 1-4|x|, & 0 \leq |x| < 1/4 \\ 4|x|-1, & 1/4 \leq |x| < 1/2 \\ 2-2|x|, & 1/2 \leq |x| < 3/4 \\ 2|x|-1, & 3/4 \leq |x| \leq 1, \end{array} \right. \qquad u(x) = \left\{ \begin{array}{ll} 0, & 0 \leq |x| < 1/4 \\ |x|-1/4, & 1/4 \leq |x| < 3/4 \\ 2|x|-1, & 3/4 \leq |x| \leq 1. \end{array} \right. \end{equation*} This example deals with $f, u \in C^{0,1}(\overline{\Omega})$, i.e. both $f$ and $u$ are Lipschitz. The contact set $\mathcal{C}(f)$ consists of two disjoint components $\{x\in \mathbb{R}^2: |x| \ge 3/4 \}$ and $\{x\in \mathbb{R}^2: |x|=1/4 \}$. See \Cref{F:Ex2} (left) that displays slices on $\{(x,0): 0\le x\le 1\}$ of $f,u$ and the numerical solution $u_{\varepsilon}$ with $h = 2^{-6}, \delta = 0.25 h^{1/2}, \theta \approx 0.25h^{1/2}$. We point out that the pointwise error is very small in the regions $\{x\in \mathbb{R}^2: |x| \ge 3/4 \}$ and $\{x\in \mathbb{R}^2: |x| \le 1/4 \}$; in the latter $u$ is linear and thus the interpolation error disappears. On the other hand, in the region $\{x\in \mathbb{R}^2: 1/4 < |x| < 3/4 \}$, where $u$ is only linear in the radial direction, we observe larger error for $u_{\varepsilon}$. Experimental convergence rates for different choices of $\delta = O(h^{\alpha})$ are plotted in \Cref{F:Ex2} (right): we see that these rates are better than those predicted in \Cref{C:convergence-rate} (convergence rate). This theoretical rate can be improved upon exploiting that both functions $f$ and $u$ are non-smooth only at $\{0\}$ and across the curves $\{|x|=1/4\}$ and $\{|x|=3/4\}$. In fact, for those $x_i \in \mathcal{N}_h^0$ satisfying $\big| |x_i|-1/4 \big| \leq \delta$ or $\big| |x_i|-3/4 \big| \leq \delta$, according to \Cref{Prop:Consistency-lowreg} (consistency for $u$ with H\"older regularity), we have \begin{equation*} T_{\varepsilon}[\mathcal I_h u;f](x_i) \leq f(x_i) - u(x_i) \leq C(u) \delta, \end{equation*} whereas for the rest of $x_i \in \mathcal{N}_h^0$ the consistency error can be estimated exactly as for $f, u \in C^{1,1}(\overline{\Omega})$. Therefore carrying out the same analysis as in \Cref{T:error-estimate} (error estimate), we end up with the error estimate \begin{equation*} \|u-u_{\varepsilon}\|_{L^\infty(\Omega_h)} \le C(u) \left( \delta + \frac{(\delta \theta)^2 + h^2}{\delta^2}\right). \end{equation*} This yields a rate $O(h^{2/3})$ provided $\delta = O(h^{2/3})$, which is twice better than the rate from \Cref{C:convergence-rate} but still worse than the experimental ones in \Cref{F:Ex2} (right). \begin{figure}[!htb] \includegraphics[width=0.48\linewidth]{figures/pic2_slice-eps-converted-to.pdf} \includegraphics[width=0.48\linewidth]{figures/pic2_rate-eps-converted-to.pdf} \caption{\small \Cref{Ex:ex2}. Left: slices of $f,u$ and numerical solution $u_{\varepsilon}$ on $\{(x,0): x \ge 0\}$ with $h = 2^{-6}, \delta = 0.25h^{1/2}, \theta \approx 0.25h^{1/2}$. Right: experimental rates of convergence upon choosing $\theta = O(h^{1/2})$ and $\delta = O(h^{\alpha})$ with $\alpha = 1/3, 1/2, 2/3, 1$. The orders are about $0.78, 0.96, 1.06, 0.74$.} \label{F:Ex2} \end{figure} \end{example} \begin{example}[Lipschitz $u \in C^{0,1}{(\overline{\Omega})}$ and nonstrictly convex $\Omega$]\label{Ex:ex3} Let $\Omega = (-1,1)^2$ and $f,u$ be as in \cite[Example 6.3]{Ob2} with $\alpha = \beta = 1$, i.e. \[ f(x,y) = xy \;,\qquad u(x,y) = |x+y| - 1 . \] We point out that the Dirichlet boundary condition $u = f$ is attained on $\partial\Omega$ although the domain $\Omega$ is not strictly convex, whence \Cref{T:error-estimate} (error estimates) still applies. In this example, $f$ is smooth but $u$ is only Lipschitz because $\Omega$ is not uniformly convex and non-smooth: $u$ exhibits a kink across the diagonal $\{(x,y): x+y=0\}$ and is piecewise linear otherwise. Moreover, $u<f$ in $\Omega$ whence the contact set $\mathcal{C}(f)$ reduces to $\partial\Omega$. \Cref{F:Ex3} (left) displays slices on $\{(x,y): x \ge 0, \ y = x\}$ of $f,u$ and the numerical solution $u_{\varepsilon}$ with $h = 2^{-6}, \delta = h^{1/2}, \theta \approx 0.25h^{1/2}$. One can observe a clear mismatch between $u_{\varepsilon}$ and $u$ near the singular set $\{(x,y): x+y=0 \}$. Compared with \Cref{Ex:ex1} (full regularity $u\in C^{1,1}(\overline\Omega)$), the lack of regularity of $u$ here entails larger consistency error and $L^{\infty}$ error between $u_{\varepsilon}$ and $u$. Experimental convergence rates for different choices of $\delta = O(h^{\alpha})$ are depicted in \Cref{F:Ex3} (right); we see that the best convergence rate $O(h^{0.58})$ is found when $\delta = O(h^{1/3})$, which is again better than the $O(h^{1/3})$ rate predicted in \Cref{C:convergence-rate} (convergence rate). \begin{figure}[!htb] \includegraphics[width=0.495\linewidth]{figures/pic3_slice-eps-converted-to.pdf} \includegraphics[width=0.495\linewidth]{figures/pic3_rate-eps-converted-to.pdf} \caption{\small \Cref{Ex:ex3}. Left: slice of numerical solution $u_{\varepsilon}$ on $\{(x,y): x \ge 0, \ y = x\}$ with $h = 2^{-6}, \delta = h^{1/2}, \theta \approx 0.25h^{1/2}$. Right: experimental rates of convergence upon choosing $\theta = O(h^{1/2})$ and $\delta = O(h^{\alpha})$ with $\alpha = 1/3, 1/2, 2/3, 1$. The orders are about $0.58, 0.45, 0.41, 0.03$.} \label{F:Ex3} \end{figure} \end{example} \begin{example}[non-attainment of Dirichlet condition] \label{Ex:ex4} Let $\Omega = (-1,1)^2$ and the function $f$ be $f(x,y) = \cos(\pi x)\cos(\pi y)$, whose restriction to $\partial\Omega$ is not convex. According to our definition \eqref{E:def-CE}, the convex envelope is given by \begin{equation*} u(x,y) = \begin{cases} -1 & \quad |x|+|y| \leq 1 \\ -\cos \big(\pi (|x|+|y| - 1) \big) & \quad 1 < |x|+|y| \leq 1 + \beta_{*} \\ -\cos \left(\pi \beta_* \right) + \pi \sin \left(\pi \beta_* \right) \big( |x|+|y| - 1 - \beta_* \big) & \quad 1+\beta_{*} < |x|+|y|, \end{cases} \end{equation*} where the constant $\beta_{*} \approx 0.2580$ satisfies the equation \[ -\cos(\pi \beta_*) + \pi \sin(\pi \beta_*)(1 - \beta_*) = 1. \] This assertion requires a brief explanation. First of all note that by symmetry it suffices to examine the first quadrant $0\le x,y \le 1$. On the edges $\{y=1\}$ and $\{x=1\}$ the function $u$ is convex by construction and definition of $\beta_*$; see \Cref{F:Ex4} (left). Since $u$ is flat along lines $x+y=\beta$ and convex along perpendicular lines, we infer that $u$ is convex. It remains to show that $u\le f$ and $\ge$ than the convex envelope. To this end, we take convex combinations of boundary values $u(\beta-1,1)$ and $u(1,\beta-1)$ along the line $x+y=\beta$ with $1\le\beta\le2$ and show that they are $\le f(x,y)$. For $\beta=1$ we realize that $u(x,y)=-1\le f(x,y)$ on $x+y=1$ and by symmetry for all $x+y\le1$. For $\beta>1$ a tedious calculation gives $u(x,y)=u(\beta-1,1) \le f(\beta-1,1) \le f(x,y)$ along $x+y=\beta$ as desired. We finally point out that the contact set $\mathcal{C}(f)$ consists of four boundary segments of length $2\beta_*$ centered at $(0,\pm 1), (\pm 1,0)$ and the four vertices $(\pm 1,\pm 1)$ of $\Omega$; see \Cref{F:Ex4} (left). \looseness=-1 \begin{figure}[!htb] \includegraphics[width=0.48\linewidth]{figures/pic4_slice-eps-converted-to.pdf} \includegraphics[width=0.48\linewidth]{figures/pic4_rate-eps-converted-to.pdf} \caption{\small \Cref{Ex:ex4}. Left: slices $f,u$ and $u_{\varepsilon}$ on the set $\{(x,1): x \ge 0 \}$ with $h = 2^{-6}, \delta = 2h^{1/2}, \theta \approx 0.5h^{1/2}$. Note that $u_{\varepsilon}$ is indistinguishable from $u$ on this part of $\partial\Omega$. Right: experimental rates of convergence upon choosing $\theta = O(h^{1/2})$ and $\delta = O(h^{\alpha})$ with $\alpha = 1/3, 1/2, 2/3, 1$; the orders of convergence are about $1.30, 1.04, 0.91, 0.20$.} \label{F:Ex4} \end{figure} We implemented the modified two-scale method \eqref{E:2ScOp-Ex4}, which first solves boundary subproblems on each edge of $\partial\Omega$ to find the trace of the discrete convex envelope $u_{\varepsilon}$ and next determines $u_{\varepsilon}$ within $\Omega$. \Cref{F:Ex4} (left) shows $f,u$ and $u_{\varepsilon}$ on the boundary set $\{(x,1): 0\le x \le 1\}$; we point out that $u(x,1)=f(x,1)$ for $|x|\le \beta_*$. \Cref{F:Ex4} (right) displays the $L^{\infty}$ error for several choices of $h$ and $\delta$: we see that the experimental convergence rate is about $O(h)$ for $\delta=O(h^{1/2})$, in agreement with theory, but the rates for $O(h^{\alpha})$ with $\alpha = 1/3, 2/3$ seem to be better than those predicted in \Cref{C:convergence-rate} (convergence rate). \end{example} \subsection{Computational performance} Thanks to the search tools provided by FELICITY \cite{Walker1,Walker2}, the process of locating the triangle of the mesh containing points $x_i \pm \delta_i v_j$ and computing the barycentric coordinates only takes a small percentage of the total computing time; this is consistent with the two-scale method for the Monge-Amp\`{e}re equation in \cite{NoNtZh1}. In \Cref{Ex:ex1} for $h = 2^{-6}, \delta = 0.25h^{1/2}, \theta \approx 2h^{1/2}$, this process is 6.7\% ($<$ 4 sec) of the total computation time (56.2 sec). The most time consuming part of the experiment is constructing and solving the linear systems, i.e. the third line in \Cref{alg:Howard}; this takes 53.2\% of the total time. We do not attempt to exploit the sparsity pattern of the matrix $B^{\bm{\alpha}}$ and simply resort to MATLAB backslash command for solving linear systems; we leave this important issue open. All of our computations are performed on an Intel Xeon E5-2630 v2 CPU (2.6 GHz), 16 GB RAM using MATLAB R2016b. \subsection{Comparison with other existing methods} In this subsection, we briefly compare our two-scale method with two other methods for the computation of convex envelopes: the wide stencil method in \cite{Ob2} and the modified version of Dolzmann's method in \cite{Bartels}. Both the wide stencil method and our two-scale method are derived from the PDE formulation \eqref{E:pde-CE}, and have a discrete operator with similar structure. As explained in \Cref{S:modified-wide-stencil}, the wide stencil method can be viewed as a two-scale method with no interpolation error but with the constraint $\theta \approx h/\delta$. Our two-scale method suffers from the interpolation error but allows some freedom in the choice of parameters and works well on unstructured grids, which provide geometric flexibility to fit the boundary $\partial\Omega$. The modified version of Dolzmann's method in \cite{Bartels}, built for the computation of rank-one convex envelopes of functions defined on $\mathbb{R}^{n \times m}$, can be applied to compute the convex envelope by simply letting $m = 1$. When applied to compute convex envelopes, the technique of \cite{Bartels} hinges on the following algorithm: if $f^{(0)} = f$, and $f^{(k)}$ for $k \ge 1$ is iteratively defined as \begin{equation}\label{E:iter-CE} \begin{aligned} f^{(k)}(x) = \inf\{ & \lambda f^{(k-1)}(x_1) + (1 - \lambda) f^{(k-1)}(x_2): \\ & \lambda \in [0,1], x_1, x_2 \in \mathbb{R}^d, \lambda x_1 + (1-\lambda) x_2 = x\}, \end{aligned} \end{equation} then the convex envelope $u = f^{(d)}$ by Carath\'{e}odory's theorem. Consequently, at the continuous level this process terminates in at most $d$ iterations. The method in \cite{Bartels} is a discrete version of this iteration on a structured grid $h\mathbb{Z}^d$ with interpolation on the finer grid $h^2 \mathbb{Z}^d$, namely $x\in h\mathbb{Z}^d$ but $x_1,x_2\in h^2\mathbb{Z}^d$ in \eqref{E:iter-CE}. This is thus a two-scale method, with coarse scale $h$, but conceptually different from ours because it does not solve a PDE but rather an algebraic iteration. Moreover, it assumes $u = f$ in a layer $\{x \in \Omega: \textrm{dist}(x,\partial\Omega) \le Ch \}$ near the boundary $\partial\Omega$ to deal with nodes in this region. Regarding convergence rates, both the method in \cite{Bartels} and our two-scale method exhibit provable linear rates with respect to the coarse scale for solutions $u\in C^{0,1}(\overline\Omega)$ according to \Cref{R:two-scenarios} (two important scenarios); moreover, \Cref{R:two-scenarios} also shows that our method is quadratic in the coarse scale $\delta$ and linear in the fine scale $h$ for $u \in C^{1,1}(\overline\Omega)$ . Performing $d$ iterations of the discrete version of \eqref{E:iter-CE} is enough for linear convergence, whereas those for Howard's method cannot be quantified a priori. However, practice reveals that $10$ iterations of Howard's method are enough for convergence, which is consistent with its superlinear structure. Our iterations are simpler than those in \cite{Bartels} because they require much fewer interpolation points. Finally, our two-scale method is designed to work on unstructured meshes and deal with the Dirichlet boundary condition in a natural fashion. The boundary layer effect is handled via discrete barrier functions. \section*{Acknowledgement} We are grateful to Dimitrios Ntogkas for allowing us to modify his codes on the two-scale method for the Monge-Amp\`{e}re equation to solve the convex envelope problems. \bibliographystyle{amsplain}
{ "timestamp": "2019-01-01T02:12:31", "yymm": "1812", "arxiv_id": "1812.11519", "language": "en", "url": "https://arxiv.org/abs/1812.11519", "abstract": "We develop two-scale methods for computing the convex envelope of a continuous function over a convex domain in any dimension.This hinges on a fully nonlinear obstacle formulation [A. M. Oberman, \"The convex envelope is the solution of a nonlinear obstacle problem\", Proc. Amer. Math. Soc. 135(6):1689--1694, 2007]. We prove convergence and error estimates in the max norm. The proof utilizes a discrete comparison principle, a discrete barrier argument to deal with Dirichlet boundary values, and the property of flatness in one direction within the non-contact set. Our error analysis extends to a modified version of the finite difference wide stencil method of [A. M. Oberman, \"Computing the convex envelope using a nonlinear partial differential equation\", Math. Models Meth. Appl. Sci, 18(05):759--780, 2008].", "subjects": "Numerical Analysis (math.NA)", "title": "Two-scale methods for convex envelopes", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9830850877244272, "lm_q2_score": 0.815232480373843, "lm_q1q2_score": 0.8014428944841218 }
https://arxiv.org/abs/1805.04427
Multi-crossing Braids
Traditionally, knot theorists have considered projections of knots where there are two strands meeting at every crossing. A multi-crossing is a crossing where more than two strands meet at a single point, such that each strand bisects the crossing. In this paper we generalize ideas in traditional braid theory to multi-crossing braids. Our main result is an extension of Alexander's Theorem. We prove that every link can be put into an $n$-crossing braid form for any even $n$, and that every link with two or more components can be put into an $n$-crossing braid form for any $n$. We find relationships between the $n$-crossing braid indices, or the number of strings necessary to represent a link in an $n$-crossing braid.
\section{Introduction} \label{sec:intro} In traditional knot theory, knots are drawn in a projection where there are two strands passing over each other at every crossing. An \emph{$n$-crossing} is a crossing where there are $n$ strands meeting at one point, with each strand bisecting the crossing. We call this crossing a \emph{multi-crossing} if $n>2$, and we call the traditional type ($n=2$) a \emph{double crossing}. The strands are labeled with the levels $1, \dots n$ from the top. In~\cite{triple crossing}, Adams proved that every link has an $n$-crossing projection for all $n \ge 3$. This fact allows us to generalize notions in traditional knot theory to their multi-crossing versions. For example, the \emph{crossing number} $c(L)$, the minimum number of crossings in any double crossing projection of the link $L$, generalizes to the \emph{multi-crossing number} $c_n(L)$, which is the minimum number of crossings in any $n$-crossing projection of the link $L$. This gives us an infinite spectrum of crossing numbers that can be explored. In this paper we similarly generalize ideas from braid theory to multi-crossing braids. Alexander's Theorem states that every link can be put into a braid form \cite{alexander}. This Theorem has since been proved in several ways \cite{yamada} \cite{braid index bound}. Is it true that every link can be put into an $n$-crossing braid, or a braid where every crossing is an $n$-crossing? If true, we can generalize the notion of \emph{braid index} $\beta(L)$, which is the minimum number of strings needed to represent the link $L$ as the closure of a double crossing braid. Can we define the $n$-crossing braid indices $\beta_n(L)$, and what are their properties? In Section~\ref{sec:evenAlexander} we prove a version of Alexander's Theorem~\cite{alexander} for even multi-crossing braids. Specifically, we prove that every link can be represented as a closed $n$-crossing braid, for all even $n$. In Section~\ref{sec:tripleAlexander} we consider an equivalent of Alexander's Theorem for triple crossing braids. We prove that every link with at least two components can be represented as a closed triple-crossing braid. In Section~\ref{sec:oddAlexander} we extend this result to all odd $n$. Finally, in Section~\ref{sec:braidIndices} we find relationships between the $n$-crossing braid indices. This paper is a part of a senior thesis completed at Williams College. I would like to thank my advisor Professor Colin Adams for his guidance throughout this year. Thanks to Daniel Vitek for suggesting this problem. He had proved Theorem~\ref{thm:evenAlexander} for $n=6,10,14,18,22,26$ using Lemma~\ref{lem:levelPosition} and computation by Mathematica. The idea to convert the problem into looking at the corresponding permutations is due to him, as is the proof of Lemma~\ref{lem:levelPosition}. \section{Even Multi-crossing Braids} \label{sec:evenAlexander} In this section, we prove a result similar to Alexander's Theorem for $n$-crossing braids, where $n$ is even. Specifically, we prove the following. \begin{theorem} \label{thm:evenAlexander} Every link can be represented as a closed $n$-crossing braid, for all even $n$. \end{theorem} We prove this theorem by starting with a link in double crossing braid form, and finding an isotopy to make it a sequence of $n$-crossings. \subsection{Level position} Given a collection of $m$ strings, label them $1, \dots, m$ from the left. We can then think of each crossing that occurs in these $m$ strings as a permutation of the strings. Specifically, we can define a homomorphism $\phi:B_n \rightarrow S_n$ by $\phi(\sigma_i) = \phi(\sigma_i^{-1}) = (i,i+1)$, and extend by linearity. A double crossing corresponds to a transposition $(i, i+1)$. An $n$-crossing corresponds to a permutation of the form $\pi_j = (j,j+n-1)(j+1,j+n-2)\cdots$; we call any permutation of this form a \emph{crossing permutation}. Note that there are $m-n+1$ possible $n$-crossing permutations, corresponding to each $j$ where $1 \le j \le m-n+1$. First we show a lemma that allows us to ignore the levels of each crossing and focus on the images under $\phi$. We say a sequence of crossings is \emph{disjoint} if each string is switched by at most one crossing in the sequence. In this section we only need this lemma for $s=2$, but we state a general version which can be used in later sections for other values of $s$. \begin{lemma} \label{lem:levelPosition} Let $\alpha$ be a sequence of disjoint $s$-crossings. Suppose that a product of $n$-crossing permutations in $S_m$ produces $\phi(\alpha)$. Then there exists a sequence of $n$-crossings over a $m$-string braid which produces $\alpha$. \end{lemma} \begin{proof} Consider the sequence of $n$-crossing permutations that produces $\phi(\alpha)$. We want to show that we can choose the levels of the corresponding $n$-crossings appropriately so that the result is equivalent to $\alpha$. We do this by placing the $m$ strands on different heights. Choose an $s$-crossing in $\alpha$. We place the $s$ strands of this crossing in the $s$ highest levels, according to their levels in the $s$-crossing. We continue this process by choosing a new $s$-crossing in $\alpha$, and placing the $s$ strands of this crossing in the next $s$ highest levels. Once we have exhausted $s$-crossings in $\alpha$, the remaining strands can be placed in any order. The heights are well defined since the crossings are disjoint. For each permutation corresponding to an $n$-crossing, we want to assign levels to the strands to make it into an $n$-crossing. We can simply choose the levels of the strands in the order of the heights assigned above. This will mean that each strand always stays on the same level, and that each $n$-crossing can be untangled easily (Fig.~\ref{fig:diagonalBox}). We call this a \emph{level position} of the braid. \begin{figure}[ht] \centering \includegraphics[scale=0.2]{diagonalBox.png} \caption{The left picture shows how we might use 6-crossings to obtain a double crossing which switches the first two strings, when the second string is the overstrand. The remaining strands have been ordered from left to right. The right picture is a view from above.} \label{fig:diagonalBox} \end{figure} Then, once we have achieved $\phi(\alpha)$ we can pull the strings taut, so that the only crossings that are left are the $s$-crossing in $\alpha$ that we were looking for (Fig.~\ref{fig:diagonalBoxAfterPull}). \begin{figure}[ht] \centering \includegraphics[scale=0.2]{diagonalBoxAfterPull.png} \caption{What happens after the strings are pulled taut. All the 6-crossings disappear and one double crossing remains.} \label{fig:diagonalBoxAfterPull} \end{figure} \end{proof} By this lemma, it suffices to show that we can use $n$-crossing permutations to obtain double crossing permutations. \subsection{Conjugation of permutations} Observe that for permutations $\pi, \sigma \in S_m$, where \[ \pi = (i_1, i_2, \dots, i_r) \cdots (i_s, i_{s+1}, \dots, i_t) \] in cycle notation, the conjugate of $\pi$ by $\sigma$ has a nice form: \[ \sigma \pi \sigma^{-1} = (\sigma(i_1), \sigma(i_2), \dots, \sigma(i_r)) \cdots (\sigma(i_s), \sigma(i_{s+1}), \dots, \sigma(i_t)). \] We consider the case when $\sigma$ is a crossing permutation. Then it is its own inverse, so $\sigma \pi \sigma^{-1} = \sigma \pi \sigma$. Thus, if we multiply both sides of $\pi$ by $\sigma$, we are essentially switching elements that appear in $\pi$ according to $\sigma$. Note that taking the conjugate does not change the cycle type, which is to say it keeps the number of cycles and the length of each cycle constant. Also note that taking conjugates by the crossing permutation $\pi_i$ can only affect elements between $i$ and $i+n-1$, and must leave all other elements fixed. Finally, observe that we can reverse this process since $\sigma$ is its own inverse; if we can obtain $\pi$ by taking conjugates of $\pi'$, then we can obtain $\pi'$ by taking conjugates of $\pi$ in reverse order. \subsection{Obtaining permutations of the same cycle type} Using this idea, we can start with $\pi_1$, and take conjugates by some $\pi_j$ to obtain different permutations with the same cycle type. In fact, we can show that for sufficiently large $m$, we can obtain any permutation with the same cycle type. Note that we can always make sure that we have enough strands (i.e. that $m$ is large enough) by taking stabilizations. First we present several lemmas which will be helpful in proving this result. In the proofs of these lemmas we repeatedly take conjugates by crossing permutations. Recall that taking conjugates by the crossing permutation $\pi_i$ can only affect elements between $i$ and $i+n-1$, and must leave all other elements fixed. We can keep track of which entries of a permutation are affected by conjugation, by checking which entries lie within or outside the range from $i$ to $i+n-1$. The first lemma shows how to obtain a permutation which sends 1 to some $N$. \begin{lemma} \label{lem:moveMaxPerm} Let $n$ be even. Then, for any $N$ with $1 \le N \le \frac{3n}{2}$, there exists a sequence of $n$-crossing permutations over $S_{3n/2}$ whose product sends 1 to $N$. \end{lemma} \begin{proof} First observe that over $S_{3n/2}$, we have the $n$-crossing permutations $\pi_1, \dots, \pi_{n/2+1}$. We consider permutations which send $r$ to some $r+j$. We call this incrementing $r$ by $j$. For $1 \le r \le \frac{n}{2}+1$, we can increment it by $n-1$ with $\pi_r$. For $r=1,2$ and $1 \le s \le \frac{n}{2}-1$, we can increment $r$ by $2s$ with $\pi_{r+s} \pi_r$. (First $\pi_r$ sends $r$ to $r+n-1$. Then observe that $r+n-1$ is $s$ away from $r+s+n-1$, which is the highest entry in $\pi_{r+s}$, and $r+2s$ is $s$ away from $r+s$, which is the lowest entry in $\pi_{r+s}$.) Also, if $r=1$, we can increment it by 1 with $\pi_{r+1} \pi_{r+n/2} \pi_r$. Hence we can obtain a permutation that sends 1 to 2, or to $n$, or to $2s+1$ for $1 \le s \le \frac{n}{2}-1$. We can also obtain a permutation that sends 1 to $2s+2$, for $1 \le s < \frac{n}{2}-1$, by first sending it to 2 and then incrementing by $2s$. Thus, for any $N \le n$, we can obtain a sequence of $n$-crossing permutations over $S_{3n/2}$ that sends 1 to $N$. If $n < N \le \frac{3n}{2}$, then we can first take a permutation that sends 1 to some element $N-n+1$, where $1 < N-n+1 \le \frac{n}{2}+1$. Then we can compose it with $\pi_{N-n+1}$, which will send 1 to $N$. \end{proof} The next lemma describes how to increment the largest entry in a permutation. \begin{lemma} \label{lem:moveMax} Let $n$ be even. Then, given a permutation of the form \[ \tau = (1, 2, \dots, l_1)(l_1+1, \dots, l_2) \cdots (l_{t-1}+1, \dots, l_t-1, l_t), \] we can conjugate it by the $n$-crossing permutations over $S_{l_t+3n-3}$ to obtain any permutation of the form \[ (1, 2, \dots, l_1) (l_1+1, \dots, l_2) \cdots (l_{t-1}+1, \dots, l_t-1, N), \] where $l_t \le N \le l_t+3n-3$. \end{lemma} \begin{proof} First observe that over $S_{l_t + 3n-3}$, we have the $n$-crossing permutations $\pi_1, \dots, \pi_{l_t + 2n-2}$. By applying Lemma~\ref{lem:moveMaxPerm} to the strings from $l_t$ to $l_t + \frac{3n}{2}-1$, we can see that for any $N$ with $l_t \le N \le l_t + \frac{3n}{2}-1$, there is a sequence of $n$-crossing permutations over $S_{l_t+3n/2-1}$ which sends $l_t$ to $N$, and leaves $1,\dots,l_t-1$ fixed. Then we can conjugate $\tau$ by this sequence of $n$-crossing permutations to obtain the permutation $(1, \dots, l_1)\cdots (l_{t-1}+1, \dots, l_t-1, N)$. If $l_t+\frac{3n}{2} \le N \le l_t+3n-3$, then we first obtain a permutation $(1, \dots, l_1)\cdots (l_{t-1}+1, \dots, l_t-1, N')$ where $N'$ is an element with $l_t \le N' \le l_t + \frac{3n}{2}-1$ that is a multiple of $n-1$ away from $N$. Then we can conjugate by $\pi_{N'}$, $\pi_{N'+n-1}$, and so on, until we reach $N$ after conjugating by $\pi_{N-n+1}$. Note that in doing so, we only need the crossing permutations $\pi_1, \dots, \pi_{l_t + 2n-2}$. Thus, the crossing permutations over $S_{l_t+3n-3}$ are sufficient to produce any permutation of the form $(1, \dots, l_1)\cdots (l_{t-1}+1, \dots, l_t-1, N)$, where $N \le l_t+3n-3$. \end{proof} The next lemma describes how to increment the remaining entries of a permutation, without changing the order of the entries. \begin{lemma} \label{lem:moveAll} Let $n$ be even. Then, given $\tau$ as in the statement of Lemma~\ref{lem:moveMax}, we can conjugate it by the $n$-crossing permutations over $S_{l_t+3n-3}$ to obtain any permutation of the form \[ (a_1, \dots, a_{l_1}) (a_{l_1+1}, \dots, a_{l_2}) \cdots (a_{l_{t-1}+1}, \dots, a_{l_t}) \] where $1 \le a_1 < a_2 < \cdots < a_{l_t} \le l_t+3n-3$. \end{lemma} \begin{proof} We prove this by iterating over each entry in $\tau$, from largest to smallest. The base case (largest entry in $\tau$) can be done by Lemma~\ref{lem:moveMax}. Suppose we have moved the $k$ largest entries to their appropriate positions, where $1 \le k \le l_t$. Let $N = a_{l_t-k+1}$, which is to say that we have a permutation of the form \[ \tau' = (1, \dots, l_1) \cdots (l_{s-1}+1, \dots, l_t-k, N, a_{l_t-k+2}, \dots, a_{l_s}) \cdots (a_{l_{t-1}+1}, \dots, a_{l_t}), \] Let $N' = a_{l_t - k}$, the new desired entry for the $k+1$st largest entry. Note that $N > N'$ by the hypothesis. We want to move $l_t-k$ to $N'$. First consider the case when $N \le l_t - k + \frac{3n}{2} - 1$. Note that in this case $N' < l_t - k + \frac{3n}{2} - 1$. Observe that since we have only moved the $k$ largest entries, one of the $k$ indices between $l_t - k + \frac{3n}{2} - 1$ and $l_t + \frac{3n}{2} - 2$ must be ``empty'', which is to say it does not appear in $\tau'$, or that it is a fixed point in $\tau'$. Let $M$ be such an element. We apply Lemma~\ref{lem:moveMaxPerm} to the $\frac{3n}{2}$ strings starting at $l_t - k + \frac{3n}{2} - 1$. Note that we need $l_t + 3n - 3$ strings for this to be possible for all $k$ with $1 \le k \le l_t$. Then, there exists a sequence $\{\pi_{b_i}\}$ of $n$-crossing permutations over $S_{l_t - k + 3n - 2}$ that sends $l_t - k + \frac{3n}{2} - 1$ to $M$, and leaves $1, \dots, l_t - k + \frac{3n}{2} - 2$ fixed. Note that this sequence in reverse order sends $M$ to $l_t-k+\frac{3n}{2} - 1$. Hence if we conjugate $\tau'$ by $\{\pi_{b_i}\}$ in reverse order, then the element $l_t-k+\frac{3n}{2} - 1$ will not appear in the resulting permutation. Thus we have some permutation of the form \[ (1, \dots, l_1) \cdots (l_{s-1}+1, \dots, l_t-k, c_{l_t-k+1}, c_{l_t-k+2} \dots, c_{l_s}) \cdots (c_{l_{t-1}+1}, \dots, c_{l_t}), \] where all $c_i$ are strictly greater than $l_t - k + \frac{3n}{2} - 1$. By Lemma~\ref{lem:moveMaxPerm}, there is a sequence of $n$-crossing permutations $\{\pi_{b_i}\}$ over $S_{l_t-k+3n/2-1}$ which sends $l_t$ to $N'$, and leaves $1,\dots,l_t-k-1$ fixed. (Note that $\{\pi_{b_i}\}$ also leave entries greater than $l_t-k+\frac{3n}{2}-1$ fixed.) Then we can take conjugates by $\{\pi_{b_i}\}$ to obtain the permutation \[ (1, \dots, l_1) \cdots (l_{s-1}+1, \dots, l_t-k-1, N', c_{l_t-k+1}, c_{l_t-k+2} \dots, c_{l_s}) \cdots (c_{l_{t-1}+1}, \dots, c_{l_t}). \] Finally, we can take conjugates by $\{\pi_{b_i}\}$ in the forward order to move the other entries back and obtain \[ (1, \dots, l_1) \cdots (l_{s-1}+1, \dots, l_t-k-1, N', N, a_{l_t-k+2} \dots, a_{l_s}) \cdots (a_{l_{t-1}+1}, \dots, a_{l_t}). \] Note that since $N' < l_t - k + \frac{3n}{2} - 1$, the conjugations by $\{\pi_{b_i}\}$ does not affect $N'$. Next, consider the case when $N' \le l_t - k + \frac{3n}{2} - 1 < N$. In this case, by Lemma~\ref{lem:moveMaxPerm} we can move $l_t-k$ to $N'$ without affecting any other elements. Finally, consider the case when $N' > l_t - k + \frac{3n}{2} - 1$. First we use Lemma~\ref{lem:moveMaxPerm} to move $l_t-k$ to some $M'$, where $M'$ is an element with $l_t - k \le N' \le l_t + \frac{3n}{2}-1$ that is a multiple of $n-1$ away from $N$. Then we can conjugate by $\pi_{M'}$, $\pi_{M'+n-1}$, and so on, until we reach $N'$. \end{proof} Finally we prove the desired result. We define the \emph{size} of a permutation to be the number of distinct entries that appear in the permutation when written in cycle notation, where we drop any 1-cycles. \begin{lemma} \label{lem:sameCycleType} Let $n$ be even. Then, given any $\tau \in S_{N+3n-3}$ with size at most $N$, we can conjugate by the $n$-crossing permutations over $S_{N+3n-3}$ to obtain any permutation in $S_{N+3n-3}$ of the same cycle type as $\tau$. \end{lemma} \begin{proof} Write \[ \tau = (d_1, \dots, d_{l_1})(d_{l_1+1}, \dots, d_{l_2}) \cdots (d_{l_{t-1}+1}, \dots, d_{l_t}) \] in cycle notation. Note that $l_t \le N$ since the size of $\tau$ is at most $N$. Then consider the permutation \[ \tau' = (1, \dots, l_1)(l_1+1, \dots, l_2) \cdots (l_{t-1}+1, \dots, l_t). \] Note $\tau'$ has the same cycle type as $\tau$. It suffices to show that we can conjugate $\tau'$ to get any permutation of the same cycle type, since we can reverse this process to obtain $\tau'$ from $\tau$. Since $l_t \le N$, we know we can use the $n$-crossing permutations over $S_{l_t + 3n - 3}$. Thus we can apply Lemma~\ref{lem:moveAll} to obtain any permutation of the form \[ (a_1, \dots, a_{l_1})(a_{l_1+1}, \dots, a_{l_2}) \cdots (a_{l_{t-1}+1}, \dots, a_{l_t}), \] where $a_1 < \cdots < a_{l_t}$. Hence it suffices to show that we can switch two entries in the permutation. We show how to switch two entries in $\tau'$, after which all entries can be moved to the appropriate positions. Suppose we want to switch $r$ and $r+1$ for $r \le l_t$; in other words, suppose we want to obtain the permutation \[ (1, \dots, l_1) \cdots (l_{s-1}+1, \dots, r-1, r+1, r, r+2, \dots, l_s) \cdots (l_{t-1}+1, \dots, l_t). \] Note that $s$ and $s+1$ may appear in different cycles in $\tau$, but the same argument applies. First we can conjugate $\tau'$ by $\pi_r$ to obtain the permutation \[ (1, \dots, l_1) \cdots (l_{s-1}+1, \dots, r-1, r+n-1, r+n-2, e_{r+2} \dots, e_{l_s}) \cdots (e_{l_{t-1}+1}, \dots, e_{l_t}), \] where the $e_i$ are strictly less than $r+n-2$ for $i \ge r+2$. Then we can conjugate by $\pi_{r+n-2}$ to obtain the permutation \[ (1, \dots, l_1) \cdots (l_{s-1}+1, \dots, r-1, r+2n-4, r+2n-3, e_{r+2}, \dots, e_{l_s}) \cdots (e_{l_{t-1}+1}, \dots, e_{l_t}). \] We can then conjugate by $\pi_{r+\frac{3n}{2}-3}$ to obtain the permutation \[ (1, \dots, l_1) \cdots (l_{s-1}+1, \dots, r-1, r+2n-3, r+2n-4, e_{r+2}, \dots, e_{l_s}) \cdots (e_{l_{t-1}+1}, \dots, e_{l_t}). \] Observe that we have switched the $r$th entry with the $r+1$st entry. Finally, we can conjugate by $\pi_{r+n-2}$ and then by $\pi_r$ to move all entries back and obtain the desired permutation. \end{proof} While this suffices to prove Theorem~\ref{thm:evenAlexander}, we wish to reduce the number $d+3n-3$ as much as possible. As we will see in Section~\ref{sec:braidIndices}, we conjecture that it can be reduced down to $n+2$ for $d \le n$. \subsection{Proof of Theorem~\ref{thm:evenAlexander}} Before proving Theorem~\ref{thm:evenAlexander}, we present one final lemma. \begin{lemma} \label{lem:evenPair} Let $n$ be even, and let $m \ge \frac{5n}{2}-1$. Then we can obtain the permutations $(1,n)(2n-1,2n)$ as a product of the $n$-crossing permutations over $S_m$. \end{lemma} \begin{proof} First note that over $S_{5n/2-1}$ we have the crossing permutations $\pi_1, \dots, \pi_{3n/2}$. Take the crossing permutation $\pi_1 = (1,n)(2,n-1)\cdots(\frac{n}{2}, \frac{n}{2}+1)$. We can conjugate by $\pi_n$ to obtain the permutation $(1,2n-1)(2,n-1)\cdots(\frac{n}{2}, \frac{n}{2}+1)$. Then conjugate by $\pi_{3n/2}$ to get $(1,2n)(2,n-1)\cdots(\frac{n}{2}, \frac{n}{2}+1)$. Thus we have moved the largest entry of the permutation to $2n$. We then perform a similar sequence to move the smallest entry to $2n-1$. This is done by conjugating by $\pi_1$ and then by $\pi_n$. We are now left with the permutation $(2n-1,2n)(2,n-1)\cdots(\frac{n}{2},\frac{n}{2}+1)$. Note that the first conjugation by $\pi_1$ also moves the other entries around, but the pairs stay the same; for example $\pi_1$ switches the entries 2 and $n-1$, but this keeps the cycle $(2,n-1)$ constant. Finally we can multiply this permutation by $\pi_1$. The transpositions in the middle would cancel, and we are left with $(1,n)(2n-1,2n)$. \end{proof} We are now ready to prove Theorem~\ref{thm:evenAlexander}. \begin{proof}[Proof of Theorem~\ref{thm:evenAlexander}] Let $L$ be a link. Put it in a double crossing braid form; call this braid $\alpha$. First consider the case when $n = 4k + 2$. Start with $\alpha$, and take stabilizations until we have at least $3n+1$ strands; call this braid $\alpha'$. By Lemma~\ref{lem:evenPair}, we can obtain the permutation $(1,n)(2n-1,2n)$. This permutation has size 4, so by Lemma~\ref{lem:sameCycleType}, we can conjugate $\pi_1$ to obtain any permutation whose cycle type is two transpositions. We start with $\pi_1$, which has an odd number of transpositions. We can cancel pairs of transpositions if we multiply by the pairs, which we can obtain by Lemma~\ref{lem:sameCycleType}. Thus we can cancel all but one of the transpositions. By Lemma~\ref{lem:sameCycleType}, we can rearrange the entries of this transposition to obtain any transposition of the form $(i,i+1)$ for any $i$ with $1 \le i < 3n+1$. Hence, by Lemma~\ref{lem:levelPosition}, we can produce any double crossing that appears in $\alpha'$ as a product of $n$-crossings. Thus, we have produced an $n$-crossing braid that is equivalent to $L$. Now consider the case when $n=4k$. Observe that the permutations obtained from the $n$-crossings are all even. Therefore these permutations cannot generate the transpositions, which are odd. However, we can start with a double crossing braid with an even number of crossings, which will be generated by all disjoint pairs of double crossings. Note that if two consecutive crossings involve a common strand, we can obtain one of them with a pair of crossings by creating an extra dummy crossing elsewhere, and obtain the other with a pair which undoes the dummy crossing. Start with $\alpha$, and take stabilizations until we have at least $3n+1$ strands. If this braid has an odd number of crossings, take an extra stabilization so we have an even number of crossings. From the top of the braid, put the crossings in disjoint pairs, adding a pair of dummy crossings as necessary. Call this braid $\alpha'$. As before, by Lemma~\ref{lem:evenPair} and Lemma~\ref{lem:sameCycleType} we can obtain any permutation whose cycle type is two transpositions. We start with $\pi_1$, which has an even number of transpositions. We can cancel pairs of transpositions as before, so we are left with a pair of transpositions. Then, by Lemma~\ref{lem:sameCycleType} we can rearrange the entries to get any pair of transpositions of the form $(a, a+1)(b,b+1)$, where the four entries are all distinct. Hence, by Lemma~\ref{lem:levelPosition}, we can produce any disjoint pair of double crossings that appears in $\alpha'$ as a product of $n$-crossings. Thus, we have produced an $n$-crossing braid that is equivalent to $L$. \end{proof} \section{Triple-crossing Braids} \label{sec:tripleAlexander} Once we know that every link can be put into an $n$-crossing braid for even $n$, a natural question to ask is: is this also true when $n$ is odd? First, consider an open braid with $m$ strings. If we label the strings $1,\dots,m$ from left to right, we can see that any odd multi-crossing only mixes strings with the same parity. The closure of the braid must therefore have at least two components, which means that we cannot put knots in an odd multi-crossing braid form. However, we claim that we can put any link with two or more components into any odd multi-crossing braid form. In this section we prove this for $n=3$. \begin{theorem} \label{thm:tripleAlexander} Every link with two or more components can be represented as a closed triple crossing braid. \end{theorem} This will be extended to other odd $n$ in the next section. \subsection{Setup} Consider an open braid with $m \ge 3$ strings. As before, we can label the strings $1,\dots,m$ from left to right, at the top of the whole braid. We call this label the \emph{index} of a string. At a given section of the braid, we can also label the strings $1, \dots, m$ from left to right at the top of this section. We call this the \emph{position} of the string in the section. Note that for a given string, the index stays the same from the top of the braid to the bottom, but the position changes every time it is involved in a crossing. First we show a lemma which takes a double crossing braid and isotopes it into a form that is easier to work with. \begin{lemma} \label{lem:evenOdd} Let $\alpha$ be a closed double crossing braid with at least 2 components. Then we can find an isotopy of $\alpha$ such that in the resulting braid, a set of components always enter and leave the braid in an odd position, while the remaining components always enter and leave the braid in an even position. \end{lemma} \begin{proof} First, look at the the string with index 1, and remember this to be an odd component. Then we check if the string with index 2 is another component, which we remember to be an even component. If it is the same component, then we can find some string from another component that has a larger index, and make it have index 2 as follows. If the string from another component has index $i$, then we can conjugate the open braid by $\sigma^{i-1}\sigma^{i-2}\cdots\sigma^2$, so that this string now has index 2. Then we continue this process, checking at every step that a string with an odd (similarly even) index $i$ is an odd (similarly even) component or a new component, and if not, finding an appropriate component and making it index $i$. Then at the top of the braid, the strings alternate between the odd and the even components. If we do not have enough strands for the odd components or the even components, then we can perform stabilizations on a component to obtain new strings. \end{proof} Note that once we go through this process and make sure that the top of the braid alternates between odd and even components, we know that the bottom of the braid follows the same pattern. This is because if a string leaves the braid in an odd position $i$, it must be the same component as the string that enters the braid at the $i$th index. This means that this string must be an odd component. Similarly a string that leaves the braid in an even position must be an even component. This also means that if a string enters the braid at an odd (similarly even) index, then it must leave the braid at an odd (similarly even) index. In other words, the parity of the position of the string at the bottom of the braid must be the same as the parity of its index. \subsection{Putting the braid in level position} Recall from Section~\ref{sec:evenAlexander} that we can put the braid in level position; we assign heights to each string, and if we put in crossings in a way that keeps each string on its level, we can pull the strings taut and then we are left with a set of crossings which represents the corresponding permutation of the strings. We can assign these heights according to the index of each string, so that the index and the level coincide. We describe a process to isotope this braid so that, starting at the top of the braid and moving down, we are left with a sequence of triple crossings and then a double crossing braid in level position. (Note that when we refer to the ``top'' or the ``bottom'' of the braid, we always mean the beginning and the end of the braid word, rather than the heights that are assigned to a braid in level position, which we refer to in terms of its level.) First we define a notion that becomes key to this argument. A \emph{clasp} is two strings that are hooked together as below (Fig.~\ref{fig:clasp}). \begin{figure}[ht] \centering \includegraphics[scale=0.4]{clasp.png} \caption{Two strings hooked together is called a clasp.} \label{fig:clasp} \end{figure} Note that in a braid, a clasp appears as $\sigma_i \sigma_i$. \begin{lemma} \label{lem:tripleLevel} Let $\alpha$ be a double crossing braid with 3 or more strings. We can find an isotopy of this braid such that we are left with a sequence of triple crossings followed by a double crossing braid in level position. \end{lemma} \begin{proof} By Lemma~\ref{lem:evenOdd}, we can find an isotopy of $\alpha$ such that the parity of the position of each string is preserved from the top of the braid to the bottom. In the course of this argument, we can ``ignore'' any sequence of triple crossings at the top of the braid, as long as we do not try to take conjugations or stabilizations. This is because triple crossings always preserve the parity of the position of each string, which is the property that is central to this argument. If we can turn some double crossing braid into a triple crossing braid, we can clearly do the same to a sequence of triple crossing braid followed by this double crossing braid. Start with a double crossing braid $\alpha$ such that the parity of the position of each component remain the same. Assign the indices $1, \dots, m$ to the $m$ strings. We can then find a new (different) double crossing braid $\alpha'$, with the same projection as $\alpha$ but different crossings, that is in level position with respect to this leveling. We find an isotopy of $\alpha$ such that we are left with a sequence of triple crossings followed by $\alpha'$. We start from the top of the braid $\alpha$, and at every step we find a crossing that is different from the braid $\alpha'$ in level position, and change this crossing so that we are closer to being in level position. Note that in changing this crossing, we may introduce triple crossings at the top of the braid to preserve the original braid type, but we are not concerned with this. This will be done recursively. Suppose we want to change a double crossing at the top of the braid. Then we can introduce two trivial crossings under this crossing. Then the two crossings at the top can be turned into triple crossings as below. Thus we obtain a double crossing braid with this top crossing flipped (Fig.~\ref{fig:tripleTop}). \begin{figure}[ht] \centering \includegraphics[scale=0.3]{tripleTop.png} \caption{Changing a crossing at the top of the braid. Introduce two trivial crossings below, and use an extra strand to turn the two top crossings into triple crossings. We are now left with two triple crossings followed by a double crossing, which is now different from the one we started with.} \label{fig:tripleTop} \end{figure} Now suppose we want to change a crossing $\sigma$ in the middle of a braid, assuming that all crossings above it have been changed so that the portion above it is in level position. Let $A$ be the portion of the braid above this crossing, not including itself. Then we can add trivial crossings and obtain $A\sigma = A\sigma \sigma A^{-1} A \sigma^{-1}$ (Fig.~\ref{fig:tripleSequence}). It suffices to turn $A \sigma \sigma A^{-1}$ into triple crossings, for this would mean we have changed $A \sigma$ into a sequence of triple crossings followed by $A \sigma^{-1}$. Note that since $A$ is in level position, we know that $A^{-1}$ must also be in level position with the same leveling of strings. \begin{figure}[ht] \centering \includegraphics[scale=0.25]{tripleSequence.png} \caption{Starting with $A \sigma$, we can add $A^{-1} A \sigma^{-1}$ below it, since it is equal to the identity. If we can turn $A \sigma \sigma A^{-1}$ into triple crossings, then we are left with $A \sigma^{-1}$ as desired.} \label{fig:tripleSequence} \end{figure} Let $\sigma$ be a crossing that switches the strings with indices $a,b$, where $a < b$. Note that the $\emph{position}$ of these strings must be adjacent on this portion of the braid. Recall that $A$ and $A^{-1}$ are in level position, so these strings are in the levels $a$, $b$ respectively in both $A$ and $A^{-1}$. The crossings $\sigma \sigma$ form a clasp that violates the level position, and may get in the way of strings that are between $a$ and $b$. For clarification, call the first crossing of the clasp $\sigma_r$, and the second one $\sigma_s$. Let $i$, $i+1$ be the positions of the two strings at the top of the clasp. Note that because $\sigma_r$ is a crossing that violates the level position, $a$ must be the understrand of this crossing. Then $a$ is the overstrand of $\sigma_s$. We consider two possible cases: $a$ could be in the $i$th position at the top of the clasp, or it could be in the $i+1$st position. If $a$ is in $i+1$st position, this means that $b$ is in the $i$th position. Now, pull the parts of $a$ and $b$ in $A$ and $A^{-1}$ taut. Then there must be a crossing between $a$ and $b$ in both $A$ and $A^{-1}$. Let $\sigma_p$ the one in $A$, and let $\sigma_q$ be the one in $A^{-1}$. Observe that $a$ is the overstrand of both of these crossings, since $a$ is in a higher level than $b$. Then $\sigma_q$ cancels with $\sigma_s$, since $a$ is the overstrand in both. We can then move $\sigma_p$ down so that $\sigma_p$ and $\sigma_r$ is a clasp between the $a$ and $b$ strings; however $a$ is now in the $i$th position at the top of the clasp, and $b$ is in the $i+1$st (Fig.~\ref{fig:tripleSigma}). Note that the portions above and below this clasp are still in level position. Hence we can assume that $a$ is the $i$th position at the top of the clasp, and $b$ is in the $i+1$st position. \begin{figure}[ht] \centering \includegraphics[scale=0.25]{tripleSigma.png} \caption{If $a$ is the $i+1$st position at the top of the clasp, then there must be a crossing $\sigma_p$ between $a$ and $b$ in $A$, and a corresponding crossing $\sigma_q$ in $A^{-1}$. Then $\sigma_q$ cancels with $\sigma_s$, and a clasp is formed by $\sigma_p$ and $\sigma_r$. The portions above and below the clasp are still in level position. Thus we can assume that $a$ is in the $i$th position at the top of the clasp.} \label{fig:tripleSigma} \end{figure} The strings $a$ and $b$ look like the following figure (Fig.~\ref{fig:tripleDiagonalMess}). Both strings stay on their levels until they reach the clasp, where they wrap around each other and move back to their original levels. \begin{figure}[ht] \centering \includegraphics[scale=0.2]{tripleDiagonalMess.png} \caption{The left picture depicts the strings $a$ and $b$. Note that the other strings are not drawn, but they would all stay on the same level. The right picture is a view from the side, where this time the other strings are also drawn. The two strings move toward the level of the other string, wrap around each other, and then go back to their original levels.} \label{fig:tripleDiagonalMess} \end{figure} Now pull all the strings taut. All strings on levels higher than $a$ or lower than $b$ (note lower numbers are on higher levels here) are not affected, so they will go straight down from the top of the braid to the bottom. The strings $a$ and $b$ create a diagonal plane between the $a$th and the $b$th levels. The strings that are in between the levels $a$ and $b$ will be either above or below this plane, depending on what its position was at the clasp. If its position was greater than $i+1$, then it will be above this diagonal plane; it its position was less than $i$, then it will be below the plane. Thus we have the following picture (Fig.~\ref{fig:tripleDiagonalTaut}). \begin{figure}[ht] \centering \includegraphics[scale=0.2]{tripleDiagonalTaut.png} \caption{The left picture shows what happens to $a$ and $b$ once the strings are pulled taut. The right picture shows the view from the side, where we see which strings lie above/below the diagonal plane created by $a$ and $b$.} \label{fig:tripleDiagonalTaut} \end{figure} Therefore when all of these strings are pulled taut we have the following picture (Fig.~\ref{fig:tripleHook}). \begin{figure}[ht] \centering \includegraphics[scale=0.3]{tripleHook.png} \caption{Once the strings are pulled taut, we are left with one clasp between the strings $a$ and $b$, and all other strings are either above or below this clasp.} \label{fig:tripleHook} \end{figure} The strings that are in between $a$ and $b$ will go over both strands of the clasp, or under both strands of the clasp. This means that all of these $b-a-1$ strings can be moved to one side of the clasp. But any pair of such strings can be turned into two triple crossings as below (Fig.~\ref{fig:tripleResolve}). \begin{figure}[ht] \centering \includegraphics[scale=0.3]{tripleResolve.png} \caption{We can move the strings between $a$ and $b$ to one side, and resolve pairs of them into triple crossings.} \label{fig:tripleResolve} \end{figure} Therefore we only need to consider the case when one string and the clasp are left, or the case when only the clasp is left. If we have one string left we can simply let it run through the middle of the clasp (Fig.~\ref{fig:tripleResolveLast}). Note that we can always do this because it can be under or over both strings of the clasp, but it is never over one and under the other. If only the clasp is left we can take a strand from another string, and pull it under both crossings of the clasp. We can always do this because we have assumed that we have at least 3 strings. \begin{figure}[ht] \centering \includegraphics[scale=0.3]{TripleResolveLast.png} \caption{If one string is left, we let it run through the middle of the clasp. The left figure depicts the case when this string goes under both strands of the clasp. If only the clasp is left, then we can take an extra string and pull it under both crossings of the clasp.} \label{fig:tripleResolveLast} \end{figure} Therefore we can recursively change every crossing to turn the braid into a sequence of triple crossings followed by a double crossing braid in level position. \end{proof} \subsection{Obtaining the triple crossing braid} All we have to do now is to turn this double crossing braid in level position into a triple crossing braid. \begin{proof}[Proof of Theorem~\ref{thm:tripleAlexander}] Let $L$ be a link with at least 2 components. Put it in double crossing braid form. If it is a 2-braid, take a stabilization so that it has at least 3 strings. By Lemma~\ref{lem:tripleLevel}, we can find an isotopy of this braid so that it becomes a sequence of triple crossings followed by a double crossing braid in level position. It remains to isotope this double crossing braid into a triple crossing braid. Recall that each string enters and leaves the braid in positions with the same parity. We can pull the strings taut so that the braid is just a permutation of the even components, and a permutation of the odd components. We can then shift the strings slightly so that the permutation of the odd components occur in the top half of the braid, and the permutation of the even components occur in the bottom half of the braid (Fig.~\ref{fig:tripleLevel}). \begin{figure}[ht] \centering \includegraphics[scale=0.3]{tripleLevel.png} \caption{Once the strings are pulled taut, we are left with a permutation of the odd strings and a permutation of the even strings. Since the braid is in level position, we can separate the two permutations as above.} \label{fig:tripleLevel} \end{figure} Now, any permutation of the odd strands is generated by a transposition of two adjacent odd strands. Note that a triple crossing switches two adjacent odd strands. We can therefore obtain any permutation of the odd strands as a sequence of triple crossings, and similarly for a permutation of the even strands. Thus we have a triple crossing braid form for $L$. \end{proof} \section{Odd Multi-crossing Braids} \label{sec:oddAlexander} We have already seen that given a link $L$ with two or more components, it can be put into a triple-crossing braid form. In this section, we extend this result to any $n$-crossings, where $n$ is odd. Our goal is to prove the following theorem: \begin{theorem} \label{thm:oddAlexander} Every link with two or more components can be represented as a closed $n$-crossing braid, for all $n$. \end{theorem} \subsection{Setup} We start by putting $L$ in a triple crossing braid form by Theorem~\ref{thm:tripleAlexander}, and finding an equivalent $n$-crossing braid. In other words, we want to show that we can produce the triple crossings from a sequence of $n$-crossings. As in the case for $n$ even, by Lemma~\ref{lem:levelPosition} it suffices to show that we can obtain each triple crossing permutation as a combination of the $n$-crossing permutations. The triple crossing permutations are all transpositions of the form $(i, i+2)$, and when $n$ is odd, the $n$-crossing permutations are $\pi_j = (j,j+n-1)(j+1,j+n-2)\cdots(j+\frac{n-3}{2},j+\frac{n+1}{2})$. Observe that there are two types of triple crossing permutations: those that switch even numbered strings, and those that switch odd numbered strings. We call them \emph{even-string transpositions} and \emph{odd-string transpositions} respectively. \subsection{Obtaining pairs of transpositions with same parity} First we show that given a sufficiently large number of strings, we can use the $n$-crossing permutations to obtain any pair of even-string transpositions, or any pair of odd-string transpositions. Note that these pairs of transpositions must be ``disjoint'' in that the two transpositions permute 4 distinct numbers, for otherwise we would have a 3-cycle or the identity function. While we refrain from using the word ``disjoint'' in this context, it is worth noting that this is different from the meaning of disjoint crossings defined for Lemma~\ref{lem:levelPosition}. For example, a triple crossing permuting the first three strands would have the corresponding transposition $(1,3)$, and a triple crossing permuting the second to fourth strands would have the corresponding transposition $(2,4)$. These two crossings are not disjoint, but the two corresponding permutations are considered to be ``disjoint'' since they involve four distinct numbers. We can always make sure that we have enough strands as follows. Take the second string from the right, and perform an Type II Reidemeister move over the rightmost string. Then we can stabilize the portion that has now become the rightmost string. We have now added three double crossings, which can be put together into a triple-crossing (Fig.~\ref{fig:3-stabilization}). We call this move a \emph{3-stabilization}. \begin{figure}[ht] \centering \includegraphics[scale=0.2]{3-stabilization.png} \caption{A 3-stabilization.} \label{fig:3-stabilization} \end{figure} We start by obtaining one pair of even-string transpositions, and one pair of odd string transpositions. \begin{lemma} \label{lem:sameParityPairs} Let $n$ be odd, and let $m \ge \frac{5n-1}{2}$. Then we can obtain the permutations $(3,n)(2n-1,2n+1)$ and $(2,n+1)(2n,2n+2)$, as a product of the $n$-crossing permutations over $S_m$. \end{lemma} \begin{proof} First note that over $S_{(5n-1)/2}$ we have the crossing permutations $\pi_1, \dots, \pi_{(3n+1)/2}$. Recall that multiplying both sides of $\pi$ by $\sigma$, which is equivalent to conjugating by $\sigma$ since $\sigma^{-1} = \sigma$, switches the entries of $\pi$ according to $\sigma$. Recall also that this does not change the cycle type, which is to say it keeps the number of cycles and the length of each cycle constant. We start with some crossing permutation $\pi_i$ and take conjugates by some $\pi_j$ to obtain different permutations with the same cycle type. We start by obtaining one pair of odd-string transpositions. Take the crossing permutation $\pi_1 = (1,n)(2,n-1)\cdots(\frac{n-1}{2},\frac{n+3}{2})$. Note that in an odd multi-crossing, the central string is fixed; in this case $\pi_1$ keeps the $\frac{n+1}{2}$ string in the same position. We can conjugate by $\pi_n$ to obtain the permutation $(1,2n-1)(2,n-1)\cdots(\frac{n-1}{2},\frac{n+3}{2})$. Then conjugate by $\pi_{\frac{3n+1}{2}}$ to get $(1,2n+1)(2,n-1)\cdots(\frac{n-1}{2},\frac{n+3}{2})$. Thus we have moved the largest entry of the permutation to $2n+1$. We then perform a similar sequence to move the smallest entry to $2n-1$. This is done by conjugating by $\pi_1$ and then by $\pi_n$. We are now left with the permutation $(2n-1,2n+1)(2,n-1)\cdots(\frac{n-1}{2},\frac{n+3}{2})$. Note that the first conjugation by $\pi_1$ also moves the other entries around, but the pairs stay the same; for example $\pi_1$ switches the entries 2 and $n-1$, but this keeps the cycle $(2,n-1)$ constant. We can then multiply this permutation by $\pi_1$. The transpositions in the middle would cancel, and we are left with $(1,n)(2n-1,2n+1)$. Now, we can similarly obtain the permutation $(2,n+1)(2n,2n+2)$, a pair of even-string transpositions, by starting with $\pi_2$. Finally, we can conjugate $(1,n)(2n-1,2n+1)$ by $\pi_{(n-1)/2}$ to obtain $(1,n-2)(2n-1,2n+1)$. We can conjugate the result by $\pi_1$ to obtain $(3,n)(2n-1,2n+1)$. \end{proof} Now we show that we can obtain any pair of even-string transpositions, and any pair of odd string transpositions. We have two cases: $n=4k+3$ and $n=4k+1$. \begin{lemma} \label{lem:sameParity4k+3} Let $n = 4k+3$, and let $m \ge 3n+5$. Then we can obtain any pair of even-string transpositions in $S_m$, or any pair of odd-string transpositions in $S_m$, as a product of the $n$-crossing permutations over $S_m$. \end{lemma} \begin{proof} First observe that $3n+5 = 3(4k+3)+5 = 12k+14 = 2(6k+7)$, so we have at least $6k+7$ odd and even strings respectively. By Lemma~\ref{lem:sameParityPairs}, we can obtain the permutations $(3,n)(2n-1,2n+1)$ and $(2,n+1)(2n,2n+2)$ as a product of the $n$-crossing permutations over $S_m$. We want to move the entries of this permutation around so that we can get any pair of odd-string transpositions, or any pair of even-string transpositions. If we focus on the odd strings and ignore the even strings, we can see that the $(4k+3)$-crossing permutations starting on odd indices function as $(2k+2)$-crossing permutations on the odd strings. Since $2k+2$ is even, we know by Lemma~\ref{lem:sameCycleType} that if we have $3(2k+2)+1 = 6k+7$ odd strings, then we can obtain any odd-string permutation of the same cycle type by taking conjugations. Therefore we can obtain any pair of odd-string transpositions. Similarly we can obtain any pair of even-string transpositions. \end{proof} \begin{lemma} \label{lem:sameParity4k+1} Let $n = 4k+1$, and let $m \ge 3n-2$. Then we can obtain any pair of even-string transpositions in $S_m$, or any pair of odd-string transpositions in $S_m$, as a product of the $n$-crossing permutations over $S_m$. \end{lemma} \begin{proof} First observe that $3n-2 = 3(4k+1)-2 = 12k+4$ strings. This means that we have $12k+2 = 2(6k+1)$ strings that are not the 1st or the $m$th string. Hence we have $6k+1$ odd strings that are not on either end of the braid, and $6k+1$ even strings that are not on either end of the braid. As in the proof of Lemma~\ref{lem:sameParity4k+3}, we start with some permutation obtained through Lemma~\ref{lem:sameParityPairs}, and move the entries of these permutations around. Observe that if we consider an $n$-crossing permutation $\pi_i$ which starts on the $i$th strand, the $(4k+1)$-crossing permutations function as $2k$-crossing permutations on the string with parity different from $i$. Note, however, that none of the $(4k+1)$-crossings can function as a $2k$-crossings that acts on the 1st string or the $m$th string. First suppose we want to obtain the permutation $(a,b)(c,d)$, where the entries are all odd, and none of them are equal to 1 or $m$. We know by Lemma~\ref{lem:sameParityPairs} that we can obtain the permutation $(3,n)(2n-1,2n+1)$. Since $2k$ is even, we know by Lemma~\ref{lem:sameCycleType} that if we have $3(2k)+1 = 6k+1$ odd strings excluding the first and the last strings, then we can take conjugations and obtain $(a,b)(c,d)$. Now, suppose we want to obtain $(1,b)(c,d)$, where $b,c,d$ are all odd, and none of them are equal to $m$. Then, using Lemma~\ref{lem:sameParityPairs} and Lemma~\ref{lem:sameCycleType} we can first obtain some $(n,b')(c',d')$ where none of the entries are equal to 1 or $m$. Then we conjugate by $\pi_1$ to obtain $(1,b'')(c'',d'')$. Finally, using the same argument as in the proof of Lemma~\ref{lem:sameCycleType}, but only on the odd strings excluding the first and the last string, we can rearrange the remaining entries to obtain $(1,b)(c,d)$. By symmetry, we can similarly obtain any permutation with an entry equal to $m$ and none equal to 1; we first obtain $(m-n+1,b')(c',d')$, conjugate by $\pi_{m-n+1}$, and then rearrange the remaining entries. If an entry is equal to 1 and another is equal to $m$, then we can first obtain some permutations where the corresponding entries are equal to $n$ and $m-n+1$ respectively, then conjugate by $\pi_1$ and $\pi_{m-n+1}$, and rearrange the remaining two entries. We can apply the same argument to obtain any pair of even-string transpositions, by starting with $(2,n+1,2n,2n+2)$. We can use the same argument for the case when one of the entries are equal to $m$. \end{proof} Now we present a couple of lemmas for when $n=8k+5$ which will be useful in the proof of Theorem~\ref{thm:oddAlexander}. \begin{lemma} \label{lem:triple8k+5} Let $n=8k+5$. Then we can obtain any odd-string triple crossing permutation in $S_{(3n-1)/2}$, by conjugating $(1,3)$ with the $n$-crossing permutations over $S_{(3n-1)/2}$. \end{lemma} \begin{proof} First observe that over $S_{(3n-1)/2}$, we have the $n$-crossing permutations $\pi_1, \dots, \pi_{(n+1)/2}$. Suppose we want to obtain the transposition $(2i-1,2i+1)$. First consider the case when $2i-1 \le n$. We conjugate $(1,3)$ by $\pi_1$ to obtain $(n-2,n)$. We then conjugate by $\pi_{(n+1)/2}$ to obtain $(n+2,n)$. We then conjugate by $\pi_{i+1}$ to obtain $(2i-1,2i+1)$. (To see this, observe that $n$ is $j$ away from $i+n$, the highest entry in $\pi_{i+1}$, and that $2j+1$ is $j$ away from $j+1$, the lowest entry in $\pi_{i+1}$.) Now consider the case when $n < 2i-1 < \frac{3n-2}{2}$. Then we can move $(1,3)$ to some $(j,j+2)$ such that $j \le n$ and $j$ is $n-1$ away from $2i-1$. We can then conjugate $(j,j+2)$ by $\pi_{j+2}$ and then by $\pi_j$ to obtain $(2i-1,2i+1)$. \end{proof} \begin{lemma} \label{lem:pair8k+5} Let $n=8k+5$, and let $m \ge 4n-4$. Then we can obtain any pair of an even-string transposition and an odd-string transposition over $S_m$, corresponding to a disjoint pair of triple crossings, as a product of the $n$-crossing permutations over $S_m$. \end{lemma} \begin{proof} Observe that when $n=8k+5$, an $n$-crossing permutation consists of $2k+1$ even-string transpositions and $2k+1$ odd-string transpositions. We now show that we can obtain any pair consisting of one odd-string transposition and one even-string transposition whose corresponding crossings are disjoint. Let $(l_a, l_a+2)(l_b, l_b+2)$ be the permutation we want to obtain, with $l_a+2 < l_b$. First suppose $l_a$ is odd. We can then start $\pi_{(n+1)/2}$, and cancel all but one of the even-string transpositions by multiplying it with pairs of even-string transpositions as before. We can similarly cancel all but one of the odd-string transpositions. Thus we are left with a single even-string transposition and a single odd-string transposition. We may choose the pairs to cancel from the outside so that we are left with a permutation $(n-2,n+2)(n-1,n+1)$. We first ``separate'' the two transpositions. We can conjugate $(n-2,n+2)(n-1,n+1)$ by $\pi_{n-1}$ to obtain $(n-2,2n-5)(2n-2,2n-4)$. When $n=5$ this is equivalent to $(n-2,n)(n+3,n+1)$. When $n>5$, we conjugate $(n-2,2n-5)(2n-2,2n-4)$ by $\pi_{2n-4}$ to obtain $(n-2,2n-5)(3n-7,3n-5)$. Then we conjugate by $\pi_{(3n-7)/2}$ to obtain $(n-2,2n-3)(3n-7,3n-5)$. Then we conjugate by $\pi_{n-1}$ again to obtain $(n-2,n)(3n-7,3n-5)$. Then we conjugate by $\pi_{2n-3}$ to obtain $(n-2,n)(2n,2n-2)$. Finally, we conjugate by $\pi_{n+1}$ to obtain $(n-2,n)(n+1,n+3)$. In both cases, we have the permutation $(n-2,n)(n+1,n+3)$. Now we can conjugate by $\pi_1$, and then by $\pi_4$ to obtain $(1,3)(4,6)$. Consider the $4n-8$ strings starting with the 4th string. Then, as in the proof of Lemma~\ref{lem:sameParity4k+1}, we can move the entries of $(4,6)$ to any even numbers between 4 and $4n-5$. Thus we can obtain the permutation $(1,3)(l_b,l_b+2)$. Now we want to move $(1,3)$ to $(l_a, l_a+2)$. First consider the case when $l_b \le \frac{3n-1}{2}$. We conjugate $(1,3)(l_b,l_b+2)$ by $\pi_{l_b}$ to obtain $(1,3)(l_b+n-1,l_b+n-3)$. We then conjugate by $\pi_{l_b+n-3}$ to obtain $(1,3)(l_b+2n-6, l_b+2n-4)$. Now by Lemma~\ref{lem:triple8k+5} we can conjugate by $n$-crossing permutations over $S_{(3n-1)/2}$ to move $(1,3)$ to $(l_a,l_a+2)$. Since $l_b+2n-6 > \frac{3n-1}{2}$ the other permutation is not affected, and we are left with $(l_a,l_a+2)(l_b+2n-6, l_b+2n-4)$ We can then conjugate back by $\pi_{l_b+n-3}$ and then by $\pi_{l_b}$ to obtain $(l_a, l_a+2)(l_b, l_b+2)$. Next, consider the case when $l_a + 2 \le \frac{3n-1}{2} < l_b$. Then by Lemma~\ref{lem:triple8k+5} we can move $(1,3)$ to $(l_a,l_a+2)$ without affecting any other elements. Finally consider the case when $l_a + 2 > \frac{3n-1}{2}$. Then by Lemma~\ref{lem:triple8k+5} we can move $(1,3)$ to some $(j,j+2)$, where $j < \frac{3n-1}{2}$ and $j$ is a multiple of $n-1$ away from $l_a$. Then we can conjugate by $\pi_{j+2}$ and then by $\pi_j$ to obtain $(j+n-1,j+n+1)$. We can continue this until we reach $(l_a,l_a+2)$. If $l_a$ is even, we start with $\pi_{(n+3)/2}$, and similarly take conjugations to obtain the permutation $(2,4)(5,7)$. We can obtain $(2,4)(l_b,l_b+2)$ using the $4n-8$ strings starting with the 5th string. Moving $(2,4)$ to $(l_a,l_a+2)$ can be done in the same way as above. \end{proof} \subsection{Proof of Theorem~\ref{thm:oddAlexander}} Now we use these pairs of transpositions to produce the desired braid. \begin{proof}[Proof of Theorem~\ref{thm:oddAlexander}] If $n$ is even, the result follows directly from Theorem~\ref{thm:evenAlexander}. If $n$ is odd, by Theorem~\ref{thm:oddAlexander} we can put the link $L$ in a triple crossing braid form. We consider three cases for when $n$ is odd: $n=4k+3$, $n=8k+5$, and $n=8k+1$. \noindent \textbf{Case 1: $n = 4k+3$.} Take 3-stabilizations until we have at least $3n+5$ strings. It suffices to show that we can obtain any triple-crossing permutation, for then by Lemma~\ref{lem:levelPosition} we can produce any triple-crossing. Observe that when $n=4k+3$, an $n$-crossing permutation consists of $k$ even-string transpositions and $k+1$ odd-string transpositions, or $k$ odd-string transpositions and $k+1$ even-string transpositions. Recall that by Lemma~\ref{lem:sameParity4k+3}, we can obtain pairs of odd-string transpositions, and pairs of even-string transpositions, as products of the $n$-crossing permutations. We can obtain one odd-string transposition as follows: start with $\pi_1 = (1,n)(2,n-1)\cdots(\frac{n-1}{2},\frac{n+3}{2})$, which has an odd number of odd-string transpositions and an even number of even-string transpositions. Then we can cancel the even number of even-string transpositions, and all but one of the odd-string transpositions, by multiplying it with pairs of odd-string transpositions and pairs of even-string transpositions. We are left with a single odd-string transposition. As before, we can focus on the odd strings and see that we can obtain any permutation of the same cycle type by Lemma~\ref{lem:sameCycleType}, since the $(4k+3)$-crossing permutations starting on odd indices function as $(2k+2)$-crossing permutations on the odd strings. Therefore we can obtain any odd-string transposition. We can similarly obtain any even-string transposition. This shows that we can produce any triple-crossing as a sequence of $n$-crossings, so $L$ can be put into an $n$-crossing braid. \noindent \textbf{Case 2: $n=8k+5$.} First observe that an $(8k+5)$-crossing permutation is an even permutation (here we mean even in the traditional sense of the word in symmetric groups, namely that it can be written as a product of an even number of transpositions.) Therefore we want to start with a triple-crossing braid with an even number of crossings, and then turn pairs of triple-crossings into sets of $n$-crossings. Take 3-stabilizations of $L$ until we have at least $3n-2$ strings. If this braid has an odd number of triple crossings at this stage, take one more 3-stabilization so that we have an even number of triple crossings. We claim we can put these triple-crossings in pairs such that each pair is disjoint, in the sense that they involve 6 distinct strings in total. This is because if two consecutive crossings involve a common strand, then we can put one of them in a pair with a dummy crossing far away, and put the other in another pair which undoes the dummy crossing (Fig.~\ref{fig:disjointPair}). \begin{figure}[ht] \centering \includegraphics[scale=0.2]{disjointPair.png} \caption{If two consecutive crossings share a common strand, we can put them in two disjoint pairs along with two dummy crossings.} \label{fig:disjointPair} \end{figure} Now it suffices to show that we can obtain any permutation corresponding to a disjoint pair of triple-crossings. By Lemma~\ref{lem:sameParity4k+1}, we can obtain pairs of transpositions of the same parity. By Lemma~\ref{lem:pair8k+5}, we can obtain any pair of an even-string transposition and an odd-string transposition, corresponding to a disjoint pair of triple crossings. Thus by Lemma~\ref{lem:levelPosition} we can produce any pair of disjoint triple-crossings using $n$-crossings. Hence $L$ can be put into a $n$-crossing braid. \noindent \textbf{Case 3: $n=8k+1$.} First observe that a $(8k+1)$-crossing permutation consists of $2k$ even-string transpositions and $2k$ odd-string transpositions. Therefore we want to start with a triple-crossing braid with an even number of even-string crossings, and an even number of odd-string crossings. Take 3-stabilizations of $L$ until we have at least $3n-2$ strings. If the number of even-string crossings and odd-string crossings are both even at this stage, then it is in the desired form. If both numbers are odd, then we can take two 3-stabilizations, which will increase both numbers by one. The remaining case is when the number of one set of crossings is odd, and the number of the other set of crossings is even. In this case, we consider two possibilities. First suppose the number of strings is even. Then note that a 3-stabilization on the second string will increase the number of even-string crossings, and a 3-stabilization on the second last string will increase the number of odd-string crossings (Fig.~\ref{fig:3-stabilization2}). Hence if the parity is off by one, we can always find the appropriate 3-stabilization. \begin{figure}[ht] \centering \includegraphics[scale=0.3]{3-stabilization2.png} \caption{A 3-stabilization on the second string will increase the number of even-string crossings, and a 3-stabilization on the second last string will increase the number of odd-string crossings.} \label{fig:3-stabilization2} \end{figure} Now suppose the number of strings is odd. Then both 3-stabilizations increase the number of even-string crossings. But once we take this stabilization, the number of even-string crossings and the number of odd-string crossings will have the same parity. We can therefore put it in the desired form with at most two more 3-stabilizations. Hence we can put $L$ into a triple-crossing braid with an even number of even-string crossings, and an even number of odd-string crossings. As before we can put these crossings into pairs that are disjoint. Note that here we must also put them in pairs such that they have the same parity. We can then produce each pair of crossings since we can obtain pairs of transpositions with the same parity by Lemma~\ref{lem:sameParity4k+1}. Therefore we can produce each pair of triple-crossings with the same parity using the $n$-crossings, so $L$ can be put into a $n$-crossing braid form. \end{proof} \section{Bounds on Braid Indices} \label{sec:braidIndices} In this section we find relationships between the multi-crossing braid indices. Many arguments in this section have been simplified; details can be found in the original thesis. Let $\beta_n(L)$ be the minimum number of strings necessary to represent the link $L$ as an $n$-crossing braid. We define $\beta_n(L) = \infty$ if $L$ cannot be represented as an $n$-crossing braid. Note this happens if and only if $L$ is a knot and $n$ is odd. A simple observation tells us the following. \begin{theorem} Let $L$ be a link. For any $n \ge 2$, we have \[ \beta_2(L) \le \beta_n(L). \] \end{theorem} \begin{proof} Observe that any multi-crossing braid can be turned into a double crossing braid with the same number of strings, by breaking up each multi-crossing into double crossings. \end{proof} \subsection{Even multi-crossing braids} For even $n$ we have the following results. \begin{theorem} \label{thm:braidEven} Let $L$ be a link that is not an unlink. Let $n \le 202$, and $m \ge n+2$. \begin{itemize} \setlength\itemsep{0.3em} \item[(i)] If $n=4k+2$, then we have $\beta_n(L) \le \beta_m(L)$. \item[(ii)] If $n=4k$, then we have $\beta_n(L) \le \beta_m(L) + 1$. Moreover, if $m=4k$ or $m=4k+1$, then we have $\beta_n(L) \le \beta_m(L)$. \end{itemize} \end{theorem} \begin{proof} \text{} \begin{itemize} \setlength\itemsep{0.3em} \item[(i)] We can show through computations in Mathematica that for all $n\le 202$ with $n=4k+2$, the $n$-crossing permutations over $S_{n+2}$ generate $S_{n+2}$. (See the original thesis for the code.) By Lemma~\ref{lem:levelPosition}, this means that we can obtain every double crossing as a product of $n$-crossings if we have $n+2$ strings. Let $m \ge n+2$. Since an $m$-crossing requires $m$ strings, and $L$ is not an unlink, we have $\beta_m(L) \ge n+2$. We can decompose an $m$-crossing braid into a double crossing braid with the same number of strings. Then each double crossing can be turned into a product of $n$-crossings, so we have an $n$-crossing braid with $\beta_m(L)$ strings. This means that there is no need for the extra stabilization, so we have $\beta_n(L) \le \beta_m(L)$ for all such $m \ge n$. \item[(ii)] We can again show with Mathematica that for all such $n \le 200$, the $n$-crossing permutations over $S_{n+2}$ generate $A_{n+2}$, the alternating group on $n$ elements. The first inequality follows as in (i), except in this case we may have to take a stabilization to ensure that the number of double crossings is even. Observe that if $m = 4k$ or $m=4k+1$, then an $m$-crossing breaks down into an even number of double crossings. Thus the double crossing braid must have an even number of crossings. Therefore $\beta_n(L) \le \beta_m(L)$ for all such $m \ge n$. \end{itemize} \end{proof} Hence we have the following relationships: \begin{align*} \beta_2&(L) \le \beta_6(L) \le \beta_{10}(L) \le \cdots \le \beta_{202}(L) \le \beta_{206}(L) \\ &\mathbin{\rotatebox[origin=c]{270}{$\le$}} \hspace{25pt} \mathbin{\rotatebox[origin=c]{270}{$\le$}} \hspace{32pt} \mathbin{\rotatebox[origin=c]{270}{$\le$}} \hspace{55pt} \mathbin{\rotatebox[origin=c]{270}{$\le$}} \\ \beta_4&(L) \le \beta_8(L) \le \beta_{12}(L) \le \cdots \le \beta_{204}(L). \end{align*} And the following: \[ \beta_4(L) \le \beta_6(L) + 1, \hspace{5pt} \beta_8(L) \le \beta_{10}(L) + 1, \hspace{5pt} \dots, \hspace{5pt} \beta_{200}(L) \le \beta_{202}(L) + 1. \] Note that the inequalities need not end at 202; this is an arbitrary choice on how far to go with the Mathematica computation. It would however be interesting to ask whether there is a general argument that could extend the inequalities to all even $n$. \begin{example} The inequality $\beta_{4k} \le \beta_{4k+2} + 1$ is strict for any link $L$ with the property that $\beta_2(L) \ge 4k+4$, and the number of crossings when it realizes the double crossing braid index is odd. We know this because Markov moves (\cite{markov}) preserve the parity of $b+c$, where $b$ is the number of strings and $c$ is the number of crossings. A conjugation does not alter either $b$ or $c$, and a stabilization increases both $b$ and $c$ by 1. This means that given such a link, it cannot be represented as any double crossing braid with $\beta_2(L)$ strings and an even number of double crossings. In order to turn it into a $4k$-crossing braid we must therefore take a stabilization to make the number of crossings even, so $\beta_{4k}(L) = \beta_2(L) + 1$. We can of course realize $L$ as a $(4k+2)$-crossing braid with $\beta_2(L)$ strings, so $\beta_{4k+2}(L) = \beta_2(L) = \beta_{4k}(L) - 1$. An artificial example of such a link is a split link consisting of the trefoil knot and an unlink with $4k+2$ components. \end{example} We also have inequalities that hold for infinitely many even $n$. \begin{theorem} \label{thm:braidEvenAll} Let $L$ be a link that is not an unlink. \begin{itemize} \setlength\itemsep{0.3em} \item[(i)] Let $n=4k+2$. Then for any $m \ge 3n+1$, we have $\beta_n(L) \le \beta_m(L)$. \item[(ii)] Let $n=4k$. Then for any $m \ge 3n+1$, we have $\beta_n(L) \le \beta_m(L) + 1$. Moreover, if $m=4k$ or $m=4k+1$, then we have $\beta_n(L) \le \beta_m(L)$. \end{itemize} \end{theorem} \begin{proof} (i) The proof of Theorem~\ref{thm:evenAlexander} required at least $3n+1$ strings. This means that for $m \ge 3n+1$, we can turn an $m$-crossing braid into an $n$-crossing braid with the same number of strings, so $\beta_n(L) \le \beta_m(L)$. (ii) can be proved similarly. \end{proof} \begin{corollary} Let $L$ be a link that is not an unlink. Then for all even $n$, \[ \beta_n(L) \le \beta_{3n+1}(L). \] \end{corollary} \begin{proof} After noting that when $n=4k$, $3n+1$ can be written in as $8k'+1$ or $8k'+5$ for some $k'$, the result follows directly from Theorem~\ref{thm:braidEvenAll}. \end{proof} \subsection{Odd multi-crossing braids} For links with two or more components, we can consider braid indices for odd $n$. First consider the case when $n=3$. \begin{theorem} \label{thm:braidn=3} Let $L$ be a link with two or more components that is not an unlink. Then \[ \beta_3(L) = \begin{cases} 3 &\mbox{ if } \beta_2(L) = 2; \\ \beta_2(L) &\mbox{ otherwise.} \end{cases} \] \end{theorem} \begin{proof} The proof of Theorem~\ref{thm:tripleAlexander} required that the double crossing braid have at least 3 strings, and gave us a triple crossing braid with the same number of strings. \end{proof} By this we can see that $\beta_3(L) \le \beta_n(L)$ for any odd $n$. However, this can also be seen by noting that any $n$-crossing can be decomposed into triple crossings; any $n$-crossing is a permutation of even strings and a permutation of odd strings, each of which can be realized as a product of triple crossings. \begin{theorem} \label{thm:braidOdd} Let $L$ be a link with two or more components that is not an unlink. Let $n \le 205$, and $m \ge n+3$. \begin{itemize} \setlength\itemsep{0.3em} \item[(i)] If $n=4k+3$, then we have $\beta_n(L) \le \beta_m(L)$. \item[(ii)] If $n=8k+5$, then we have $\beta_n(L) \le \beta_m(L) + 1$. Moreover, if $m=4k+1$, then we have $\beta_n(L) \le \beta_m(L)$. \item[(iii)] If $n=8k+1$, then we have $\beta_n(L) \le \beta_m(L) + 3$. Moreover, if $m=8k+1$, then we have $\beta_n(L) \le \beta_m(L)$. \end{itemize} \end{theorem} \clearpage \begin{proof} \text{} \begin{itemize} \setlength\itemsep{0.3em} \item[(i)] We can show through computations in Mathematica that for all $n\le 203$ with $n=4k+3$, the $n$-crossing permutations over $S_{n+3} = S_{4k+6}$ generate $S_{2k+3} \times S_{2k+3}$. Each $S_{2k+3}$ corresponds to permutations of odd strings and permutations of even strings. By Lemma~\ref{lem:levelPosition}, this means that we can obtain every triple crossing as a product of $n$-crossings if we have $n+2$ or more strings. The rest of the proof follows as in the proof of Theorem~\ref{thm:braidEven} (i), but by decomposing the $m$-crossing braid into triple crossings. \item[(ii)] We can again show with Mathematica that for all such $n \le 205$, the $n$-crossing permutations $S_{n+3} = S_{8k+8}$ generate half of $S_{4k+4} \times S_{4k+4}$. Since we know that the $n$-crossing permutations must be even, this shows that the $n$-crossing permutations generate all pairs of triple crossings permutations. Thus, if we have $n+3$ or more strings, we can obtain every pair of triple crossings as a product of $n$-crossings. The rest of the proof follows as in the proof of Theorem~\ref{thm:braidEven} (ii), but by decomposing the $m$-crossing braid into triple crossings. \item[(iii)] We can again show with Mathematica that for all such $n \le 201$, the $n$-crossing permutations $S_{n+3} = S_{8k+4}$ generate $A_{4k+2} \times A_{4k+2}$. This shows that the $n$-crossing permutations generate all pairs of even-string transpositions, and all pairs of odd-string transpositions. Thus, if we have $n+3$ or more strings, we can obtain every pair of odd triple crossings and every pair of even triple crossings as a product of $n$-crossings. The rest of the proof follows as in (ii). \end{itemize} \end{proof} As for the case with $n$ even, we suspect that the inequalities can be extended for all $n$. The following inequalities hold for infinitely many $n$. This is done as in the proof of Theorem~\ref{thm:braidEvenAll}, by checking the number of strings that were necessary for the proof of Theorem~\ref{thm:oddAlexander}. \begin{theorem} \label{thm:braidOddAll} Let $L$ be a link with two or more components that is not an unlink. \begin{itemize} \setlength\itemsep{0.3em} \item[(i)] Let $n=4k+3$. Then for any $m \ge 3n+5$, we have $\beta_n(L) \le \beta_m(L)$. \item[(ii)] Let $n=8k+5$. Then for any $m \ge 3n-2$, we have $\beta_n(L) \le \beta_m(L) + 1$. Moreover, if $m=4k+1$, then we have $\beta_n(L) \le \beta_m(L)$. \item[(iii)] Let $n=8k+1$. Then for any $m \ge 3n-2$, we have $\beta_n(L) \le \beta_m(L) + 3$. Moreover, if $m=8k+1$, then we have $\beta_n(L) \le \beta_m(L)$. \end{itemize} \end{theorem} Finally we present an inequality that holds for all $n \ge 8$, even or odd. \begin{corollary} Let $L$ be a link that is not an unlink. Then for all $n \ge 8$, \[ \beta_n(L) \le \beta_{4n-3}(L). \] \end{corollary} \begin{proof} Observe that when $n \ge 8$, we have $4n-3 \ge 3n+5$. Observe the following about $4n-3$. When $n=4k$, we have $4n-3 = 16k-3 = 8(2k-1) + 5$. When $n=8k+5$, we have $4n-3 = 32k+20-3 = 8(4k+2) + 1$. Finally, when $n=8k+1$, we have $4n-3 = 32k+4-3 = 8(4k) + 1$. Then the result follows directly from Theorem~\ref{thm:braidEvenAll} and Theorem~\ref{thm:braidOddAll}. \end{proof} \FloatBarrier
{ "timestamp": "2018-05-14T02:09:42", "yymm": "1805", "arxiv_id": "1805.04427", "language": "en", "url": "https://arxiv.org/abs/1805.04427", "abstract": "Traditionally, knot theorists have considered projections of knots where there are two strands meeting at every crossing. A multi-crossing is a crossing where more than two strands meet at a single point, such that each strand bisects the crossing. In this paper we generalize ideas in traditional braid theory to multi-crossing braids. Our main result is an extension of Alexander's Theorem. We prove that every link can be put into an $n$-crossing braid form for any even $n$, and that every link with two or more components can be put into an $n$-crossing braid form for any $n$. We find relationships between the $n$-crossing braid indices, or the number of strings necessary to represent a link in an $n$-crossing braid.", "subjects": "Geometric Topology (math.GT)", "title": "Multi-crossing Braids", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9888419690807407, "lm_q2_score": 0.8104789109591832, "lm_q1q2_score": 0.8014355622112931 }
https://arxiv.org/abs/2003.06457
Interactions between Hlawka Type-1 and Type-2 Quantities
The classical Hlawka inequality possesses deep connections with zonotopes and zonoids in convex geometry, and has been related to Minkowski space. We introduce Hlawka Type-1 and Type-2 quantities, and establish a Hlawka-type relation between them, which connects a vast number of strikingly different variants of the Hlawka inequalities, such as Serre's reverse Hlawka inequality in the future cone of the Minkowski space, the Hlawka inequality for subadditive function on abelian group by Ressel, and the integral analogs by Takahasi et al. Besides, we announce several enhanced results, such as the Hlawka inequality for the power of measure function. Particularly, we give a complete study of the Hlawka inequality for quadratic form which relates to a work of Serre.
\section{Introduction} Hlawka's inequality saying for any $x,y,z$ in a inner product space \begin{equation}\label{eq:Hlawka-classical} \|x\|+\|y\|+\|z\|+\|x+y+z\|\ge \|x+y\|+\|y+z\|+\|z+x\|, \end{equation} was proved firstly by Hlawka and originally appeared in 1942 in a paper of Hornich \cite{Hornich42}. It has a long series of investigations and extensions, such as the Hlawka inequality in integral form \cite{TTW00,TTW09} and abelian group \cite{R15}. The readers can also find an excellent summary of related works in \cite{Wl}, and the beautiful relations to discrete and convex geometry like zonotopes as well as zonoids by Witsenhausen \cite{W78,W73,SW1983}. Recently, Serre consider the pseudo-norm for the future cone of the Minkowski space \cite{Serre15}. There he presented the reverse Hlawka-type inequality. According to these beautiful works, the classical Hlawka inequality has deep connections with zonotopes and zonoids in convex geometry, and relates to the geometry on timelike cone of Minkowski space. Note that the proof of \eqref{eq:Hlawka-classical} depends on the identity \begin{equation}\label{eq:Hlawka-classical-2form} \|x\|^2+\|y\|^2+\|z\|^2+\|x+y+z\|^2= \|x+y\|^2+\|y+z\|^2+\|z+x\|^2, \end{equation} which is a quadratic form equality. To some extend, the one-homogeneous inequality \eqref{eq:Hlawka-classical} essentially relates to the two-homogeneous equality \eqref{eq:Hlawka-classical-2form}. In this work, we introduce Hlawka one-form and Hlawka two-form, and establish a Hlawka-type relation which encodes the signatures of them (see Theorem \ref{thm:operator-main}). This helps us to give the Hlawka inequality for a class of functions on semigroups. By this result, we can connect a vast number of Hlawka inequalities in the literature, even though they come from various perspectives and are very different from each other. Furthermore, we announce several exciting results, such as the Hlawka inequality for the power of measure function. Particularly, we investigate the Hlawka inequality on quadratic form thoroughly. For a glimpse of these results, we give some remarkable notes here: \begin{itemize} \item For a reversed version of the Hlawka inequality, Serre gave a demonstration in the future cone of Minkowski space \cite{Serre15}. In that paper, he shows: if $q$ is a quadratic form on $\mathbb{R}^n$ with signature $(1, n-1)$, then the length $l=\sqrt{q}$ satisfies \begin{equation}\label{eq:Serre} l(x)+l(y)+l(z)+l(x+y+z)\leq l(x+y)+l(y+z)+l(z+x) \end{equation} for every vectors $x,y,z$ in the future cone with respect to $q$. In the present paper, we show a simple proof for \eqref{eq:Serre}, and give a systematic study for the Hlawka inequality on quadratic form (see Section \ref{sec:quadratic}). \item Ressel \cite{R15} shows a generalization of the Hlawka inequality for subadditive functions on abelian group. In this work, we extend his result to the setting of sub/super-additive functions on semigroup (see Section \ref{sec:semigroup}). For convenience, Ressel's result is provided in Example \ref{cor:R15} as an application. \item Takahasi et al \cite{TTW00,TTW09} study the integral analogs of the Hlawka inequality. In Section \ref{sec:integral}, we generalize this integral inequality to the form of positive linear operator, and their main theorem is rewritten in Example \ref{cor:TTW00}. \end{itemize} This paper provides a theorem combining the above different progresses together in a unified form (see Theorem \ref{thm:operator-main}), which also produces several other promotive results. \section{ The Hlawka-type relation of Hlawka one-form and two-form}\label{sec:Hlawka-relation} \noindent \textbf{Basic setting}: Given nonempty sets $\Omega$ and $X$, the ring $\mathbb{R}^\Omega$ of all real valued functions on $\Omega$ is a real linear space equipped with a product operator `$\cdot$'. Take a linear subspace $\mathcal{S}\subset \mathbb{R}^\Omega$ equipped with a linear and signature-preserving function $T: \mathcal{S} \to \mathbb{R}$ (i.e., for $\zeta\in \mathcal{S}$ satisfying $\forall\omega\in\Omega,\,\zeta(\omega)\geq0$, there holds $T(\zeta)\geq0$). We use $1$ to denote the constant function in $\mathcal{S}$ satisfying $1(\omega)=1$, $\forall\omega\in\Omega$. \begin{thm}\label{thm:operator-main} Given $a,b\in \mathbb{R}$ with $a\ne0$, $a+b >0$ and $\eta, \xi:\Omega\to X$, for $f: X\rightarrow \mathbb{R}$ satisfying $(f\circ\xi)(\omega)-(f\circ\eta)(\omega) \leq b$, $\forall\omega\in\Omega$, where `$\circ$' represents the composition operator, we have the following: \item (I) If $(f\circ\xi)(\omega)+(f\circ\eta)(\omega) \leq a$, $\forall\omega\in\Omega$, then $C_2\ge 0$ implies $C_1\ge 0$, and $C_1\le 0$ implies $C_2\le 0$; \item(II) If $(f\circ\xi)(\omega)+(f\circ\eta)(\omega) \geq a$, $\forall\omega\in\Omega$, then $C_1\ge 0$ implies $C_2\ge 0$, and $C_2\le 0$ implies $C_1\le 0$. \noindent Here $$C_1=(T(1)-c)b+T(f\circ\eta)- T(f\circ\xi)$$ is the {\sl Hlawka one-form}, and $$C_2=(T(1)-c)b^2+T(f^2\circ\eta) - T(f^2\circ\xi)$$ is the corresponding {\sl Hlawka two-form}, in which $c:=\frac{2}{a}T(f\circ\eta)$. \end{thm} In summary, Theorem \ref{thm:operator-main} says that under suitable `summation control' and `difference control', the signatures of Hlawka one-form $C_1$ and Hlawka two-form $C_2$ are essentially depended on each other in some way. \begin{proof} First, the relation among the quantities in Theorem \ref{thm:operator-main} can be shown in the following diagram: \begin{center} \begin{spacing}{0.5} $$ \xymatrix{ \Omega\ar[rr]^{\xi,\eta}\ar@/_1pc/[rrrr]_{ \scriptstyle\begin{array}{c}\scriptstyle f\circ \xi,f\circ\eta \\{ \scriptstyle \in}\end{array} }& & X\ar[rr]^f & &\mathbb{R}\\ & & {\mathcal{S} \ar[rru]_{T} } & & } $$ \end{spacing} \end{center} We note the following identities: \begin{align*} &\;\;\;\; T\left((a-f\circ\eta-f\circ\xi)(f\circ\eta+b-f\circ\xi)\right) \\&= aT(f\circ\eta)-T(f^2\circ\eta)+T(f^2\circ\xi)+T\left((f\circ\eta) \cdot (f\circ\xi)\right)-T((f\circ\xi )\cdot (f\circ\eta)) \\&\;\;\;\; -aT(f\circ\xi)-bT(f\circ\eta) -bT(f\circ\xi)+abT(1) \\&= aT(f\circ\eta)+\left(T(1)-c\right)b^2+ T(f\circ\eta)b + (T(f^2\circ\xi) -T(f^2\circ\eta) -(T(1)-c)b^2) \\&\;\;\;\; -T(f\circ\xi)(a+b)+(T(1)a-2T(f\circ\eta))b +T((f\circ\eta) \cdot (f\circ\xi))-T((f\circ\xi )\cdot (f\circ\eta)) \\&= T(f\circ\eta)(a+b)+\left(T(1)-c\right)b^2 + (T(f^2\circ\xi) -T(f^2\circ\eta) -(T(1)-c)b^2) \\&\;\;\;\; -T(f\circ\xi) (a+b)+(T(1)-c)ab \\&= \left((T(1)-c) b+T(f\circ\eta) -T(f\circ\xi)\right)(a+b) + (T(f^2\circ\xi) -T(f^2\circ\eta) -(T(1)-c)b^2), \end{align*} where the notation $f^2\circ\eta:=(f\circ\eta)\cdot(f\circ\eta)$ is used. Therefore, we obtain \begin{align} &\;\;\;\; T\left((a-f\circ\eta-f\circ\xi)(f\circ\eta+b-f\circ\xi)\right) \notag \\&= \left((T(1)-c) b+T(f\circ\eta) -T(f\circ\xi)\right)(a+b) - ((T(1)-c)b^2+T(f^2\circ\eta) -T(f^2\circ\xi)). \label{eq:T} \end{align} (I). For any $ \omega\in \Omega$, $f\circ\eta(\omega)+b\geq f\circ\xi(\omega)$ and $a\geq f\circ\eta(\omega)+f\circ\xi(\omega)$. By the assumption, $\forall \omega\in\Omega$, $$(a-f\circ\eta(\omega)-f\circ\xi(\omega))(f\circ\eta(\omega)+b-f\circ\xi(\omega))\geq 0.$$ This deduces that $$T\left((a-f\circ\eta-f\circ\xi)(f\circ\eta+b-f\circ\xi)\right)\geq0.$$ Accordingly, Eq.~\eqref{eq:T} gives $$\left((T(1)-c)b+T(f\circ\eta) -T(f\circ\xi)\right)(a+b)\geq (T(1)-c)b^2+T(f^2\circ\eta) -T(f^2\circ\xi)$$ which arrives the final result. (II). Since only the assumption $a\leq f\circ\eta(\omega)+f\circ\xi(\omega)$ is reversed, similar process gives $$\left((T(1)-c)b+T(f\circ\eta) -T(f\circ\xi)\right)(a+b)\leq (T(1)-c)b^2+T(f^2\circ\eta) -T(f^2\circ\xi)$$ and then the reversed case could be verified immediately. \end{proof} \begin{remark} From the proof of Theorem \ref{thm:operator-main}, it is obvious that the conditions could be weaken as follows: Given $a\in \mathbb{R}\setminus\{0\}, b\in \mathbb{R}$ with $ a+b >0 $ and $\eta, \xi \in \mathcal{S}$, $f: X\rightarrow \mathbb{R}$, let $\tilde{\Omega}=\{\omega\in\Omega|b+f\circ\eta(\omega)-f\circ\xi(\omega)\neq 0\}$. If $a \geq f\circ\eta(\omega)+f\circ\xi(\omega),\; b \geq f\circ\xi(\omega)-f\circ\eta(\omega) $, $\forall\omega\in \tilde{\Omega}$, then $C_2\ge 0$ $\Rightarrow$ $C_1\ge 0$, and $C_1\le 0$ $\Rightarrow$ $C_2\le 0$. If $a \leq f\circ\eta(\omega)+f\circ\xi(\omega),\; b \geq f\circ\xi(\omega)-f\circ\eta(\omega)$, $\forall\omega\in \tilde{\Omega}$, then $C_1\ge 0$ $\Rightarrow$ $C_2\ge 0$, and $C_2\le 0$ $\Rightarrow$ $C_1\le 0$. \end{remark} \begin{remark} To some extent, the two key controls for $f\circ\eta+f\circ\xi$ and $f\circ\xi-f\circ\eta$ via constants $a$ and $b$ are indeed `summation control' and `difference control'. The identities used in the proof of Theorem \ref{thm:operator-main} is inspired by product-to-sum formulas. \end{remark} \vspace{0.5cm} \section{Applications to variant Hlawka inequalities} \subsection{Applications to quadratic form}\label{sec:quadratic} Given a nondegenerate quadratic form $q$, i.e., $q(x)=x^\top Qx$, where $x^\top$ is the transpose of $x$ and $Q$ is a matrix of dimension $n$. Henceforth a pair $(k,n-k)$ is said to be the signature of $Q$, if $Q$ has $k$ positive eigenvalues and $(n-k)$ negative eigenvalues. Consider $l=\sqrt{q}$, then we have the following: \begin{pro}\label{quadratic} (P1) If $Q$ is of the signature $(1,n-1)$, then $l$ satisfies the reversed Hlawka inequality in the closure of the future cone; (P2) If $Q$ is of the signature $(n,0)$, then $l$ satisfies the Hlawka inequality in $\mathbb{R}^n$. \end{pro} \begin{proof} We will apply Theorem \ref{thm:operator-main} to this setting, where the symbols appearing in Theorem \ref{thm:operator-main} can be concretely chosen (see Table~\ref{tab:pro1}). Firstly, according to the definition of $q$, there is \begin{equation}\label{quadratic1} q(x+y+z)+q(x)+q(y)+q(z)=q(x+y)+q(x+z)+q(y+z). \end{equation} (P1) Since $Q$ is of the signature $(1,n-1)$, we may assume without loss of generality that $Q=\mathrm{diag}(1,-1,\cdots,-1)$ and let $X=\{x=(x_1,\cdots,x_n)|q(x)>0,x_1> 0\}$, i.e., the future cone in Minkowski space. \begin{table} \centering \caption{\small The concrete quantities of Theorem \ref{thm:operator-main} used in the proof of Proposition \ref{quadratic} (1). Besides, for Proposition \ref{quadratic} (2), only let $X$ change to $\mathbb{R}^n$.} \begin{tabular}{|l|l|} \hline Terminologies in Theorem \ref{thm:operator-main} & Concrete choices in Proposition \ref{quadratic} (1) for fixed $x,y,z\in X$ \\ \hline $\Omega=$ & $\{1,2,3\}$\\ \hline $X=$ & $\{x=(x_1,\cdots,x_n)|q(x)>0,x_1> 0\}$\\ \hline $\mathcal{S}=$ & $\mathbb{R}^{\{1,2,3\}}$\\ \hline $T=$&$\sum_{\omega\in\{1,2,3\}}$, i.e., $T(g)=g(1)+g(2)+g(3)$, $\forall g\in\mathcal{S}$\\ \hline $\eta=$&$x,y,z$ for $\omega=1,2,3$ respectively\\ \hline $\xi=$&$\sum_{\omega=1}^3\eta(\omega)-\eta$, i.e., $y+z, z+x, x+y$ for $\omega=1,2,3$ respectively\\ \hline $f=$&$\sqrt{q}$\\ \hline $a=$&$\sqrt{q(x)}+\sqrt{q(y)}+\sqrt{q(z)}$\\ \hline $b=$&$\sqrt{q(x+y+z)}$\\ \hline \end{tabular} \label{tab:pro1} \end{table} Indeed, there is no subtraction `$-$' in $X$ and it is closed under addition. Because $q(x)=x^2_1-x^2_2-\cdots-x^2_n>0, \ \ q(y)=y^2_1-y^2_2-\cdots-y^2_n>0$, i.e., $$ x^2_1>x^2_2+\cdots+x^2_n, \ \ y^2_1>y^2_2+\cdots+y^2_n, $$ by Cauchy inequality, the following inequality holds: $$ x^2_1y^2_1>(x^2_2+\cdots+x^2_n)(y^2_2+\cdots+y^2_n)\geq (x_2y_2+\cdots+x_ny_n)^2. $$ Due to $x,y\in X$, there is $x_1y_1> 0$, so $x_1y_1>x_2y_2+\cdots+x_ny_n$, i.e., $x^\top Qy=y^\top Qx=x_1y_1-x_2y_2-\cdots-x_ny_n>0$. Hence $q(x+y)=q(x)+q(y)+x^\top Qy+y^\top Qx> 0$ which implies $x+y\in X$. According to the Azteca inequality (i.e., a reversed version of Cauchy inequality), for any $x,y \in X$, there is $$(x_1y_1-\sum_{j\geq2}x_jy_j)^2\geq (x^2_1-\sum_{j\geq2}x^2_j)(y^2_1-\sum_{j\geq2}y^2_j),$$ i.e., \begin{equation}\label{quadratic2} (x^\top Qy)^2\geq x^\top Qx\cdot y^\top Qy. \end{equation} By further elementary computation, \eqref{quadratic2} is equivalent to \begin{equation}\label{quadratic3} \sqrt{q(x+y)}\geq\sqrt{q(x)}+\sqrt{q(y)} \end{equation} whenever $x,y\in X$. By \eqref{quadratic3}, for any $x,y,z\in X$, there is $$\sqrt{q(x)}+\sqrt{q(y+z)}\geq a,\ \ \sqrt{q(y+z)}-\sqrt{q(x)}\leq b.$$ By the parameters shown in Table~\ref{tab:pro1}, we further have $c=2$ in Theorem \ref{thm:operator-main}, $C_2=q(x+y+z)+q(x)+q(y)+q(z)-q(x+y)-q(x+z)-q(y+z)=0$ (by Eq.~\eqref{quadratic1}) and $$C_1=l(x)+l(y)+l(z)+l(x+y+z)- l(x+y)-l(y+z)-l(z+x).$$ According to Theorem \ref{thm:operator-main} (II), $C_1\le0$, thus $$ l(x)+l(y)+l(z)+l(x+y+z)\leq l(x+y)+l(y+z)+l(z+x), $$ whenever $x,y,z\in X$. By taking limit, one can find that the reversed Hlawka inequality also holds on the boundary of the future cone. \vspace{0.2cm} (P2) If $Q$ is $(n,0)$, we may assume without loss of generality that $Q=\mathrm{diag}(1,1,\cdots,1)$ and let $X=\mathbb{R}^n$. In this case, the inner product $\langle x, y\rangle:=x^\top Qy$ satisfies Cauchy inequality, i.e., $(x^\top Qy)^2\leq x^\top Qx\cdot y^\top Qy$. By elementary computation, there is $\sqrt{q(x+y)}\leq\sqrt{q(x)}+\sqrt{q(y)}$. From this, we have $$\sqrt{q(x)}+\sqrt{q(y+z)}\leq a,\ \ \sqrt{q(y+z)}-\sqrt{q(x)}\leq b.$$ In case $a+b> 0$, similar to (P1), according to Theorem \ref{thm:operator-main} (I) and Eq.~\eqref{quadratic1}, we have $l(x)+l(y)+l(z)+l(x+y+z)\geq l(x+y)+l(y+z)+l(z+x)$. In the case of $a=0$ or $a+b=0$, i.e., $x=y=z=0$, it is obvious that $l(x)+l(y)+l(z)+l(x+y+z)= l(x+y)+l(y+z)+l(z+x)=0$. Consequently, $l$ satisfies the Hlawka inequality. \end{proof} Proposition \ref{quadratic} contains Hlawka-type inequalities in the settings of both Euclidean case and Minkowski case. Moreover, by using Theorem \ref{thm:operator-main}, here we indeed provide an alternative but easier proof of the reverse Hlawka inequality in Minkowski space (Theorem 1.1 in \cite{Serre15}). However, there is no similar conclusion on other cases that $Q$ is of the signature $(k,n-k)$ for $2\le k\le n-1$, and we will give an example to show this. \begin{example} If $Q$ is of the signature $(k,n-k)$ for $2\le k\le n-1$, we may assume without loss of generality that $Q=\textrm{diag}(\mathop{\underbrace{1,\cdots,1}}\limits_k,\mathop{\underbrace{-1,\cdots,-1}}\limits_{n-k})$. By finding suitable cone $X\subset \{x=(x_1,\cdots,x_n)\in\mathbb{R}^n|\,q(x)>0\}$, one may obtain that both the Hlawka inequality and the reversed Hlawka inequality fail for $l$. Indeed, take $0<\epsilon \ll 1$, let $$ \vec v_1=\vec v_2=(\mathop{\underbrace{1,1,\epsilon,\cdots,\epsilon}}\limits_k, \mathop{\underbrace{1,\epsilon,\cdots,\epsilon}}\limits_{n-k}) ,\ \ \vec v_3=(1,1,\epsilon,\cdots,\epsilon) $$ and $$ \vec v_4=(2,1,\epsilon,\cdots,\epsilon),\; \vec v_5=(1,2,\epsilon,\cdots,\epsilon). $$ Consider $X=\{t_1\vec v_1+t_2\vec v_2+t_3\vec v_3+t_4\vec v_4+t_5\vec v_5|t_i>0, 1\leq i \leq 5\}$. It is clear that $X\subset\{x|\,q(x)>0\}$. By computation, $l(\vec v_1)+l(\vec v_2)+l(\vec v_3)+l(\vec v_1+\vec v_2+\vec v_3)< l(\vec v_1+\vec v_2)+l(\vec v_2+\vec v_3)+l(\vec v_3+\vec v_1)$. While, $l(\vec v_5)+l(\vec v_4)+l(\vec v_3)+l(\vec v_3+\vec v_4+\vec v_5)> l(\vec v_4+\vec v_5)+l(\vec v_3+\vec v_5)+l(\vec v_3+\vec v_4)$. So in $X$ both the Hlawka inequality and the reversed Hlawka inequality fail for $l$. \end{example} \hspace{0.5cm} \subsection{Applications to sub/super -additive functions on semigroup}\label{sec:semigroup} Let $X$ in Theorem \ref{thm:operator-main} be an abelian semigroup $(G,+)$, and let $F:G\rightarrow \mathbb{R}$ be a non-negative real-valued function. We will consider the Hlawka inequality in form \begin{equation}\label{eq:2^k} F(x+y)^{1/2^k}+F(y+z)^{1/2^k}+F(z+x)^{1/2^k}\le F(x)^{1/2^k}+F(y)^{1/2^k}+F(z)^{1/2^k}+F(x+y+z)^{1/2^k}, \end{equation} $\forall x,y,z\in G$, where $k$ is an integer. \begin{pro}\label{pro:2^k} Let $G$ be an abelian semigroup, and let $x\mapsto F(x)$ be a non-negative real-valued function on $G$. If $F$ is {\sl strong subadditive} (i.e., $F(x)+F(y)\ge F(x+y)$ and $F(x)+F(x+y)\ge F(y)$, $\forall x,y\in G$), and \eqref{eq:2^k} holds for some $k_0\ge -1$, then \eqref{eq:2^k} holds for all $k\ge k_0$. If $F$ is assumed to be superadditive (i.e., $F(x)+F(y)\le F(x+y)$, $\forall x,y\in G$), and \eqref{eq:2^k} holds for some $k_0\le 0$, then \eqref{eq:2^k} holds for all $k\le k_0$. \end{pro} \begin{proof} Given $a,b>0$, the function $(a^t+b^t)^{1/t}$ is decreasing on $(0,\infty)$. \begin{table} \centering \caption{\small The concrete quantities of Theorem \ref{thm:operator-main} used in the proof of Proposition \ref{pro:2^k}.} \begin{tabular}{|l|l|} \hline Terminologies in Theorem \ref{thm:operator-main} & Concrete choices in Proposition \ref{pro:2^k} for fixed $x,y,z\in X$ \\ \hline $\Omega=$ & $\{1,2,3\}$\\ \hline $X=$ & abelian semigroup $G$\\ \hline $\mathcal{S}=$ & $\mathbb{R}^{\{1,2,3\}}$\\ \hline $T=$& $\sum_{\omega\in\{1,2,3\}}$, i.e., $T(g)=g(1)+g(2)+g(3)$, $\forall g\in\mathcal{S}$\\ \hline $\eta=$& $x,y,z$ for $\omega=1,2,3$ respectively\\ \hline $\xi=$&$\sum_{\omega=1}^3\eta(\omega)-\eta$, i.e., $y+z, z+x, x+y$ for $\omega=1,2,3$ respectively\\ \hline $f=$& $F^{\frac 1{2^{k}}}$\\ \hline $a=$&$F(x)^{\frac 1{2^{k}}}+F(y)^{\frac 1{2^{k}}}+F(z)^{\frac 1{2^{k}}}$\\ \hline $b=$&$F(x+y+z)^{\frac 1{2^{k}}}$\\ \hline \end{tabular} \label{tab:pro2} \end{table} Case (1). $F$ is strong non-negative subadditive. For any $0<\alpha\le 1$, $$ F(x+y)^\alpha\leq (F(x)+F(y))^\alpha\leq F(x)^\alpha+F(y)^\alpha. $$ Suppose \eqref{eq:2^k} holds for some $k_0\ge -1$. Then for any $k> k_0$, and any $x,y,z\in G$, $$ F(x+y)^{\frac 1{2^{k}}}+F(z)^{\frac 1{2^{k}}}\leq F(x)^{\frac 1{2^{k}}}+F(y)^{\frac 1{2^{k}}}+F(z)^{\frac 1{2^{k}}} $$ and $$ F(x+y)^{\frac 1{2^{k}}}-F(z)^{\frac 1{2^{k}}}\leq F(x+y+z)^{\frac 1{2^{k}}}. $$ Here, let $a$ and $b$ in Theorem \ref{thm:operator-main} be $F(x)^{\frac 1{2^{k}}}+F(y)^{\frac 1{2^{k}}}+F(z)^{\frac 1{2^{k}}}$ and $F(x+y+z)^{\frac 1{2^{k}}}$, respectively. The detailed parameters are shown in Table~\ref{tab:pro2}. If $a\neq 0$ and $a+b>0$, then the proof is finished by Theorem \ref{thm:operator-main} (I). If $a=0$, then $F(x)=F(y)=F(z)=F(x+y)=F(x+z)=F(y+z)=F(x+y+z)=0$ and \eqref{eq:2^k} is obvious. \vspace{0.3cm} \noindent Case (2). $F$ is non-negative superadditive, i.e., $F(x)+F(y)\leq F(x+y)$. Note that for any $\alpha \geq 1$, $$ F(x+y)^\alpha\geq(F(x)+F(y))^\alpha\geq F(x)^\alpha+F(y)^\alpha. $$ In consequence, for any $k\leq0$ and any $x,y,z\in G$, $$ F(x+y)^{\frac 1{2^{k}}}+F(z)^{\frac 1{2^{k}}}\geq F(x)^{\frac 1{2^{k}}}+F(y)^{\frac 1{2^{k}}}+F(z)^{\frac 1{2^{k}}} $$ and $$ F(x+y+z)^{\frac 1{2^{k}}}\geq F(x+y)^{\frac 1{2^{k}}}+F(z)^{\frac 1{2^{k}}}\geq F(x+y)^{\frac 1{2^{k}}}-F(z)^{\frac 1{2^{k}}}. $$ If $a\neq 0$ and $a+b>0$, by Theorem \ref{thm:operator-main}, the result is proved. If $a+b=0$, then $F(x)=F(y)=F(z)=F(x+y)=F(x+z)=F(y+z)=F(x+y+z)=0$, the result is obvious. If $a=0$, then $F(x)=F(y)=F(z)=0$. According to the condition, we have $$ F(x+y)^{1/2^{k}}+F(y+z)^{1/2^{k}}+F(z+x)^{1/2^{k}}\le F(x+y+z)^{1/2^{k}}$$ for some $k\leq 0$. Take the square of above inequality, there is $$ F(x+y)^{1/2^{k-1}}+F(y+z)^{1/2^{k-1}}+F(z+x)^{1/2^{k-1}}\le F(x+y+z)^{1/2^{k-1}}. $$ Hereto, the prove is completed. \end{proof} Now we show an interesting example even though this result seems to be elementary. \begin{example} Taking $G=L^p$ and $F=\|\cdot\|_p$, together with Corollary 2.1 in \cite{W73} and Proposition \ref{pro:2^k}, we have $$\|a+b\|_p^{\frac 1{2^k}}+\|b+c\|_p^{\frac 1{2^k}}+\|c+a\|_p^{\frac 1{2^k}}\le \|a\|_p^{\frac 1{2^k}}+\|b\|_p^{\frac 1{2^k}}+\|c\|_p^{\frac 1{2^k}}+\|a+b+c\|_p^{\frac 1{2^k}}$$ for any $a,b,c\in L^p$, and $k\in \mathbb{N}$, where $1\le p\le 2$. Replacing $a,b,c$ respectively by $a^{2^k}$, $b^{2^k}$, $c^{2^k}$, one gets $$\|(a^{2^k}+b^{2^k})^{\frac 1{2^k}}\|_{2^kp}+\|(b^{2^k}+c^{2^k})^{\frac 1{2^k}}\|_{2^kp}+\|(c^{2^k}+a^{2^k})^{\frac 1{2^k}}\|_{2^kp} \le \|a\|_{2^kp}+\|b\|_{2^kp}+\|c\|_{2^kp}+\|(a^{2^k}+b^{2^k}+c^{2^k})^{\frac 1{2^k}}\|_{2^kp}.$$ For convenience, we define an operation $\Diamond_k$ by $a\Diamond_k b=(a^{2^k}+b^{2^k})^{\frac 1{2^k}}$ for $1\le k<+\infty$, $a\Diamond_0 b:= a+b$ and $a\Diamond_\infty b:=|a|\vee |b|:=\max\{|a|,|b|\}$. Then using this notation, we obtain $$\|a \Diamond_k b\|_{2^kp}+\|b\Diamond_k c\|_{2^kp}+\|c\Diamond_k a\|_{2^kp} \le \|a\|_{2^kp}+\|b\|_{2^kp}+\|c\|_{2^kp}+\|a\Diamond_k b\Diamond_k c\|_{2^kp}$$ for any $p\in[1,2]$ and any $k\in \mathbb{N}\cup\{+\infty\}$. Thus $$\|a\Diamond_k b\|_p+\|b\Diamond_k c\|_p+\|c\Diamond_k a\|_p \le \|a\|_p+\|b\|_p+\|c\|_p+\|a\Diamond_k b\Diamond_k c\|_p$$ holds for $p\in[2^k,2^{k+1}]$, and $$ \|\max\{|a|,|b|\}\|_\infty+\|\max\{|b|,|c|\}\|_\infty+\|\max\{|c|,|a|\}\|_\infty\le \|a\|_\infty+\|b\|_\infty+\|c\|_\infty+\|\max\{|a|,|b|,|c|\}\|_\infty $$ by taking $k\to+\infty$. \end{example} A direct application of Proposition \ref{pro:2^k} is the following Hlawka inequality on abelian group. \begin{example}[Theorem 2 in \cite{R15}]\label{cor:R15} Let $G$ be an abelian group, $x\mapsto |x|$ a non-negative symmetric and subadditive function on $G$ (i.e., $|-x|=|x|$ and $|x|+|y|\ge|x+y|$, $\forall x,y\in G$), and let $S:[0,\infty)\to [0,\infty)$ be concave. Then, if $\forall x,y,z\in G$, $$S^2(|x+y|)+S^2(|y+z|)+S^2(|z+x|)\le S^2(|x|)+S^2(|y|)+S^2(|z|)+S^2(|x+y+z|),$$ so does $S$. In fact, the function $F(\cdot):=S(|\cdot|)$ must be non-negative and strong subadditive. So, Proposition \ref{pro:2^k} is applicable here. \end{example} The following measure-type Hlawka inequality is non-trivial and it cannot be deduced from Theorem 2 in \cite{R15} (i.e., Example \ref{cor:R15} above), because a measure space equipped with any set operation is not a group. But it can be obtained straightforward by Proposition \ref{pro:2^k} since a measure space with any set operation becomes a semigroup. \begin{example} Let $G$ be a measure space and $F=\mu$ be the measure. For the case of $k=0$, note that $\mu(A)+\mu(B)+\mu(C)-\mu(A\cup B)-\mu(B\cup C)-\mu(C\cup A)+\mu(A\cup B\cup C)=\mu(A\cap B\cap C)\ge 0$ and for symmetric difference $\triangle$, $\mu(A)+\mu(B)+\mu(C)-\mu(A\triangle B)-\mu(B\triangle C)-\mu(C\triangle A)+\mu(A\triangle B\triangle C)=3\mu(A\cap B\cap C)\ge0$. According to Proposition \ref{pro:2^k}, we have for any $k\ge 0$, $$ \mu(A)^{1/2^k}+\mu(B)^{1/2^k}+\mu(C)^{1/2^k}+\mu(A\cup B\cup C)^{1/2^k}\ge \mu(A\cup B)^{1/2^k}+\mu(B\cup C)^{1/2^k}+\mu(C\cup A)^{1/2^k} $$ and $$ \mu(A)^{1/2^k}+\mu(B)^{1/2^k}+\mu(C)^{1/2^k}+\mu(A\triangle B\triangle C)^{1/2^k}\ge \mu(A\triangle B)^{1/2^k}+\mu(B\triangle C)^{1/2^k}+\mu(C\triangle A)^{1/2^k} $$ because $ \mu(A)+\mu(B)\ge \mu(A\cup B)\ge \mu(A\triangle B)$ and $\mu(A)+\mu(A\cup B)\ge \mu(A)+\mu(A\triangle B)\ge \mu(B)$. \end{example} \vspace{0.5cm} \subsection{Applications to integral form}\label{sec:integral} Next, we would pay our attention to the following setting. Let $\Omega$ be a nonempty set and let $G$ be an abelian group, and let $x\mapsto |x|$ be a non-negative symmetric and subadditive function on $G$ (i.e., $|-x|=|x|$ and $|x|+|y|\ge|x+y|$, $\forall x,y\in G$). The function spaces $G^\Omega$ and $\mathbb{R}^\Omega$ are also abelian groups under the natural operation `$+$'. Take abelian subgroups ${\mathcal F}\subset G^\Omega$ and linear subspace $\widehat{{\mathcal F}}\subset \mathbb{R}^\Omega$ equipped with $T:\widehat{{\mathcal F}}\to\mathbb{R}$ satisfying the \textbf{basic setting} in the beginning of Section \ref{sec:Hlawka-relation}. Moreover,\footnote{Here, for $f\in {\mathcal F}$, $|f|$ is a function mapping $\Omega$ to $[0,\infty)$.} $\forall f\in {\mathcal F}$, $|f|\in \widehat{{\mathcal F}}$, $1\in \widehat{{\mathcal F}}$. \begin{table} \centering \caption{\small The concrete quantities of Theorem \ref{thm:operator-main} used in the proof of Proposition \ref{thm:groupmain}. \begin{tabular}{|l|l|} \hline Terminologies in Theorem \ref{thm:operator-main} & Concrete choices in Proposition \ref{thm:groupmain} \\ \hline $X=$ & abelian group $G$\\ \hline $\mathcal{S}=$ & $\widehat{{\mathcal F}}$\\ \hline $\eta=$& $\hat{g}$ \\ \hline $\xi=$& $\mathcal{T}g-\hat{g}$\\ \hline $f=$& $S|\cdot|$\\ \hline $a=$& $A$\\ \hline $b=$& $T(S|\hat{g}|)$\\ \hline \end{tabular} \label{tab:pro3} \end{table} Applying Theorem \ref{thm:operator-main} to the above restricted situation, we have: \begin{pro}\label{thm:groupmain} Given $A\neq 0 \in \mathbb{R}$, an operator ${\mathcal T}: {\mathcal F}\to G$, two maps $g,\hat{g}\in {\mathcal F}$, and a concave function $S:[0,+\infty)\to [0,+\infty)$ such that $A>0$ or $S(|{\mathcal T} g|)>0$ and when $x$ satisfies $S|\hat{g}(x)|+S|{\mathcal T} g|\ne S|\hat{g}(x)-{\mathcal T} g|$, there is $A\ge S|\hat{g}(x)|+S|{\mathcal T} g-\hat{g}(x)|$, then \begin{equation}\label{eq:HlawkaS2} \left(T(1)-C\right)S^2|{\mathcal T} g|+T(S^2|\hat{g}|)\ge T(S^2|\hat{g}-{\mathcal T} g|) \end{equation} implies \begin{equation}\label{eq:Hlawka} \left(T(1)-C\right)S|{\mathcal T} g|+T(S|\hat{g}|)\ge T(S|\hat{g}-{\mathcal T} g|), \end{equation} where $C=2T(S|\hat{g}|)/A$. \end{pro} \begin{proof} Taking $\xi=\mathcal{T}g-\hat{g}$, $\eta=\hat{g}$ and $f(\cdot)=S|\cdot|$ in Theorem \ref{thm:operator-main} (see Table~\ref{tab:pro3} for details), we immediately complete the proof. \end{proof} The main theorems of \cite{TTW00,R15} can be seen as direct conclusions of Proposition \ref{thm:groupmain}. \begin{proof}[A proof of Example \ref{cor:R15} (i.e., main theorem in \cite{R15}) via Proposition \ref{thm:groupmain}] Take $\Omega=\{1,2,3\}$, and for given $x,y,z\in G$, let $\hat{g}=g$ be defined as $g(1)=x$, $g(2)=y$ and $g(3)=z$. Let $T(S|g|)=S|g(1)|+S|g(2)|+S|g(3)|$ and ${\mathcal T} g=g(1)+g(2)+g(3)$ in Proposition \ref{thm:groupmain}. Then $T(1)=3$, ${\mathcal T} g-g(i)=g(j)+g(k)$, where $\{i,j,k\}=\{1,2,3\}$. Hence, the result is easy to check. \end{proof} \vspace{0.3cm} Given an inner product space $(H,\langle\cdot,\cdot\rangle)$, suppose ${\mathcal F}\subset H^\Omega$ and $\widehat{{\mathcal F}}\subset \mathbb{R}^\Omega$ are linear spaces equipped with linear operators ${\mathcal T}: {\mathcal F}\to H$ and $T:\widehat{{\mathcal F}}\to\mathbb{R}$, then we have the following: \begin{cor}\label{pro:inner} Given ${\mathcal T},T$ and a map $f\in {\mathcal F}$ with $T(|f|)>0$ and for any $a\in H$, there is $T\langle f,a\rangle=\langle {\mathcal T} f,a\rangle$, where $|\cdot|$ is the norm induced by the inner product. If $T(|f|)\ge |f(x)|+|{\mathcal T} f-f(x)|$ whenever $x$ satisfies $-f(x)\ne \alpha {\mathcal T} f$ for any $\alpha\ge 0$, then the following hold: $$\left(T(1)-2\right)|{\mathcal T} f|+T(|f|)\ge T(|f-{\mathcal T} f|).$$ \end{cor} \begin{proof} By the basic equality for inner product, we have \begin{align*} T\left(|f-{\mathcal T} f|^2\right)&=T\left(|f|^2+|{\mathcal T} f|^2-2\langle f,{\mathcal T} f\rangle\right) \\ ~&=T(|f|^2)+|{\mathcal T} f|^2T(1)-2\langle {\mathcal T} f,{\mathcal T} f\rangle \\ ~&=T(|f|^2)+|{\mathcal T} f|^2(T(1)-2). \end{align*} Let $S$ in Proposition \ref{thm:groupmain} be the identity operator and the rest conditions in Proposition \ref{thm:groupmain} are easy to verify. The prove is completed. \end{proof} It is clear that $T$ and ${\mathcal T}$ are uniquely determined by each other according to Riesz representation theorem. \begin{cor}\label{pro:inner2integral} Let $(\Omega,\mu)$ be a finite measurable space and let $(H,\|\cdot\|)$ be an inner product space. Suppose $f,g:\Omega\to H$ are two nonzero integrable functions satisfying $$\frac{\int_\Omega fd\mu}{\int_\Omega \|f\|d\mu}=\frac{\int_\Omega gd\mu}{\int_\Omega \|g\|d\mu}.$$ Assume that for $x$ with $-g(x)\ne\alpha \int_\Omega fd\mu$ for any $\alpha\ge0$, there is $$\int_\Omega \|f\|d\mu\ge \|g(x)\|+\left\|g(x)-\int_\Omega fd\mu\right\|.$$ Then we have the following Hlawka's inequality $$\left(\mu(\Omega)-C\right)\left\| \int_\Omega fd\mu\right\|+\int_\Omega \|g\|d\mu\ge \int_\Omega\left\|g -\int_\Omega fd\mu\right\|d\mu,$$ where $C=2\int_\Omega \|g\|d\mu/\int_\Omega \|f\|d\mu$. \end{cor} \begin{proof} Take ${\mathcal T} f= \int_\Omega f(\omega)d\mu$ and $T (\|f\|) = \int_\Omega\|f(t)\|d\mu(t)$. Now it is ready to apply Proposition \ref{thm:groupmain} to finish the proof. \end{proof} \begin{cor}\label{cor:t=lambda} Suppose for $x$ with $- f(x)\ne\alpha \int_\Omega fd\mu$ for any $\alpha\ge0$, there is $$\int_\Omega \|f\|d\mu\ge t\| f(x)\|+\left\| t f(x)-\int_\Omega fd\mu\right\|$$ for some $t\ge 0$. Then we have the following Hlawka's inequality $$\left(\mu(\Omega)-2t\right)\left\| \int_\Omega fd\mu\right\|+t\int_\Omega \|f\|d\mu\ge \int_\Omega\left\| t f -\int_\Omega fd\mu\right\|d\mu.$$ \end{cor} \begin{cor}\label{bar} If $\bar f $ is a rearrangement of $f$ with the same distribution, and for $x$ with $-\bar f(x)\ne\alpha \int_\Omega fd\mu$ for any $\alpha\ge0$, there is $$\int_\Omega \|f\|d\mu\ge \|\bar f(x)\|+\left\|\bar f(x)-\int_\Omega fd\mu\right\|.$$ Then we have the following Hlawka's inequality $$\left(\mu(\Omega)-2\right)\left\| \int_\Omega fd\mu\right\|+\int_\Omega \|\bar f\|d\mu\ge \int_\Omega\left\| \bar f -\int_\Omega fd\mu\right\|d\mu.$$ \end{cor} \begin{proof} Clearly, the properties of rearrangement mean that $ \int_\Omega fd\mu=\int_\Omega \bar fd\mu$ and $\int_\Omega \| f\|d\mu=\int_\Omega \|\bar f\|d\mu$. Hence, Corollary \ref{pro:inner2integral} is applicable here. \end{proof} Theorem 1 in \cite{TTW00} could be viewed as a corollary of Corollary \ref{bar}. In fact, take $\bar f=f$ in Corollary \ref{bar}, it is easy to verify the following. \begin{example}[Theorem 1 in \cite{TTW00}]\label{cor:TTW00} Let $H$ be a Hilbert space, $(\Omega,\mu)$ a finite measure space and let $f$ be a Bochner integrable $H$-valued function on $(\Omega,\mu)$. Suppose that $$\int_\Omega\|f(t)\|d\mu(t)\ge \left\|f(\omega)-\int_\Omega f(t)d\mu(t)\right\|+\|f(\omega)\|\;\;(a.e., \omega\in\Omega_f),$$ where $\Omega_f=\{\omega\in\Omega:-f(\omega)\ne \alpha \int_\Omega f(t)d\mu(t)\text{ for any }\alpha\ge 0\}$. Then $$(\mu(\Omega)-2)\left\|\int_\Omega f(\omega)d\mu\right\|+\int_\Omega\|f(\omega)\|d\mu\ge \int_\Omega\left\|f(\omega)-\int_\Omega fd\mu\right\|d\mu.$$ \end{example} Next remark contains some interesting examples as corollaries of Proposition \ref{thm:groupmain}. \begin{remark}Given an inner product space $H$, we have: \begin{itemize} \item For any $\lambda\in[0,1]$, $x,y,z\in H$, $$(1-\lambda)(\|x\|+\|y\|+\|z\|)+(1+2\lambda)\|x+y+z\|\ge \|\lambda x+y+z\|+\|x+ \lambda y+z\|+\|x+\lambda y+z\|.$$ {\sl It is deduced by taking $\Omega=\{1,2,3\}$ and $t=(1-\lambda)$ in Corollary \ref{cor:t=lambda}, which is rather different from Corollary 2 in \cite{TTW00}.} \item Let $\mu_i,\lambda\ge 0$ such that $\sum_{i=1}^n\mu_i\|x_i\|\ge \lambda\mu_i\|x_i\|+\|\lambda x_i-\sum_{j=1}^n\mu_j x_j\|$ for any $1\leq i \leq n$. Then $$\left(\sum_{i=1}^n\mu_i-2\lambda\right)\left\|\sum_{i=1}^n\mu_ix_i\right\|+ \lambda\sum_{i=1}^n\mu_i\|x_i\|\ge \sum_{i=1}^n\mu_i\left\|\lambda x_i-\sum_{j=1}^n\mu_jx_j\right\|.$$ {\sl It is deduced by taking $\Omega=\{1,\cdots,n\}$ and $\mu(i)=\mu_i$ in Corollary \ref{cor:t=lambda}, which is an improved version of Corollary 2 in \cite{TTW00} and Proposition 11 in \cite{TTW09}.} \end{itemize} \end{remark} \vspace{1cm} {\bf Acknowledgements.} This research is supported by grant from the Project funded by China Postdoctoral Science Foundations (No. 2019M660829). The author thanks her husband for interesting discussions.
{ "timestamp": "2020-03-17T01:01:11", "yymm": "2003", "arxiv_id": "2003.06457", "language": "en", "url": "https://arxiv.org/abs/2003.06457", "abstract": "The classical Hlawka inequality possesses deep connections with zonotopes and zonoids in convex geometry, and has been related to Minkowski space. We introduce Hlawka Type-1 and Type-2 quantities, and establish a Hlawka-type relation between them, which connects a vast number of strikingly different variants of the Hlawka inequalities, such as Serre's reverse Hlawka inequality in the future cone of the Minkowski space, the Hlawka inequality for subadditive function on abelian group by Ressel, and the integral analogs by Takahasi et al. Besides, we announce several enhanced results, such as the Hlawka inequality for the power of measure function. Particularly, we give a complete study of the Hlawka inequality for quadratic form which relates to a work of Serre.", "subjects": "Functional Analysis (math.FA); Operator Algebras (math.OA)", "title": "Interactions between Hlawka Type-1 and Type-2 Quantities", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9859363717170516, "lm_q2_score": 0.8128673246376009, "lm_q1q2_score": 0.801435460740543 }
https://arxiv.org/abs/2206.14911
Minimum Weight Euclidean $(1+\varepsilon)$-Spanners
Given a set $S$ of $n$ points in the plane and a parameter $\varepsilon>0$, a Euclidean $(1+\varepsilon)$-spanner is a geometric graph $G=(S,E)$ that contains, for all $p,q\in S$, a $pq$-path of weight at most $(1+\varepsilon)\|pq\|$. We show that the minimum weight of a Euclidean $(1+\varepsilon)$-spanner for $n$ points in the unit square $[0,1]^2$ is $O(\varepsilon^{-3/2}\,\sqrt{n})$, and this bound is the best possible. The upper bound is based on a new spanner algorithm in the plane. It improves upon the baseline $O(\varepsilon^{-2}\sqrt{n})$, obtained by combining a tight bound for the weight of a Euclidean minimum spanning tree (MST) on $n$ points in $[0,1]^2$, and a tight bound for the lightness of Euclidean $(1+\varepsilon)$-spanners, which is the ratio of the spanner weight to the weight of the MST. Our result generalizes to Euclidean $d$-space for every constant dimension $d\in \mathbb{N}$: The minimum weight of a Euclidean $(1+\varepsilon)$-spanner for $n$ points in the unit cube $[0,1]^d$ is $O_d(\varepsilon^{(1-d^2)/d}n^{(d-1)/d})$, and this bound is the best possible.For the $n\times n$ section of the integer lattice in the plane, we show that the minimum weight of a Euclidean $(1+\varepsilon)$-spanner is between $\Omega(\varepsilon^{-3/4}\cdot n^2)$ and $O(\varepsilon^{-1}\log(\varepsilon^{-1})\cdot n^2)$. These bounds become $\Omega(\varepsilon^{-3/4}\cdot \sqrt{n})$ and $O(\varepsilon^{-1}\log(\varepsilon^{-1})\cdot \sqrt{n})$ when scaled to a grid of $n$ points in the unit square. In particular, this shows that the integer grid is \emph{not} an extremal configuration for minimum weight Euclidean $(1+\varepsilon)$-spanners.
\section{Introduction} \label{sec:intro} For a set $S$ of $n$ points in a metric space, a graph $G=(S,E)$ is a \emph{$t$-spanner} if $G$ contains, between any two points $p,q\in S$, a $pq$-path of weight at most $t\cdot \|pq\|$, where $t\geq 1$ is the \emph{stretch factor} of the spanner. In other words, a $t$-spanner approximates the true distances between the ${n\choose 2}$ pairs of points up to a factor $t$ distortion. Several optimization criteria have been developed for $t$-spanners for a given parameter $t\geq 1$. Natural parameters are the \emph{size} (number of edges), the \emph{weight} (sum of edge weights), the \emph{maximum degree}, and the \emph{hop-diameter}. Specifically, the \emph{sparsity} of a spanner is the ratio $|E|/|S|$ between the number of edges and vertices; and the \emph{lightness} is the ratio between the weight of a spanner and the weight of an MST on $S$. In the geometric setting, $S$ is a set of $n$ points in Euclidean $d$-space in constant dimension $d\in \N$. For every $\eps>0$, there exist $(1+\eps)$-spanners with $O_d(\eps^{1-d})$ sparsity and $O_d(\eps^{-d})$ lightness, and both bounds are the best possible~\cite{LeS19}. In particular, the $\Theta$-graphs, Yao-graphs~\cite{ruppert1991approximating}, gap-greedy and path-greedy spanners provide $(1+\eps)$-spanners of sparsity $O_d(\eps^{1-d})$. For lightness, Das et al.~\cite{DasHN93,DasNS95,NS-book} were the first to construct $(1+\eps)$-spanners of lightness $\eps^{-O(d)}$. Gottlieb~\cite{Gottlieb15} generalized this result to metric spaces with doubling dimension $d$; see also~\cite{BorradaileLW19,FiltserS20}. Recently, Le and Solomon~\cite{LeS19} showed that the greedy $(1+\eps)$-spanner in $\R^d$ has lightness $O(\eps^{-d})$; and so it simultaneously achieves the best possible bounds for both lightness and sparsity. The greedy $(1+\eps)$-spanner algorithm~\cite{althofer1993sparse} generalizes Kruskal's MST algorithm: It sorts the ${n\choose 2}$ edges of $K_n$ by nondecreasing weight, and incrementally constructs a spanner $H$: it adds an edge $uv$ if $H$ does not contain an $uv$-path of weight at most $(1+\eps)\|uv\|$. \subparagraph{Lightness versus Minimum Weight.} Lightness is a convenient optimization parameter, as it is invariant under scaling. It also provides an approximation ratio for the minimum weight $(1+\eps)$-spanner, as the weight of a Euclidean MST (for short, EMST) is a trivial lower bound on the spanner weight. However, minimizing the lightness is not equivalent to minimizing the spanner weight for a given input instance, as the EMST is highly sensitive to the distribution of the points in $S$. Given that worst-case tight bounds are now available for the lightness, it is time to revisit the problem of approximating the \emph{minimum weight} of a Euclidean $(1+\eps)$-spanner, without using the EMST as an intermediary. \subparagraph{Euclidean Minimum Spanning Trees.} For $n$ points in the unit cube $[0,1]^d$, the weight of the EMST is $O_d(n^{1-1/d})$, and this bound is also best possible~\cite{Fe55,SteeleS89}. In particular, a suitably scaled section of the integer lattice attains these bounds up to constant factors. Supowit et al.~\cite{SupowitRP83} proved similar bounds for the minimum weight of other popular graphs, such as spanning cycles and perfect matchings on $n$ points in the unit cube $[0,1]^d$. \subparagraph{Extremal Configurations for Euclidean $(1+\eps)$-Spanners.} The tight $O_d(\eps^{-d})$ bound on lightness~\cite{LeS19} implies that for every set of $n$ points in $[0,1]^d$, there is a Euclidean $(1+\eps)$-spanner of weight $O(\eps^{-d} n^{1-1/d})$. However, the combination of two tight bounds need not be tight; and it is unclear which $n$-point configurations require the heaviest $(1+\eps)$-spanners. We show that this bound can be improved to $O(\eps^{-3/2}\sqrt{n})$ in the plane. Furthermore, the extremal point configurations are not an integer grid, but an asymmetric grid. \subparagraph{Contributions.} We obtain a tight upper bound on the minimum weight of a Euclidean $(1+\eps)$-spanner for $n$ points in $[0,1]^d$. \begin{theorem}\label{thm:cube} For constant $d\geq 2$, every set of $n$ points in the unit cube $[0,1]^d$ admits a Euclidean $(1+\eps)$-spanner of weight $O_d(\eps^{(1-d^2)/d}n^{(d-1)/d})$, and this bound is the best possible. \end{theorem} The upper bound is established by a new spanner algorithm, \textsc{SparseYao}, that sparsifies the classical Yao-graph using novel geometric insight (Section~\ref{sec:alg}). The weight analysis is based on a charging scheme that charges the weight of the spanner to empty regions (Section~\ref{sec:square}). The lower bound construction is the scaled lattice with basis vectors of weight $\sqrt{\eps}$ and $\frac{1}{\sqrt{\eps}}$ (Section~\ref{sec:lower}); and not (scaled copies of) the integer lattice $\Z^d$. We analyze the minimum weight of Euclidean $(1+\eps)$-spanners for the integer grid in the plane. \begin{theorem}\label{thm:grid} For every $n\in \N$, the minimum weight of a $(1+\eps)$-spanner for the $n\times n$ section of the integer lattice is between $\Omega(\eps^{-3/4}n^2)$ and $O(\eps^{-1}\log(\eps^{-1})\cdot n^2)$. \end{theorem} When scaled to $n$ points in $[0,1]^2$, the upper bound confirms that a scaled section of the integer lattice does not maximize the weight of Euclidean $(1+\eps)$-spanners. \begin{corollary}\label{cor:grid} For every $n\in \N$, the minimum weight of a $(1+\eps)$-spanner for $n$ points in a scaled section of the integer grid in $[0,1]^2$ is between $\Omega(\eps^{-3/4}\sqrt{n})$ and $O(\eps^{-1}\log(\eps^{-1})\sqrt{n})$. \end{corollary} The lower bound is derived from two elementary criteria (the empty ellipse condition and the empty slab condition) for an edge to be present in every $(1+\eps)$-spanner (Section~\ref{sec:lower}). The upper bound is based on analyzing the \textsc{SparseYao} algorithm from Section~\ref{sec:alg}, combined with results from number theory on Farey sequences (Section~\ref{sec:grid}). Closing the gap between the lower and upper bounds in Theorem~\ref{thm:grid} remains an open problem. Higher dimensional generalizations are also left for future work. In particular, multidimensional variants of Farey sequences are currently not well understood. \subparagraph{Further Related Previous Work.} Many algorithms have been developed for constructing $(1+\eps)$-spanners for $n$ points in $\R^d$~\cite{ABUAFFASH2022101807,ChanHJ20,DasHN93,DasNS95,ElkinS15,GudmundssonLN02,LeSolomon-unified1,LevcopoulosNS02,RaoS98}, designed for one or more optimization criteria (lightness, sparsity, hop diameter, maximum degree, and running time). A comprehensive survey up to 2007 is in the book by Narasinham and Smid~\cite{NS-book}. We briefly review previous constructions pertaining to the \emph{minimum weight} for $n$ points the unit square (i.e., $d=2$). As noted above, the recent worst-case tight bound on the lightness~\cite{LeS19} implies that the greedy algorithm returns a $(1+\eps)$-spanner of weight $O(\eps^{-2}\|\MST\|)=O(\eps^{-2}\sqrt{n})$. A classical method for constructing a $(1+\eps)$-spanners uses \emph{well-separated pair decompositions} (\emph{WSPD}) with a hierarchical clustering (e.g., quadtrees); see~\cite[Chap.~3]{Sariel}. Due to a hierarchy of depth $O(\log n)$, this technique has been adapted broadly to dynamic, kinetic, and reliable spanners~\cite{BuchinHO20,ChanHJ20,gao2006deformable,Roditty12}. However, the weight of the resulting $(1+\eps)$-spanner for $n$ points in $[0,1]^2$ is $O(\eps^{-3}\sqrt{n}\cdot\log n)$~\cite{gao2006deformable}. The $O(\log n)$ factor is due to the depth of the hierarchy; and it cannot be removed for any spanner with hop-diameter $O(\log n)$~\cite{AgarwalWY05,DinitzES10,SolomonE14}. Yao-graphs and $\Theta$-graphs are geometric proximity graphs, defined as follows. For a constant $k\geq 3$, consider $k$ cones of aperture $2\pi/k$ around each point $p\in S$; in each cone, connect $p$ to the ``closest'' point $q\in S$. For Yao-graphs, $q$ minimizes the Euclidean distance $\|pq\|$, and for $\Theta$-graphs $q$ is the point that minimizes the length of the orthogonal projection of $pq$ to the angle bisector of the cone. It is known that both $\Theta$- and Yao-graphs are $(1+\eps)$-spanners for a parameter $k\in \Theta(\eps^{-1})$, and this bound is the best possible~\cite{NS-book}. However, if we place $\lfloor n/2\rfloor$ and $\lceil n/2\rceil$ equally spaced points on opposite sides of the unit space, then the weight of both graphs with parameter $k=\Theta(\eps^{-1})$ will be $\Theta(\eps^{-1}\, n)$. \subparagraph{Organization.} We start with lower bound constructions in the plane (Section~\ref{sec:lower}) as a warm-up exercise. The two elementary geometric criteria build intuition and highlight the significance of $\sqrt{\eps}$ as the ratio between the two axes of an ellipse of all paths of stretch at most $1+\eps$ between the foci. Section~\ref{sec:alg} presents Algorithm \textsc{SparseYao} and its stretch analysis in the plane. Its weight analysis for $n$ points in $[0,1]^2$ is in Section~\ref{sec:square}. The generalization of the algorithm and its analysis are fairly straightforward, and sketched in Section~\ref{sec:d-space}. We analyze the performance of Algorithm \textsc{SparseYao} for the $n\times n$ grid, after a brief review of Feray sequences, in Section~\ref{sec:grid}. We conclude with a selection of open problems in Section~\ref{sec:con}. \section{Lower Bounds in the Plane} \label{sec:lower} We present lower bounds for the minimum weight of a $(1+\eps)$-spanner for the $n\times n$ section of the integer lattice (Section~\ref{ssec:gridLB}); and for $n$ points in a unit square $[0,1]^2$ (Section~\ref{ssec:squareLB}). Let $S\subset \R^2$ be a finite point set. We observe two elementary conditions that guarantee that an edge $ab$ is present \emph{in every} $(1+\eps)$-spanner for $S$. Two points, $a,b\in S$, determine a (closed) line segment $ab=\conv \{a,b\}$; the relative interior of $ab$ is denoted by $\mathrm{int}(ab)=ab\setminus \{a,b\}$. let $\mathcal{E}_{ab}$ denote the ellipse with foci $a$ and $b$, and great axis of weight $(1+\eps)\|ab\|$, $\mathcal{L}_{ab}$ be the slab bounded by two lines parallel to $ab$ and tangent lines to $\mathcal{E}_{ab}$; see Fig.~\ref{fig:ellipse}. Note that the width of $\mathcal{L}_{ab}$ equals the minor axes of $\mathcal{E}_{ab}$, which is $((1+\eps)^2-1^2)^{1/2}\|ab\|=(2\eps+\eps^2)^{1/2}\|ab\|>\sqrt{2\eps}\|ab\|$. \begin{itemize}\itemsep 0pt \item \textbf{Empty ellipse condition}: $S\cap \mathcal{E}_{ab}=\{a,b\}$. \item \textbf{Empty slab condition}: $S\cap \mathrm{int}(ab)=\emptyset$ and all points in $S\cap \mathcal{L}_{ab}$ are on the line $ab$. \end{itemize} \begin{figure}[htbp] \begin{center} \includegraphics[width=0.8\textwidth]{ellipse} \end{center} \caption{The ellipse $\mathcal{E}_{ab}$ with foci $a$ and $b$, and great axis of weight $(1+\eps)\|ab\|$.} \label{fig:ellipse} \end{figure} \begin{observation}\label{obs:elementary} Let $S\subset \R^2$, $G=(S,E)$ a $(1+\eps)$-spanner for $S$, and $a,b\in S$. \begin{enumerate} \item If $ab$ meets the empty ellipse condition, then $ab\in E$. \item If $S$ is a section of $\Z^2$, $\eps<1$, and $ab$ meets the empty slab condition, then $ab\in E$. \end{enumerate} \end{observation} \begin{proof} The ellipse $\mathcal{E}_{ab}$ contains all points $p\in \R^2$ satisfying $\|ap\|+\|pb\|\leq (1+\eps)\|ab\|$. Thus, by the triangle inequality, $\mathcal{E}_{ab}$ contains every $ab$-path of weight at most $(1+\eps)\|ab\|$. The empty ellipse condition implies that such a path cannot have interior vertices. If $S$ is the integer lattice, then $S\cap \mathrm{int}(ab)=\emptyset$ implies that $\overrightarrow{ab}$ is a primitive vector (i.e., the $x$- and $y$-coordinates of $\overrightarrow{ab}$ are relatively prime), hence the distance between any two lattice points along the line $ab$ is at least $\|ab\|$. Given that $\mathcal{E}_{ab}\subset \mathcal{L}_{ab}$, the empty slab condition now implies the empty ellipse condition. \end{proof} \subsection{Lower Bounds for the Grid} \label{ssec:gridLB} \begin{lemma}\label{lem:LBgrid} For every $n\in \N$ with $n\geq 2\,\eps^{-1/4}$, the weight of every $(1+\eps)$-spanner for the $n\times n$ section of the integer lattice is $\Omega(\eps^{-3/4}n^2)$. \end{lemma} \begin{proof} Let $S=\{(s_1,s_2)\in \Z^2: 0\leq s_1,s_2<n\}$ and $A=\{(a_1,a_2)\in \Z^2: 0\leq a_1,a_2<\lceil \eps^{-1/4}\rceil/2\}$. Denote the origin by $o=(0,0)$. For every grid point $a\in A$, we have $\|oa\|\leq \eps^{-1/4}/\sqrt{2}$. A vector $\overrightarrow{oa}$ is \emph{primitive} if $a=(a_1,a_2)$ and $\mathrm{gcd}(a_1,a_2)=1$. We show that every primitive vector $\overrightarrow{oa}$ with $a\in A$ satisfies the empty slab condition. It is clear that $S\cap \mathrm{int}(oa)=\emptyset$. Suppose that $s\in S$ but it is not on the line spanned by $oa$. By Pick's theorem, $\area(\Delta(oas))\geq \frac12$. Consequently, the distance between $s$ and the line $oa$ is at least $\|oa\|^{-1}\geq \sqrt{2}\cdot \eps^{1/4} \geq 2\, \eps^{1/2}\, \|oa\|$; and so $s\notin \mathcal{L}_{oa}$, as claimed. By elementary number theory, $\overrightarrow{oa}$ is primitive for $\Theta(|A|)$ points $a\in A$. Indeed, every $a_1\in \N$ is relatively prime to $N\varphi(a_1)/a_1$ integers in every interval of length $N$, where $\varphi(.)$ is Euler totient function, and $\varphi(a_1)=\Theta(a_1)$. Consequently, the total weight of primitive vectors $\overrightarrow{oa}$, $a\in A$, is $\Theta(|A|\cdot \eps^{-1/4})=\Theta(\eps^{-3/4})$. The primitive edges $oa$, $a\in A$, form a star centered at the origin. The translates of this star to other points $s\in S$, with $0\leq s_1,s_2\leq \frac{n}{2}\leq n - \lceil \eps^{-1/4}\rceil$ are also present in every $(1+\eps)$-spanner for $S$. As every edge is part of at most two such stars, summation over $\Theta(n^2)$ stars yields a lower bound of $\Omega(\eps^{-3/4}n^2)$. \end{proof} \begin{remark} The lower bound in Lemma~\ref{lem:LBgrid} derives from the total weight of primitive vectors $\overrightarrow{oa}$ with $\|oa\|\leq O(\eps^{-1/4})$, which satisfy the empty slab condition. There are additional primitive vectors that satisfy the empty ellipse condition (e.g., $\overrightarrow{oa}$ with $a=(1,a_2)$ for all $|a_2|<\eps^{-1/3}$). However, it is unclear how to account for all vectors satisfying the empty ellipse condition, and whether their overall weight would improve the lower bound in Lemma~\ref{lem:LBgrid}. \end{remark} \begin{remark} The empty ellipse and empty slab conditions each imply that an edge \emph{must} be present in every $(1+\eps)$-spanner for $S$. It is unclear how the total weight of such ``must have'' edges compare to the the minimum weight of a $(1+\eps)$-spanner. \end{remark} \subsection{Lower Bounds in the Unit Square} \label{ssec:squareLB} \begin{lemma}\label{lem:squareLB} For every $n\in \N$ and $\eps\in (0,1]$, there exists a set $S$ of $n$ points in $[0,1]$ such that every $(1+\eps)$-spanner for $S$ has weight $\Omega(\eps^{-3/2}\sqrt{n})$. \end{lemma} \begin{proof} First let $S_0$ be a set of $2m$ points, where $m=\lfloor \eps^{-1}/2\rfloor$, with $m$ equally spaced points on two opposite sides of a unit square. By the empty ellipse property, every $(1+\eps)$-spanner for $S_0$ contains a complete bipartite graph $K_{m,m}$. The weight of each edge of $K_{m,m}$ is between $1$ and $\sqrt{2}$, and so the weight of every $(1+\eps)$-spanner for $S_0$ is $\Omega(\eps^{-2})$. For $n\geq \eps^{-1}$, consider an $\lfloor \sqrt{\eps n}\rfloor\times \lfloor \sqrt{\eps n}\rfloor$ grid of unit squares, and insert a translated copy of $S_0$ in each unit square. Let $S$ be the union of these $\Theta(\eps n)$ copies of $S_0$; and note that $|S|=\Theta(n)$. A $(1+\eps)$-spanner for each copy of $S_0$ still requires a complete bipartite graph of weight $\Omega(\eps^{-2})$. Overall, the weight of every $(1+\eps)$-spanner for $S$ is $\Omega(\eps^{-1}n)$. Finally, scale $S$ down by a factor of $\lfloor \sqrt{\eps n}\rfloor$ so that it fits in a unit square. The weight of every edge scales by the same factor, and the weight of a $(1+\eps)$-spanner for the resulting $n$ points in $[0,1]^2$ is $\Omega(\eps^{-3/2}\, \sqrt{n})$, as claimed. \end{proof} \begin{remark} The points in the lower bound construction above lie on $O(\sqrt{\eps n})$ axis-parallel lines in $[0,1]^2$, and so the weight of their MST is $O(\sqrt{\eps n})$. Recall that the lightness of the greedy $(1+\eps)$-spanner is $O(\eps^{-d}\log \eps^{-1})$~\cite{LeS19}. For $d=2$, it yields a $(1+\eps)$-spanner of weight $O(\eps^{-2}\log \eps^{-1})\cdot \|\MST(S)\|=O(\eps^{-3/2}\log (\eps^{-1}) \sqrt{n} )$. \end{remark} \section{Spanner Algorithm: Sparse Yao-Graphs} \label{sec:alg} Let $S$ be a set of $n$ points in the plane and $\eps\in (0,\frac19)$. As noted above, the Yao-graph $Y_k(S)$ with $k=\Theta(\eps^{-1})$ cones per vertex is a $(1+\eps)$-spanner for $S$. We describe an new algorithm, \textsc{SparseYao}$(S,\eps)$, that computes a subgraph of a Yao-graph $Y_k(S)$ (Section~\ref{ssec:alg}); and show that it returns a $(1+\eps)$-spanner for $S$ (Section~\ref{ssec:stretch}). Later, we use this algorithm for $n$ points in the unit square (Section~\ref{sec:square}; and for an $n\times n$ section of the integer lattice (Section~\ref{sec:grid}). Our algorithm starts with a Yao-graph that is a $(1+\frac{\eps}{2})$-spanner, in order to leave room for minor loss in the stretch factor due to sparsification. The basic idea is that instead of cones of aperture $2\pi/k=\Theta(\eps)$, cones of much larger aperture $\Theta(\sqrt{\eps})$ suffice in some cases. (This is idea is flashed out in Section~\ref{ssec:stretch}). The angle $\sqrt{\eps}$ then allows us to charge the weight of the resulting spanner to the area of empty regions (specifically, to an empty section of a cone) in Section~\ref{sec:square}. \subsection{Sparse Yao-Graph Algorithm} \label{ssec:alg} We present an algorithm that computes a subgraph of a Yao-graph for $S$. It starts with cones of aperture $\Theta(\sqrt{\eps})$, and refines them to cones of aperture $\Theta(\eps^{-1})$. We connect each point $p\in S$ to the closest points in the larger cones, and use the smaller cones only when ``necessary.'' To specify when exactly the smaller cones are used, we define two geometric regions that will also play crucial roles in the stretch and weight analyses. \begin{figure}[htbp] \begin{center} \includegraphics[width=0.7\textwidth]{wedge2} \end{center} \caption{Wedges $W_1$ and $W_2$, line segment $A(p,q)$, and regions $\widehat{A}(p,q)$, $B(p,q)$, and $\widehat{B}(p,q)$ for $p,q\in S$.} \label{fig:wedge} \end{figure} \subparagraph{Definitions.} Let $p, q\in S$ be distinct points; refer to Fig.~\ref{fig:wedge}. Let $A(p,q)$ be the line segment of weight $\frac{\sqrt{\eps}}{2}\,\|pq\|$ on the line $pq$ with one endpoint at $p$ but interior-disjoint from the ray $\overrightarrow{pq}$; and $\widehat{A}(p,q)$ the set of points in $\R^2$ within distance $\frac{\eps}{16}\,\|pq\|$ from $A(p,q)$. Let $W_1$ be the cone with apex $p$, aperture $\frac12\cdot \sqrt{\eps}$, and symmetry axis $\overrightarrow{pq}$; and let $W_2$ be the cone with apex $q$, aperture $\sqrt{\eps}$, and the same symmetry axis $\overrightarrow{pq}$. Let $B(p,q)=W_1\cap W_2$. Finally, let $\widehat{B}(p,q)$ be the set of points in $\R^2$ within distance at most $\frac{\eps}{8}\,\|pq\|$ from $B(p,q)$. We show below (cf.~Lemma~\ref{lem:technical+}) that if we add edge $pq$ to the spanner, then we do not need any of the edges $ab$ with $a\in \widehat{A}(p,q)$ and $b\in \widehat{B}(p,q)$. We can now present our algorithm. \subparagraph{Algorithm \textsc{SparseYao}$(S,\eps)$.} Input: a set $S\subset \R^2$ of $n$ points, and $\eps\in (0,\frac19)$. \noindent\textbf{Preprocessing Phase: Yao-graphs.} Subdivide $\R^2$ into $k:=\lceil 16\,\pi/\sqrt{\eps}\rceil$ congruent cones of aperture $2\pi/k\leq \frac18\cdot \sqrt{\eps}$ with apex at the origin, denoted $C_1,\ldots ,C_k$. For $i\in \{1,\ldots ,k\}$, let $\overrightarrow{r}_i$ be the symmetry axis of $C_i$, directed from the origin towards the interior of $C_i$. For each $i\in \{1,\ldots , k\}$, subdivide $C_i$ into $k$ congruent cones of aperture $2\pi/k^2\leq \eps/8$, denoted $C_{i,1},\ldots , C_{i,k}$; see Fig.~\ref{fig:Yao}. For each point $s\in S$, let $C_i(s)$ and $C_{i,j}(s)$, resp., be the translates of cones $C_i$ and $C_{i,j}$ to apex $s$. For all $s\in S$ and $i\in \{1,\ldots ,k\}$, let $q_i(s)$ be a closest point to $s$ in $C_i(s)\cap (S\setminus \{s\})$; and for all $j\in \{1,\ldots , k\}$, let $q_{i,j}(s)$ be a closest point in $C_{i,j}(s)\cap (S\setminus \{s\})$; if such points exist. For each $i\in\{1,\ldots , k\}$, let $L_i$ be the list of all ordered pairs $(s,q_i(s))$ sorted in decreasing order of the orthogonal projection of $s$ to the directed line $\overrightarrow{r}_i$; ties are broken arbitrarily. \noindent\textbf{Main Phase: Computing a Spanner.} Initialize an empty graph $G=(S,E)$ with $E:=\emptyset$. \begin{enumerate} \item For all $i\in \{1,\ldots , k\}$, do: \begin{itemize} \item While the list $L_i$ is nonempty, do: \begin{enumerate} \item Let $(p,q)$ be the first ordered pair in $L_i$. \item Add (the unordered edge) $pq$ to $E$. \item For all $i'\in \{i-2,\ldots , i+2\}$ and $j\in \{1,\ldots ,k\}$, do:\\ If $\|pq_i(p)\|\leq \|pq_{i',j}(p)\|$ and $q_{i',j}(p)\notin B(p,q)$, then add $p q_{i',j}(p)$ to $E$. \item For all $s\in \widehat{A}(p,q)$, including $s=p$, delete the pair $(s,q_i(s))$ from $L_i$. \end{enumerate} \end{itemize} \item Return $G=(S,E)$. \end{enumerate} \begin{figure}[htbp] \begin{center} \includegraphics[width=0.8\textwidth]{Yao} \end{center} \caption{Cones $C_i(s)$ and $C_{i,j}(s)$ for a point $s\in S$, with $k=6$.} \label{fig:Yao} \end{figure} It is clear that the runtime of Algorithm \textsc{SparseYao} is polynomial in $n$ in the RAM model of computation. In particular, the runtime is dominated preprocessing phase that constructs the Yao-graph with $O(\eps^{-1}n)$ edges: finding the closest points $q_i(s)$ and $q_{i,j}(s)$ is supported by standard range searching data structures~\cite{Aga17a}. The main phase of the algorithm computes a subgraph of $Y_{k^2}(S)$ in $O(\eps^{-1}n)$ time. Optimizing the runtime, however, is beyond the scope of this paper. \subsection{Stretch Analysis} \label{ssec:stretch} In this section, we show that $G=\textsc{SparseYao}(S,\eps)$ is a $(1+\eps)$-spanner for $S$. In the preprocessing phase, Algorithm \textsc{SparseYao} computes a Yao-graph with $k^2=\Theta(\eps^{-1})$ cones. The following four lemmas justifies that we can omit some of the edges $sq_{i,j}$ from $G$. The first two technical lemmas show that \emph{if} $G$ already contains $(1+\eps)$-paths from $a$ to $p$ and from $q$ to $b$, then we can concatenate them with the edge $pq$ to obtain a $(1+\eps)$-path from $a$ to $b$. In Lemma~\ref{lem:technical}, we assume that $a\in A(p,q)$ and $b\in B(p,q)$; and Lemma~\ref{lem:technical+} handles the general case where $a\in \widehat{A}(p,q)$ and $b\in \widehat{B}(p,q)$. In inequality~\eqref{eq:technical} below, we use $\left(1+\frac{\eps}{4}\right)\|pq\|$ instead of $\|pq\|$ to absorb further error terms in the general case. \begin{lemma}\label{lem:technical} For all $a\in A(p,q)$ and $b\in B(p,q)$, we have \begin{equation}\label{eq:technical} (1+\eps)\|ap\|+\left(1+\frac{\eps}{3}\right)\|pq\|+(1+\eps)\|qb\|\leq (1+\eps)\|ab\|. \end{equation} \end{lemma} \begin{proof} We start with three simplifying assumptions. \noindent (i) We may assume that $p$ is the origin, $q$ is on the positive $x$-axis, and $b$ is on or above the $x$-axis, by applying a suitable congruence, if necessary. In particular, this implies that $A(p,q)$ is a line segment on the nonpositive $x$-axis; see Fig.~\ref{fig:wedge}. \noindent (ii) We may assume w.l.o.g.\ that $b$ is in the boundary $\partial B(p,q)$ of $B(p,q)$, since if we rotate the segment $qb$ around $q$, then the left hand side of \eqref{eq:technical} does not change, but the right hand side is minimized for $b\in\partial B(p,q)$. \noindent (iii) Furthermore, we may assume w.l.o.g.\ that $a=p$ if we establish the following a slightly stronger inequality for $a=p$ and $b\in B(p,q)$: \begin{equation}\label{eq:technical-} \left(1+\frac{\eps}{2}\right)\|pq\|+(1+\eps)\|qb\|\leq (1+\eps)\|pb\|. \end{equation} Indeed, if $a\neq p$, we easily show that \eqref{eq:technical-} implies Lemma~\ref{lem:technical}. Note that $a\in A(p,q)$ implies $B(p,q)\subseteq B(a,q)$ since $a\in A(p,q)$ lies on the symmetry axis of $B(p,q)$, as well as the wedges $W_1$ and $W_2$. Then \eqref{eq:technical-} becomes \begin{align}\label{eq:reduction} \left(1+\frac{\eps}{2}\right)\|aq\|+(1+\eps)\|qb\| &\leq (1+\eps)\|ab\| \nonumber\\ \left(1+\frac{\eps}{2}\right)\left(\|ap\| + \|pq\|\right)+(1+\eps)\|qb\|&\leq (1+\eps)\|ab\| \nonumber\\ \left(\|ap\|+\frac{\eps}{2}\, \|ap\| + \frac{\eps}{6}\, \|pq\|\right) + \left(1+\frac{\eps}{3}\right)\|pq\|+(1+\eps)\|qb\|&\leq (1+\eps)\|ab\|\nonumber\\ (1+\eps)\|ap\|+\left(1+\frac{\eps}{3}\right)\|pq\|+(1+\eps)\|qb\|&\leq (1+\eps)\|ab\|, \end{align} as $\|ap\|\leq \|A(pq)\| = \sqrt{\eps} \|pq\| \leq \frac13\,\|pq\|$ for $\eps<\frac19$. Let us review the Taylor estimates of some of the trigonometric functions. For the secant, we use the upper bound $\sec x = \frac{1}{\cos x}= 1+\frac{x^2}{2}+\frac{5x^4}{24}+\ldots < 1+x^2$ for $0<x<\frac12$. For the tangent, we use both upper and lower bounds $x\leq \tan x\leq x+\frac{x^3}{3}+\frac{2x^5}{15}+\ldots < x+\frac{x^3}{2}$. To prove~\eqref{eq:technical-}, we distinguish between two cases based on whether $b\in \partial W_1$ or $b\in \partial W_2$. Let $c$ be the intersection point of $\partial W_1$ and $\partial W_2$ above the $x$-axis. Since $\angle pcq = \angle qpc = \sqrt{\eps}/4$, then $\Delta{pqc}$ is an isosceles triangle with $\|pq\|=\|qc\|$. \subparagraph{Case~1: $b\in qc$ (Fig.~\ref{fig:cases}(left)).} Note that $0\leq \angle qpb\leq \sqrt{\eps}/4$. Assume $\angle qpb = t\cdot \sqrt{\eps}/4$ for some $t\in [0,1]$. Since the interior angles of triangle $\Delta{pqb}$ add up to $\pi$, then $\angle qbp = (2-t)\sqrt{\eps}/4$. Let $q^\perp$ be the orthogonal projection of $q$ to $pb$. Then $\|pb\|=\|pq^\perp\|+\|q^\perp b\|$. Since $\angle qpb \leq \angle qbp$ implies $\angle qpq^\perp \leq \angle qb q^\perp$, then we have $\|q^\perp b\|\leq \|p q^\perp\|$, and so $\|q^\perp b\|\leq \frac12\, \|pb\|$. We are now ready to prove \eqref{eq:technical-} in Case~1: \begin{align*} \left(1+\frac{\eps}{2}\right)\|pq\|+(1+\eps)\|qb\| &= \left(1+\frac{\eps}{2}\right) \|pq^\perp \| \sec \angle qpb +(1+\eps) \|q^\perp b\| \sec \angle qbp \\ &= \left(1+\frac{\eps}{2}\right) \|p q^\perp\| \sec \frac{t\,\sqrt{\eps}}{4} +(1+\eps) \|q^\perp b\| \sec \frac{(2-t)\sqrt{\eps}}{4}\\ &< \left(1+\frac{\eps}{2}\right) \left(1+\frac{t^2\eps}{16}\right)\|pq^\perp\| +(1+\eps)\left(1+ \frac{(2-t)^2 \eps}{16}\right)\|q^\perp b\| \\ &< \left(1+\frac{(t^2+9)\eps}{16}\right)\|pq^\perp\| +\left(1+\frac{(t^2-4t+21)\eps}{16}\right)\|q^\perp b\| \\ &= \left(1+\frac{(t^2+9)\eps}{16}\right) \left(\|p q^\perp\|+\|q^\perp b\|\right) + \frac{(-4t+12)\eps}{16}\, \|q^\perp b\| \\ &\leq \left(1+\frac{(t^2+9)\eps}{16}\right) \|p b\| + \frac{(-2t+6)\eps}{8}\cdot \frac{\|pb\|}{2}\\ &\leq \left(1+ \eps\cdot \frac{t(t-2)+15}{16}\right) \|pb\| \\ &< \left(1+ \eps \right) \|pb\|, \end{align*} as required. \begin{figure}[htbp] \begin{center} \includegraphics[width=0.9\textwidth]{cases} \end{center} \caption{Left: Case~1 where $b\in qc$. Right: Case~2, where $b$ lies to the right of $c$ on the boundary of $B(p,q)$.} \label{fig:cases} \end{figure} \subparagraph{Case~2: $b\in \partial B(p,q)\cap \partial W_1$ (Fig.~\ref{fig:cases}(right)).} In this case, $\angle qpb=\sqrt{\eps}/4$ is fixed. Note that $0\leq \angle pbq \leq \angle pcq\leq \sqrt{\eps}/4$. Assume $\angle pbq = t\cdot \sqrt{\eps}/4$ for some $t\in [0,1]$. Since $\angle qpb \geq \angle qbp$ implies $\angle qp q^\perp \geq qb q^\perp$, then we get $\|p q^\perp\|\leq \|q^\perp b\|$ hence $\|p q^\perp\|\leq \frac12\, \|pb\|$. Furthermore, the right triangles $\Delta{pq q^\perp}$ and $\Delta{bq q^\perp}$ yield \[ \|q q^\perp\|=\|p q^\perp\| \tan \angle qpb = \|q^\perp b\| \tan \angle qbp. \] This further implies \[ \|q^\perp b\| =\|p q^\perp\| \frac{\tan \angle qpb}{\tan \angle qbp} = \|p q^\perp\| \frac{\tan \left(\frac{\sqrt{\eps}}{4}\right) }{\tan\left(t\cdot \frac{\sqrt{\eps}}{4}\right)} \leq \|pq^\perp \| \frac{\frac{\sqrt{\eps}}{4}+\frac12 \left(\frac{\sqrt{\eps}}{4}\right)^3}{t\cdot \frac{\sqrt{\eps}}{4}} \leq \|p q^\perp \| \frac{1+\eps/32}{t}. \] We are now ready to prove \eqref{eq:technical-} in Case~2: \begin{align*} \left(1+\frac{\eps}{2}\right)\|pq\|+(1+\eps)\|qb\| &= \left(1+\frac{\eps}{2}\right) \|pq^\perp\| \sec \angle qpb +(1+\eps) \|q^\perp b\|\sec \angle qbp \\ &= \left(1+\frac{\eps}{2}\right) \|p q^\perp \|\sec \frac{\sqrt{\eps}}{4} +(1+\eps)\|q^\perp b\| \sec \frac{t\,\sqrt{\eps}}{4}\\ &< \left(1+\frac{\eps}{2}\right)\left(1+\frac{\eps}{16}\right) \|p q^\perp \| +(1+\eps)\left(1+\frac{t^2\eps}{16}\right)\|q^\perp b\| \\ &< \left(1+\frac{10\eps}{16}\right) \|p q^\perp \| +\left(1+\eps+\frac{t^2\eps(1+\eps)}{16}\right)\|q^\perp b\| \\ &=(1+\eps) \left(\|p q^\perp\|+\|q^\perp b\|\right) + \frac{\eps}{16}\left(t^2(1+\eps)\|q^\perp b\|-6 \|p q^\perp \| \right)\\ &\leq (1+\eps) \|pb\| + \frac{\eps}{16}\left(t(1+\eps)\left(1+\frac{\eps}{32}\right) -6 \right) \|p q^\perp \|\\ &< (1+\eps) \|pb\|, \end{align*} as required, since $0<t\leq 1$ and $0<\eps<1/9$. We have confirmed \eqref{eq:technical-} in both cases. This completes the proof of Lemma~\ref{lem:technical}. \end{proof} In the general case, we have $a\in \widehat{A}(p,q)$ and $b\in \widehat{B}(p,q)$. However, for technical reasons, we use a slightly larger neighborhood instead of $\widehat{A}(p,q)$. Let $\tilde{A}(p,q)$ be the set of points in $\R^2$ within distance at most $\frac{\eps}{5}$ from $A(p,q)$. \begin{lemma}\label{lem:technical+} For all $a\in \tilde{A}(p,q)$ and $b\in \widehat{B}(p,q)$, we have \begin{equation}\label{eq:technical+} (1+\eps)\|ap\|+\|pq\|+(1+\eps)\|qb\|\leq (1+\eps)\|ab\|. \end{equation} \end{lemma} \begin{proof} Since $a\in \tilde{A}(p,q)$ and $b\in \widehat{B}(p,q)$, then there exist $a'\in A(p,q)$ and $b'\in B(p,q)$ with $\|aa'\|\leq \frac{\eps}{5}\,\|pq\|$ and $\|bb'\|\leq \frac{\eps}{8}\,\|pq\|$. By the triangle inequality, we have $\|ap\|\leq \|aa'\|+\|a'p\|\leq \|a'p\|+\frac{\eps}{5}\,\|pq\|$ and $\|qb\|\leq \|qb'\|+\|b'b\|\leq \|qb'\|+\frac{\eps}{8}\,\|pq\|$. Combining these inequalities with Lemma~\ref{lem:technical} for points $a'$ and $b'$, we obtain \begin{align*} (1+\eps)\|ap\|+\|pq\|+(1+\eps)\|qb\| &\leq \Big((1+\eps)\|a'p\|+\|pq\|+(1+\eps)\|qb'\|\Big) + (1+\eps)\Big(\|aa'\|+\|bb'\|\Big)\\ &\leq \left( (1+\eps)\|a'b'\| - \frac{\eps}{4}\, \|pq\|\right) + (1+\eps)\frac{13\eps}{40}\, \|pq\| \\ &\leq (1+\eps)\Big( \|a'a\|+\|ab\| +\|bb'\|\Big) + \left(\frac{3(1+\eps)}{16}-\frac14\right)\eps\,\|pq\| \\ &\leq (1+\eps)\|ab\| + \left((1+\eps)\frac{13}{40}-\frac13\right)\eps\,\|pq\| \\ &< (1+\eps)\|ab\|, \end{align*} for $0<\eps<1/9$, as claimed. \end{proof} \subparagraph{Relation between $B(p,q)=W_1\cap W_2$ and $W_1\setminus W_2$.} The following two lemmas help analyze step~2c of Algorithm \textsc{SparseYao} that adds some of the edges $pq_{i',j}$ to the spanner. \begin{lemma}\label{lem:WWW} For points $p,q\in S$, recall that $B(p,q)=W_1\cap W_2$, where $W_1$ and $W_2$ are wedges of aperture $\frac12\cdot\sqrt{\eps}$ and $\sqrt{\eps}$, resp.; see Fig.~\ref{fig:wedge}. For every point $q'\in W_1\setminus W_2$, we have $\|pq'\|\leq 2\,\|pq\|$. \end{lemma} \begin{proof} The line segment $pq$ decomposes $W_1\setminus W_2$ into two isosceles triangles. By the triangle inequality, the diameter of each isosceles triangle is less than $2\|pq\|$. This implies that for any point $q'\in W_1\setminus W_2$, we have $\|pq' \|< 2\, \|pq_i\|$. \end{proof} The following lemma justifies the role of the regions $\widehat{B}(p,q_i)$. \begin{figure}[htbp] \centering \includegraphics[width=0.4\textwidth]{wedge25} \caption{If $q_{i,j}\in B(p,q_i)$ but $q\notin B(p,q_i)$, then $q\in \widehat{B}(p,q_i)$.} \label{fig:wedge25} \end{figure} \begin{lemma}\label{lem:bbb} Let $p,q\in S$, and assume that $q\in C_{i',j}(p)$ for some $i,j\in \{1,\ldots , k\}$ and $i'\in \{i-1,i,i+1\}$, where $q_{i',j}=q_{i',j}(p)$ is a closest point to $p$ in $C_{i',j}(p)$. If $q\notin {B}(p,q_i)$ but $q_{i',j}\in B(p,q_i)$, then $q\in \widehat{B}(p,q_i)$. \end{lemma} \begin{proof} Since $q\in C_{i-1}(p)\cup C_i(p)\cup C_{i+1}(p)$ and the aperture of $C_i(p)$ is $2\pi/k\leq \frac18\cdot \sqrt{\eps}$, then $\angle qpq_i\leq \frac38\cdot \sqrt{\eps}$, and so $q\in W_1\setminus W_2$. Lemma~\ref{lem:WWW} yields $\|pq\|\leq 2\, \|pq_i\|$. Consider the circle of radius $\|pq\|$ centered at $p$; see Fig.\ref{fig:wedge25}. Since $\|pq\|\geq \|pq_{i',j}\|$ and $q_{i',j}\in B(p,q_i)$, this circle intersects $\partial B(p,q_{i'})$ in the cone $C_{i,j}(p)$. Denote by $q'$ the intersection point. Now $q,q'\in C_{i',j}(p)$ implies that $\angle q'pq_{i',j}\leq 2\pi/k^2\leq \eps/128$. The distance $\|qq'\|$ is bounded above by the length of the circular arc between them: $\|qq'\|\leq \|pq\| \angle qpq' \leq \|pq\| \angle qpq_{i',j} \leq \frac{\eps}{128}\,\|pq\| <\frac{\eps}{128}\,2\,\|pq_i\|$. As $q$ is at distance less than $\frac{\eps}{8}$ from $q'\in B(p,q_i)$, then $q\in \widehat{B}(p,q_{i'})$, as required. \end{proof} We can also clarify the relation between $\widehat{A}(p,q)$ and $\tilde{A}(p,q)$ in the setting used in the stretch analysis. \begin{lemma}\label{lem:tilde} Let $p,q\in S$, and assume that $q\in C_{i',j}(p)$ for some $i,j\in \{1,\ldots , k\}$ and $i'\in \{i-1,i,i+1\}$, where $q_{i',j}=q_{i',j}(p)$ is a closest point to $p$ in $C_{i',j}(p)$. Then $\widehat{A}(p,q_i)\subset \tilde{A}(p,q_{i',j})$. \end{lemma} \begin{proof} Since the aperture of $C_i(p)$ is $\frac18 \cdot \sqrt{\eps}$ and $q_i\in C_i(p)$, then $\angle q_ipq_{i',j}\leq \frac14 \, \sqrt{\eps}$. Since $\|pq_i\|\leq \|pq_{i',j}\|$, then $\|A(p,q_i)\|\leq \|A(p,q_{i',j})\|$. Consequently, every point in $A(p,q_i)$ is within distance at most $\|A(p,q_i)\|\sin \angle q_ipq_{i',j})\leq \frac{\sqrt{\eps}}{2}\, \|pq_i\|\cdot \frac14 \, \sqrt{\eps} \leq \frac{\eps}{8}\, \|pq_i\|$ from $A(p,q_{i',j})$. By the triangle inequality, the $(\frac{\eps}{16}\, \|pq_i\|)$-neighborhood of $A(p,q_i)$ is within distance at most $(\frac{\eps}{8}+\frac{\eps}{16})\|pq_i\|< \frac{\eps}{5}\,\|pq_i\|$ from $A(p,q_{i',j})$. \end{proof} \subparagraph{Completing the Stretch Analysis.} We are now ready to present the stretch analysis for \textsc{SparseYao}$(S,\eps)$. \begin{theorem}\label{thm:twostage} For every finite point set $S\subset \R^2$ and $\eps\in (0,\frac19)$, the graph $G=\textsc{SparseYao}(S,\eps)$ is a $(1+\eps)$-spanner. \end{theorem} \begin{proof} Let $S$ be a set of $n$ points in the plane. Let $L_0$ be the list of all $\binom{n}{2}$ edges of the complete graph on $S$ sorted by Euclidean weight (ties broken arbitrarily). For $\ell=1,\ldots , \binom{n}{2}$, let $e_\ell$ be the $\ell$-th edge in $L_0$, and let $E(\ell)=\{e_1,\ldots ,e_\ell\}$. We show the following claim, by induction, for every $\ell=1,\ldots , \binom{n}{2}$: \begin{claim}\label{cl:induction} For every edge $ab\in E(\ell)$, $G=(S,E)$ contains an $ab$-path of weight at most $(1+\eps)\|ab\|$. \end{claim} For $\ell=1$, the claim clearly holds, as the shortest edge $pq$ is necessarily the shortest in some cones $C_i(p)$ and $C_{i'}(q)$, as well, and so the algorithm adds $pq$ to $E$. Assume that $1<\ell\leq \binom{n}{2}$ and the claim holds for $\ell-1$. If the algorithm added edge $e_\ell$ to $E$, then the claim trivially holds for $\ell$. Suppose that $e_\ell\notin E$. Let $e_\ell=pq$, and $q\in C_{i,j}(p)$ for some $i,j\in \{1,\ldots , k\}$. Recall that $q_i=q_i(p)$ is a closest point to $p$ in the cone $C_i$; and $q_{i,j}=q_{i,j}(p)$ is a closest point to $p$ in the cone $C_{i,j}(p)$. We distinguish between two cases. \smallskip\noindent\textbf{(1) The algorithm added the edge $pq_i$ to $E$.} Note that $\|q_i q\|< \|pq\|$ and $\|q_{i,j} q\|<\|pq\|$. By the induction hypothesis, $G$ contains a $q_iq$-path $P_i$ of weight at most $(1+\eps)\|q_iq\|$ and a $q_{i,j}q$-path $P_{i,j}$ of weight at most $(1+\eps)\|q_{i,j} q\|$. If $q\in \widehat{B}(p,q_i)$, then $pq_i+P_i$ is a $pq$-path of weight at most $(1+\eps)\|pq\|$ by Lemma~\ref{lem:technical+}. Otherwise, $q\notin \widehat{B}(p,q_i)$. In this case, $q_{i,j}\notin B(p,q_i)$ by Lemma~\ref{lem:bbb}. This means that the algorithm added the edge $pq_{i,j}$ to $E$. We have $q\in \widehat{B}(p,q_{i,j})$ by Lemma~\ref{lem:bbb}, and so $pq_{i,j}+P_{i,j}$ is a $pq$-path of weight at most $(1+\eps)\|pq\|$ by Lemma~\ref{lem:technical+}. \smallskip\noindent\textbf{(2) The algorithm did not add the edge $pq_i$ to $E$.} Then the algorithm deleted $(p,q_i)$ from the list $L_i$ in a step in which it added another edge $p'q_i'$ to $E$. This means that $p\in \widehat{A}(p',q_i')$, where $q_i'$ is the closest point to $p'$ in the cone $C_i(p')$. As $\diam(A(p_i,q_i')) < (\sqrt{\eps}+2\cdot \frac{\eps}{16})\|p'q_i'\|< \frac14\, \|p'q_i'\|$ for $\eps\in (0,\frac19)$, then $p\in \widehat{A}(p',q_i')$ implies $\|pp'\|\leq \frac14\, \|p'q_i'\|$. Since $L_i$ is sorted by weight, then $\|p'q_i'\|\leq \|pq_i\|$. Although we have $q\in C_i(p)$, the point $q$ need not be in the cone $C_i(p')$; see Fig.~\ref{fig:wedge6}. We claim that $q$ lies in the union of three consecutive cones: $q\in C_{i-1}(p')\cup C_i(p')\cup C_{i+1}(p')$. Let $D_i(p')$ be part of the cone $C_i(p')$ outside of the circle of radius $\|p' q_i'\|$ centered at $p'$. Since $q\in C_i(p)$ and $\|p'q_i'\|<\|pq\|$, then $q$ lies in the translate $D_i(p')+\overrightarrow{p'p}$ of $D_i(p')$. Consider the union of translates: \[ D =D_i(p)+\{\overrightarrow{p'a}: a\in A_{p_i,q_i'}\}, \] and note that $q\in D$. We have $\diam(\widehat{A}(p',q_i'))\leq (\frac{\sqrt{\eps}}{2}+2\cdot \frac{\eps}{16})\|p'q_i'\| < \frac14\, \|p'q_i'\|$ for $\eps\in (0,\frac19)$; and recall that the aperture of $C_i(p')$ is $\gamma:=2\pi/k\leq \frac{1}{8}\cdot \sqrt{\eps}$. We can now approximate $\angle qp'q_i'$ as follows; refer to Fig.~\ref{fig:wedge6}: $\tan \angle qp'q_i' \leq \|p'q_i'\| \tan \gamma / (\|p'q_i'\| - 2\diam(A(p',q_i)))\leq 2\, \tan\alpha$. Consequently, $\angle qp'q_i' < 2\, \gamma$. It follows that $q\in \bigcup_{i'=i-2}^{i+2}C_{i'}(p')$. \begin{figure}[htbp] \centering \includegraphics[width=0.9\textwidth]{wedge4} \caption{The relative position of $pq$ and $p' q_i'$. Specifically, $p\in \widehat{A}(p',q_i')$ and $q\in C_i(p)$. Left: the region $D_i(p')$ and translates of $\widehat{A}(p',q_i')$ to two critical points of $D_i(p')$. Right: $q\in C_{i-1}(p')\cup C_i(p')\cup C_{i+1}(p')$ and the region $B(p'q_i')$.} \label{fig:wedge6} \end{figure} We distinguish between two subcases: \smallskip\noindent\textbf{(2a) $q\in \widehat{B}(p',q_i')$.} By induction, $G$ contains $(1+\eps)$-paths between $p$ and $p'$, and between $q$ and $q_i'$. By Lemma~\ref{lem:technical+} (with $a=p$ and $b=q$), the concatenation of these paths and the edge $p'q_i'$ is a $pq$-path of weight at most $(1+\eps)\|pq\|$. \smallskip\noindent\textbf{(2b) $q\notin \widehat{B}(p',q_i')$.} Then $q\in C_{i',j'}(p')$ for some $i'\in \{i-1,i,i+1\}$ and $j'\in \{1,\ldots ,k\}$. By Lemma~\ref{lem:bbb}, we have $q_{i',j'}\notin B(p',q_i)$. and so the algorithm added the edge $p'q_{i',j'}$, where $q_{i',j'}$ is the closest point to $p'$ in the cone $ C_{i',j'}(p')$. We have $p\in \widehat{A}(p',q_i')\subset \tilde{A}(p',q_{i',j})$ by Lemma~\ref{lem:tilde}, and $q\in \widehat{B}(p',q_{i',j})$ by Lemma~\ref{lem:bbb}. By induction, $G$ contains $(1+\eps)$-paths between $p$ and $p'$, and between $q_{i',j'}$ and $q$. The concatenation of these paths and the edge $p'q_{i',j'}$ is a $pq$-path of weight at most $(1+\eps)\|pq\|$ by Lemma~\ref{lem:technical+}. \end{proof} \section{Spanners in the Unit Square} \label{sec:square} In this section, we show that for a set $S\subset [0,1]^2$ of $n$ points in the unit square and $\eps\in (0,\frac19)$, Algorithm~\textsc{SparseYao} returns a $(1+\eps)$-spanner of weight $O(\eps^{-3/2}\sqrt{n})$ (cf.~Theorem~\ref{thm:UBsqaure}). The spanner \textsc{SparseYao}$(S,\eps)$ is a subgraph of the Yao-graph with cones of aperture $2\pi/k^2=O(\eps)$, and so it has $O(\eps^{-1}n)$ edges. Recall that for all $p\in S$ and all $i\in \{1,\ldots ,k\}$, there is at most one edge $pq_i(p)$ in $G$, where $q_i(p)$ is the closest point to $p$ in the cone $C_i(p)$ of aperture $\frac18\,\sqrt{\eps}$. Let \[F=\big\{p q_i(p)\in E(G): p\in S, i\in \{1,\ldots , k\}\big\}.\] We first show that the weight of the edges in $F$ approximates the weight of all other edges. \begin{lemma}\label{lem:deltoid} If Algorithm~\textsc{SparseYao} adds $pq_i(p)$ and $pq_{i',j}(p)$ to $G$ in the same iteration, then $\|pq_{i',j}(p)\|< 2\,\|pq_i(p)\|$. \end{lemma} \begin{proof} For short, we write $q_i=q_i(p)$ and $q_{i',j}=q_{i',j}(p)$, where $i\in \{i-1,i,i+1\}$. Since \textsc{SparseYao} added $pq_{i',j}$ to $G$, then $q_{i',j}\notin B(p,q_i)$. Recall (cf.\ Fig.~\ref{fig:wedge}) that $B(p,q_i) = W_1\cap W_2$, where $W_1$ and $W_2$ are cones centered at $p$ and $q_i$, resp., with apertures $\frac12\, \sqrt{\eps}$ and $\sqrt{\eps}$. Since the aperture of the cone $C_i(p)$ is $\frac18\,\sqrt{\eps}$, then $C_{i-1}(p)\cup C_i(p)\cup C_{i+1}(p)\subset W_1$. Consequently, $\left(C_{i-1}(p)\cup C_i(p)\cup C_{i+1}(p)\right)\setminus B(p,q_i)\subset W_1\setminus W_2$. Lemma~\ref{lem:WWW} gives $\|pq_{i',j}(p)\|< 2\,\|pq_i(p)\|$. as claimed. \end{proof} \begin{lemma}\label{lem:factor} For $G=\textsc{SparseYao}(S,\eps)$, we have $\|G\|=O(\eps^{-1/2}) \cdot \|F\|$. \end{lemma} \begin{proof} Fix $p$ and $i\in \{1,\ldots , k\}$, let $q_i=q_i(p)$ for short, and suppose $pq_i\in E(G)$. Consider one step of the algorithm that adds the edge $pq_i$ to $G$, together with up to $3k=\Theta(\eps^{-1/2})$ edges of type $p q_{i',j}$, where $q_{i',j}\notin B(p,q_i)$ and $i'\in \{i-1,i,i+1\}$. By Lemma~\ref{lem:deltoid}, $\|pq_{i',j}\|< 2 \|pq_i\|$. The total weight of all edges $p q_{i',j}$ added to the spanner is \[ \|p q_i(p)\|+\sum_{i'=i-1}^{i+1}\sum_{j=1}^k \|pq_{i',j}\| \leq \|pq_i\|+3k\cdot 2\, \|pq_i\| \leq O(k\|pq_i\|) \leq O(\eps^{-1/2})\|pq_i\|). \] Summation over all edges in $F$ yields $\|G\|=O(\eps^{-1/2}) \cdot \|F\|$. \end{proof} It remains to show that $\|F\|\leq O(\eps^{-1}\sqrt{n})$. For $i=1,\ldots , k$, let \[ F_i=\{pq_i(p)\in E(G): p\in S\}, \] that is, the set of edges in $G$ between points $p$ and the closest point $q_i(p)$ in cone $C_i(p)$ of aperture $\sqrt{\eps}$. We prove that $\|F_i\|\leq O(\eps^{-1/2}\, \sqrt{n})$ (Lemma~\ref{lem:Fi} below). Since $k=\Theta(\eps^{-1/2})$ this will immediately imply $\|F\|=\sum_{i=1}^k \|F_i\| =O(k\eps^{-1/2} \sqrt{n})=O(\eps^{-1}\sqrt{n})$. \subparagraph{Charging Scheme.} Let $i\in \{1,\ldots , k\}$ be fixed. Assume w.l.o.g.\ that the symmetry axis of the cone $C_i$ is horizontal, and the apex is the leftmost point of $C_i$. Refer to Fig.~\ref{fig:drops}(left). For each edge $pq_i(p)\in F_i$, let $R_i(p)$ be the intersection of cone $C_i(p)$ and the disk of radius $\|pq_i(p)\|$ centered at $p$. Note that $R_i(p)$ is a sector of the disk, and the sectors $R_i(p)$ are pairwise homothetic. The sector $R_i(p)$ has three vertices: Its leftmost vertex is $p$, and the other two vertices are the endpoints of a circular arc, which have the same $x$-coordinate (by symmetry). As $q_i(p)$ is the closest point to $p$ in $C_i(p)$, then $S\cap \mathrm{int}(R_i(p))=\emptyset$. \begin{figure}[htbp] \begin{center} \includegraphics[width=0.75\textwidth]{drops} \end{center} \caption{Left: a sector $R_i(p)$. Right: two intersecting sectors $R_i(p)$ and $R_i(p')$.} \label{fig:drops} \end{figure} Let $\mathcal{R}_i$ be the set of sectors for all edges $pq_i(p)$ in $G$. These sectors are not necessarily disjoint; but we can still give a lower bound on the area of their union. We first study their intersection pattern. \begin{lemma}\label{lem:intersect} Assume that $R_i(p),R_i(p')\in \mathcal{R}_i$ and $R_i(p)\cap R_i(p')\neq \emptyset$, Then $p$ and $p'$ have smaller $x$-coordinates than any other vertices of the two sectors. \end{lemma} \begin{proof} Denote the vertices of $R_i(p)$ and $R_i(p')$ by $a,b,p$ and $a',b',p'$, resp.; see Fig.~\ref{fig:drops}(right). Point $p'$ is in the exterior of $R_i(p)$, since $S\cap \mathrm{int}((R_i(p))=\emptyset$. Now $R_i(p)\cap R_i(p')\neq \emptyset$ implies that boundaries of $R_i(p)$ and $R_i(p')$ intersect. Suppose, to the contrary, $p'$ to the right of the vertical line $ab$. Since $p'$ is the leftmost point of $R_i(p')$, then all boundary intersections $\partial R_i(p)\cap \partial R_i(p')$ are on the circular arc $ab$. If $a'p'$ or $b'p'$ intersects the circular arc $ab$, then $p'\in \mathrm{int}(R_i(p))$, a contradiction. Otherwise only the circular arcs $ab$ and $a'b'$ intersect: Then both $p$ and $p'$ lie on the orthogonal bisector of the two intersection points; and we arrive again at a contradiction $p'\in \mathrm{int}(R_i(p))$. \end{proof} We partition the sectors $R_i(p)$ according to their sizes: For all $j\in \N$, let $\mathcal{R}_{i,j}$ be the set of sectors $R_i(p)$ such that $2^{-j}\leq \|p q_i(p)\| < 2^{1-j}$. We show that the sectors $\mathcal{R}_{i,j}$ do not overlap too heavily. \begin{lemma}\label{lem:density} For all $j\in \N$, any point $g\in [0,1]^2$ is contained in $O(\eps^{-1/2})$ sectors in $\mathcal{R}_{i,j}$. \end{lemma} \begin{proof} Let $\mathcal{R}_{i,j}(g)=\{R\in \mathcal{R}_{i,j}: g\in R\}$ be the set of sectors that contain $g$; these sectors pairwise intersect. By Lemma~\ref{lem:intersect}, the leftmost vertices of the sectors have smaller $x$-coordinates than any other vertices (i.e., endpoints of circular arcs). Let $\ell$ be a vertical line that separates the leftmost vertices of these sectors from all other vertices; and let $\ell^-$ be the left halfplane bounded by $\ell$; see Fig.~\ref{fig:terrain}(left). \begin{figure}[htbp] \begin{center} \includegraphics[width=0.9\textwidth]{terrain} \end{center} \caption{Left: the set of sectors that contains point $g$; and the vertical line $\ell$. Right: regions $A(p,q_i(p))$ and $\widehat{A}(p,q_i(p))$ for a vertex $p\in \gamma$.} \label{fig:terrain} \end{figure} Recall that for every sector $R_i(p)\in \mathcal{R}_{i,j}(g)$, we have $\|gp\|\geq 2^{-j}$. As the aperture of $C_i$ is $\frac18\, \sqrt{\eps}$, with $\eps\in (0,\frac19)$, then the weight of the vertical segment $\ell\cap R_i(p)$ is at most $\|\ell\cap R_i(p)\| \leq 2^{-j}\ \cdot 2\sin (\frac{1}{16}\, \sqrt{\eps})\leq O(2^{-j} \sqrt{\eps})$. For every sector $R_i(p)\in \mathcal{R}_{i,j}(g)$, $R_i(p)\cap \ell^-$ is an isosceles triangle with two legs of slopes $\pm \tan (\frac{1}{16}\sqrt{\eps}) =\pm \Theta(\sqrt{\eps})$. Consider the union of these isosceles triangles, $(\bigcup_{R\in \mathcal{R}(g)} R)\cap \ell^-$. Its right boundary is a vertical segment of weight $O(2^{-j}\sqrt{\eps})$; and its left boundary is a $y$-monotone curve, that we denote by $\gamma$. The local $x$-minima of $\gamma$ are the leftmost vertices of the sectors $R_i(p)\in \mathcal{R}(g)$. Suppose $p,p'\in S$ are two consecutive local $x$-minima along $\gamma$; see Fig.~\ref{fig:terrain}(right). We claim that the $y$-coordinates of $p$ and $p'$ differ by at least $\frac{\eps}{32}\cdot 2^{-j}$. Suppose, to the contrary, that $|y(p)-y(p')|< \frac{\eps}{32}\cdot 2^{-j}$. Due to the slopes of $\gamma$, this implies \[ |x(p)-x(p')| <\frac{\eps}{32}\cdot 2^{-j} / \tan \left(\frac{1}{16}\sqrt{\eps}\right) \leq \frac{\sqrt{\eps}}{2}\,2^{-j}. \] Assume w.l.o.g.\ that Algorithm~\textsc{SparseYao} added edge $pq_i(p)$ before edge $p' q_i(p')$. Then $p'$ is to the left of $p$ (i.e., $p'$ has smaller $x$-coordinate than $p$). When the algorithm added edge $p q_i(p)$ to $G$, it deleted the pairs $(s, q_i(s))$ from $L_i$ for all $s\in \widehat{A}(p,q_i(p))$. Since $\|p q_i(p)\|\geq 2^{-j}$ and $pq_i(p)\subset C_i(P)$, then $A(p,q_i(p))$ is a line segment of weight at least $2^{-j}\, \sqrt{\eps}$ and slope at most $\tan(\frac{1}{16}\,\sqrt{\eps})$. This implies that $p'$ lies in the horizontal slab spanned by $A(p,q_i(p))$. Furthermore, the vertical distance between $p'$ and $A(p,q_i(p))$ is at most $2\, |y(p)-y(p')|< \frac{\eps}{16}\cdot 2^{-j}$. Recall that region $\widehat{A}(p,q_i(p))$ contains every point in an $(\frac{\eps}{16}\cdot 2^{-j})$-neighborhood of $A(p,q_i(p))$. Consequently, $p'\in \widehat{A}(p,q_i(p))$, and the algorithm deleted the pair $(p',q_i(p)')$ from $L_i$. This contradicts the assumption $p'q_i(p')\in E(G)$, and proves the claim. As the height of $\gamma$ is $O(2^{-j} \sqrt{\eps})$, it can accommodate at most $O(2^{-j}\sqrt{\eps})/ (\frac{\eps}{32}\cdot 2^{-j}) =O(\eps^{-1/2})$ local $x$-minima. This implies that $|\mathcal{R}(g)|\leq O(\eps^{-1/2})$, as claimed. \end{proof} In order to obtain disjoint regions, we define the \emph{core} of a sector $R_i(p)$, denoted $\widehat{R}_i(p)$; see Fig.~\ref{fig:core}. Label the vertices of $R_i(p)$ by $a$, $b$, and $p$, where $\|pa\|=\|pb\|=\|pq_i(p)\|$. Let $\mathbf{v}$ be a vector along the angle bisector of $\angle apb$ of weight $\|\mathbf{v}\|=\frac23 \|p q_i(p)\|$. Now let $\widehat{R}_i(p)=R_i(p)\cap (R_i(p)+\mathbf{v})$. Note that $\area(\widehat{R}_i(p))\geq \frac19\,\area(R_i(p))$. \begin{figure}[htbp] \begin{center} \includegraphics[width=0.8\textwidth]{core} \end{center} \caption{Left: the core $\widehat{R}_i(p)$ of a sector $R_i(p)$. Right: The cores $\widehat{R}_i(p)$ and $\widehat{R}_i(p')$ are disjoint.} \label{fig:core} \end{figure} \begin{lemma}\label{lem:disjoint} If $j+2\leq j'$, then any two sectors in $\mathcal{R}_{i,j}$ and $\mathcal{R}_{i,j'}$ have disjoint cores. \end{lemma} \begin{proof} Let $R_i(p)\in \mathcal{R}_{i,j}$ and $R_i(p')\in \mathcal{R}_{i,j'}$ with $j+2\leq j'$. Label their vertices by $a,b,p$ and $a',b',p'$, respectively. Note that \[ \|a'p'\| \leq 2^{-j'} \leq 2^{-(j+2)} \leq \frac12\,\|ap\|. \] Suppose, for the sake of contradiction, that $\widehat{R}_i(p)\cap \widehat{R}_i(p')\neq \emptyset$. Then $R_i(p)\cap R_i(p')\neq \emptyset$. By Lemma~\ref{lem:intersect}, $p'$ lies to the left of the vertical line $ab$. To maximize the intersection $R_i(p)\cap R_i(p')$, we may assume that $R_i(p')$ has maximal size (that is, $\|a'p'\| = \frac12\,\|ap\|$); and $p'\in ap\cup bp$. In this extremal case, however, $R_i(p')$ is still disjoint from the core $\widehat{R}_i(p)$; cf.~Fig.~\ref{fig:core}. Consequently, $\mathcal{R}_{i,j}$ and $\mathcal{R}_{i,j'}$ are disjoint. \end{proof} The combination of Lemmas~\ref{lem:density} and~\ref{lem:disjoint} gives an upper bound on the total area of all sectors in $\bigcup_{j\in \N}\mathcal{R}_{i,j}$. \begin{corollary}\label{cor:volume} For every $i$, we have $\sum_{R\in \mathcal{R}_i}\area(R)=\sum_{j\in \N} \sum_{R\in \mathcal{R}_{i,j}} \area(R)\leq O(\eps^{-1/2})$. \end{corollary} \begin{proof} For every sector $R\in \mathcal{R}_i(p)$, we have $\area(R)=\Theta(\area(\widehat{R}))$. Define the function $f_{i,j}:[0,1]^2\rightarrow \N$ such that for all $g\in [0,1]^2$, $f(g)$ is the number of cores $\widehat{R}$, $R\in \mathcal{R}_{i,j}$, that contain $g$ Then $\sum_{R\in \mathcal{R}_{i,j}} \area(\widehat{R}) =\int_{[0,1]^2} f_{i,j}(g)$. By Lemma~\ref{lem:density}, we have $f(g)\leq O(\eps^{-1/2})$ for all $g\in [0,1]$. By Lemma~\ref{lem:disjoint}, $\sum_{j\in\N} f_{i,3j+\ell}(g) \leq O(\eps^{-1/2})$ for all $\ell\in \{0,1,2\}$ and $g\in [0,1]$. Consequently, \begin{align*} \sum_{R\in \mathcal{R}_i}\area(R) &\leq O\left(\sum_{R\in \mathcal{R}_i}\area(\widehat{R}) \right) = O\left(\sum_{\ell=0}^2 \sum_{j\in \N} \sum_{R\in \mathcal{R}_{i,3j+\ell}} \area(\widehat{R}) \right)\\ &= O\left(\sum_{\ell=0}^2 \sum_{j\in \N} \int_{[0,1]^2} f_{i,3j+\ell}(g) \right) = O\left(\sum_{\ell=0}^2 \int_{[0,1]^2} \sum_{j\in \N} f_{i,3j+\ell}(g) \right)\\ &\leq O\left(\sum_{\ell=0}^2 \int_{[0,1]^2} \eps^{-1/2} \right) = O\left(\sum_{\ell=0}^2 \eps^{-1/2}\area([0,1]^2)\right) =O(\eps^{-1/2}), \end{align*} as claimed. \end{proof} \begin{lemma}\label{lem:Fi} For every $i\in \{1,\ldots, k\}$, we have $\|F_i\|\leq O(\eps^{-1/2}\sqrt{n})$. \end{lemma} \begin{proof} For every $R_i(p)\in \bigcup_{j\in \N}\mathcal{R}_{i,j}$, we have \[ \area(\widehat{R}_i(p)) \geq \Omega(\area(R_i(p))) \geq \Omega(\|p q_i(p)\|^2 \sqrt{\eps}). \] In particular, summation over all $e\in F_i$ and Jensen's inequality gives \[ \sum_{j\in \N} \sum_{R\in \mathcal{R}_{i,j}} \area(R) \geq \Omega\left(\sqrt{\eps}\, \sum_{e\in F_i} \|e\|^2\right) \geq \Omega\left(\sqrt{\eps}\cdot |F_i|\left(\frac{1}{|F_i|}\sum_{e\in F_i} \|e\|\right)^2\right) \geq \Omega\left(\sqrt{\eps}\cdot |F_i| \cdot w_i^2\right), \] where $w_i=\frac{1}{|F_i|}\sum_{e\in F_i} \|e\| = \|F_i\|/|F_i|$ is the average weight of an edge in $F_i$. Combined with Corollary~\ref{cor:volume}, we obtain \later{ \begin{align} \sqrt{\eps}\cdot |F_i|\cdot w_i^2 &\leq O(\eps^{-1/2})\label{eq:key}\\ w_i^2 &\leq O\left(\frac{1}{\eps\, |F_i|}\right)\nonumber \\ w_i &\leq O\left(\frac{1}{\sqrt{\eps\, |F_i|}}\right). \nonumber \end{align} } \begin{equation}\label{eq:key} \sqrt{\eps}\cdot |F_i|\cdot w_i^2 \leq O(\eps^{-1/2}) \hspace{3mm}\Rightarrow\hspace{3mm} w_i^2 \leq O\left(\frac{1}{\eps\, |F_i|}\right) \hspace{3mm}\Rightarrow\hspace{3mm} w_i \leq O\left(\frac{1}{\sqrt{\eps\, |F_i|}}\right). \end{equation} Finally, $\|F_i\|\leq |F_i|\cdot w_i \leq O(|F_i| /\sqrt{\eps\, |F_i|}) = O(\eps^{-1/2}\sqrt{|F_i|}) \leq O(\eps^{-1/2}\sqrt{n})$, as required. \end{proof} \begin{theorem}\label{thm:UBsqaure} For every set of $n$ points in $[0,1]^2$ and every $\eps>0$, Algorithm \textsc{SparseYao} returns a Euclidean $(1+\eps)$-spanner of weight $O(\eps^{-3/2}\,\sqrt{n})$. \end{theorem} \begin{proof} Let $G=\textsc{SparseYao}(S,\eps)$, and define $F\subset E(G)$ and $F_1,\ldots, F_k$ as above. By Lemma~\ref{lem:Fi}, $\|F\|=\sum_{i=1}^k \|F_i\|= O(k\,\eps^{-1/2}\,\sqrt{n}) =O(\eps^{-1}\,\sqrt{n})$. By Lemma~\ref{lem:factor}, $\|G\|\leq O(\eps^{-1/2})\cdot (\|F\|+\sqrt{2})\leq O(\eps^{-3/2}\,\sqrt{n})$, as claimed. \end{proof} \section{Spanners for the Integer Grid} \label{sec:grid} We briefly review known results from analytic number theory in Section~\ref{ssec:Farey}, and derive upper bounds on the minimum weight of a $(1+\eps)$-spanner for the $n\times n$ grid: First we analyze the weight of Yao-graphs (Section~\ref{ssec:gridUB}), and then refine the analysis for Sparse Yao-graphs in (Section~\ref{ssec:next}). \subsection{Preliminaries: Farey Sequences} \label{ssec:Farey} Two points in the integer lattice $p,q\in \Z^2$ are \emph{visible} if the line segment $pq$ does not pass through any lattice point. An integer point $(i,j)\in \Z^2$ is visible from the origin $(0,0)$ if $i$ and $j$ are relatively prime, that is, $\mathrm{gcd}(i,j)=1$. The \emph{slope} of a segment between $(0,0)$ and $(i,j)$ is $j/i$. For every $n\in \N$, the \emph{Farey set of order $n$}, \[ F_n=\left\{\frac{a}{b} : 0\leq a\leq b\leq n\right\}, \] is the set of slopes of the lines spanned by the origin and lattice points $(b,a)\in [0,n]^2$ with $a\leq b$. The \emph{Farey sequence} is the sequence of elements in $F_n$ in increasing order. Note that $F_n\subset [0,1]$. Farey sets and sequences have fascinating properties, and the distribution of $F_n$, as $n\rightarrow \infty$ is not fully understood. It is known that \[ |F_n|=1+\sum_{1\leq i\leq n} \varphi(i) = \frac{3n^2}{\pi^2}+O(n \log n), \] where $\varphi(i)$ is Euler's totient function (i.e., $\varphi(i)$ is the number of nonnegative integers $j$ with $1\leq j\leq i$ and $\mathrm{gcd}(i,j)=1$). Furthermore, if $\frac{p_1}{q_1}$ and $\frac{p_2}{q_2}$ are consecutive terms of the Farey sequence in reduced form (i.e., $\mathrm{gcd}(p_1,q_1)=1$ and $\mathrm{gcd}(p_2,q_2)=1$), then $|p_1q_2-p_2q_1|=1$ \cite{HW79}. The Farey sequence is uniformly distributed on $[0,1]$ in the sense that for any fixed subinterval $[\alpha,\beta]\subset [0,1]$, the asymptotic frequency of the Farey set in $[\alpha,\beta]$ is known converge as $n$ tends to infinity~\cite{Dress99}: \[ \frac{|F_n\cap [\alpha,\beta]|}{|F_n|} = \beta-\alpha +o(1). \] The error term is bounded by $O(n^{-1+\varepsilon})$, for some $\eps>0$. However, determining the rate of convergence is known to be equivalent to the Riemann hypothesis~\cite{Franel24,Landau24}; see also~\cite{Ledoan18} and references therein. The key result we use is a bound on the average distance to a Farey set $F_n$. For every $x\in [0,1]$, let \[\rho_n(x)= \min_{\frac{p}{q}\in F_n} \left| \frac{p}{q} - x\right| \] denote the distance between $x$ and the Farey set $F_n$. Kargaev and Zhigljavsky~\cite{KZ96} proved that \begin{equation}\label{eq:KZ} \int_0^1 \rho_n(x) \, dx = \frac{3}{\pi^2} \, \frac{\ln n}{n^2} + O\left(\frac{1}{n^2}\right), \hspace{5mm}\textrm{as} \hspace{5mm} n\rightarrow \infty. \end{equation} \subsection{The Weight of Yao-Graphs for the Grid} \label{ssec:gridUB} \begin{lemma}\label{lem:num} For a positive integer $k\in \N$, consider the subdivision of the unit interval $[0,1]$ into $k$ subintervals, $[0,1]=\bigcup_{i=1}^k [\frac{i-1}{k},\frac{i}{k}]$. For every $i=1,\ldots , k$, let $q_i$ be the smallest positive integer such that $\frac{i-1}{k}\leq \frac{p_i}{q_i}\leq \frac{i}{k}$ for some integer $p_i$. Then $\sum_{i=1}^k q_i =O(k^{3/2}\log^{1/2} k)$. \end{lemma} \begin{proof} Let $[\alpha,\beta]\subset [0,1]$ be an interval of length $\beta-\alpha=\frac{1}{k}$. If $[\alpha,\beta]$ contains a point $\frac{p_i}{q_i}\in F_n$, then $q_i\leq n$. Otherwise, $[\alpha,\beta]$ is disjoint from the Farey set $F_n$, and then $\rho_n(x)\geq \min\{|x-\alpha|,|x-\beta|\}$ for all $x\in [\alpha,\beta]$. In this case, \begin{equation}\label{eq:blip} \int_{[\alpha,\beta]} \rho_n(x) \, dx \geq \int_{[\alpha,\beta]}\min\{|x-\alpha|,|x-\beta|\} \, dx = \frac{1}{4\cdot |\beta-\alpha|^2} = \frac{1}{4k^2}. \end{equation} For every positive integer $n$, let $a(n)$ be the number of intervals in $\{[\frac{i-1}{k},\frac{i}{k}]: i=1,\ldots , k\}$ that are disjoint from $F_n$. Note that $a(n)=0$ for $n\geq k$, since all $k$ (closed) intervals contain a rational of the form $\frac{p}{k}$. In particular, we have $q_i\leq k$ for all $i=1,\ldots , k$. The combination of \eqref{eq:KZ} and \eqref{eq:blip} yields $a(n)\leq O(k^2 \log n / n^2)$. If we set $n=\sqrt{ck\log k}$ for a sufficiently large constant $c\geq 2$, then \[ a \left(\sqrt{ck\log k}\right) \leq O \left( \frac{k^2 \log(\sqrt{ck \log k})}{ck\log k} \right) = O \left( \frac{k}{2} \cdot \frac{\log(ck\log k)}{c\log k} \right) = O \left( \frac{k}{2} \cdot \frac{\log(c) + 2\log(k)}{c\log k} \right) \leq \frac{k}{2}. \] That is, at most $\frac{k}{2}$ of $k$ intervals are disjoint from $F_n$. For the remaining at least $\frac{k}{2}$ intervals, we have $q_i\leq n\leq O(\sqrt{k\log k})$, and the sum of these $q_i$s is $O(k^{3/2}\log^{1/2}k)$. It remains to bound the sum of $q_i$s in the intervals disjoint form $F_n$. We use a standard diadic partition. Let $A=\sqrt{ck\log k}$. Then \begin{align*} \sum_{i=1}^k q_i &=O(k^{3/2}\log^{1/2} k)+\sum_{j=\lfloor \log A\rfloor}^{\lfloor \log k\rfloor} \big(a(2^j) - a(2^{j-1})\big)2^j \\ &=O(k^{3/2}\log^{1/2} k)+\sum_{j=\lfloor \log A\rfloor}^{\lfloor \log k\rfloor} a(2^j) \big(2^j-2^{j-1}\big)\\ &\leq O(k^{3/2}\log^{1/2} k)+\sum_{j=\lfloor \log A\rfloor}^{\lfloor \log k\rfloor} O\left(\frac{k^2 j}{2^{2j}}\right) 2^{j-1}\\ &=O\left(k^{3/2}\log^{1/2} k+ k^2 \sum_{j=\lfloor \log A\rfloor}^{\lfloor \log k\rfloor} \frac{j}{2^j}\right)\\ &=O\left(k^{3/2}\log^{1/2} k + k^2 \frac{\log A}{A} \right)\\ &=O\left(k^{3/2}\log^{1/2} k + k^2 \frac{\log k}{\sqrt{k \log k}} \right)\\ &= O(k^{3/2}\log^{1/2} k), \end{align*} as claimed. \end{proof} \begin{lemma}\label{lem:num2} In the setting of Lemma~\ref{lem:num}, we have $\sum_{i=1}^k q_i^3 =O(k^3\log k)$. \end{lemma} \begin{proof} If we set $n=\sqrt{ck\log k}$ for a sufficiently large constant $c\geq 2$, then $a(n)\leq \frac{k}{2}$. That is, at most $\frac{k}{2}$ of $k$ intervals are disjoint from $F_n$. For the remaining at least $\frac{k}{2}$ intervals, we have $q_i\leq n\leq O(\sqrt{k\log k})$, and the sum of $q_i^3$ for these values is $O(k^{5/2}\log^{3/2}k)$. It remains to bound the sum of $q_i$s in the intervals disjoint form $F_n$. Let $A=\sqrt{ck\log k}$. Then \begin{align*} \sum_{i=1}^k q_i^3 &=O(k^{5/2}\log^{3/2} k)+\sum_{j=\lfloor \log A\rfloor}^{\lfloor \log k\rfloor} \big(a(2^j) - a(2^{j-1})\big) 2^{3j}\\ &=O(k^{5/2}\log^{3/2} k)+\sum_{j=\lfloor \log A\rfloor}^{\lfloor \log k\rfloor} a(2^j) \big(2^{3j}-2^{3(j-1)}\big)\\ &\leq O(k^{5/2}\log^{3/2} k)+\sum_{j=\lfloor \log A\rfloor}^{\lfloor \log k\rfloor} O\left(\frac{k^2 j}{2^{2j}}\right) 2^{3j}\\ &=O\left(k^{5/2}\log^{3/2} k+ k^2 \sum_{j=\lfloor \log A\rfloor}^{\lfloor \log k\rfloor} j\cdot 2^{j}\right)\\ &=O\left(k^{5/2}\log^{3/2} k + k^2 \cdot k\log k \right)\\ &= O(k^3\log k), \end{align*} as claimed. \end{proof} \begin{lemma}\label{lem:grid} For a positive integer $k\in \N$, consider the subdivision of the plane into $k$ cones, each with apex at the origin $o$, and aperture $2\pi/k$. For each $i=1,\ldots , k$, let $p_i\in \Z^2\setminus \{0\}$ be a point in the $i$th cone that lies closest to the origin. Then $\sum_{i=1}^k \|op_i\| =O(k^{3/2}\log^{1/2} k)$. \end{lemma} \begin{proof} The function $\tan(x)$ is a monotone increasing bijection from the interval $[0,\frac{\pi}{4}]$ to $[0,1]$. Its derivative is bounded by $1\leq \tan'(x)\leq 2$. Consequently, it distorts the length of any interval by a factor of at most 2. That is, it maps any interval $[\alpha,\beta]\subset [0,\frac{\pi}{4}]$ to an interval $[\tan(\alpha), \tan(\beta)]\subset [0,1]$ with $\beta-\alpha\leq \tan(\beta)-\tan(\alpha)\leq 2(\beta-\alpha)$. Recall that Algorithm \textsc{SparseYao} operates on $k=\Theta(\eps^{-1/2})$ cones $C_1,\ldots ,C_k$. The two coordinate axes and the lines $y=\pm y$ subdivide the plane into eight cones with aperture $\pi/4$. These four lines contain a set $R$ of eight rays emanating from the origin. A cone $C_i$ that contains any ray in $R$ necessarily contains a lattice point $p_i$ with $\|op_i\|\leq \sqrt{2}$. Consider the cones $C_i$ between two consecutive rays in $R$; there are $O(k)$ such cones. Rotate these cones by a multiple of $\frac{\pi}{4}$ so that they correspond to slopes in the interval $[0,1]$. Due to the rotational symmetry of $\Z^2$, the rotation is an isometry on $\Z^2$. Each cone $C_i$ is an interval of angles $[\alpha_i,\beta_i]\subset [0,\frac{\pi}{4}]$ with $\beta_i-\alpha_i=\frac{2\pi}{k}$. As noted above, this corresponds to an interval of slopes $[\tan(\alpha_i),\tan(\beta_i)]\subset [0,1]$ with $\frac{2\pi}{k}\leq \tan \beta_i-\tan\alpha_i\leq \frac{4\pi}{k}$. Subdivide the interval $[0,1]$ into $k$ subintervals of length $\frac{1}{k}$. Each interval $[\tan(\alpha_i),\tan(\beta_i)]\subset [0,1]$ contains at least one intervals of this subdivision. Let $a_i$ be the smallest positive integer such that there exists a rational $\frac{b_i}{a_i}$ in the $i$th interval. Then the lattice point $(a_i,b_i)$ lies in $C_i$. Since $\frac{b_i}{a_i}\leq 1$, then $b_i\leq a_i$, and so the distance between $(a_i,b_i)$ and the origin is at most $\sqrt{2}\cdot a_i$. Let $p_i$ be a point in $C_i$ closest to the origin, hence $\|op_i\|\leq \sqrt{2}\cdot a_i$. By Lemma~\ref{lem:num}, we conclude that $\sum_{i} \|op_i\| \leq \sqrt{2} \sum_i a_i = O(k^{3/2}\log^{1/2} k)$, where the summation is over cones between two consecutive rays in $R$. Summation over all octants readily yields $\sum_{i=1}^k \|op_i\| \leq O(k^{3/2}\log^{1/2} k)$. \end{proof} \begin{corollary}\label{cor:Yao} For positive integers $k$ and $n$, let $G_{k,n}$ be the Yao-graph with $k$ cones on the $n^2$ points in the $n\times n$ section of the integer lattice. Then $\|G_{k,n}\| \leq O(nk^{3/2}\log^{1/2} k)$. \end{corollary} \begin{proof} Recall the Yao-graph is defined as a union of $n$ stars, one for each vertex. For each vertex $p$, a star $S_p$ is obtained by subdividing the plane into $k$ cones with apex $p$ and aperture $2\pi/k$, and connecting $p$ to the closes point in the cone (if such a point exists). By Lemma~\ref{lem:grid}, $\|S_p\|\leq O(k^{3/2}\log^{1/2} k)$ for any $p\in \Z^2$. Summation over $n^2$ vertices yields $\|G_{k,n}\| \leq O(nk^{3/2}\log^{1/2} k)$. \end{proof} For every finite point set $S\subset \R^2$, the Yao-graph $Y_k(s)$ with $k=\Theta(\eps^{-1})$ cones per vertex is a $(1+\eps)$-spanner. Corollary~\ref{cor:Yao} readily implies that the $n\times n$ section of the integer lattice admits a $(1+\eps)$-spanner of weight $O(\eps^{-3/2}\log^{1/2}(\eps^{-1})\cdot n^2)$. \subsection{The Weight of Sparse Yao-Graphs for the Grid} \label{ssec:next} \begin{theorem}\label{thm:UBgrid} Let $S$ be the $n\times n$ section of the integer lattice for some positive integer $n$. Then the graph $G=$\textsc{SparseYao}$(S,\eps)$ has weight $O(\eps^{-1}\log(\eps^{-1})\cdot n^2)$. \end{theorem} \begin{proof} The edges $p q_i(p)$ in Algorithm \textsc{SparseYao} form a Yao-graph $Y_k(S)$ with $k=\Theta(\eps^{-1/2})$ cones per vertex. By Corollary~\ref{cor:Yao}, we have $\|Y_k(S)\|\leq O(\eps^{-3/4}\log^{1/2}(\eps^{-1})\cdot n^2)$. Note, though, that Algorithm \textsc{SparseYao} does not necessarily add all these edges to the spanner. Algorithm \textsc{SparseYao} refines each cone $C_i(p)$ of apex $p$ and aperture $2\pi/k$ into $k$ cones $C_{i,j}(p)$, and adds some of the edges between $p$ and a closest point $q_{i,j}(p)\in C_{i,j}(p)$. It remains to bound the weight of the edges $pq_{i,j}$ added to the spanner. Fix $p\in S$ and $i\in \{1,\ldots , k\}$. Suppose that the algorithm adds $pq_i(p)$ together with $m$ edges of the form $pq_{i',j}$ for $i'\in \{i-1,i,i+1\}$ and $j\in \{1,\ldots , k\}$. We may assume w.l.o.g.\ that $m'\geq \lfloor m/2\rfloor$ of these points lie on the left side of the ray $\overrightarrow{pq_i}$. Label these $m'$ points in ccw order about $p$ as $t_1,\ldots ,t_{m'}$, and let $t_0=q_i$. Then the triangles $\Delta(p,t_{h-1},t_h)$, for $h=1,\ldots , m'$, are interior-disjoint, contained in $W_i\setminus W_2$, and spanned by lattice points in the $\Z^2$. By Pick's theorem, $\area(\Delta(p,t_{h-1},t_h))\geq \frac12$ for all $h=1,\ldots , m'$. Consequently, $\area(W_1\setminus W_2)\geq m'/2\geq \Omega(m)$. We also derive an upper bound on $\area(W_1\setminus W_2)$. Recall (cf.\ Fig.~\ref{fig:wedge}) that $B(p,q_i) = W_1\cap W_2$, where $W_1$ and $W_2$ are cones centered at $p$ and $q_i$, with apertures $\frac12\, \sqrt{\eps}$ and $\sqrt{\eps}$, respectively. Note that $pq_i$ partitions $W_1\setminus W_2$ into two congruent isosceles triangles, each with two sides of weight $\|pq_i\|$, and so $\area(W_1\setminus W_2)=2\|pq_i\|^2\sin (\pi-\sqrt{\eps}/2) = O(\eps^{1/2}\,\|pq_i\|^2 )$. By contrasting the lower and upper bounds for $\area(W_1\setminus W_2)$, we obtain \[\Omega(m)\leq \area(W_1\setminus W_2)\leq O(\eps^{1/2}\,\|pq_i\|^2 ) \,\,\, \Rightarrow \,\,\, m\leq O(\eps^{1/2} \|pq_i\|^2). \] By Lemma~\ref{lem:deltoid}, if an edge $pq_{i',j}(p)$ was added in the same iteration as $pq_i(p)$, then $\|pq_{i',j}(p)\|< 2\,\|pq_i(p)\|$. Overall, the total weight of all edges added together with $pq_i$ is $O(m\, \|pq_i\|)\leq O(\eps^{1/2}\, \|pq_i\|^3)$. Summation over $i=1,\ldots ,k$ yields $O(\eps^{1/2}\, \sum_{i=1}^k \|pq_i\|^3)$. Using Lemma~\ref{lem:num2} with $k=\Theta(\eps^{-1/2})$, the total weight of the edges added for a vertex $p$ is $O(\eps^{1/2}\cdot k^3\log k) = O(\eps^{-1}\log \eps^{-1})$. Summation over all $n^2$ vertices $p\in S$, gives an overall weight of $\|G\|=O(\eps^{-1}\log(\eps^{-1})\cdot n^2)$, as claimed. \end{proof} The combination of Lemma~\ref{lem:LBgrid} and Theorem~\ref{thm:UBgrid} (lower and upper bounds) establishes Theorem~\ref{thm:grid}. We restate Theorem~\ref{thm:UBgrid} for $n$ points in the unit square $[0,1]^2$. \begin{corollary} Let $S$ be the $\lfloor \sqrt{n}\rfloor \times \lfloor\sqrt{n}\rfloor$ section of the scaled lattice $\frac{1}{\lfloor \sqrt{n}\rfloor}\,\Z^2$, contained in $[0,1]^2$. Then $G=\textsc{SparseYao}(S,\eps)$ has weight $O(\eps^{-1}\log(\eps^{-1})\cdot \sqrt{n})$. \end{corollary} \begin{proof} Apply Theorem~\ref{thm:UBgrid} for the $\lfloor \sqrt{n}\rfloor \times \lfloor\sqrt{n}\rfloor$ section of the integer lattice $\Z^2$, and then scale down all weights by a factor of $\frac{1}{\lfloor \sqrt{n}\rfloor}$. \end{proof} \section{Generalization to Higher Dimensions} \label{sec:d-space} \later{ Algorithm \textsc{SparseYao} and its analysis generalizes to $\R^d$ for constant $d\geq 2$. } \subparagraph{Upper Bound.} We sketch the necessary adjustments for a point set $S\subset [0,1]^d$. In the plane, we partitioned $\R^2$ into $k=\Theta(\eps^{-1/2})$ cones $C_1,\ldots , C_k$, of aperture $2\pi/k=\frac18\,\sqrt{\eps}$. In $d$-dimensions, we can \emph{cover} $\R^d$ with $k=\Theta(\eps^{(1-d)/2})$ cones of aperture $\frac18\,\sqrt{\eps}$. With these cones, Algorithm \textsc{SparseYao} and its stretch analysis goes through almost verbatim. Standard volume argument shows that every cone of aperture $\frac18\,\sqrt{\eps}$ is covered by $O(\eps^{(1-d)/2})$ cones of aperture $\Theta(\eps^{-1})$. Therefore, the generalization of Lemma~\ref{lem:factor} yields $\|G\|=O_d(\eps^{(1-d)/2})\cdot \|F\|$. We can partition $F$ into $k=\Theta(\eps^{(1-d)/2})$ subsets $F=\bigcup_{i=1}^k F_i$. In the generalization of Lemma~\ref{lem:density}, every generic point $g\in [0,1]^d$ is contained in $O(\eps^{(1-d)/2})$ regions $R_i(p)$: In the proof, the $y$-monotote curve $\gamma$ is replaced by a $(d-1)$-dimensional surface/terrain. In the weight analysis for $\|F_i\|$, we need to charge the weight of each edge $e\in F_i$ to an empty region $\widehat{R}_i$, which is the intersection of a cone of aperture $\Theta(\sqrt{\eps})$ and a ball of radius $\|e\|$. The volume of such a region is $\Theta_d(\|e\|^d\cdot \eps^{(d-1)/2})$. Finally, in the proof of Lemma~\ref{lem:Fi}, Jensen's inequality is used for the function $x\rightarrow x^d$. In particular, for the average weight of edge in $F_i$, $w_i=\|F_i\|/ |F_i|$, inequality~\eqref{eq:key} becomes \begin{align} \eps^{(d-1)/2}\cdot |F_i|\cdot w_i^d &\leq O(\eps^{(1-d)/2})\label{eq:key+d}\\ w_i^d &\leq O\left(\frac{1}{\eps^{d-1} |F_i|}\right)\nonumber \\ w_i &\leq O\left(\frac{1}{\eps^{1-1/d}\, |F_i|^{1/d}}\right), \nonumber \end{align} \later{ \[ \eps^{(d-1)/2}\cdot |F_i|\cdot w_i^d \leq O(\eps^{(1-d)/2}) \hspace{3mm}\Rightarrow\hspace{3mm} w_i^d \leq O\left(\frac{1}{\eps^{d-1} |F_i|}\right) \hspace{3mm}\Rightarrow\hspace{3mm} w_i \leq O\left(\frac{1}{\eps^{1-1/d}\, |F_i|^{1/d}}\right), \] } and $\|F_i\| =|F_i|\cdot w_i \leq O( \eps^{1/d-1}|F_i|)^{1-1/d}) \leq O( \eps^{1/d-1} n^{1-1/d})$. Overall, \[ \|G\| \leq \eps^{(1-d)/2}\sum_i \|F_i\| \leq O(\eps^{(1-d)/2}\cdot \eps^{(1-d)/2}\cdot \eps^{1/d-1} n^{1-1/d}) \leq O(\eps^{-d+1/d} n^{1-1/d}). \] \subparagraph{Lower Bound.} The lower bound construction readily generalizes to every dimension $d\geq 2$. Let $S_0$ be a set of $2m$ points, where $m=\lfloor (\eps/d)^{1-d}\rfloor$, with $m$ points arranged in a grid on two opposite faces of a unit cube $[0,1]^d$. Every $(1+\eps)$-spanner for $S_0$ contains a complete bipartite graph $K_{m,m}$, of weight $\Omega_d(\eps^{2(1-d)})$. By arranging $\Theta_d(\eps^{d-1}n)$ translated copies of $S_0$ in a $(\eps^{d-1}n)^{1/d}\times \ldots \times (\eps^{d-1}n)^{1/d}$ grid, we obtain a set $S$ of $\Theta(n)$ points, and a lower bound of $\Omega_d(\eps^{1-d} n)$. Scaling by a factor of $(\eps^{d-1}n)^{1/d}$ yields a set of $\Theta(n)$ points in $[0,1]^d$ for which any $(1+\eps)$-spanner has weight $\Omega_d(\eps^{-d+1/d}\, n^{1-1/d})$. \section{Outlook} \label{sec:con} Our \textsc{SparseYao} algorithm combines features of Yao-graphs and greedy spanners. It remains an open problem whether the celebrated greedy algorithm~\cite{althofer1993sparse} always returns a $(1+\eps)$-spanner of weight $O(\eps^{-3/2}\sqrt{n})$ for $n$ points in the unit square (and $O(\eps^{(1-d^2)/d}n^{(d-1)/d})$ for $n$ points in $[0,1]^d$). The analysis of the greedy algorithm is known to be notoriously difficult~\cite{FiltserS20,LeS19}. It is also an open problem whether \textsc{SparseYao} or the greedy algorithm achieves an approximation ratio better than the tight lightness bound of $O(\eps^{-d})$ for $n$ points in $\R^d$ (where the approximation ratio compares the weight of the output with the instance-optimal weight of a $(1+\eps)$-spanner). All results in this paper pertain to the Euclidean distance (i.e., $L_2$-norm). Generalizations to $L_p$-norms for $p\geq 1$ (or Minkowski norms with respect to a centrally symmetric convex body in $\R^d$) would be of interest. It is unclear whether some or all of the machinery developed here generalizes to other norms. Finally, we note that Steiner points can substantially improve the weight of a $(1+\eps)$-spanner in Euclidean space~\cite{BhoreT21b,LeS19,LeS20}. It is left for future work to study the minimum weight of a Euclidean Steiner $(1+\eps)$-spanner for $n$ points in the unit square $[0,1]^2$ or unit cube $[0,1]^d$; and for an $n\times n$ section of the integer lattice. \bibliographystyle{plainurl}
{ "timestamp": "2022-07-01T02:03:43", "yymm": "2206", "arxiv_id": "2206.14911", "language": "en", "url": "https://arxiv.org/abs/2206.14911", "abstract": "Given a set $S$ of $n$ points in the plane and a parameter $\\varepsilon>0$, a Euclidean $(1+\\varepsilon)$-spanner is a geometric graph $G=(S,E)$ that contains, for all $p,q\\in S$, a $pq$-path of weight at most $(1+\\varepsilon)\\|pq\\|$. We show that the minimum weight of a Euclidean $(1+\\varepsilon)$-spanner for $n$ points in the unit square $[0,1]^2$ is $O(\\varepsilon^{-3/2}\\,\\sqrt{n})$, and this bound is the best possible. The upper bound is based on a new spanner algorithm in the plane. It improves upon the baseline $O(\\varepsilon^{-2}\\sqrt{n})$, obtained by combining a tight bound for the weight of a Euclidean minimum spanning tree (MST) on $n$ points in $[0,1]^2$, and a tight bound for the lightness of Euclidean $(1+\\varepsilon)$-spanners, which is the ratio of the spanner weight to the weight of the MST. Our result generalizes to Euclidean $d$-space for every constant dimension $d\\in \\mathbb{N}$: The minimum weight of a Euclidean $(1+\\varepsilon)$-spanner for $n$ points in the unit cube $[0,1]^d$ is $O_d(\\varepsilon^{(1-d^2)/d}n^{(d-1)/d})$, and this bound is the best possible.For the $n\\times n$ section of the integer lattice in the plane, we show that the minimum weight of a Euclidean $(1+\\varepsilon)$-spanner is between $\\Omega(\\varepsilon^{-3/4}\\cdot n^2)$ and $O(\\varepsilon^{-1}\\log(\\varepsilon^{-1})\\cdot n^2)$. These bounds become $\\Omega(\\varepsilon^{-3/4}\\cdot \\sqrt{n})$ and $O(\\varepsilon^{-1}\\log(\\varepsilon^{-1})\\cdot \\sqrt{n})$ when scaled to a grid of $n$ points in the unit square. In particular, this shows that the integer grid is \\emph{not} an extremal configuration for minimum weight Euclidean $(1+\\varepsilon)$-spanners.", "subjects": "Computational Geometry (cs.CG)", "title": "Minimum Weight Euclidean $(1+\\varepsilon)$-Spanners", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9947798732231231, "lm_q2_score": 0.8056321936479701, "lm_q1q2_score": 0.8014266914615943 }
https://arxiv.org/abs/1408.5180
Improvements on the infinity norm bound for the inverse of Nekrasov matrices
We focus on the estimating problem of the infinity norm of the inverse of Nekrasov matrices, give new bounds which involve a parameter, and then determine the optimal value of the parameter such that the new bounds are better than those in L. Cvetkovic et al. (2013). Numerical examples are given to illustrate the corresponding results.
\section{Introduction} A matrix $A=(a_{ij})\in C^{n, n}$ is called an $H$-matrix if its comparison matrix $ <A>=[m_{ij}]$ defined by \[ <A>=[m_{ij}]\in C^{n, n},m_{ij}=\left\{ \begin{array}{cc} |a_{ii}|, &i=j \\ -|a_{ij}|, &i\neq j, \end{array} \right.\] is an $M$-matrix, i.e., $<A>^{-1}\geq 0$ \cite{Ba,Cv1,Cv2}. $H$-matrices has a large number of applications. One special interest problem among them is to find upper bounds of the infinity norm of $H$-matrices, since it can be used for proving the convergence of matrix splitting and matrix multisplitting iteration methods for solving large sparse systems of linear equations, see \cite{Ba,Cv3,Hu,Hu1}. Many researchers gave some well-known bounds. In 1975, J.M. Varah\cite{Va} provided the follwoing upper bound for strictly diagonally dominant (SDD) matrices as one most important subclass of $H$-matrices. Here a matrix $A=[a_{ij}]\in C^{n, n}$ is called SDD if for each $i\in N=\{1,2,\ldots,n\}$, \[|a_{ii}|>r_i(A),\] where $r_i(A)=\sum\limits_{j\neq i} |a_{ij}|$. \begin{thm} \emph{\cite{Va}} \label{th1.1} Let $A=[a_{ij}]\in C^{n, n}$ be SDD. Then \[ ||A^{-1}||_\infty \leq \frac{1}{\min\limits_{i\in N} (|a_{ii}|-r_i(A))}.\] \end{thm} We call the bound in Theorem \ref{th1.1} the Varah's bound. As Cvetkovi\'c et al. \cite{Cv3} said, the Varah's bound works only for SDD matrices, and even then it is not always good enough. Hence, it can be useful to obtain new upper bounds for a wider class of matrices which sometimes works better in the SDD case. In \cite{Cv3}, Cvetkovi\'c et al. study the class of Nekrasov matrices which contains SDD matrices and is a subclass of $H$-matrices, and give the following bounds (see Theorem \ref{th1.2}). \begin{definition} \cite{Cv2,Cv3} A matrix $A=[a_{ij}]\in C^{n, n}$ is called a Nekrasov matrix if for each $i\in N$, \[|a_{ii}|>h_i(A),\] where $h_1(A)=r_1(A)=\sum\limits_{j\neq 1} |a_{1j}|$ and $h_i(A)=\sum\limits_{j=1}^{i-1}\frac{|a_{ij}|}{|a_{jj}|}h_j(A)+\sum\limits_{j=i+1}^{n}|a_{ij}|$, $i=2,3,\ldots,n$. \end{definition} \begin{thm} \emph{\cite[Theorem 2]{Cv3}} \label{th1.2} Let $A=[a_{ij}]\in C^{n, n}$ be a Nekrasov matrix. Then \begin{equation} \label{eq1.1} ||A^{-1}||_\infty \leq \frac{\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}} { 1-\max\limits_{i\in N} \frac{h_i(A)}{|a_{ii}|}},\end{equation} and \begin{equation} \label{eq1.10} ||A^{-1}||_\infty \leq \frac{\max\limits_{i\in N}z_i(A)} { \min\limits_{i\in N}(|a_{ii}|- h_i(A))},\end{equation} where $z_1(A)=1$ and $z_i(A)=\sum\limits_{j=1}^{i-1} \frac{|a_{ij}|}{|a_{jj}|}z_j(A)+1,$ $i=2,3\ldots,n$. \end{thm} Since an SDD matrix is a Nekrasov matrices \cite{Cv2,Li}, the bounds (\ref{eq1.1}) and (\ref{eq1.10}) can be also applied to SDD matrices. However, the Varah's bound can not be used to estimate the infinity norm of the inverse of Nekrasov matrices. Furthermore, when we use both bounds to estimate the infinity norm of the inverse of SDD matrices, the bound (\ref{eq1.1}) or (\ref{eq1.10}) works better than the Varah's bound in some cases (for details, see \cite{Cv3}). In this paper, we also focus on the estimating problem of the infinity norm of the inverse of Nekrasov matrices, and give new bounds which involve a parameter $\mu$ based on the bounds in Theorem \ref{th1.2}, and then determine the optimal value of $\mu$ such that the new bounds are better than those in Theorem \ref{th1.2} (Theorem 2 in \cite {Cv3}). Numerical examples are given to illustrate the corresponding results. \section {New bounds for the infinity norm of the inverse of Nekrasov matrices} First, some lemmas and notations are listed. Given a matrix $A=[a_{ij}]$, by $A=D-L- U$ we denote the standard splitting of $A$ into its diagonal $(D)$, strictly lower $(-L)$ and strictly upper $(-U)$ triangular parts. And by $[A]_{ij}$ we denote the $(i,j)$-entry of $A$, that is, $[A]_{ij}=a_{ij}$. \begin{lem}\label{le2.0} \emph{\cite{Be}} Let $A=[a_{ij}]\in C^{n,n}$ be a nonsingular $H$-matrix. Then \[ |A^{-1}|\leq ~<A>^{-1}.\] \end{lem} \begin{lem}\emph{\cite{Ro}}\label{le 2.1} Given any matrix $A=[a_{ij}]\in C^{n,n}$, $n\geq 2$, with $a_{ii}\neq 0$ for all $i\in N$, then \[h_i(A)= |a_{ii}|\left[(|D|-|L|)^{-1}|U|e\right]_i,\] where $e\in C^{n,n}$ is the vector with all components equal to 1. \end{lem} \begin{lem}\label{le 2.2} \emph{\cite{Sz}} A matrix $A=[a_{ij}]\in C^{n,n}$, $n\geq 2$ is a Nekrasov matrix if and only if \[(|D|-|L|)^{-1}|U|e<e,\] i.e., if and only if $E-(|D|-|L|)^{-1}|U|$ is an SDD matrix, where $E$ is the identity matrix. \end{lem} Let\[C=E-(|D|-|L|)^{-1}|U|=[c_{ij}]\] and \[B=|D|C= |D|-|D|(|D|-|L|)^{-1}|U|=[b_{ij}]\] and Then from Lemma \ref{le 2.2}, $B$ and $C$ are SDD when $A$ is a Nekrasov matrix. Note that $c_{11}=1$, $c_{k1}=0$, $k=2,3,\ldots,n$, and $c_{1k}=-\frac{|a_{1k}|}{|a_{11}|}$, $k=2,3,\ldots,n$, and that $b_{11}=|a_{11}|$, $b_{k1}=0$, $k=2,3,\ldots,n$, and $b_{1k}=-|a_{1k}|$, $k=2,3,\ldots,n$, which lead to the following lemma. \begin{lem}\label{le 2.3} Let $A=[a_{ij}] \in C^{n,n}$ be a Nekrasov matrix, \begin{equation} \label{eq2.0} C(\mu)=CD(\mu)=\left(E-(|D|-|L|)^{-1}|U|\right)D(\mu),\end{equation} and \begin{equation} \label{eq2.00} B(\mu)=BD(\mu)=\left(|D|-|D|(|D|-|L|)^{-1}|U|\right)D(\mu),\end{equation} where $D(\mu)=diag(\mu,1,\cdots,1)$ and $\mu> \frac{r_1(A)}{|a_{11}|}$. Then $C(\mu)$ and $B(\mu)$ are SDD, \begin{equation} \label{eq2.1} ||C(\mu)^{-1}||_\infty \leq \max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\},\end{equation} and \begin{equation} \label{eq2.10} ||B(\mu)^{-1}||_\infty \leq \frac{1}{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}}. \end{equation} \end{lem} \begin{proof} We first prove (\ref{eq2.1}) holds. It is not difficult from (\ref{eq2.0}) to see that $[C(\mu)]_{k1}=\mu c_{k1}$ for all $k\in N$ and $[C(\mu)]_{kj}=c_{kj}$ for all $k\in N$ and $j\neq 1$. Hence \[[C(\mu)]_{11}=\mu,~ r_1(C(\mu))=r_1(C)=\frac{r_1(A)}{|a_{11}|}\] and for $i=2,\ldots,n$, \[ [C(\mu)]_{ii}=c_{ii},~ r_i(C(\mu))=r_i(C).\] From $C$ is SDD and $\mu> \frac{r_1(A)}{|a_{11}|}$, we have that $C(\mu) $ is SDD. Moreover, by applying the Varah's bound to estimate the infinity norm of its inverse matrix, we can obtain \[|| C(\mu)^{-1}||_\infty \leq \max\limits_{i\in N} \frac{1}{|[C(\mu)]_{ii}|- r_i(C(\mu) )}=\max\left\{\frac{1}{\mu - r_1(C)}, \max\limits_{i\neq 1} \frac{1}{|c_{ii}|- r_i(C)}\right\}.\] Note that $C=E-(|D|-|L|)^{-1}|U|=[c_{ij}]$ and all diagonal entries of matrix $(|D|-|L|)^{-1}|U|$ are less than 1. Then we have that for $i\in N, ~i\neq 1, $ \[|c_{ii}|= 1-\left[(|D|-|L|)^{-1}|U| \right]_{ii}\] and that for each $i\in N$, \[ r_i(C)=\sum\limits_{k\neq i}\left[(|D|-|L|)^{-1}|U| \right]_{ik}.\] These lead to that (also see the proof of Theorem 2 in \cite{Cv3}) for $i\in N,~i\neq 1$, \[ |c_{ii}|-r_i(C)=1-\sum\limits_{k\in N} \left[(|D|-|L|)^{-1}|U| \right]_{ik}= 1-\left[(|D|-|L|)^{-1}|U|e \right]_{i}=1-\frac{h_i(A)}{|a_{ii}|}.\] Since $r_1(C)=\frac{r_1(A)}{|a_{11}|}=\frac{h_1(A)}{|a_{11}|}$, we have \begin{eqnarray*}|| C(\mu)^{-1}||_\infty&\leq & \max\left\{\frac{1}{\mu - r_1(C)}, \max\limits_{i\neq 1} \frac{1}{|c_{ii}|- r_i(C)}\right\}\\ &=& \max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \max\limits_{i\neq 1}\frac{1}{ 1-\frac{h_i(A)}{|a_{ii}|}}\right\}\\ &=&\max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\}. \end{eqnarray*} We prove easily that (\ref{eq2.10}) holds in an analogous way. The proof is completed. \end{proof} The main result of this paper is the following theorem: \begin{thm}\label{th 2.1} Let $A=[a_{ij}] \in C^{n,n}$ be a Nekrasov matrix. Then for $\mu>\frac{r_1(A)}{|a_{11}|}$, \begin{equation} \label{m1}||A^{-1}||_\infty \leq \max\{ \mu,1\} \max\limits_{i\in N} \frac{z_i(A)}{|a_{ii}|}\max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\},\end{equation} and \begin{equation}\label{m10} ||A^{-1}||_\infty \leq \frac{ \max\{ \mu,1\} \max\limits_{i\in n} z_i(A) }{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}}.\end{equation} \end{thm} \begin{proof}We only prove that (\ref{m1}) holds, and in an analogous way, (\ref{m10}) is proved easily. Let $C(\mu)=CD(\mu)=\left(E-(|D|-|L|)^{-1}|U|\right)D(\mu)$, where $D(\mu)=diag(\mu,1,\cdots,1)$. From (\ref{eq2.0}), we have \[C(\mu)=\left(E-(|D|-|L|)^{-1}|U|\right)D(\mu)=(|D|-|L|)^{-1}<A>D(\mu),\] which implies that \[<A>=(|D|-|L|)C(\mu) D(\mu)^{-1}.\] Furthermore, since a Nekrasov matrix is an $H$-matrix, we have from Lemma \ref{le2.0}, \begin{equation}\label{eq2.2}|| A^{-1}||_\infty\leq || <A>^{-1}||_\infty \leq ||D(\mu) ||_\infty||C(\mu)^{-1} ||_\infty || (|D|-|L|)^{-1}||_\infty.\end{equation} Note that $|D|-|L| $ is an $M$-matrix, and then similar to the proof of Theorem 2 in \cite{Cv3}, we can easily obtain \begin{equation}\label{eq2.3} ||(|D|-|L|)^{-1}||_\infty= ||y ||_\infty=\max\limits_{i\in n} \frac{z_i(A)}{|a_{ii}|}, \end{equation} where $y= (|D|-|L|)^{-1}e=[y_1,y_2,\ldots,y_n]^T$ and $z_i(A)=|a_{ii}|y_i$, i.e., \[ z_1(A)=1, ~and ~z_i(A)=\sum\limits_{j=1}^{i-1} \frac{|a_{ij}|}{|a_{jj}|}z_j(A)+1, i=2,\ldots,n.\] From (\ref{eq2.1}), (\ref{eq2.2}), (\ref{eq2.3}) and the fact that $||D(\mu) ||_\infty= \max\{ \mu,1\}$, we have \[||A^{-1}||_\infty \leq \max\{ \mu,1\} \max\limits_{i\in N} \frac{z_i(A)}{|a_{ii}|}\max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\}.\] The conclusions follows. \end{proof} \begin{example} \label{ex1} Consider the Nekrasov matrix $A_1$ in \cite{Cv3}, where \[A_1 = \left[ \begin{array}{cccc} -7 &1 &-0.2 &2 \\ 7 &88 &2 &-3 \\ 2 &0.5 &13 &-2\\ 0.5 & 3.0 &1 & 6 \end{array} \right].\] By computation, $h_1(A)= 3.2000,~h_2(A)= 8.2000,~h_3(A)= 2.9609,~h_4(A)= 0.7359,$ $ z_1(A)=1,$ $z_2(A)=2,$ $z_3(A)=1.2971$ and $z_4(A)=1.2394$. By Theorem \ref{th1.2} (Theorem 2 in \cite{Cv3}), we have \[||A_1^{-1}||_\infty \leq 0.3805, ~(The~ bound ~(\ref{eq1.1}) ~of~Theorem ~\ref{th1.2})\] and \[||A_1^{-1}||_\infty \leq 0.5263. ~(The~ bound ~(\ref{eq1.10}) ~of~Theorem ~\ref{th1.2})\] By the bound (\ref{m1}) of Theorem \ref{th 2.1}, we have \begin{eqnarray*}||A_1^{-1}||_\infty \leq & 4.8198 & ~(Taking ~\mu=0.5),\\ ||A_1^{-1}||_\infty \leq & 0.6025 & ~(Taking ~\mu=0.8),\\||A_1^{-1}||_\infty \leq & 0.3535 & ~(Taking ~\mu=1.1), \\||A_1^{-1}||_\infty \leq & 0.3745 & ~(Taking ~\mu=1.4),\\||A_1^{-1}||_\infty \leq & 0.4547 & ~(Taking ~\mu=1.7),\end{eqnarray*} and by the bound (\ref{m10}) of Theorem \ref{th 2.1}, we have \begin{eqnarray*}||A_1^{-1}||_\infty \leq & 2.0000 & ~(Taking ~\mu=0.6),\\ ||A_1^{-1}||_\infty \leq & 0.6452 & ~(Taking ~\mu=0.9),\\||A_1^{-1}||_\infty \leq & 0.4615 & ~(Taking ~\mu=1.2), \\||A_1^{-1}||_\infty \leq & 0.5699 & ~(Taking ~\mu=1.5),\\||A_1^{-1}||_\infty \leq & 0.6839 & ~(Taking ~\mu=1.8).\end{eqnarray*} In fact, $||A_1^{-1}||_\infty= 0.1921,$ \begin{figure}[bpt] \centering \begin{minipage}[b]{0.5\textwidth} \centering \includegraphics[width=2.4in]{Fig1.eps} \caption{The bounds (\ref{eq1.1}) and (\ref{m1})} \end{minipage}% \begin{minipage}[b]{0.5\textwidth} \centering \includegraphics[width=2.4in]{Fig2.eps} \caption{The bounds (\ref{eq1.10}) and (\ref{m10})} \end{minipage} \end{figure} \end{example} \begin{rmk} Example \ref{ex1} shows that by choosing the value of $\mu$, the bound (\ref{m1}) ((\ref{m10}), resp.) of Theorem \ref{th 2.1} is better than the bound (\ref{eq1.1}) ((\ref{eq1.10}), resp.) of Theorem \ref{th1.2} in some cases. We further observe the bounds in Theorem \ref{th 2.1} by Figures 1 and 2, and find that there is an interval such that for any $\mu$ in this interval, the bound (\ref{m1}) ((\ref{m10}), resp.) of Theorem \ref{th 2.1} for the matrix $A_1$ is always smaller than the bound (\ref{eq1.1}) ((\ref{eq1.10}), resp.) of Theorem \ref{th1.2}. An interesting problem arises: whether there is an interval of $\mu$ such that the bound (\ref{m1}) ((\ref{m10}), resp.) of Theorem \ref{th 2.1} for any Nekrasov matrix is smaller than the bound (\ref{eq1.1}) ((\ref{eq1.10}), resp.) of Theorem \ref{th1.2}? In the following section, we will study this problem. \end{rmk} \section{The choice of $\mu$} In this section, we determine the value of $\mu$ such that our bounds for $||A^{-1}||_\infty$ are less or equal to those of \cite{Cv3}. \subsection{\textbf{the optimal value of $\mu$ for the bound (\ref{m1})}} First, we consider the Nekrasov matrix $A=[a_{ij}] \in C^{n,n}$ with \begin{equation} \label{eq2.4}\frac{h_1(A)}{|a_{11}|}> \max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|},\end{equation} and give the following lemma. \begin{lem}\label{le 2.4} Let $A=[a_{ij}] \in C^{n,n}$ be a Nekrasov matrix with \[ \frac{h_1(A)}{|a_{11}|}> \max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}.\] Then \begin{equation} \label{eq2.6} 1<1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}<\frac{1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}{1 - \frac{h_1(A)}{|a_{11}|} }.\end{equation} \end{lem} \begin{proof} Obviously, the first Inequality in (\ref{eq2.6}) holds. We only prove that the second holds. From Inequality (\ref{eq2.4}), we have that \[ \frac{h_1(A)}{|a_{11}|}\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}- \left(\frac{r_1(A)}{|a_{11}|}\right)^2<0.\] Equivalently, \[1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}+\frac{h_1(A)}{|a_{11}|}-\frac{h_1(A)} {|a_{11}|}+\frac{h_1(A)}{|a_{11}|}\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}- \left(\frac{h_1(A)}{|a_{11}|}\right)^2<1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|},\] i.e., \[ \left(1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}+\frac{h_1(A)}{|a_{11}|} \right) \left( 1-\frac{h_1(A)}{|a_{11}|}\right)< 1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}.\] Note that $1-\frac{h_1(A)}{|a_{11}|}>0 $, then \[ 1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}+\frac{h_1(A)}{|a_{11}|}<\frac{1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}{1 - \frac{h_1(A)}{|a_{11}|} }.\] The conclusion follows. \end{proof} We now give an interval of $\mu$ such that the bound (\ref{m1}) of Theorem \ref{th 2.1} is less than the bound (\ref{eq1.1}) of Theorem \ref{th1.2}. \begin{lem}\label{le 2.5} Let $A=[a_{ij}] \in C^{n,n}$ be a Nekrasov matrix with \[\frac{h_1(A)}{|a_{11}|}> \max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}.\]Then for each $\mu \in \left(1, \frac{1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}{1 - \frac{h_1(A)}{|a_{11}|} }\right)$, \begin{eqnarray} \label{eq2.5}||A^{-1}||_\infty &\leq& \max\{ \mu,1\} \max\limits_{i\in N} \frac{z_i(A)}{|a_{ii}|}\max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\}\nonumber \\& <&\frac{\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}}{ 1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}}.\nonumber\end{eqnarray} \end{lem} \begin{proof} From Lemma \ref{le 2.4}, we have \[\mu \in\left(1,1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}\right]\bigcup \left[1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}, \frac{1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}{1 - \frac{h_1(A)}{|a_{11}|} }\right).\] and $ \max\{ \mu,1\}=\mu$. (I) For $ \mu \in \left(1,1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}\right],$ then \[ \mu -\frac{h_1(A)}{|a_{11}|}\leq 1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|},\] that is, \[ \frac{1}{\mu -\frac{h_1(A)}{|a_{11}|}}\geq \frac{1}{1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|} }.\] Therefore, \[ \max\{ \mu,1\} \max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\}= \frac{\mu}{ \mu - \frac{h_1(A)}{|a_{11}|}}.\] Consider the function $f(x)=\frac{x}{x-\frac{h_1(A)}{|a_{11}|}}, ~x\in \left[1,1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}\right]$. It is easy from $\frac{h_1(A)}{|a_{11}|}<1$ to prove that $f(x)$ is a monotonically decreasing function of $x$. Hence, for any $ \mu \in \left(1,1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}\right]$, \[f(\mu)< f(1),\] i.e., \[ \frac{\mu}{ \mu - \frac{h_1(A)}{|a_{11}|}}< \frac{1}{ 1 - \frac{h_1(A)}{|a_{11}|}}=\frac{1}{1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}},\] which implies that \[\frac{\mu\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}}{ \mu - \frac{h_1(A)}{|a_{11}|}}< \frac{\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}}{1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}}. \] Hence, \[\max\{ \mu,1\} \max\limits_{i\in N} \frac{z_i(A)}{|a_{ii}|}\max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\}\nonumber <\frac{\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}}{ 1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}}.\] (II) For $ \mu \in \left[1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}, \frac{1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}{1 - \frac{h_1(A)}{|a_{11}|} }\right),$ then \[ \mu -\frac{h_1(A)}{|a_{11}|}\geq 1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|},\] that is, \[ \frac{1}{\mu -\frac{h_1(A)}{|a_{11}|}}\leq \frac{1}{1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|} }.\] Therefore, \[ \max\{ \mu,1\} \max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\}= \frac{\mu}{1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|} }.\] Consider the function $g(x)=\frac{x}{1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}}, ~x\in \left[1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}, \frac{1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}{1 - \frac{h_1(A)}{|a_{11}|} }\right]$. Obviously, $g(x)$ is a monotonically increasing function of $x$. Hence, for any $ \mu \in \left[1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}, \frac{1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}{1 - \frac{h_1(A)}{|a_{11}|} } \right)$, \[g(\mu)< g\left( \frac{1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}{1 - \frac{h_1(A)}{|a_{11}|} }\right),\] that is, \[ \frac{\mu}{1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}} <\frac{1}{1- \frac{h_1(A)}{|a_{11}|}} =\frac{1}{1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}}, \] which implies that \[ \frac{\mu\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}} <\frac{\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}}{ 1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}}.\]Hence, \[\max\{ \mu,1\} \max\limits_{i\in N} \frac{z_i(A)}{|a_{ii}|}\max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\}\nonumber <\frac{\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}}{ 1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}}.\] The conclusion follows from (I) and (II). \end{proof} Lemma \ref{le 2.5} provides an interval of $\mu$ such that the bound (\ref{m1}) in Theorem \ref{th 2.1} is better than the bound (\ref{eq1.1}) in Theorem \ref{th1.2}. Moreover, we can determine the optimal value of $\mu$ by the following theorem. \begin{thm}\label{th2.2} Let $A=[a_{ij}] \in C^{n,n}$ be a Nekrasov matrix with \[\frac{h_1(A)}{|a_{11}|}> \max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}.\] Then \begin{eqnarray} \label{eq2.7} &\min\left\{ \max\{ \mu,1\} \max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\}: \mu \in \left(1, \frac{1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}{1 - \frac{h_1(A)}{|a_{11}|} }\right) \right\}\nonumber& \\&= \frac{1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}.&\end{eqnarray} Furthermore, \begin{equation} \label{eq2.8}||A^{-1}||_\infty \leq \frac{\max\limits_{i\in N} \frac{z_i(A)}{|a_{ii}|}\left(1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}\right)}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}} <\frac{\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}}{ 1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}}.\end{equation} \end{thm} \begin{proof} From the proof of Lemma \ref{le 2.5}, we have that \[f(x)=\frac{x}{x-\frac{h_1(A)}{|a_{11}|}}, ~x\in \left[1,1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}\right]\] is decreasing, and that \[g(x)=\frac{x}{1-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}}, ~x\in \left[1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}, \frac{1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}{1 - \frac{h_1(A)}{|a_{11}|} }\right]\] is increasing. Therefore, the minimum of $ f(x)$, which is equal to that of $g(x)$, is \[ f\left(1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}\right)=g\left(1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}\right)=\frac{1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}},\] which implies that (\ref{eq2.7}) holds. Again by Lemma \ref{le 2.5}, (\ref{eq2.8}) follows easily. \end{proof} \begin{rmk} Theorem \ref{th2.2} provides a method to determine the optimal value of $\mu$ for a Nekrasov matrix $A=[a_{ij}] \in C^{n,n}$ with \[\frac{h_1(A)}{|a_{11}|}> \max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}.\] Also consider the matrix $A_1$. By computation, we get \[ \frac{h_1(A_1)}{|a_{11}|}= 0.4571> 0.2278=\max\limits_{i\neq 1}\frac{h_{i}(A_1)}{|a_{ii}|}.\] Hence, by Theorem \ref{th2.2}, we can obtain that the bound (\ref{m1}) in Theorem \ref{th 2.1} reaches its minimum \[ \frac{\max\limits_{i\in N} \frac{z_i(A_1)}{|a_{ii}|}\left(1+\frac{h_1(A_1)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A_1)}{|a_{ii}|}\right)}{ 1-\max\limits_{i\neq 1}\frac{h_i(A_1)}{|a_{ii}|}} = 0.3288\] at $\mu=1+\frac{h_1(A_1)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A_1)}{|a_{ii}|}=1.2294$ (also see Figure 1). \end{rmk} Next, we study the bound in Theorem \ref{th 2.1} for the Nekrasov matrix $A=[a_{ij}] \in C^{n,n}$ with \[\frac{h_1(A)}{|a_{11}|}\leq \max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}.\] \begin{thm} \label{th2.3} Let $A=[a_{ij}] \in C^{n,n}$ be a Nekrasov matrix with \[\frac{h_1(A)}{|a_{11}|}\leq \max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}.\] Then we can take $\mu =1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}$ such that \begin{eqnarray} ||A^{-1}||_\infty &\leq& \max\{ \mu,1\} \max\limits_{i\in N} \frac{z_i(A)}{|a_{ii}|}\max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\}\nonumber \\& =&\frac{\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}}{ 1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}}.\nonumber\end{eqnarray} \end{thm} \begin{proof} Since $\frac{h_1(A)}{|a_{11}|}\leq \max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}$, we have $ \mu =1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}\leq 1$, $ \max\{ \mu,1\}=1$ and \[\max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\}=\frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}=\frac{1}{ 1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}}.\] Hence, \[\max\{ \mu,1\} \max\limits_{i\in N} \frac{z_i(A)}{|a_{ii}|}\max\left\{\frac{1}{\mu - \frac{h_1(A)}{|a_{11}|}}, \frac{1}{ 1-\max\limits_{i\neq 1}\frac{h_i(A)}{|a_{ii}|}}\right\} =\frac{\max\limits_{i\in N}\frac{z_i(A)}{|a_{ii}|}}{ 1-\max\limits_{i\in N}\frac{h_i(A)}{|a_{ii}|}}.\] The proof is completed. \end{proof} \subsection{\textbf{the optimal value of $\mu$ for the bound (\ref{m10})}} First, we consider the Nekrasov matrix $A=[a_{ij}] \in C^{n,n}$ with \[ \label{eq2.40}|a_{11}| - h_1(A) < \min\limits_{i\neq 1}(|a_{ii}|-h_i(A)),\] and give the following lemmas. \begin{lem}\label{le2.000} Let $a, b$ and $c$ be positive real numbers, and $0<a-b<c$. Then \[ \frac{b+c}{a}<\frac{c}{a-b}.\] \end{lem} \begin{proof} we only prove that $ \frac{c}{a-b}-\frac{b+c}{a}>0.$ In fact, \begin{eqnarray*} \frac{c}{a-b}-\frac{b+c}{a}&=& \frac{ac-(a-b)(b+c)}{a(a-b)}\\ &=&\frac{ac-(ab+ac-b^2-bc)}{a(a-b)}\\ &=&\frac{-ab+b^2+bc}{a(a-b)}\\ &=&\frac{b(c-(a-b))}{a(a-b)}>0. \end{eqnarray*} The proof is completed. \end{proof} \begin{lem}\label{le 2.40} Let $A=[a_{ij}] \in C^{n,n}$ be a Nekrasov matrix with \[|a_{11}| - h_1(A) < \min\limits_{i\neq 1}(|a_{ii}|-h_i(A)).\] Then \begin{equation} \label{eq2.60} 1<\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}<\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A)) }{|a_{11}|- h_1(A)}.\end{equation} \end{lem} \begin{proof} Since $A$ is a Nekrasov matrix, we have $|a_{11}| - h_1(A)>0$, consequently, the first Inequality in (\ref{eq2.60}) holds. Moreover, Let $a=|a_{11}|$, $ b=h_1(A)$ and $ c=\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))$. Then from Lemma \ref{le2.000}, the second holds. \end{proof} We now give an interval of $\mu$ such that the bound (\ref{m10}) of Theorem \ref{th 2.1} is less than the bound (\ref{eq1.10}) of Theorem \ref{th1.2}. \begin{lem}\label{le 2.50} Let $A=[a_{ij}] \in C^{n,n}$ be a Nekrasov matrix with \[|a_{11}| - h_1(A) < \min\limits_{i\neq 1}(|a_{ii}|-h_i(A)).\] Then for each $\mu \in \left(1, \frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A)) }{|a_{11}|- h_1(A)}\right)$, \begin{eqnarray} \label{eq2.5}||A^{-1}||_\infty &\leq&\frac{ \max\{ \mu,1\} \max\limits_{i\in n} z_i(A) }{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}}\nonumber \\& <&\frac{\max\limits_{i\in N}z_i(A)} { \min\limits_{i\in N}(|a_{ii}|- h_i(A))}.\nonumber\end{eqnarray} \end{lem} \begin{proof} From Lemma \ref{le 2.40}, we have \[\mu \in \left(1,\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}\right]\bigcup \left[\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}, \frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A)) }{|a_{11}|- h_1(A)}\right).\] and $ \max\{ \mu,1\}=\mu$. (I) For $ \mu \in \left(1,\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}\right],$ then \[ \mu|a_{11}|\leq \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) ,\] that is, \[\mu|a_{11}| - h_1(A) \leq \min\limits_{i\neq 1}(|a_{ii}|-h_i(A)) .\] Therefore, \[ \frac{ \max\{ \mu,1\} }{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}}= \frac{\mu}{\mu|a_{11}|-h_1(A)}.\] Consider the function $f(x)=\frac{x}{|a_{11}|x-h_1(A)}, ~x\in \left[1,\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}\right]$. It is easy to prove that $f(x)$ is a monotonically decreasing function of $x$. Hence, for any $ \mu \in \left(1,\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}\right]$, \[f(\mu)< f(1),\] i.e., \[ \frac{\mu}{\mu|a_{11}|-h_1(A)}< \frac{1}{ |a_{11}| -h_1(A)}= \frac{1}{\min\limits_{i\in N}(|a_{ii}|-h_i(A))},\] which implies that \[ \frac{\mu \max\limits_{i\in N}z_i(A)}{\mu|a_{11}|-h_1(A)}< \frac{ \max\limits_{i\in N}z_i(A)}{\min\limits_{i\in N}(|a_{ii}|-h_i(A))}.\] Hence, \[\frac{ \max\{ \mu,1\} \max\limits_{i\in n} z_i(A) }{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}} <\frac{\max\limits_{i\in N}z_i(A)} { \min\limits_{i\in N}(|a_{ii}|- h_i(A))}.\] (II) For $ \mu \in \left[\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}, \frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A)) }{|a_{11}|- h_1(A)}\right),$ then \[ \mu|a_{11}|\geq \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+h_1(A),\] that is, \[ \mu|a_{11}|-h_1(A)\geq \min\limits_{i\neq 1}(|a_{ii}|-h_i(A)).\] Therefore, \[ \frac{ \max\{ \mu,1\} }{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}}= \frac{\mu}{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))}.\] Consider the function $g(x)=\frac{x}{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))}, ~x\in \left[\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}, \frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A)) }{|a_{11}|- h_1(A)}\right]$. Obviously, $g(x)$ is a monotonically increasing function of $x$. Hence, for any $ \mu \in \left[\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}, \frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A)) }{|a_{11}|- h_1(A)}\right)$, \[g(\mu)< g\left( \frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A)) }{|a_{11}|- h_1(A)}\right),\] that is, \[ \frac{\mu}{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))} <\frac{1}{ |a_{11}| -h_1(A)}= \frac{1}{\min\limits_{i\in N}(|a_{ii}|-h_i(A))}, \] which implies that \[ \frac{\mu \max\limits_{i\in N}z_i(A)}{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))} <\frac{ \max\limits_{i\in N}z_i(A)}{\min\limits_{i\in N}(|a_{ii}|-h_i(A))}.\]Hence, \[\frac{ \max\{ \mu,1\} \max\limits_{i\in n} z_i(A) }{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}} <\frac{\max\limits_{i\in N}z_i(A)} { \min\limits_{i\in N}(|a_{ii}|- h_i(A))}.\] The conclusion follows from (I) and (II).\end{proof} Similar to the proof of Theorem \ref{th2.2}, we can easily determine the optimal value of $\mu$ by Lemma \ref{le 2.50}. \begin{thm}\label{th2.20} Let $A=[a_{ij}] \in C^{n,n}$ be a Nekrasov matrix with \[|a_{11}| - h_1(A) < \min\limits_{i\neq 1}(|a_{ii}|-h_i(A)).\] Then \begin{eqnarray} \label{eq2.70} &\min\left\{ \frac{ \max\{ \mu,1\} }{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}}: \mu \in \left(1, \frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A)) }{|a_{11}|- h_1(A)}\right)\right\}\nonumber& \\&= \frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))}.&\end{eqnarray} Furthermore, \begin{equation} \label{eq2.80}||A^{-1}||_\infty \leq \frac{ \max\limits_{i\in N}z_i(A) \left(\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) \right)}{|a_{11}|\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))} <\frac{\max\limits_{i\in N}z_i(A)} { \min\limits_{i\in N}(|a_{ii}|- h_i(A))}.\end{equation} \end{thm} \begin{rmk} Theorem \ref{th2.20} provides a method to determine the optimal value of $\mu$ for a Nekrasov matrix $A=[a_{ij}] \in C^{n,n}$ with \[|a_{11}| - h_1(A) < \min\limits_{i\neq 1}(|a_{ii}|-h_i(A)).\] Also consider the matrix $A_1$. By computation, we get \[ |a_{11}| - h_1(A)= 3.8000< 5.2641= \min\limits_{i\neq 1}(|a_{ii}|-h_i(A)).\] Hence, by Theorem \ref{th2.20}, we can obtain that the bound (\ref{m10}) in Theorem \ref{th 2.1} reaches its minimum\[ \frac{\max\limits_{i\in N}z_i(A)\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))} = 0.4594\] at $\mu=\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}= 1.2092$ (also see Figure 2). \end{rmk} Next, we study the bound (\ref{m10}) in Theorem \ref{th 2.1} for the Nekrasov matrix $A=[a_{ij}] \in C^{n,n}$ with \[|a_{11}| - h_1(A) \geq \min\limits_{i\neq 1}(|a_{ii}|-h_i(A)).\] \begin{thm} \label{th2.30} Let $A=[a_{ij}] \in C^{n,n}$ be a Nekrasov matrix with \[|a_{11}| - h_1(A) \geq \min\limits_{i\neq 1}(|a_{ii}|-h_i(A)).\] Then we can take $\mu =\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}$ such that \begin{eqnarray} ||A^{-1}||_\infty &\leq&\frac{ \max\{ \mu,1\} \max\limits_{i\in n} z_i(A) }{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}}\nonumber \\& =&\frac{\max\limits_{i\in N}z_i(A)} { \min\limits_{i\in N}(|a_{ii}|- h_i(A))}.\nonumber\end{eqnarray} \end{thm} \begin{proof} since $|a_{11}| - h_1(A) \geq \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))$, we have \[ \mu =\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}\leq 1,\] $ \max\{ \mu, ~1\}=1$, and \[\frac{ \max\{ \mu,1\} }{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}}= \frac{1 }{ \min\limits_{i\in N}(|a_{ii}|- h_i(A))}.\] Hence, \[\frac{ \max\{ \mu,1\} \max\limits_{i\in n} z_i(A) }{\min\left\{\mu|a_{11}| - h_1(A), \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))\right\}} =\frac{\max\limits_{i\in N}z_i(A)} { \min\limits_{i\in N}(|a_{ii}|- h_i(A))}. \] The proof is completed. \end{proof} \begin{rmk} (I) Theorems \ref{th2.2} and \ref{th2.3} provide the value of $\mu$, i.e., \[\mu=1+\frac{h_1(A)}{|a_{11}|}-\max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}\] such that the bound (\ref{m1}) in Theorem \ref{th 2.1} is not worse than the bound (\ref{eq1.1}) in theorem \ref{th1.2} for a Nekrasov matrix $A=[a_{ij}] \in C^{n,n}$. In particular, for the Nekrasov matrix $A$ with $\frac{h_1(A)}{|a_{11}|}> \max\limits_{i\neq 1} \frac{h_i(A)}{|a_{ii}|}$, the bound (\ref{m1}) is better than the bound (\ref{eq1.1}). (II) Theorems \ref{th2.20} and \ref{th2.30} provide the value of $\mu$, i.e., \[\mu=\frac{\min\limits_{i\neq 1}(|a_{ii}|-h_i(A))+ h_1(A) }{|a_{11}|}\] such that the bound (\ref{m10}) in Theorem \ref{th 2.1} is not worse than the bound (\ref{eq1.10}) in theorem \ref{th1.2} for a Nekrasov matrix $A=[a_{ij}] \in C^{n,n}$. In particular, for the Nekrasov matrix $A$ with $|a_{11}| - h_1(A) < \min\limits_{i\neq 1}(|a_{ii}|-h_i(A))$, the bound (\ref{m10}) is better than the bound (\ref{eq1.10}). \end{rmk} \section{Numerical Examples} \begin{example} Consider the following five Nekrasov matrices in \cite{Cv3}: \[A_2 = \left[ \begin{array}{cccc} 8 &1 &-0.2 &3.3 \\ 7 &13 &2 &-3 \\ -1.3 &6.7 &13 &-2\\ 0.5 & 3 &1 & 6 \end{array} \right],~ A_3=\left[\begin{array}{cccc} 21 & -9.1 &-4.2&-2.1 \\ -0.7& 9.1 &-4.2&-2.1 \\ -0.7& -0.7& 4.9&-2.1 \\ -0.7&-0.7&-0.7&2.8 \end{array} \right],\] \[A_4 = \left[ \begin{array}{cccc} 5 &1 &0.2 &2 \\ 1 &21 &1 &-3 \\ 2 &0.5 &6.4 &-2\\ 0.5 & -1 &1 & 9 \end{array} \right],~ A_5=\left[\begin{array}{ccc} 6 & -3 &-2 \\ -1& 11 &-8 \\ -7& -3&10 \end{array} \right],\] \[A_6= \left[ \begin{array}{cccc} 8 &-0.5 &-0.5 &-0.5 \\ -9 &16 &-5 &-5 \\ -6 &-4 &15 &-3\\ -4.9 & -0.9 &-0.9 & 6 \end{array} \right].\] Obviously, $A_2$, $A_3$ and $A_4$ are SDD. And it is not difficult to verify that $A_4,~A_5$ satisfy the conditions in Theorems \ref{th2.2} and \ref{th2.20} and $ A_2,~A_3,~A_6$ satisfy the conditions in Theorems \ref{th2.3} and \ref{th2.30}. We compute by Matlab 7.0 the upper bounds for the infinity norm of the inverse of $A_i$, $i=2,\ldots,6$, which are showed in Table 1. It is easy to see from Table 1 that this example illustrates Theorems \ref{th2.2}, \ref{th2.3}, \ref{th2.20} and \ref{th2.30},. \hspace{0.2cm} \noindent \begin{tabular}[htbp]{cccccc}\hline Matrix & $A_2$ &$A_3$ &$A_4$ & $A_5$ &$A_6$ \\\hline Exact $||A^{-1}||_\infty$ &0.2390 &0.8759 &0.2707 & 1.1519 &0.4474 \\\hline Varah &1 &1.4286 &0.5556 & -- &--\\\hline The bound (\ref{eq1.1}) &0.8848 &1.8076 &0.6200 & 1.4909 &1.1557 \\\hline Theorems \ref{th2.2} or \ref{th2.3} &0.8848 &1.8076 &0.5270 & 1.4266 &1.1557 \\\hline The bound (\ref{eq1.10}) &0.6885 &0.9676 &0.7937 & 2.4848 &0.5702 \\\hline Theorems \ref{th2.20} or \ref{th2.30}&0.6885 &0.9676 &0.5895 & 1.5923 &0.5702 \\\hline \end{tabular} \hspace{0.3cm} \noindent Table 1. The upper bounds for $||A_{i}^{-1}||_\infty$, $i=2,\ldots,6$. \end{example} \section*{Acknowledgements} This work is supported by National Natural Science Foundations of China (11361074, 11326242) and Natural Science Foundations of Yunnan Province (2013FD002).
{ "timestamp": "2014-08-25T02:04:18", "yymm": "1408", "arxiv_id": "1408.5180", "language": "en", "url": "https://arxiv.org/abs/1408.5180", "abstract": "We focus on the estimating problem of the infinity norm of the inverse of Nekrasov matrices, give new bounds which involve a parameter, and then determine the optimal value of the parameter such that the new bounds are better than those in L. Cvetkovic et al. (2013). Numerical examples are given to illustrate the corresponding results.", "subjects": "Numerical Analysis (math.NA)", "title": "Improvements on the infinity norm bound for the inverse of Nekrasov matrices", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9719924761487653, "lm_q2_score": 0.8244619242200082, "lm_q1q2_score": 0.8013707872129814 }
https://arxiv.org/abs/1905.12560
On the equivalence between graph isomorphism testing and function approximation with GNNs
Graph Neural Networks (GNNs) have achieved much success on graph-structured data. In light of this, there have been increasing interests in studying their expressive power. One line of work studies the capability of GNNs to approximate permutation-invariant functions on graphs, and another focuses on the their power as tests for graph isomorphism. Our work connects these two perspectives and proves their equivalence. We further develop a framework of the expressive power of GNNs that incorporates both of these viewpoints using the language of sigma-algebra, through which we compare the expressive power of different types of GNNs together with other graph isomorphism tests. In particular, we prove that the second-order Invariant Graph Network fails to distinguish non-isomorphic regular graphs with the same degree. Then, we extend it to a new architecture, Ring-GNN, which succeeds in distinguishing these graphs and achieves good performances on real-world datasets.
\section{Conclusions} In this work we address the important question of organizing the fast-growing zoo of GNN architectures in terms of what functions they can and cannot represent. We follow the approach via the graph isomorphism test, and show that is equivalent to the other perspective via function approximation. We leverage our graph isomorphism reduction to augment order-$2$ G-invariant nets with the ring of operators associated with matrix multiplication, which gives provable gains in expressive power with complexity $O(n^3)$, and is amenable to efficiency gains by leveraging sparsity in the graphs. Our general framework leaves many interesting questions unresolved. First, a more comprehensive analysis on which elements of the algebra are really needed depending on the application. Next, our current GNN taxonomy is still incomplete, and in particular we believe it is important to further discern the abilities between spectral and neighborhood-aggregation-based architectures. Finally, and most importantly, our current notion of invariance (based on permutation symmetry) defines a topology in the space of graphs that is too strong; in other words, two graphs are either considered equal (if they are isomorphic) or not. Extending the theory of symmetric universal approximation to take into account a weaker metric in the space of graphs, such as the Gromov-Hausdorff distance, is a natural next step, that will better reflect the stability requirements of powerful graph representations to small graph perturbations in real-world applications. \paragraph{Acknowledgements} We would like to thank Haggai Maron for fruitful discussions and for pointing us towards $G$-invariant networks as powerful models to study representational power in graphs. This work was partially supported by NSF grant RI-IIS 1816753, NSF CAREER CIF 1845360, the Alfred P. Sloan Fellowship, Samsung GRP and Samsung Electronics. SV was partially funded by EOARD FA9550-18-1-7007 and the Simons Collaboration Algorithms and Geometry. \section{Graph G-invariant Networks with maximum tensor order 2} \label{app.Ginvariant} In this section we prove Theorem \ref{prop.Ginvariant} that says that graph G-invariant Networks with tensor order 2 cannot distinguish between non-isomorphic regular graphs with the same degree. First, we need to state our definition of the order-2 Graph $G$-invariant Networks. In general, given $G \in \mathbb{R}^{n \times n}$, we let $A^{(0)} = G$, $d^{(0)} = 1$, and \[ A^{(t+1)} = \sigma(L^{(t)}(A^{(t)})) \] and outputs $m \circ h \circ A^{(L)}$, where each $L^{(t)}$ is an equivariant linear layer from $\mathbb{R}^{n \times n \times d^{(t)}}$ to $\mathbb{R}^{n \times n \times d^{(t+1)}}$, $\sigma$ is a point-wise activation function, $h$ is an invariant linear layer from $\mathbb{R}^{n \times n}$ to $\mathbb{R}$, and $m$ is an MLP. $d^{(t)}$ is the feature dimension in layer $t$, interpreted as the dimension of the hidden state attached to each pair of nodes. For simplicity of notations, in the following proof we assume that $d^{(t)} = 1, \forall t = 1, ..., L$, and thus each $A^{(t)}$ is essentially a matrix. The following results can be extended to the cases where $d^{(t)} > 1$, by adding more subscripts in the proof. Given an unweighted graph $G$, let $E \subseteq [n]^2$ be the edge set of $G$, i.e., $(u, v) \in E$ if $u \neq v$ and $G_{uv} = 1$; set $S \subseteq [n]^2$ to be $\{ (u, u) \}_{u \in [n]^2}$; and let $N = [n]^2 \setminus (E \cup S)$. Thus, $E \cup N \cup S = [n]^2$. \begin{lemma} Let $G, G'$ be the adjacency matrices of two unweighted regular graphs with the same degree $d$, and let $A^{(t)}, E, N, S$ and $A'^{(t)}, E', N', S'$ be defined as above for $G$ and $G'$, respectively. Then $\forall n \leq L, \exists \xi_1^{(t)}, \xi_2^{(t)}, \xi_3^{(t)} \in \mathbb{R}$ such that $A^{(t)}_{uv} = \xi_1^{(t)} \mathds{1}_{(u, v) \in E} + \xi_2^{(t)} \mathds{1}_{(u, v) \in N} + \xi_3^{(t)} \mathds{1}_{(u, v) \in S}$, and $A'^{(t)}_{uv} = \xi_1^{(t)} \mathds{1}_{(u, v) \in E'} + \xi_2^{(t)} \mathds{1}_{(u, v) \in N'} + \xi_3^{(t)} \mathds{1}_{(u, v) \in S'}$ \end{lemma} \begin{proof} We prove this lemma by induction. For $t=0$, $A^{(0)} = G$ and $A'^{(0)} = G'$. Since the graph is unweighted, $G_{uv} = 1$ if $u \neq v$ and $(u, v) \in E$, and $0$ otherwise. Similar is true for $G'$. Therefore, we can set $\xi_1^{(0)} = 1$ and $\xi_2^{(0)} = \xi_3^{(0)} = 0$. Next, we consider the inductive steps. Assume that the conditions in the lemma are satisfied for layer $t-1$. To simplify the notation, we use $A, A'$ to stand for $A^{(t-1)}, A'^{(t-1)}$, and we assume to satisfy the inductive hypothesis with $\xi_1, \xi_2$ and $\xi_3$. We thus want to show that if $L$ is any equivariant linear, then $\sigma(L(A)), \sigma(L(A'))$ also satisfies the inductive hypothesis. Also, in the following, we use $p_1, p_2, q_1, q_2$ to refer to nodes, $a, b$ to refer to pairs of nodes, $\lambda$ to refer to any equivalence class of 2-tuples (i.e. pairs) of nodes, and $\mu$ to refer to any equivalence class of 4-tuples of nodes. $\forall a = (p_1, p_2), b = (q_1, q_2) \in [n]^2$, let $\mathcal{E}(a, b)$ denote the equivalence class of 4-tuples containing $(p_1, p_2, q_1, q_2)$, and let $\mathcal{E}(b)$ represent the equivalence class of 2-tuples containing $(q_1, q_2)$. Two 4-tuples $(u, v, w, x), (u', v', w', x')$ are considered equivalent if $\exists \pi \in S_n$ such that $\pi(u) = u', \pi(v) = v', \pi(w) = w', \pi(x) = x'$. Similarly is equivalence between 2-tuples defined. By equation 9(b) in \cite{maron2018invariant}, using the notations of $T, B, C, w, \beta$ defined there, $L$ is described by, given $A$ as an input as $b$ as the subscript index on the output, \begin{equation} \begin{split} L(A)_{b} &= \sum_{a = (p_1, p_2) = (1, 1)}^{(n, n)} T_{a, b} A_a + Y_b \\ &= \sum_{a, \mu} w_{\mu} B_{a, b}^{\mu} A_a + \sum_\lambda \beta_\lambda C_b^\lambda\\ &= \sum_\mu (\mathop{\sum_{a \in [n]^2}}_{(a, b) \in \mu} A_a) w_\mu + \beta_{\mathcal{E}(b)} \end{split} \end{equation} First, let \[S_{\mu}^b = \mathop{\sum_{a \in [n]^2}}_{(a, b) \in \mu} A_a\] By the inductive hypothesis, \begin{equation} \begin{split} S_{\mu}^b &= \mathop{\mathop{\sum_{a \in [n]^2}}_{(a, b) \in \mu}}_{a \in E} A_a + \mathop{\mathop{\sum_{a \in [n]^2}}_{(a, b) \in \mu}}_{a \in N} A_a + \mathop{\mathop{\sum_{a \in [n]^2}}_{(a, b) \in \mu}}_{a \in S} A_a \\ &= \mathop{\mathop{\sum_{a \in [n]^2}}_{(a, b) \in \mu}}_{a \in E} \xi_1 + \mathop{\mathop{\sum_{a \in [n]^2}}_{(a, b) \in \mu}}_{a \in N} \xi_2 + \mathop{\mathop{\sum_{a \in [n]^2}}_{(a, b) \in \mu}}_{a \in S} \xi_3 \\ &= m_E(b, \mu) \xi_1 + m_N(b, \mu) \xi_2 + m_S(b, \mu) \xi_3 \end{split} \end{equation} where $m_E(b, \mu)$ is defined as the total number of distinct $a \in [n]^2$ that satisfies $(a, b) \in \mu$ and $a \in E$, and similarly for $m_N(b, \mu)$ and $m_S(b, \mu)$. Formally, for example, $m_E(b, \mu) = card\{ a \in [n]^2 : (a, b) \in \mu, a \in E\}$. Since $E \cup N \cup S = [n]^2$, $b$ belongs to one of $E, N$ and $S$. Thus, let $\tau(b) = E$ if $b \in E$, $\tau(b) = N$ if $b \in N$ and $\tau(b) = S$ if $b \in S$. It turns out that if $A$ is the adjacency matrix of a undirected regular graph with degree $d$, then $m_E(b, \mu), m_N(b, \mu), m_S(b, \mu)$ can be instead written (with an abuse of notation) as $m_E(\tau(b), \mu), m_N(\tau(b), \mu), m_S(\tau(b), \mu)$, meaning that for a fixed $\mu$, the values of $m_E, m_N$ and $m_S$ only depend on which of the three sets ($E, N$ or $S$) $b$ is in, and changing $b$ to a different member in the set $\tau(b)$ won't change the three numbers. In fact, for each $\tau(b)$ and $\mu$, the three numbers can be computed as functions of $n$ and $d$ using simple combinatorics, and their values are seen in the three tables \ref{table:mE}, \ref{table:mN} and \ref{table:mS}. An illustration of these numbers is given in Figure \ref{coloredreg}. \begin{figure} \label{coloredreg} \centering \includegraphics[width=0.25\textwidth,trim={6cm 4cm 20cm 7cm},clip]{skl2color2} \includegraphics[width=0.25\textwidth,trim={6cm 4cm 20cm 7cm},clip]{skl3color2} \caption{$m_E(E, \mathcal{E}(1, 2, 3, 4))$, $m_E(E, \mathcal{E}(1, 2, 3, 2))$, $m_E(E, \mathcal{E}(1, 2, 3, 1))$, $m_E(E, \mathcal{E}(1, 2, 2, 3))$ and $m_E(E, \mathcal{E}(1, 2, 1, 3))$ of $G_{8, 2} $ and $G_{8, 3}$. In either graph, twice the total number of black edges equal $m_E(E, \mathcal{E}(1, 2, 3, 4)) = 18$ (it is twice because each undirected edge corrspond to two pairs $(p_1, p_2)$ and $(p_2, p_1)$, which combined with $(q_1, q_2)$ both belongs to $\mathcal{E}(1, 2, 3, 4)$); the total number of of red edges, $3$, equals both $m_E(E, \mathcal{E}(1, 2, 2, 3))$ and $m_E(E, \mathcal{E}(1, 2, 1, 3))$; the total number of green edges, also $3$, equals both $m_E(E, \mathcal{E}(1, 2, 3, 2))$, $m_E(E, \mathcal{E}(1, 2, 3, 1))$.} \label{fig:my_label} \end{figure} \begin{table}[h] \centering \begin{tabular}{llll} \multicolumn{1}{l|}{$\mu$} & $m_E(E, \mu)$ & $m_E(N, \mu)$ & $m_E(S, \mu)$ \\ \hline \multicolumn{1}{l|}{(1, 2, 3, 4)} & $(n-4)d+2$ & $(n-4)d$ & 0 \\ \multicolumn{1}{l|}{(1, 1, 2, 3)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 2, 3)} & $d-1$ & $d$ & 0 \\ \multicolumn{1}{l|}{(1, 2, 1, 3)} & $d-1$ & $d$ & 0 \\ \multicolumn{1}{l|}{(1, 2, 3, 2)} & $d-1$ & $d$ & 0 \\ \multicolumn{1}{l|}{(1, 2, 3, 1)} & $d-1$ & $d$ & 0 \\ \multicolumn{1}{l|}{(1, 1, 1, 2)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 1, 2, 1)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 1, 2)} & 1 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 2, 1)} & 1 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 3, 3)} & 0 & 0 & $(n-2)d$ \\ \multicolumn{1}{l|}{(1, 1, 2, 2)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 2, 2)} & 0 & 0 & $d$ \\ \multicolumn{1}{l|}{(1, 2, 1, 1)} & 0 & 0 & $d$ \\ \multicolumn{1}{l|}{(1, 1, 1, 1)} & 0 & 0 & 0 \\ \hline \multicolumn{1}{l|}{Total} & $nd$ & $nd$ & $nd$ \end{tabular} \caption{$m_E$} \label{table:mE} \end{table} \begin{table}[h] \centering \begin{tabular}{llll} \multicolumn{1}{l|}{$\mu$} & $m_N(E, \mu)$ & $m_N(N, \mu)$ & $m_N(S, \mu)$ \\ \hline \multicolumn{1}{l|}{(1, 2, 3, 4)} & $(n-4)(n-d-1)$ & $(n-4)(n-d-1) + 2$ & 0 \\ \multicolumn{1}{l|}{(1, 1, 2, 3)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 2, 3)} & $n-d-1$ &$ n-d-2$ & 0 \\ \multicolumn{1}{l|}{(1, 2, 1, 3)} & $n-d-1$ & $n-d-2$ & 0 \\ \multicolumn{1}{l|}{(1, 2, 3, 2)} & $n-d-1$ & $n-d-2$ & 0 \\ \multicolumn{1}{l|}{(1, 2, 3, 1)} & $n-d-1$ & $n-d-2$ & 0 \\ \multicolumn{1}{l|}{(1, 1, 1, 2)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 1, 2, 1)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 1, 2)} & 0 & 1 & 0 \\ \multicolumn{1}{l|}{(1, 2, 2, 1)} & 0 & 1 & 0 \\ \multicolumn{1}{l|}{(1, 2, 3, 3)} & 0 & 0 & $(n-2)(n-d-1)$ \\ \multicolumn{1}{l|}{(1, 1, 2, 2)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 2, 2)} & 0 & 0 & $n-d-1$ \\ \multicolumn{1}{l|}{(1, 2, 1, 1)} & 0 & 0 & $n-d-1$ \\ \multicolumn{1}{l|}{(1, 1, 1, 1)} & 0 & 0 & 0 \\ \hline \multicolumn{1}{l|}{Total} & $n(n-d-1)$ & $n(n-d-1)$ & $n(n-d-1)$ \end{tabular} \caption{$m_N$} \label{table:mN} \end{table} \begin{table}[h] \centering \begin{tabular}{llll} \multicolumn{1}{l|}{$\mu$} & $m_S(E, \mu)$ & $m_S(N, \mu)$ & $m_S(S, \mu)$ \\ \hline \multicolumn{1}{l|}{(1, 2, 3, 4)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 1, 2, 3)} & $n-2$ & $n-2$ & 0 \\ \multicolumn{1}{l|}{(1, 2, 2, 3)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 1, 3)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 3, 2)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 3, 1)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 1, 1, 2)} & 1 & 1 & 0 \\ \multicolumn{1}{l|}{(1, 1, 2, 1)} & 1 & 1 & 0 \\ \multicolumn{1}{l|}{(1, 2, 1, 2)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 2, 1)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 3, 3)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 1, 2, 2)} & 0 & 0 & $n-1$ \\ \multicolumn{1}{l|}{(1, 2, 2, 2)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 2, 1, 1)} & 0 & 0 & 0 \\ \multicolumn{1}{l|}{(1, 1, 1, 1)} & 0 & 0 & 1 \\ \hline \multicolumn{1}{l|}{Total} & $n$ & $n$ & $n $ \end{tabular} \caption{$m_S$} \label{table:mS} \end{table} Therefore, we have $L(A)_b = \sum_\mu w_\mu (m_E(\tau(b), \mu) + m_N(\tau(b), \mu) + m_S(\tau(b), \mu)) + \beta_{\mathcal{E}(b)}$. Moreover, notice that $\tau(b)$ determines $\mathcal{E}(b)$: if $\tau(b) = E$ or $N$, then $\mathcal{E}(b) = \mathcal{E}(1, 2); $ if $\tau(b) = S$, then $\mathcal{E}(b) = \mathcal{E}(1, 1)$. Hence, we can write $\beta_{\tau(b)}$ instead of $\beta_{\mathcal{E}(b)}$ without loss of generality. Then in particular, this means that $L(A)_b = L(A)_{b'}$ if $\tau(b) = \tau(b')$. Therefore, $L(A)_b = \overline{\xi}_1 \mathds{1}_{b \in E} + \overline{\xi}_2 \mathds{1}_{b \in N} + \overline{\xi}_3 \mathds{1}_{b \in S}$, where $\overline{\xi}_1 = \sum_\mu w_\mu (m_E(E, \mu) + m_N(E, \mu) + m_S(E, \mu)) + \beta_E$, $\overline{\xi}_2 = \sum_\mu w_\mu (m_E(N, \mu) + m_N(N, \mu) + m_S(N, \mu)) + \beta_N$, and $\overline{\xi}_3 = \sum_\mu w_\mu (m_E(S, \mu) + m_N(S, \mu) + m_S(S, \mu)) + \beta_S$. Similarly, $L(A')_b = \overline{\xi'}_1 \mathds{1}_{b \in E'} + \overline{\xi'}_2 \mathds{1}_{b \in N'} + \overline{\xi'}_3 \mathds{1}_{b \in S'}$. But importantly, $\forall$ equivalence class of 4-tuples, $\mu$, and $\forall \lambda_1, \lambda_2 \in \{E, N, S \}, m_{\lambda_1}(\lambda_2, \mu) = m'_{\lambda_1}(\lambda_2, \mu)$, as both of them can be obtained from the same entry of the same table. Therefore, $\overline{\xi}_1 = \overline{\xi'}_1, \overline{\xi}_2 = \overline{\xi'}_2$, $\overline{\xi}_3 = \overline{\xi'}_3$. Finally, let $\xi^*_1 = \sigma(\overline{\xi}_1), \xi^*_2 = \sigma(\overline{\xi}_2)$, and $\xi^*_3 = \sigma(\overline{\xi}_3)$. Then, there is $\sigma(L(A))_b = \xi^*_1 \mathds{1}_{b \in E} + \xi^*_2 \mathds{1}_{b \in N} + \xi^*_3 \mathds{1}_{b \in S}$, and $\sigma(L(A'))_b = \xi^*_1 \mathds{1}_{b \in E'} + \xi^*_2 \mathds{1}_{b \in N'} + \xi^*_3 \mathds{1}_{b \in S'}$, as desired. \end{proof} Since $h$ is an invariant function, $h$ acting on $A^{(L)}$ essentially computes the sum of all the diagonal terms (i.e., for $b \in S$) and the sum of all the off-diagonal terms (i.e., for $b \in E \cup N$) of $A^{(L)}$ separately and then adds the two sums with two weights. If $G, G'$ are regular graphs with the same degree, then $|E| = |E'|, |S| = |S'|$ and $|N| = |N'|$. Therefore, by the lemma, there is $h(A^{(L)}) = h(A'^{(L)})$, and as a consequence $m(h(A^{(L)})) = m(h(A'^{(L)}))$. \section{Universal approximation, graph isomorphism and motif counting} \section{Introduction} Graph structured data naturally occur in many areas of knowledge, including computational biology, chemistry and social sciences. Graph neural networks, in all their forms, yield useful representations of graph data partly because they take into consideration the intrinsic symmetries of graphs, such as invariance and equivariance with respect to a relabeling of the nodes \cite{scarselli2008graph, duvenaud2015convolutional, kipf2016semi, gilmer2017neural, hamilton2017representation, velivckovic2017graph, bronstein2017geometric}. All these different architectures are proposed with different purposes (see \cite{wu2019comprehensive} for a survey and references therein), and a priori it is not obvious how to compare their power. The recent work \cite{xu2018powerful} proposes to study the representation power of GNNs via their performance on graph isomorphism tests. They developed the Graph Isomorphism Networks (GINs) that are as powerful as the one-dimensional Weisfeiler-Lehman (1-WL or just WL) test for graph isomorphism \cite{weisfeiler1968reduction}, and showed that no other neighborhood-aggregating (or message passing) GNN can be more powerful than the 1-WL test. Variants of message passing GNNs include \cite{scarselli2008graph, hamilton2017inductive}. On the other hand, for feed-forward neural networks, many results have been obtained regarding their ability to approximate continuous functions, commonly known as the universal approximation theorems, such as the seminal works of \cite{cybenko1989approximation, hornik1991hornik}. Following this line of work, it is natural to study the expressivity of graph neural networks in terms of function approximation. Since we could argue that many if not most functions on a graph that we are interested in are invariant or equivariant to permutations of the nodes in the graph, GNNs are usually designed to be invariant or equivariant, and therefore the natural question is whether certain classes GNNs can approximate any continuous and invariant or equivariant functions. Recent work \cite{maron2019universality} showed the universal approximation of $G$-invariant networks, constructed based on the linear invariant and equivariant layers studied in \cite{maron2018invariant}, if the order of the tensor involved in the networks can grow as the graph gets larger. Such a dependence on the graph size was been theoretically overcame by the very recent work \cite{keriven2019universal}, though there is no known upper bound on the order of the tensors involved. With potentially very-high-order tensors, these models that are guaranteed of univeral approximation are not quite feasible in practice. The foundational part of this work aims at building the bridge between graph isomorphism testing and invariant function approximation, the two main perspectives for studying the expressive power of graph neural networks. We demonstrate an equivalence between the the ability of a class of GNNs to distinguish between any pairs of non-isomorphic graph and its power of approximating any (continuous) invariant functions, for both the case with finite feature space and the case with continuous feature space. Furthermore, we argue that the concept of sigma-algebras on the space of graphs is a natural description of the power of graph neural networks, allowing us to build a taxonomy of GNNs based on how their respective sigmas-algebras interact. Building on this theoretical framework, we identify an opportunity to increase the expressive power of order-$2$ $G$-invariant networks with computational tractability, by considering a ring of invariant matrices under addition and multiplication. We show that the resulting model, which we refer to as \emph{Ring-GNN}, is able to distinguish between non-isomorphic regular graphs where order-$2$ $G$-invariant networks provably fail. We illustrate these gains numerically in synthetic and real graph classification tasks. Summary of main contributions: \begin{itemize} \item We show the equivalence between graph isomorphism testing and approximation of permutation-invariant functions for analyzing the expressive power of graph neural networks. \item We introduce a language of sigma algebra for studying the representation power of graph neural networks, which unifies both graph isomorphism testing and function approximation, and use this framework to compare the power of some GNNs and other methods. \item We propose Ring-GNN, a tractable extension of order-2 Graph $G$-invariant Networks that uses the ring of matrix addition and multiplication. We show this extension is necessary and sufficient to distinguish Circular Skip Links graphs. \end{itemize} \section{Related work} \paragraph{Graph Neural Networks and graph isomorphism.} Graph isomorphism is a fundamental problem in theoretical computer science. It amounts to deciding, given two graphs $A,B$, whether there exists a permutation $\pi$ such that $\pi A = B\pi$. There exists no known polynomial-time algorithm to solve it, but recently Babai made a breakthrough by showing that it can be solved in quasi-polynomial-time \cite{babai2016graph}. Recently \cite{xu2018powerful} introduced graph isomorphism tests as a characterization of the power of graph neural networks. They show that if a GNN follows a neighborhood aggregation scheme, then it cannot distinguish pairs of non-isomorphic graphs that the 1-WL test fails to distinguish. Therefore this class of GNNs is at most as powerful as the 1-WL test. They further propose the Graph Isomorphism Networks (GINs) based on approximating injective set functions by multi-layer perceptrons (MLPs), which can be as powerful as the 1-WL test. Based on $k$-WL tests \cite{cai1992optimal}, \cite{morris2019higher} proposes $k$-GNN, which can take higher-order interactions among nodes into account. Concurrently to this work, \cite{maron2019provably} proves that order-$k$ invariant graph networks are at least as powerful as the $k$-WL tests, and similarly to us, it and augments order-2 networks with matrix multiplication. They show they achieve at least the power of 3-WL test. \cite{murphy2019relational} proposes relational pooling (RP), an approach that combines \textit{permutation-sensitive} functions under all permutations to obtain a permutation-invariant function. If RP is combined with permutation-sensitive functions that are sufficiently expressive, then it can be shown to be a universal approximator. A combination of RP and GINs is able to distinguish certain non-isomorphic regular graphs which GIN alone would fail on. A drawback of RP is that its full version is intractable computationally, and therefore it needs to be approximated by averaging over randomly sampled permutations, in which case the resulting functions is not guaranteed to be permutation-invariant. \paragraph{Universal approximation of functions with symmetry.} Many works have discussed the function approximation capabilities of neural networks that satisfy certain symmetries. \cite{bloemreddy2019probabilistic} studies the symmetry in neural networks from the perspective of probabilistic symmetry and characterizes the deterministic and stochastic neural networks that satisfy certain symmetry. \cite{ravanbakhsh2017sharing} shows that equivariance of a neural network corresponds to symmetries in its parameter-sharing scheme. \cite{yarotsky2018universal} proposes a neural network architecture with polynomial layers that is able to achieve universal approximation of invariant or equivariant functions. \cite{maron2018invariant} studies the spaces of all invariant and equivariant linear functions, and obtained bases for such spaces. Building upon this work, \cite{maron2019universality} proposes the $G$-invariant network for a symmetry group $G$, which achieves universal approximation of $G$-invariant functions if the maximal tensor order involved in the network to grow as $\frac{n (n-1)}{2}$, but such high-order tensors are prohibitive in practice. Upper bounds on the approximation power of the $G$-invariant networks when the tensor order is limited remains open except for when $G = A_n$ \cite{maron2019universality}. The very recent work \cite{keriven2019universal} extends the result to the equivariant case, although it suffers from the same problem of possibly requiring high-order tensors. Within the computer vision literature, this problem has also been addressed, in particular \cite{herzig2018mapping} proposes an architecture that can potentially express all equivariant functions. To the best our knowledge, this is the first work that shows an explicit connection between the two aforementioned perspectives of studying the representation power of graph neural networks - graph isomorphism testing and universal approximation. Our main theoretical contribution lies in showing an equivalence between them, for both finite and continuous feature space cases, with a natural generalization of the notion of graph isomorphism testing to the latter case. Then we focus on the Graph $G$-invariant network based on \cite{maron2018invariant,maron2019universality}, and showed that when the maximum tensor order is restricted to be 2, then it cannot distinguish between non-isomorphic regular graphs with equal degrees. As a corollary, such networks are not universal. Note that our result shows an upper bound on order 2 $G$-invariant networks, whereas concurrently to us, \cite{maron2019provably} provides a lower bound by relating to $k$-WL tests. Concurrently to \cite{maron2019provably}, we propose a modified version of order-2 graph networks to capture higher-order interactions among nodes without computing tensors of higher-order. \section{Graph isomorphism testing and universal approximation} In this section we show that there exists a very close connection between the universal approximation of permutation-invariant functions by a class of functions, and its ability to perform graph isomorphism tests. We consider graphs with nodes and edges labeled by elements of a compact set $\mathcal{X}\subset \mathbb R$. We represent graphs with $n$ nodes by an $n$ by $n$ matrix $G \in \mathcal{X}^{n \times n}$, where a diagonal term $G_{ii}$ represents the label of the $i$th node, and a non-diagonal $G_{ij}$ represents the label of the edge from the $i$th node to the $j$th node. An undirected graph will then be represented by a symmetric $G$. Thus, we focus on analyzing a collection $\mathcal{C}$ of functions from $\mathcal{X}^{n \times n}$ to $\mathbb{R}$. We are especially interested in collections of \textit{permutation-invariant functions}, defined so that $f(\pi^\intercal G \pi) = f(G)$, for all $G \in \mathcal{X}^{n \times n}$, and all $\pi \in S_n$, where $S_n$ is the permutation group of $n$ elements. For classes of functions, we define the property of being able to discriminate non-isomorphic graphs, which we call \textit{GIso-discriminating}, which as we will see generalizes naturally to the continuous case. \begin{definition} \label{pd} Let $\mathcal{C}$ be a collection of permutation-invariant functions from $\mathcal{X}^{n \times n}$ to $\mathbb{R}$. We say $\mathcal C$ is \textbf{GIso-discriminating} if for all non-isomorphic $G_1, G_2 \in \mathcal X^{n\times n}$ (denoted $G_1 \not\simeq G_2$), there exists a function $ h \in \mathcal{C}$ such that $h(G_1) \neq h(G_2)$. This definition is illustrated by figure 2 in the appendix. \end{definition} \begin{definition} Let $\mathcal{C}$ be a collection of permutation-invariant functions from $\mathcal{X}^{n \times n}$ to $\mathbb{R}$. We say $\mathcal C$ is \textbf{universally approximating} if for all permutation-invariant function $f$ from $\mathcal{X}^{n \times n}$ to $\mathbb{R}$, and for all $\epsilon > 0$, there exists $h_{f,\epsilon} \in \mathcal{C}$ such that $\| f - h_{f,\epsilon} \|_{\infty} := \sup_{G \in \mathcal X^{n\times n}} |f(G) - h(G)| < \epsilon$ \end{definition} \subsection{Finite feature space} As a warm-up we first consider the space of graphs with a finite set of possible features for nodes and edges, $\mathcal X=\{1,\ldots, M\}$. \begin{theorem} \label{UA2PD} Universally approximating classes of functions are also GIso-discriminating. \end{theorem} \vspace{-1.0em} \begin{proof} Given $G_1, G_2 \in \mathcal X^{n\times n}$, we consider the permutation-invariant function $\mathds{1}_{\simeq G_1}:\mathcal X^{n\times n}\to \mathbb R$ such that $\mathds{1}_{\simeq G_1}(G)=1$ if $G$ is isomorphic to $G_1$ and 0 otherwise. Therefore, it can be approximated with $\epsilon = 0.1$ by a function $h \in \mathcal{C}$. Then $h$ is a function that distinguishes $G_1$ from $G_2$, as in Definition~\ref{pd}. Hence $\mathcal{C}$ is GIso-discriminating. \end{proof} To obtain a result on the reverse direction, we first introduce the concept of an augmented collection of functions, which is especially natural when $\mathcal{C}$ is a collection of neural networks. \begin{definition} Given $\mathcal{C}$, a collection of functions from $\mathcal{X}^{n \times n}$ to $\mathbb{R}$, we consider an augmented collection of functions also from $\mathcal{X}^{n \times n}$ to $\mathbb{R}$ consisting of functions that map an input graph $G$ to $\mathcal{NN}([h_1(G), ..., h_d(G)])$ for some finite $d$, where $\mathcal{NN}$ is a feed-forward neural network / multi-layer perceptron, and $h_1, ..., h_d \in \mathcal{C}$. When $\mathcal{NN}$ is restricted to have $L$ layers, we denoted this augmented collection by $\mathcal{C}^{+L}$. In this work, we consider ReLU as the nonlinear activation function in the neural networks. \end{definition} \begin{remark} If $\mathcal{C}_{L_0}$ is the collection of feed-forward neural networks with $L_0$ layers, then $\mathcal{C}_{L_0}^{+L}$ represents the collection of feed-forward neural networks with $L_0 + L$ layers. \end{remark} \begin{remark} If $\mathcal{C}$ is a collection of permutation-invariant functions, so is $\mathcal{C}^{+L}$. \end{remark} \begin{theorem} \label{PD2UAfin} If $\mathcal{C}$ is GIso-discriminating, then $\mathcal{C}^{+2}$ is universal approximating. \end{theorem} The proof is simple and it is a consequence of the following lemmas that we prove in Appendix~\ref{app.universal}. \begin{lemma} \label{lemma1} If $\mathcal{C}$ is GIso-discriminating, then for all $G \in \mathcal{X}^{n \times n}$, there exists a function $\tilde{h}_G \in \mathcal{C}^{+1}$ such that for all $G', \tilde{h}_G(G') = 0$ if and only if $G \simeq G'$. \end{lemma} \begin{lemma} \label{lemma2} Let $\mathcal C$ be a class of permutation-invariant functions from $\mathcal X^{n\times n}$ to $\mathbb R$ satisfying the consequences of Lemma \ref{lemma1}, then $\mathcal{C}^{+1}$ is universally approximating. \end{lemma} \subsection{Extension to the case of continuous (Euclidean) feature space} Graph isomorphism is an inherently discrete problem, whereas universal approximation is usually more interesting when the input space is continuous. With our definition \ref{pd} of \textit{GIso-discriminating}, we can achieve a natural generalization of the above results to the scenarios of continuous input space. All proofs for this section can be found in Appendix~\ref{app.universal}. Let $\mathcal{X}$ be a compact subset of $\mathbb{R}$, and we consider graphs with $n$ nodes represented by $G \in K = \mathcal{X}^{n \times n}$; that is, the node features are $\{G_{ii}\}_{i = 1, \ldots, n}$ and the edge features are $\{G_{ij}\}_{i, j = 1,\ldots, n; i \neq j}$. \begin{theorem}\label{ua2pdinf} If $\mathcal{C}$ is universally approximating, then it is also GIso-discriminating \end{theorem} The essence of the proof is similar to that of Theorem~\ref{UA2PD}. The other direction - showing that pairwise discrimination can lead to universal approximation - is less straightforward. As an intermediate step between, we make the following definition: \begin{definition} \label{locate} Let $\mathcal{C}$ be a class of functions $K \to \mathbb{R}$. We say it is able to \textbf{locate every isomorphism class} if for all $G \in K$ and for all $\epsilon > 0$ there exists $h_G \in \mathcal{C}$ such that: \begin{itemize} \item for all $G' \in K, h_G(G') \geq 0$; \item for all $G' \in K$, if $G' \simeq G$, then $h_G(G') = 0$; and \item there exists $\delta_G > 0$ such that if $h_G < \delta_G$, then $\exists \pi \in S_n$ such that $d(\pi(G'), G) < \epsilon$, where $d$ is the Euclidean distance defined on $\mathbb{R}^{n \times n}$ \end{itemize} \end{definition} \begin{lemma} \label{lemma.C+1} If $\mathcal{C}$, a collection of continuous permutation-invariant functions from $K$ to $\mathbb{R}$, is GIso-discriminating, then $\mathcal{C}^{+1}$ is able to locate every isomorphism class. \end{lemma} Heuristically, we can think of the $h_G$ in the definition above as a ``loss function'' that penalizes the deviation of $G'$ from the equivalence class of $G$. In particular, the second condition says that if the loss value is small enough, then we know that $G'$ has to be close to the equivalence class of $G$. \begin{lemma} \label{lemma.locate.approx} Let $\mathcal{C}$ be a class of permutation-invariant functions $K \to \mathbb{R}$. If $\mathcal{C}$ is able to locate every isomorphism class, then $\mathcal{C}^{+2}$ is universally approximating. \end{lemma} Combining the two lemmas above, we arrive at the following theorem: \begin{theorem} If $\mathcal{C}$, a collection of continuous permutation-invariant functions from $K$ to $\mathbb{R}$, is GIso-discriminating, then $\mathcal{C}^{+3}$ is universaly approximating. \end{theorem} \section{A framework of representation power based on sigma-algebra} \label{sec.sigma} \subsection{Introducing sigma-algebra to this context} Let $K = \mathcal{X}^{n \times n}$ be a finite input space. Let $Q_K:=K/_{\simeq}$ be the set of isomorphism classes under the equivalence relation of graph isomorphism. That is, for all $\tau \in Q_K, \tau = \{ \pi^\intercal G \pi : \pi \in \Gamma_n \}$ for some $G \in K$. Intuitively, a maximally expressive collection of permutation-invariant functions, $\mathcal{C}$, will allow us to know exactly which isomorphism class $\tau$ a given graph $G$ belongs to, by looking at the outputs of certain functions in the collection applied to $G$. Heuristically, we can consider each function in $\mathcal{C}$ as a ``measurement'', which partitions that graph space $K$ according to the function value at each point. If $\mathcal{C}$ is powerful enough, then as a collection it will partition $K$ to be as fine as $Q_K$. If not, it is going to be coarser than $Q_K$. These intuitions motivate us to introduce the language of sigma-algebra. Recall that an algebra on a set $K$ is a collection of subsets of $K$ that includes $K$ itself, is closed under complement, and is closed under finite union. Because $K$ is finite, we have that an algebra on $K$ is also a sigma-algebra on $K$, where a sigma-algebra further satisfies the condition of being closed under countable unions. Since $Q_K$ is a set of (non-intersecting) subsets of $K$, we can obtain the algebra generated by $Q_K$, defined as the smallest algebra that contains $Q_K$, and use $\sigma(Q_K)$ to denote the algebra (and sigma-algebra) generated by $Q_K$. \begin{observation} If $f : \mathcal{X}^{n \times n} \to \mathbb{R}$ is a permutation-invariant function, then $f$ is measurable with respect to $\sigma(Q_K)$, and we denote this by $f \in \mathcal{M}[\sigma(Q_K)]$ \end{observation} Now consider a class of functions $\mathcal{C}$ that is permutation-invariant. Then for all $f \in \mathcal{C}, f \in \mathcal{M}[\sigma(Q_K)]$. We define the sigma-algebra generated by $f$ as the set of all the pre-images of Borel sets on $\mathbb{R}$ under $f$, and denote it by $\sigma(f)$. It is the smallest sigma-algebra on $K$ that makes $f$ measurable. For a class of functions $\mathcal{C}$, $\sigma(\mathcal{C})$ is defined as the smallest sigma-algebra on $K$ that makes all functions in $\mathcal{C}$ measurable. Because here we assume $K$ is finite, it does not matter whether $\mathcal{C}$ is a countable collection. \subsection{Reformulating graph isomorphism testing and universal approximation with sigma-algebra} \label{sec.reformulating} We restrict our attention to the case of finite feature space. Given a graph $G \in \mathcal{X}^{n \times n}$, we use $\mathcal{E}(G)$ to denote its isomorphism class, $\{ G' \in \mathcal{X}^{n \times n}: G' \simeq G \}$. The following results are proven in Section~\ref{sec.proofs.reformulating} \begin{theorem}\label{teo5} If $\mathcal{C}$ is a class of permutation-invariant functions on $\mathcal{X}^{n \times n}$ and $\mathcal{C}$ is GIso-discriminating, then $\sigma(\mathcal{C}) = \sigma(Q_K)$ \end{theorem} Together with Theorem \ref{UA2PD}, the following is an immediate consequence: \begin{corollary} If $\mathcal{C}$ is a class of permutation-invariant functions on $\mathcal{X}^{n \times n}$ and $\mathcal{C}$ achieves universal approximation, then $\sigma(\mathcal{C}) = \sigma(Q_K)$. \end{corollary} \begin{theorem} \label{teo6} Let be $\mathcal{C}$ a class of permutation-invariant functions on $\mathcal{X}^{n \times n}$ with $\sigma(\mathcal{C}) = \sigma(Q_K)$. Then $\mathcal{C}$ is GIso-discriminating. \end{theorem} Thus, this sigma-algebra language is a natural notion for characterizing the power of graph neural networks, because as shown above, generating the finest sigma-algebra $\sigma(Q_K)$ is equivalent to being GIso-discriminating, and therefore to universal approximation. Moreover, when $\mathcal{C}$ is not GIso-discriminating or universal, we can evaluate its representation power by studying $\sigma(\mathcal{C})$, which gives a measure for comparing the power of different GNN families. Given two classes of functions $\mathcal{C}_1, \mathcal{C}_2$, there is $\sigma(\mathcal{C}_1) \subseteq \sigma(\mathcal{C}_2)$ if and only if $\mathcal{M}[\sigma(\mathcal{C}_1)] \subseteq \mathcal{M}[\sigma(\mathcal{C}_2)]$ if and only if $\mathcal{C}_1$ is less powerful than $\mathcal{C}_2$ in terms of representation power. In Appendix \ref{app.comparison} we use this notion to compare the expressive power of different families of GNNs as well as other algorithms like 1-WL, linear programming and semidefinite programming in terms of their ability to distinguish non-isomorphic graphs. We summarize our findings in Figure 1. \begin{figure}[ht] \label{diagram_main_text} \small \centering \begin{tikzcd} & \text{sGNN}(I,A) \arrow[d, hook] \arrow[ddr, hook]& \\ & LP\equiv 1-WL \equiv GIN \arrow[d, hook] \arrow[dl, hook] & \\ \text{SDP}\arrow[dd,hook]& \text{MPNN}^* \arrow[d,hook] & \text{sGNN}(I,D,A,\{\min\{A^t,1\}\}_{t=1}^T) \arrow[ddl,hook]\\ & \text{order 2 $G$-invariant networks}^* \arrow[d,hook] & \text{spectral methods}\arrow[dl,hook] \\ \text{SoS hierarchy}& \text{Ring-GNN}& \end{tikzcd} \caption{\small Relative comparison of function classes in terms of their ability to solve graph isomorphism. \newline $^*$Note that, on one hand GIN is defined by \cite{xu2018powerful} as a form of message passing neural network justifying the inclusion GIN $\hookrightarrow$ MPNN. On the other hand \cite{maron2018invariant} shows that message passing neural networks can be expressed as a modified form of order 2 $G$-invariant networks (which may not coincide with the definition we consider in this paper). Therefore the inclusion GIN $\hookrightarrow$ order 2 $G$-invariant networks has yet to be established rigorously. \vspace{-15pt} } \label{fig.diagram} \end{figure} \section{Ring-GNN: a GNN defined on the ring of equivariant functions} We now investigate the $G$-invariant network framework proposed in \cite{maron2019universality} (see Appendix~\ref{app.Ginvariant} for its definition and a description of an adapted version that works on graph-structured inputs, which we call the \textit{Graph $G$-invariant Networks}). The architecture of $G$-invariant networks is built by interleaving compositions of equivariant linear layers between tensors of potentially different orders and point-wise nonlinear activation functions. It is a powerful framework that can achieve universal approximation if the order of the tensor can grow as $\frac{n(n-1)}{2}$, where $n$ is the number of nodes in the graph, but less is known about its approximation power when the tensor order is restricted. One particularly interesting subclass of $G$-invariant networks is the ones with maximum tensor order 2, because \cite{maron2018invariant} shows that it can approximate any Message Passing Neural Network \cite{gilmer2017neural}. Moreover, it is both mathematically cumbersome and computationally expensive to include equivariant linear layers involving tensors with order higher than 2. Our following result shows that the order-2 Graph $G$-invariant Networks subclass of functions is quite restrictive. The proof is given in Appendix \ref{app.Ginvariant}. \begin{theorem} \label{prop.Ginvariant} Order-2 Graph $G$-invariant Networks cannot distinguish between non-isomorphic regular graphs with the same degree. \end{theorem} Motivated by this limitation, we propose a GNN architecture that extends the family of order-2 Graph $G$-invariant Networks without going into higher order tensors. In particular, we want the new family to include GNNs that can distinguish some pairs of non-isomorphic regular graphs with the same degree. For instance, take the pair of Circular Skip Link graphs $G_{8, 2}$ and $G_{8,3}$, illustrated in Figure \ref{cslfig}. Roughly speaking, if all the nodes in both graphs have the same node feature, then because they all have the same degree, the updates of node states in both graph neural networks based on neighborhood aggregation and the WL test will fail to distinguish the nodes. However, the \textit{power graphs}\footnote{If $A$ is the adjacency matrix of a graph, its power graph has adjacency matrix $\min(A^2, 1)$. The matrix $\min(A^2, 1)$ has been used in \cite{chen2019cdsbm} in graph neural networks for community detection and in \cite{nowak2017note} for the quadratic assignment problem.} of $G_{8, 2}$ and $G_{8,3}$ have different degrees. Another important example comes from spectral methods that operate on \emph{normalized} operators, such as the normalized Laplacian $\Delta = I - D^{-1/2} A D^{-1/2}$, where $D$ is the diagonal degree operator. Such normalization preserves the permutation symmetries and in many clustering applications leads to dramatic improvements \cite{von2007tutorial}. This motivates us to consider a polynomial ring generated by the matrices that are the outputs of permutation-equivariant linear layers, rather than just the linear space of those outputs. Together with point-wise nonlinear activation functions such as ReLU, power graph adjacency matrices like $\min(A^2, 1)$ can be expressed with suitable choices of parameters. We call the resulting architecture the \textit{Ring-GNN} \footnote{We call it Ring-GNN since the main object we consider is the ring of matrices, but technically we can express an associative algebra since our model includes scalar multiplications.}. \begin{definition}[Ring-GNN] Given a graph in $n$ nodes with both node and edge features in $\mathbb R^d$, we represent it with a matrix $A\in \mathbb R^{n\times n\times d}$. \cite{maron2018invariant} shows that all linear equivariant layers from $\mathbb{R}^{n \times n}$ to $\mathbb{R}^{n \times n}$ can be expressed as $L_\theta(A)=\sum_{i=1}^{15} \theta_i L_i(A) + \sum_{i=16}^{17} \theta_i \overline{L}_i$, where the $\{L_i\}_{i = 1, ..., 15}$ are the 15 basis functions of all linear equivariant functions from $\mathbb{R}^{n \times n}$ to $\mathbb{R}^{n \times n}$, $\overline{L}_{16}$ and $\overline{L}_{17}$ are the basis for the bias terms, and $\theta \in \mathbb{R}^{17}$ are the parameters that determine $L$. Generalizing to an equivariant linear layer from $\mathbb{R}^{n \times n \times d}$ to $\mathbb{R}^{n \times n \times d'}$, we set $L_{\theta}(A)_{\cdot, \cdot, k'} = \sum_{k=1}^d \sum_{i=1}^{15} \theta_{k, k', i} L_i(A_{\cdot, \cdot, i}) + \sum_{i=16}^{17} \theta_{k, k', i} \overline{L}_i$, with $\theta \in \mathbb{R}^{d \times d' \times 17}$. With this formulation, we now define a Ring-GNN with $T$ layers. First, set $A^{(0)}=A$. In the $t^{th}$ layer, let \begin{eqnarray*} B_1^{(t)}&=& \rho(L_{\alpha^{(t)}}(A^{(t)})) \\ B_2^{(t)}&=& \rho(L_{\beta^{(t)}}(A^{(t)})\cdot L_{\gamma^{(t)}}(A^{(t)})) \\ A^{(t+1)}&=& k_1^{(t)} B_1^{(t)} + k_2^{(t)} B_2^{(t)} \end{eqnarray*} where $k_1^{(t)}, k_2^{(t)} \in \mathbb{R}$, $\alpha^{(t)}, \beta^{(t)}, \gamma^{(t)} \in \mathbb{ R}^{d^{(t)} \times {d'}^{(t)} \times 17}$ are learnable parameters. If a scalar output is desired, then in the general form, we set the output to be $\theta_S \sum_{i,j} A_{ij}^{(T)} + \theta_D \sum_{i,i} A_{ii}^{(T)} + \sum_{i} \theta_{i} \lambda_{i}(A^{(T)})$, where $\theta_S, \theta_D, \theta_1, \ldots, \theta_n \in \mathbb{R}$ are trainable parameters, and $\lambda_{i}(A^{(T)})$ is the $i$-th eigenvalue of $A^{(L)}$. \end{definition} Note that each layer is equivariant, and the map from $A$ to the final scalar output is invariant. A Ring-GNN can reduce to an order-2 Graph $G$-invariant Network if $k_2^{(t)} = 0$. With $J+1$ layers and suitable choices of the parameters, it is possible to obtain $\min(A^{2^{J}}, 1)$ in the $(J+1)^{th}$ layer. Therefore, we expect it to succeed in distinguishing certain pairs of regular graphs that order-2 Graph $G$-invariant Networks fail on, such as the Circular Skip Link graphs. Indeed, this is verified in the synthetic experiment presented in the next section. The normalized Laplacian can also be obtained, since the degree matrix can be inverted by taking the reciprocal on the diagonal, and then entry-wise inversion and square root on the diagonal can be approximated by MLPs. The terms in the output layer involving eigenvalues are optional, depending on the task. For example, in community detection spectral information is commonly used \cite{krzakala2013spectral}. We could also take a fixed number of eigenvalues instead of the full spectrum. In the experiments, Ring-GNN-SVD includes the eigenvalue terms while Ring-GNN does not, as explained in appendix \ref{archi}. Computationally, the complexity of running the forward model grows as $O(n^3)$, dominated by matrix multiplications and possibly singular value decomposition for computing the eigenvalues. We note also that Ring-GNN can be augmented with matrix inverses or more generally with functional calculus on the spectrum of any of the intermediate representations \footnote{When $A=A^{(0)}$ is an undirected graph, one easily verifies that $A^{(t)}$ contains only symmetric matrices for each $t$.} while keeping $O(n^3)$ computational complexity. Finally, note that a Graph $G$-invariant Network with maximal tensor order $d$ will have complexity at least $O(n^d)$. Therefore, the Ring-GNN explores higher-order interactions in the graph that order-2 Graph $G$-invariant Networks neglects while remaining computationally tractable. \begin{figure} \label{cslfig} \centering \includegraphics[width=0.15\textwidth,trim={6cm 4.8cm 20cm 7.8cm},clip]{skl2black} \includegraphics[width=0.15\textwidth,trim={6cm 4.8cm 20cm 7.8cm},clip]{skl3black} \caption{The Circular Skip Link graphs $G_{n,k}$ are undirected graphs in $n$ nodes $q_0,\ldots, q_{n-1}$ so that $(i,j)\in E$ if and only if $|i-j|\equiv 1 \text{ or } k \pmod n$. In this figure we depict (left) $G_{8,2}$ and (right) $G_{8,3}$. It is very easy to check that $G_{n,k}$ and $G_{n',k'}$ are not isomorphic unless $n=n'$ and $k\equiv \pm k' \pmod n$. Both 1-WL and $G$-invariant networks fail to distinguish them. \vspace{-10pt}} \label{fig.skiplength} \end{figure} \section{Experiments} \label{experiments} The different models and the detailed setup of the experiments are discussed in Appendix \ref{archi}. \subsection{Classifying Circular Skip Links (CSL) graphs} \label{cslexp} The following experiment on synthetic data demonstrates the connection between function fitting and graph isomorphism testing. The Circular Skip Links graphs are undirected regular graphs with node degree 4 \cite{murphy2019relational}, as illustrated in Figure \ref{cslfig}. Note that two CSL graphs $G_{n,k}$ and $G_{n',k'}$ are not isomorphic unless $n=n'$ and $k\equiv \pm k' \pmod n$. In the experiment, which has the same setup as in \cite{murphy2019relational}, we fix $n=41$, and set $k \in \{2, 3, 4, 5, 6, 9, 11, 12, 13, 16 \}$, and each $k$ corresponds to a distinct isomorphism class. The task is then to classify a graph $G_{n, k}$ by its skip length $k$. Note that since the 10 classes have the same size, a naive uniform classifier would obtain $0.1$ accuracy. As we see from Table \ref{table.synthetic}, both GIN and $G$-invariant network with tensor order 2 do not outperform the naive classifier. Their failure in this task is unsurprising: WL tests are proved to fall short of distinguishing such pairs of non-isomorphic regular graphs \cite{cai1992optimal}, and hence neither can GIN \cite{xu2018powerful}; by the theoretical results from the previous section, order-2 Graph $G$-invariant network are unable to distinguish them either. Therefore, their failure as graph isomorphism tests is consistent with their failure in this classification task, which can be understood as trying to approximate the function that maps the graph to their class labels. It should be noted that, since graph isomorphism tests are not entirely well-posed as classfication tasks, the performance of GNN models could vary due to randomness. But the fact that Ring-GNNs achieve a relatively high maximum accuracy (compared to RP for example) demonstrates that as a class of GNNs it is rich enough to contain functions that distinguish the CSL graphs to a large extent. \begin{table}[ht] \centering \begin{tabular}{l|lll||ll|ll} \hline & \multicolumn{3}{|c||}{Circular Skip Links} & \multicolumn{2}{c|}{IMDBB} & \multicolumn{2}{c}{IMDBM} \\ GNN architecture & max & min & std & mean & std & mean & std \\ \hline \hline RP-GIN $\dagger$ & 53.3 & 10 & 12.9 & - & - & - & - \\ GIN $\dagger$ & 10 & 10 & 0 & 75.1 & 5.1 & 52.3 & 2.8 \\ Order 2 G-invariant $\dagger$ & 10 & 10 & 0 & 71.27 & 4.5 & 48.55 & 3.9 \\ sGNN-5 & 80 & 80 & 0 & 72.8 & 3.8 & 49.4 & 3.2 \\ sGNN-2 & 30 & 30 & 0 & 73.1 & 5.2 & 49.0 & 2.1 \\ sGNN-1 & 10 & 10 & 0 & 72.7 & 4.9 & 49.0 & 2.1 \\ LGNN \cite{chen2019cdsbm} & 30 & 30 & 0 & 74.1 & 4.6 & 50.9 & 3.0 \\ Ring-GNN & 80 & 10 & 15.7 & 73.0 & 5.4 & 48.2 & 2.7 \\ Ring-GNN-SVD & 100 & 100 & 0 & 73.1 & 3.3 & 49.6 & 3.0 \\ \hline \end{tabular} \vspace{5pt} \caption{\textbf{(left)} Accuracy of different GNNs at classifying CSL (see Section \ref{cslexp}). We report the best performance and worst performance among 10 experiments. \textbf{(right)} Accuracy of different GNNs at classifying real datasets (see Section \ref{cslexp}). We report the best performance among all epochs on a 10-fold cross validation dataset, as was done in \cite{xu2018powerful}. $\dagger$: Reported performance by \cite{murphy2019relational}, \cite{xu2018powerful} and \cite{maron2018invariant}. } \vspace{-1.5em} \label{table.synthetic} \end{table} \subsection{IMDB datasets} \label{sec.imbdb} We use the two IMDB datasets (IMDBBINARY, IMDBMULTI) to test different models in real-world scenarios. Since our focus is on distinguishing graph structures, these datasets are suitable as they do not contain node features, and hence the adjacency matrix contains all the input data. IMDBBINARY dataset has 1000 graphs, with average number of nodes 19.8 and 2 classes. The dataset is randomly partitioned into 900/100 for training/validation. IMDBMULTI dataset has 1500 graphs, with average number of nodes 13.0 and 3 classes. The dataset is randomly partitioned into 1350/150 for training/validation. All models are evaluated via 10-fold cross validation and best accuracy is calculated through averaging across folds followed by maximizing along epochs~\cite{xu2018powerful}. Importantly, the architecture hyper-parameter of Ring-GNN we use is close to that provided in \cite{maron2018invariant} to show that order-2 $G$-invariant Network is included in model family we propose. The results show that Ring-GNN models achieve higher performance than Order-2 G-invariant networks in both datasets. Admittedly its accuracy does not reach that of the state-of-the-art. However, the main goal of this part of our work is not necessarily to invent the best-performing GNN through hyperparameter optimization, but rather to propose Ring-GNN as an augmented version of order-2 Graph $G$-invariant Networks and show experimental results that support the theory. \input{conclusions.tex} \bibliographystyle{plain}
{ "timestamp": "2019-05-30T02:21:13", "yymm": "1905", "arxiv_id": "1905.12560", "language": "en", "url": "https://arxiv.org/abs/1905.12560", "abstract": "Graph Neural Networks (GNNs) have achieved much success on graph-structured data. In light of this, there have been increasing interests in studying their expressive power. One line of work studies the capability of GNNs to approximate permutation-invariant functions on graphs, and another focuses on the their power as tests for graph isomorphism. Our work connects these two perspectives and proves their equivalence. We further develop a framework of the expressive power of GNNs that incorporates both of these viewpoints using the language of sigma-algebra, through which we compare the expressive power of different types of GNNs together with other graph isomorphism tests. In particular, we prove that the second-order Invariant Graph Network fails to distinguish non-isomorphic regular graphs with the same degree. Then, we extend it to a new architecture, Ring-GNN, which succeeds in distinguishing these graphs and achieves good performances on real-world datasets.", "subjects": "Machine Learning (cs.LG); Machine Learning (stat.ML)", "title": "On the equivalence between graph isomorphism testing and function approximation with GNNs", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9693241947446616, "lm_q2_score": 0.8267117962054048, "lm_q1q2_score": 0.8013517461427168 }
https://arxiv.org/abs/1711.08791
Cantor set arithmetic
Every element $u$ of $[0,1]$ can be written in the form $u=x^2y$, where $x,y$ are elements of the Cantor set $C$. In particular, every real number between zero and one is the product of three elements of the Cantor set. On the other hand the set of real numbers $v$ that can be written in the form $v=xy$ with $x$ and $y$ in $C$ is a closed subset of $[0,1]$ with Lebesgue measure strictly between $\tfrac{17}{21}$ and $\tfrac89$. We also describe the structure of the quotient of $C$ by itself, that is, the image of $C\times (C \setminus \{0\})$ under the function $f(x,y) = x/y$.
\section{Introduction.} One of the first exotic mathematical objects encountered by the post-calculus student is the Cantor set \begin{equation}\label{Steinhaus} C = \left\{\sum_{k= 1}^\infty \alpha_k3^{-k}, \alpha_k \in \{0,2\}\right\}. \end{equation} (See {Section 2} for several equivalent definitions of $C$.) One of its most beautiful properties is that \begin{equation}\label{C+C} C+C := \{x+y: x,y \in C\} \end{equation} is equal to $[0,2]$. (The whole interval is produced by adding dust to itself!) The first published proof of \eqref{C+C} was by Hugo Steinhaus \cite{steinhaus:sums} in 1917. The result was later rediscovered by John Randolph in 1940 \cite{ran:cantor}. We remind the reader of the beautiful constructive proof of \eqref{C+C}. It is enough to prove the containment $C+C \supset [0,2]$. Given $u \in [0,2]$, consider the ternary representation for $u/2$: \begin{equation} \frac u2= \sum_{k= 1}^{\infty} \frac{\epsilon_k}{3^k}, \quad \epsilon_k \in \{0,1,2\}. \end{equation} Define pairs $(\alpha_k,\beta_k)$ to be $(0,0), (2,0), (2,2)$ according to whether $\epsilon_k = 0, 1, 2$, respectively, and define elements $x, y \in C$ by \begin{equation} x = \sum_{k= 1}^{\infty} \frac{\alpha_k}{3^k}, \qquad y = \sum_{k= 1}^{\infty} \frac{\beta_k}{3^k}. \end{equation} Since $\alpha_k + \beta_k = 2\epsilon_k$, $x+y = 2\cdot\frac u2 = u$. While presenting this proof in a class, one of the authors (BR) wondered what would happen if addition were replaced by the other arithmetic operations. Another author (JT) immediately pointed out that subtraction is easy, because of a symmetry of $C$: \begin{equation*} x = \sum_{k=1}^\infty \frac{\epsilon_k}{3^k} \in C \iff 1-x = \sum_{k=1}^\infty \frac{2-\epsilon_k}{3^k} \in C. \end{equation*} Thus \begin{equation*} C-C := \{x-y: x,y \in C\} = \{x-(1-z): x,z \in C\} = C+C -1 = [-1,1]. \end{equation*} More generally, to understand the structure of linear combinations $aC+bC$, $a,b\in{\mathbb R}$, it suffices to consider the case $a=1$ and $0\le b\le 1$. (If $a>b>0$, then $aC+bC=a(C+(b/a)C)$; the remaining cases are left to the reader.) The precise structure of the linear combination set $$ aC+bC $$ was obtained by Paw{\l}owicz \cite{paw:cantor}, who extended an earlier result by Utz \cite{utz:cantor}, published in 1951. Utz's result states that \begin{equation}\label{eq:utz} C + b C = [0,1+b] \qquad \mbox{for every $\tfrac13 \le b \le 3$.} \end{equation} Multiplication is trickier. The inclusion $C \subset [0,\frac13] \cup [\frac23, 1]$ implies that any element of the interval $(\frac 13,\frac49)$ cannot be written as a product of two elements from $C$. { Thus the measure of the product of $C$ with itself is at most $\tfrac89$.} This paper grew out of a study of multiplication on $C$. In this paper we will prove the following results. \begin{theorem} \label{maintheorem} \ \begin{enumerate} \item Every $u \in [0,1]$ can be written as $u = x^2y$ for some $x,y \in C$. \item The set of quotients from $C$ can be described as follows: \begin{equation}\label{E:quot} \left\{ \frac xy: x, y \in C, y\neq 0\right\} = \bigcup_{m = -\infty}^\infty \left[\frac 23 \cdot 3^m , \frac 32 \cdot 3^m\right]. \end{equation} \item The set $\{xy:x,y \in C\}$ is a closed set with Lebesgue measure { strictly greater than} $\frac{17}{21}$. \end{enumerate} \end{theorem} In particular, part (1) of Theorem \ref{maintheorem} implies that every real number in the interval $[0,1]$ is the product of three elements of $C$. In words, \eqref{E:quot} says that each positive real number is a quotient of two elements of the Cantor set if and only if either the left-most nonzero digit in the ternary representation of $u$ is ``2,'' or the left-most nonzero digit is ``1,'' but the first subsequent non-``1'' digit is ``0,'' not ``2.'' This paper is organized as follows. We begin in section \ref{sec:tools} with several different descriptions of the Cantor set, and then the key tools, all of which are accessible to { students in} a good undergraduate analysis class. As a warmup, we use these tools to give a short proof of Utz's result \eqref{eq:utz} in Section \ref{subsec:plus-and-minus}. Sections \ref{subsec:times}, \ref{subsec:divide} and \ref{subsec:times-again} are devoted to the proofs of {parts (1), (2) and (3) of} Theorem~\ref{maintheorem}, respectively. Sprinkled throughout are relevant open questions. This article began as a standard research paper, but we realized that many of our main results might be of interest to a wider audience. In Section \ref{sec:final-remarks}, we discuss some of our other results, which will be published elsewhere, with different combinations of co-authors. \section{Tools.}\label{sec:tools} We begin by recalling several equivalent and well-known definitions of the Cantor set. See Fleron \cite{fl:cantor} and the references within for an excellent overview of the history of the Cantor set and the context in which several of these definitions first arose. The standard ternary representation of a real number $x$ in $[0,1]$ is \begin{equation}\label{E:ternrep} x = \sum_{k=1}^{\infty} \frac {\alpha_k(x)}{3^k},\qquad \alpha_k(x) \in \{0,1,2\}. \end{equation} This representation is unique, except for the {\it ternary rationals}, $\{ \frac m{3^n}, m,n \in \mathbb N\}$, which have two ternary representations. Supposing $\alpha_n > 0$ and $m \neq 0 \mod 3$, so that $\alpha_n \in \{1,2\}$ below, we have \begin{equation}\label{E:ternexc} \begin{split} \frac m{3^n} &= \sum_{k=1}^{n-1} \frac {\alpha_k}{3^k} + \frac{\alpha_n}{3^n} + \sum_{k=n+1}^{\infty} \frac {0}{3^k}\\ &= \sum_{k=1}^{n-1} \frac {\alpha_k}{3^k} + \frac{\alpha_n-1}{3^n} + \sum_{k=n+1}^{\infty} \frac {2}{3^k}. \end{split} \end{equation} The {\it Cantor set} $C$ consists those $x \in [0,1]$ { admitting} a ternary representation as in \eqref{E:ternrep} with $\alpha_k(x) \in \{0,2\}$ for all $k$. Note that $C$ also contains those ternary rationals as in \eqref{E:ternexc} whose final digit is ``1.'' These may be transformed as above into a representation in which $\alpha_n(x) = 0$ and $\alpha_k(x) = 2$ for $k > n$. As noted earlier, $x \in C$ if and only if $1-x \in C$. Further, if $k \in \mathbb N$, then \begin{equation} x \in C \implies 3^{-k}x \in C. \end{equation} This definition arises in dynamical systems, as the Cantor set $C$ can be viewed as an invariant set for the map $x \mapsto 3x \bmod 1$, or equivalently, as the image of an invariant set $C'$ for the one-sided shift map $\sigma$ acting on $\Omega = \{0,1,2\}^{\mathbb N}$. Given a sequence $\omega = (\omega_n)_{n=1}^\infty = (\omega_1,\omega_2,\ldots)$ in $\Omega$, $$ \sigma\omega = (\omega_2,\omega_3,\ldots)\,. $$ Letting $C' = \{0,2\}^{\mathbb N} \subset \Omega$, we realize the Cantor set $C$ as the image of $C'$ under the coding map $T:\Omega \to [0,1]$ given by $$ T(\omega) = \sum_{n=1}^\infty \omega_n 3^{-n}. $$ We now present the usual ``middle-third'' definition of the Cantor set: Define $C_n = \{x: \alpha_k(x) \in \{0,2\}, 1 \le k \le n\}$, which is a union of $2^n$ closed intervals of length $3^{-n}$, written as \begin{equation} C_n = \bigcup_{i=1}^{2^n} I_{n,i}. \end{equation} The left-hand endpoints of the $I_{n,i}$'s comprise the set \begin{equation} \left \{\sum_{k=1}^{n} \frac{\epsilon_k}{3^k}: \epsilon_k \in \{0,2\}\right\}. \end{equation} The right-hand endpoints have ``1'' as their final nonzero ternary digit when written as a finite ternary expansion. The more direct definition of $C$ is as a nested intersection of closed sets: \begin{equation} C = \bigcap_{n=1}^\infty C_n;\qquad C_1 \supset C_2 \supset C_3 \supset \cdots. \end{equation} This definition is standard in fractal geometry, where the Cantor set $C$ is seen as the invariant set for the pair of contractive linear mappings $f_1(x) = \tfrac13 x$ and $f_2(x) = \tfrac13 x + \tfrac23$ acting on the real line. That is, $C$ is the unique nonempty compact set that is fully invariant under $f_1$ and $f_2$: $$ C = f_1(C) \cup f_2(C). $$ Observe that each ``parent'' interval $I_{n,i} = [a,a+\frac 1{3^n}]$ in $C_n$ has two ``child'' intervals \begin{equation} I_{n+1,2i-1} = \left[ a,a+\tfrac 1{3^{n+1}}\right],\qquad I_{n+1,2i} = \left[a+\tfrac2{3^{n+1}},a+\tfrac3{3^{n+1}}\right] \end{equation} in $C_{n+1}$, and $C_{n+1}$ is the union of all children intervals whose parents are in $C_n$. It is useful to introduce the following notation to represent the omission of the middle third: \begin{equation} I = [a,a+3t] \implies \ddot I = [a,a+t] \cup [a+2t,a+3t]. \end{equation} Using this notation, \begin{equation} C_{n+1} = \bigcup_{i=1}^{2^{n+1}} I_{n+1,i} = \bigcup_{i=1}^{2^n} \ddot I_{n,i}. \end{equation} It will also be useful, for studying products and quotients, to give a third definition. Let $\widetilde C$ = $C \cap [1/2,1] = C \cap [2/3,1] = \frac 23 + \frac 13\cdot C$. Then by examining the smallest $k$ for which $\epsilon_k = 2$, we see that \begin{equation} C = \{0\} \cup \bigcup_{k=0}^{\infty} 3^{-k}\widetilde C. \end{equation} Similarly, let $\widetilde C_n$ = $C_n \cap [1/2,1]$, consisting of $2^{n-1}$ closed intervals of length $3^{-n}$: \begin{equation} \widetilde C_n = \bigcup_{i=2^{n-1}+1}^{2^n} I_{n,i}. \end{equation} Then \begin{equation} \widetilde C = \bigcap_{n=1}^\infty \widetilde C_n. \end{equation} And, analogously, \begin{equation} \widetilde C_{n+1} = \bigcup_{i=2^{n}+1}^{2^{n+1}} I_{n+1,i} = \bigcup_{i=2^{n-1}+1}^{2^n} \ddot I_{n,i}. \end{equation} By looking at the left-hand endpoints, we see that each interval $I_{n,i} \neq [0,\frac 1{3^n}]$ can be written as $\frac 1{3^{n-k}} I_{k,j}$ for some $I_{k,j} \in \widetilde C_k$; hence \begin{equation} \begin{gathered} C_n = \left[0,\frac 1{3^n}\right] \cup \bigcup_{k=1}^{n} \frac 1{3^{n-k}} \widetilde C_k. \end{gathered} \end{equation} The keys to our method lie in two lemmas which might appear on a first serious analysis exam. (Please do not send such exams to the authors!) \begin{lemma}\label{lemma1} Suppose $\{K_i\} \subset \mathbb R$ are nonempty compact sets, $K_1 \supseteq K_2 \supseteq K_3 \supseteq \cdots$, and $K = \cap K_i$. \noindent (i) If $(x_j) \to x$, $x_j \in K_j$, then $x \in K$. \noindent (ii) If $F:\mathbb R^m \to \mathbb R$ is continuous, then $F(K^{m}) = \cap F(K_i^{m})$. \end{lemma} \begin{proof} (i). If $x \notin K$, then $x \notin K_r$ for some $r$. Since $K_r^c$ is open, there exists $\epsilon > 0$ so that $(x - \epsilon, x+ \epsilon) \subseteq K_r^c \subseteq K_{r+1}^c \cdots$ and hence $|x_j - x| \ge \epsilon$ for $j \ge r$, a contradiction to $x_j \to x$. (ii). Since $K \subseteq K_i$, { we have} $F(K^{m}) \subseteq \cap F(K_i^{m})$. Conversely, suppose $u \in \cap F(K_i^{m})$. We need to find $x \in K^{m}$ such that $F(x) = u$. For each $i$, choose $x_i = (x_{i,1}, \dots, x_{i,m}) \in K_i^m$ so that $F(x_i) = u$. Since $K_1^{m}$ is compact, the Bolzano--Weierstrass theorem implies { that} the sequence $(x_i)$ has a convergent subsequence $x_{r_j} = (x_{r_j,1}, \dots x_{r_j,m}) \to y = (y_1,\dots, y_m)$. Applying (i) to the subsequence $K_{r_1} \supseteq K_{r_2} \supseteq K_{r_3} \supseteq \cdots$, we see that each $y_k \in K$ and since $F$ is continuous, $F(y) = u$, as desired. \end{proof} If we perform the middle-third construction with an initial interval of $[a,b]$, it is easy to see that the limiting object is a translate of the Cantor set, specifically $C_{a,b}:= a +(b-a)C$. \begin{lemma}\label{lemma2} Suppose $F: \mathbb R^m \to \mathbb R$ is continuous, and suppose that for every choice of disjoint or identical subintervals $I_k \subset [a,b]$ of common length, \begin{equation} F(I_1,\dots,I_m) = F(\ddot I_1,\dots, \ddot I_m). \end{equation} Then $F(C_{a,b}^{m}) = F([a,b]^{m})$. \end{lemma} \begin{proof} We prove the result for $[a,b] = [0,1]$; the result follows generally by composing $F$ with an appropriate linear function. Let \begin{equation} C_k = \bigcup_{j=1}^{2^k} I_{k,j}, \end{equation} where each interval $I_{k,j}$ has length $3^{-k}$. It follows that \begin{equation} F(C_k^{m}) = \bigcup_{1\le j_1,\dots,j_m\le 2^k} F(I_{k,j_1},\dots, I_{k,j_m}), \end{equation} where for each pair $(\ell,\ell')$, $I_{k,j_\ell}$ and $I_{k,j_\ell'}$ are either identical or disjoint. Since \begin{equation} C_{k+1} = \bigcup_{j=1}^{2^k} \ddot I_{k,j}, \end{equation} the hypothesis implies that $F(C_k^{m}) = F(C_{k+1}^{m})$, and the result then follows from Lemma \ref{lemma1}(ii). \end{proof} We only apply Lemma \ref{lemma2} in the cases that $m=2$ and $[a,b] = [0,1]$ and $[\frac23,1]$. (In the latter case, when $F(x,y) = xy$ or $x/y$, it is helpful to have control of the ratio $x/y$.) \section{Arithmetic on the Cantor set.} \subsection{Addition and subtraction}\label{subsec:plus-and-minus} Sums and differences of Cantor sets have been widely studied in connection with dynamical systems. In this section we give a brief proof of the following result of Utz \cite{utz:cantor}. Further information about sums of Cantor sets can be found in \cite{chm:sums} and \cite{chm:sums2}. \begin{theorem}[Utz] If $\lambda \in [\frac 13,3]$, then every element $u$ in $[0,1+\lambda]$ can be written in the form $u = x + \lambda y$ for $x,y \in C$. \end{theorem} We include this proof in order to introduce the main ideas in the proof of Theorem \ref{maintheorem} in a simpler context. The key tool is Lemma \ref{lemma2}. Let $f_{\lambda}(x,y) = x + \lambda y$; we wish to show that $f_{\lambda}(C^{2}) = [0,1+\lambda]$. Observe that $C + \lambda C = \lambda(C + \lambda^{-1}C)$ for $\lambda \neq 0$, so it suffices to consider $\tfrac13 \le \lambda \le 1$. \begin{proof} We apply Lemma \ref{lemma2} and show that for any two closed intervals $I_1,I_2$ of the same length in $[a,b]$, $f_{\lambda}(I_1,I_2) = f_{\lambda}(\ddot I_1,\ddot I_2)$. For clarity, we write $I_1 = [r,r+3t]$, $I_2 = [s,s+3t]$, and $w = r + \lambda s$, so that $f_{\lambda}(I_1,I_2) = [w, w+ 3(1+\lambda)t]$. Observe that $\ddot I_1 = [r,r+t]\cup[r+2t,r+3t]$ and $\ddot I_2 = [s,s+t]\cup[s+2t,s+3t]$, so \begin{equation}\begin{split} f_{\lambda}(\ddot I_1,\ddot I_2) &= \left(\ [w,w+(1+\lambda)t] \cup [w+2\lambda t,w+(1+3\lambda)t]\ \right) \\ &\cup \left(\ [w+2t,w+(3+\lambda)t] \cup [w+(2+2\lambda t,w+(3+3\lambda)t]\ \right) \,. \end{split}\end{equation} Since $\lambda \le 1$, $1+\lambda \ge 2\lambda$ and $3+\lambda \ge 2+ 2\lambda$, the pairs of intervals coalesce into \begin{equation} f_{\lambda}(\ddot I_1,\ddot I_2) = [w,w+(1+3\lambda)t]) \cup [w+2t,w+(3+3\lambda)t]. \end{equation} Since $\lambda \ge \frac 13$, { we have} $2 \le 1+3\lambda$. { Hence} $f_{\lambda}(\ddot I_1,\ddot I_2) =f_{\lambda}(I_1,I_2)$, completing the proof. \end{proof} \begin{remark} Unlike the folklore proof for $\lambda = 1$, there seems to be no obvious algorithmic proof, save for $\lambda = \frac 13$. In this case, suppose $u \in [0,\frac 43]$. If $u = \frac 43$, then $u = 1 + \frac 13\cdot 1 \in C + \frac 13 C$. If $u < \frac 43$, then we can write $u = x + y$, $x, y \in C$, and assume $x\ge y$. Since $y \le \frac u2 < \frac 23$ is in $C$, $y \le \frac 13$, hence $y = \frac 13 z$, $z \in C$. This produces the desired construction. \end{remark} The case of subtraction, {that is, the case of} $f_{\lambda}$ when $\lambda < 0$, is easily handled. \begin{theorem} If $\beta = -\lambda <0$, then \begin{equation} f_{\beta}(C^{2}) = -\lambda + f_{\lambda}(C^{2}). \end{equation} \end{theorem} \begin{proof} If $\beta < 0$, $x,y \in C$, we have \begin{equation} x + \beta y = x + \beta(1-z) = -\lambda + x + \lambda z \end{equation} for $x$ and $z$ in $C$. \end{proof} \begin{remark}\label{rem:sums} Arithmetic sums of Cantor sets and more general compact sets have been studied intensively. For instance, Mendes and Oliveira \cite{mo:topological} discuss the topological structure of sums of Cantor sets, while Schmeling and Shmerkin \cite{SS:dimension} characterize those nondecreasing sequences $0\le d_1 \le d_2\le d_3\le \cdots \le 1$ { that} can arise as the sequence of Hausdorff dimensions of iterated sumsets $A,A+A,A+A+A,\ldots$ for a compact subset $A$ of ${\mathbb R}$. Recent work of Gorodetski and Northrup \cite{gn:sums} involves the Lebesgue measure of sumsets of Cantor sets and other compact subsets of the real line. We refer the interested reader to these papers and the references therein for more information. \end{remark} \subsection{Multiplication}\label{subsec:times} We let $f(x,y) = x^2 y$ and shall show that $f(C^2) = [0,1]$. We begin by showing that it suffices to consider $f(\widetilde C^2)$. \begin{lemma} If $f(\widetilde C^2) = [\frac 8{27},1]$, then $f(C^2) = [0,1]$. \end{lemma} \begin{proof} Suppose $u \in [0,1]$. If $u = 0$, then $u = 0^2\cdot 0$. If $u > 0$, then there exists a unique integer $r \ge 0$ so that $u = 3^{-r}v$, where $v \in (\frac 13,1]$. Since $\frac 8{27} < \frac 13$, $v = x^2y$ for $x,y\in \widetilde C \subset C$, and since $x, 3^{-r}y \in C$, $u = x^2 (3^{-r}y)$ is the desired representation. \end{proof} Accordingly, we confine our attention to $\widetilde C$. \begin{lemma} If $I = [a,a+3t]$ and $J = [b,b+3t]$ are in $[\frac 23, 1]$, then $f(I,J) = f(\ddot I, \ddot J)$. \end{lemma} \begin{proof} We first define \begin{equation} \begin{gathered} \ [a^2b, (a+t)^2 (b + t)]=: [u_1,v_1]; \\ [a^2(b+2t), (a+t)^2 (b + 3t)]=: [u_2,v_2]; \\ [(a+2t)^2b, (a+3t)^2 (b + t)]=: [u_3,v_3]; \\ [(a+2t)^2(b+2t), (a+3t)^2 (b +3 t)]=: [u_4,v_4]. \end{gathered} \end{equation} Evidently, $f(I,J) = [u_1,v_4]$, and also, $u_1 < u_2, v_1 < v_2$, and $u_3 < u_4, v_3 < v_4$. If we can first show that $v_1 > u_2$ and $v_3 > u_4$, then $[u_1,v_1] \cup [u_2,v_2] = [u_1, v_2]$ and $[u_3,v_3] \cup [u_4,v_4] = [u_3, v_4]$. Second, since $u_1 < u_3$ and $v_2 < v_4$, if we can show that $v_2 > u_3$, then $[u_1,v_2] \cup [u_3,v_4] = [u_1, v_4]$, and the proof will be complete. We compute \begin{equation}\label{eq:36} \begin{gathered} v_1 - u_2 = a(2b-a)t + (2a + b) t^2 + t^3; \\ v_2 - u_3 = a(3a - 2b)t + (6a - 3b)t^2 + 3t^3; \\ v_3 - u_4 = a(2b - a)t + (5b - 2a) t^2 + t^3. \end{gathered} \end{equation} Since $2b - a \ge 2\cdot \frac 23 - 1 > 0$, $3a-2b \ge 3\cdot \frac 23 - 2 \ge 0$, $6a - 3b \ge 6 \cdot \frac 23 -3 > 0$, and $5b-2a \ge 5\cdot \frac 23 - 2 > 0$, each of the quantities in \eqref{eq:36} is positive and the proof is complete. \end{proof} \begin{theorem}\label{mult} \begin{equation} f(\widetilde C^2) = [\tfrac 8{27},1]. \end{equation} \end{theorem} \begin{proof} Apply Lemma \ref{lemma2}, noting that $f([\frac 23,1]^2) = [\tfrac 8{27},1]$. \end{proof} \begin{remark} Observe that $u \in [0,1]$, written as $x^2y$, where $x,y \in C$, is also $x\cdot x \cdot y$, so this implies that every element in $[0,1]$ is a product of three elements from the Cantor set. (This can also be proved in a more ungainly way, by taking $m=3$ in Lemma \ref{lemma2} with $f(x_1,x_2,x_3) = x_1x_2x_3$.) \end{remark} \begin{remark} We do not know an algorithm for expressing $u \in [0,1]$ in the form of $x^2y$ or $x_1x_2x_3$ as a product of elements of the Cantor set. \end{remark} \begin{remark} More generally, one can look at $f_{a,b}(C^2)$ where $f_{a,b}(x,y) := x^ay^b$. Since $u = x^ay^b$ if and only if $u^{1/a} = x y^{b/a}$, for $u \in (0,1)$, it suffices to consider $a=1$. By looking at $C_1 = [0,\frac 13] \cup [\frac23, 1]$, it is not hard to see that if $f(x,y) = x y^t$, $t \ge 1$, and $(\frac 23)^{1+t} > \frac 13$, then $f(C_1^2)$ is already missing an interval from $[0,1]$. This condition occurs when $t < \frac{\log 2}{\log 3/2} \approx 1.7095$. \end{remark} \begin{remark} By taking logarithms, we can convert the question about products of elements of $C$ into a question about sums. (We can omit the point $0$ since its multiplicative behavior is trivial.) Of course, the underlying set is no longer the standard self-similar Cantor set but is a more general (``non-linear'') closed subset of ${\mathbb R}$. Some conclusions about the number of factors needed to recover all of $[0,1]$ can be obtained from the general results in the papers of Cabrelli--Hare--Molter \cite{chm:sums}, \cite{chm:sums2}, but it does not appear that one obtains the precise conclusion of part 2 of Theorem \ref{maintheorem} in this fashion. \end{remark} \subsection{Division}\label{subsec:divide} In this section, we complete our arithmetic discussion by considering quotients. \begin{theorem}\label{div} \begin{equation}\label{E:div} \left\{ \frac uv: u, v \in C\right\} = \bigcup_{m = -\infty}^{\infty} \left[\frac 23 \cdot 3^m , \frac 32 \cdot 3^m\right]. \end{equation} \end{theorem} \begin{proof} As with multiplication, it suffices to consider $\widetilde C$. \begin{lemma} Theorem \ref{div} is implied by the identity \begin{equation}\label{E:div-c-tilde} \left\{ \frac uv: u, v \in \widetilde C\right\} = \left[\frac 23 , \frac 32 \right]\,. \end{equation} \end{lemma} \begin{proof} Write $u, v \in C$ as $u = 3^{-s}\tilde u$, $v = 3^{-t}\tilde v$ for integers $c,d \ge 0$ and $\tilde u, \tilde v \in \widetilde C$. Then $u/v = 3^{d-c}\tilde u/\tilde v$, where $m=d-c$ can attain any integer value. \end{proof} We now prove \eqref{E:div-c-tilde}. Consider $\widetilde C_1 = [\frac 23,1]$ and apply Lemma \ref{lemma2}. Clearly, $\{\frac uv: u, v \in \widetilde C_1\} = [\frac 23, \frac 32]$. Consider two intervals in $\widetilde C_n$, $I_1 = [a,a+3t]$ and $I_2=[b,b+3t]$. { These intervals are either identical or disjoint.} Since $x = \frac uv$ implies $\frac 1x = \frac vu$, there is no harm in assuming $a \le b$, and either $I_1 = I_2$ and $a=b$, or the intervals are disjoint and $a + 3t \le b$. The quotients from these intervals will lie in \[ J_0:= \left[ \frac a{b+3t}, \frac {a+3t}b \right]:= [r_0,s_0]. \] Since $\ddot I_1 = [a,a+t] \cup [a+2t,a+3t]$ and $\ddot I_2 = [b,b+t] \cup [b+2t,b+3t]$, we obtain four subintervals \[ \begin{gathered} J_1 =\left[ \frac a{b+3t}, \frac {a+t}{b+2t} \right] = [r_1,s_1], \\ J_2 =\left[ \frac a{b+t}, \frac {a+t}{b} \right] = [r_2,s_2], \\ J_3 =\left[ \frac {a+2t}{b+3t}, \frac {a+3t}{b+2t} \right] = [r_3,s_3], \\ J_4 =\left[ \frac {a+2t}{b+t}, \frac {a+3t}{b} \right] = [r_4,s_4]. \\ \end{gathered} \] We need to see how $J_0 = J_1 \cup J_2 \cup J_3 \cup J_4$. There are two cases, depending on whether $a=b$ or $a<b$. We first record some algebraic relations. We have $r_1 = r_0$ and $s_4 = s_0$, and, evidently, $r_1 < r_2$, $s_1 < s_2$, $r_3 < r_4$, $s_3 < s_4$. Further, \[ \begin{gathered} r_3 - r_2 = \frac {a+2t}{b+3t} - \frac a{b+t} = \frac{2t(b-a+t)}{(b+t)(b+3t)}, \\ s_3 - s_2 = \frac {a+3t}{b+2t} - \frac {a+t}{b} = \frac{2t(b-a-t)}{b(b+2t)}, \\ s_1 - r_2 = \frac {a+t}{b+2t} - \frac a{b+t} = \frac{t(b-a+t)}{(b+t)(b+2t)}, \\ s_2 - r_3 = \frac {a+t}{b} - \frac {a+2t}{b+3t} = \frac{t(3a + 3t - b)}{b(b+3t)} \ge \frac {t(3\cdot\frac23 + 0 - 1)}{b(b+3t)} > 0, \\ s_3 - r_4 = \frac {a+3t}{b+2t} - \frac {a+2t}{b+t} = \frac{t(b-a-t)}{(b+t)(b+2t)}. \end{gathered} \] Suppose first that $a < b$, so $ a+3t < b$. Then each of the differences above is positive, so $r_1 < r_2 < r_3 < r_4$ and $s_1 < s_2 < s_3 < s_4$; further, the intervals overlap: $s_1 > r_2$, $s_2 > r_3$ and $s_3 > r_4$. Thus $J_0 = J_1 \cup J_2 \cup J_3 \cup J_4$. If $a = b$, then $J_3 =\left[ \frac {a+2t}{a+3t}, \frac {a+3t}{a+2t} \right] \subset \left[ \frac a{a+t}, \frac {a+t}{a} \right] = J_2$, so we may drop $J_3$ from consideration. We have $r_1 < r_2 < r_4$ and $s_1 < s_2 < s_4$ and need only show that $s_1 > r_2$ and $s_2 > r_4$. The first is clear, and for the second, \begin{equation} s_2 - r_4 = \frac {a+t}{a} - \frac {a+2t}{a+t} = \frac{t^2}{a(a+t)} > 0, \end{equation} so { $J_0 = J_1 \cup J_2 \cup J_4$,} and we are done. \end{proof} \begin{remark} We do not know an algorithm for expressing a feasible $u$ as a quotient of elements { in} $C$. \end{remark} \subsection{Multiplication, revisited}\label{subsec:times-again} Let $g(x,y) = xy$. As noted earlier, $g(C^2)$ is not the full interval $[0,1]$, though $g(C^2) = \cap g(C_i^2)$ is the intersection of a descending chain of closed sets and so is closed. In order to gain some information about $g(C^2)$, we look carefully at how Lemma \ref{lemma2} fails. \begin{lemma}\label{middle} Let $I = [a,a+3t]$ and $J = [b,b+3t]$, with { $\tfrac23 \le a \le b \le 1$,} be either identical or disjoint intervals. Then \begin{equation*} \begin{gathered} a < b \implies g(\ddot I, \ddot J) = g(I, J)\,; \\ a = b \implies g(\ddot I, \ddot I) = g(I,I) \setminus ((a+2t)^2-t^2,(a+2t)^2). \end{gathered} \end{equation*} \end{lemma} { \begin{proof} We have $$ g([a,a+3t],[b,b+3t]) = [ab, ab + 3(a+b)t + 9t^2] $$ and \begin{equation}\label{eq:gab} \begin{gathered} g([a,a+t]\cup[a+2t,a+3t],[b,b+t]\cup[b+2t,b+3t]) = \\ [ab, ab + (a + b)t + t^2] \cup [ab + 2at, ab + (3a + b)t + 3t^2] \\ \cup [ab + 2b t, ab + (a + 3b)t + 3t^2] \\ \cup[ ab + (2a+2b)t + 4t^2, ab + 3(a+b)t + 9t^2]. \end{gathered} \end{equation} Since $a \le b$, it follows that $ab + 2at \le ab + (a + b)t + t^2$, and the first two intervals coalesce into $[ab, ab + (3a + b)t + 3t^2]$. Suppose that $a < b$, and recall that we have assumed $\tfrac23 \le a < b \le 1$. Since $a+t \le b$, it follows that $ab+(2a+2b)t+4t^2 \le ab + (a+3b)t +3t^2$ and the last two intervals coalesce into $[ab+2bt,ab+3(a+b)t+9t^2]$. Thus the right-hand side of \eqref{eq:gab} reduces to \begin{equation}\label{eq:gab2} [ab, ab + (3a + b)t + 3t^2] \cup [ ab + 2bt, ab + 3(a+b)t + 9t^2] \end{equation} Moreover, \begin{equation*} ab+ (3a+b)t + 3t^2 - (ab + 2bt) = t(3a+3t - b) \ge t (3 \cdot \tfrac23 + 0 - 1) = t \ge 0, \end{equation*} which shows that the pair of intervals in \eqref{eq:gab2} coalesces to a single interval. This proves the first statement. If $a = b$, then the middle two intervals in \eqref{eq:gab} are the same, and \begin{equation*}\begin{split} &g(([a,a+t]\cup[a+2t,a+3t])^2) \\ &\quad = [a^2, a^2+4at + 3t^2] \cup [a^2 + 4at + 4t^2, a^2 + 6at + 9t^2] \\ &\quad = [a^2,(a+3t)^2] \setminus ((a+2t)^2-t^2,(a+2t)^2)). \end{split}\end{equation*} \end{proof} } This leads to the following estimate for the Lebesgue measure of $g(C^2)$. \begin{theorem}\label{meas} \[ \mu(g(C^2)) \ge \frac {17}{21}. \] \end{theorem} \begin{proof} First note that $g(\widetilde C^2) \subset g(\widetilde C_1^2) = [\frac 49,1]$. It follows as before \begin{equation} g(C^2) = \{0\} \cup \bigcup_{k=0}^\infty 3^{-k}g(\widetilde C^2), \end{equation} and since $\frac 13 < \frac 49$, the sets $3^{-k}g(\widetilde C^2)$ are disjoint. Therefore, \begin{equation} \mu(g(C^2)) = \sum_{k=0}^\infty 3^{-k}\mu(g(\widetilde C^2)) = \tfrac 32 \mu(g(\widetilde C^2)). \end{equation} Since $\widetilde C_n$ consists of $2^{n-1}$ intervals of length $3^{-n}$, it follows from Lemma \ref{middle} that \begin{equation}\begin{split} &\mu(g(\widetilde C_{n+1}^2)) \ge \mu(g(\widetilde C_{n}^2)) - \frac {2^{n-1}}{3^{2n+2}} \\ &\implies \mu(g(\widetilde C^2)) \ge \left(1-\frac 49\right) - \sum_{n=1}^{\infty} \frac {2^{n-1}}{3^{2n+2}} = \frac{34}{63} \\ &\implies \mu(g(C^2)) \ge \frac 32\cdot \frac{34}{63} = \frac {17}{21}. \end{split}\end{equation} \end{proof} \begin{remark} This argument shows that for all $m$, \begin{equation} \begin{gathered} \mu(g(\widetilde C_m^2)) \ge \mu(g(\widetilde C^2)) \ge\mu(g(\widetilde C_m^2)) - \sum_{n=m+1}^{\infty} \frac {2^{n-1}}{3^{2n+2}}. \end{gathered} \end{equation} \end{remark} The reason that Theorem \ref{meas} is only an estimate is that there is no guarantee that intervals missing from $f(\ddot I^2)$ cannot be covered elsewhere. The first instance in which this occurs is for $n=4$: one of the intervals in $\widetilde C_4$ is $I_0 = [\frac {62}{81},\frac {63}{81}]$ = $[.2022_3,.21_3]$. By Lemma \ref{middle}, \begin{equation} (\tfrac{188^2-1}{243^2}, \tfrac{188^2}{243^2}) = (\tfrac{35343}{59049},\tfrac{35344}{59049}) \approx (.5985368,.5985537) \notin f(\ddot I_0^2). \end{equation} However, $\widetilde C_5$ contains the intervals \begin{equation} J_1 = [\tfrac {162}{243}, \tfrac {163}{243}] = [.2_3,.20001_3], \quad J_2 = [\tfrac {216}{243}, \tfrac {217}{243}] = [.22_3,.22001_3] \end{equation} and \[ J_1J_2 = [\tfrac{34992}{59049},\tfrac{35371}{59049}] \approx (.5925926,.59901099) \] covers the otherwise-missing interval. A more detailed {\it Mathematica} computation, using $m=11$, gives the first eight decimal digits for $\mu(g(C^2))$: \begin{equation} \mu(g(C^2)) = .80955358\dots \approx \frac{17}{21} + 2.97 \times 10^{-5}. \end{equation} \section{Final remarks.}\label{sec:final-remarks} As mentioned in the introduction, this paper is part of a larger project. { We discuss a few results from this project whose proofs will appear elsewhere, written by various combinations of the authors and their students.} \subsection{Self-similar Cantor sets} The Cantor set easily generalizes to sets defined with different ``middle-fractions'' removed. Consider the self-similar Cantor set $D^{(t)}$ obtained as the invariant set for the pair of contractive mappings $$ f_1(x) = tx, \qquad f_2(x) = tx+(1-t) $$ acting on the real line. Thus $$ D^{(t)} = \bigcap_{n\ge 0} D_n^{(t)}, $$ where, for each $n\ge 0$, $D_n^{(t)}$ is a union of $2^n$ intervals, each of length $t^n$, contained in $[0,1]$. For instance, $$ D_1^{(t)} = [0,t] \cup [1-t,1], $$ $$ D_2^{(t)} = [0,t^2]\cup[t(1-t),t]\cup[1-t,1-t+t^2]\cup[1-t^2,1]\,. $$ For an integer $m\ge 2$, let $t_m$ be the unique solution to $(1-t)^m=t$ in $[0,1]$. Then $$ \cdots < t_4 < t_3 < t_2 < \frac12 $$ and $t_m \to 0$ as $m\to\infty$. Numerical values are \begin{eqnarray*} &t_2 = \tfrac{3-\sqrt5}2 \approx 0.381966 \dots \\ &t_3 \approx 0.317672 \dots \\ &t_4 \approx 0.275508 \dots \\ &t_5 \approx 0.245122 \dots \\ &t_6 \approx 0.22191 \dots \\ &t_7 \approx 0.203456 \dots \end{eqnarray*} If we choose $t$ such that $t_m \le t < t_{m-1}$ (that is, $(1-t)^m \le t < (1-t)^{m-1}$) and let $g_n(x_1,\dots,x_n) = x_1x_2\cdots x_n$, then \[ g_m((D^{(t)})^m) = [0,1], \] but $g_{m-1}((D^{(t)})^{m-1})$ has Lebesgue measure strictly less than one. In particular, if a Cantor set $D$ in which a middle fraction $\lambda$ is taken with $\lambda \le 1 - 2t_2 = \sqrt 5 - 2 \approx .23607$, then every element in $[0,1]$ can be written as a product of two elements of $D$. \subsection{Number Theory} { There is much more to be said about the representation of specific numbers in $C/C$. For example, if $v = 2u$ for $u,v \in C$, then \begin{equation*} u = \frac 1{3^n},\ v = \frac 2{3^n},\quad n \ge 1; \end{equation*} if $v = 11u$ for $u,v \in C$, then \begin{equation*} u = \frac 1{4\cdot 3^n},\ v = \frac {11}{4\cdot 3^n},\quad n \ge 1. \end{equation*} By contrast, if $v = 4u$ for $u,v \in C$, then there exists a finite or infinite sequence of integers $(n_k)$ with $n_1 \ge 2$ and $n_{k+1}-n_k \ge 2$, so that \begin{equation*} u = \sum_k \frac 2{3^{n_k}},\qquad v = \sum_k\left(\frac 2{3^{n_k-1}} + \frac 2{3^{n_k}}\right). \end{equation*} The proof of the second result is trickier than the other two.} We also mention a conjecture for which there is strong numerical evidence. \begin{conjecture} Every $u \in [0,1]$ can be written as $x_1^2 + x_2^2 + x_3^2 + x_4^2$, $x_i \in C$. \end{conjecture} We need a minimum of four squares, since $(\frac 13)^2 + (\frac 13)^2+(\frac 13)^2 < (\frac 23)^2$, so the open interval $(\frac 13, \frac 49)$ will be missing from the sum of three squares. \section{Acknowledgments} The authors wish to thank the referees and editors for their rapid, sympathetic and extremely useful suggestions for improving the manuscript. BR wants to thank Prof. W. A. J. Luxemburg's Math 108 at Caltech in 1970--1971 for introducing him to the beauties of analysis.
{ "timestamp": "2017-11-27T02:09:00", "yymm": "1711", "arxiv_id": "1711.08791", "language": "en", "url": "https://arxiv.org/abs/1711.08791", "abstract": "Every element $u$ of $[0,1]$ can be written in the form $u=x^2y$, where $x,y$ are elements of the Cantor set $C$. In particular, every real number between zero and one is the product of three elements of the Cantor set. On the other hand the set of real numbers $v$ that can be written in the form $v=xy$ with $x$ and $y$ in $C$ is a closed subset of $[0,1]$ with Lebesgue measure strictly between $\\tfrac{17}{21}$ and $\\tfrac89$. We also describe the structure of the quotient of $C$ by itself, that is, the image of $C\\times (C \\setminus \\{0\\})$ under the function $f(x,y) = x/y$.", "subjects": "Metric Geometry (math.MG); Number Theory (math.NT)", "title": "Cantor set arithmetic", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9886682471364099, "lm_q2_score": 0.8104789109591832, "lm_q1q2_score": 0.8012947642390421 }
https://arxiv.org/abs/2112.04654
Unimodular triangulations of sufficiently large dilations
An integral polytope is a polytope whose vertices have integer coordinates. A unimodular triangulation of an integral polytope in $\mathbb{R}^d$ is a triangulation in which all simplices are integral with volume $1/d!$. A classic result of Knudsen, Mumford, and Waterman states that for every integral polytope $P$, there exists a positive integer $c$ such that $cP$ has a unimodular triangulation. We strengthen this result by showing that for every integral polytope $P$, there exists $c$ such that for every positive integer $c' \ge c$, $c'P$ admits a unimodular triangulation. This answers a longstanding question in the area.
\section{Introduction} Unimodular triangulations are elementary objects which arise naturally in algebra and combinatorics. We refer to the paper by Haase et al.\ \cite{HPPS} for an extensive survey on the subject. In this paper we answer a longstanding question on the existence of certain unimodular triangulations. An integral polytope is a polytope with integer coordinates. Let $P$ be a $d$-dimensional integral polytope in $\mathbb R^d$. A unimodular triangulation of $P$ is a triangulation of $P$ into simplices each of which has the minimum possible volume $1/d!$. For $d \ge 3$, not every integral polytope has a unimodular triangulation. For example, the simplex with vertices $(0,0,0)$, $(1,0,0)$, $(0,1,0)$, $(1,1,99)$ does not have any nontrivial triangulation, but has volume $>1/6$. On the other hand, every polytope has a \emph{dilation} which admits a unimodular triangulation, as described below. \begin{thm}[\cite{KKMS}, 1973] \label{thm:KMW} For every integral polytope $P$, there is a positive integer $c$ such that $cP$ admits a unimodular triangulation. \end{thm} This theorem is known as the Kempf--Knudsen--Mumford--Saint-Donat theorem (KKMS) or the Knudsen--Mumford--Waterman theorem (KMW).\footnote{The result appears in the book \cite{KKMS} which is authored by Kempf, Knudsen, Mumford, and Saint-Donat. The individual chapter in which it appears is authored by Knudsen, who also credits Mumford and Waterman. As a result, both naming conventions have appeared in the literature. In this paper we use the convention ``KMW theorem''.} It is one of the earliest considerations of unimodular triangulations, and was proved in the context of algebraic geometry in order to prove semistable reduction for families of varieties over a curve. A more general version of semistable reduction was conjectured by Abramovich and Karu in 2000 \cite{AK00} and was reduced to proving the existence of certain unimodular triangulations of maps. The conjecture was recently proven by Adiprasito, Temkin, and the author in \cite{ALT18}. Understanding what values of $c$ work in Theorem~\ref{thm:KMW} is an old and difficult problem. The answer is almost completely known in dimensions $\le 3$ \cite{SZ13}, and a general upper bound for the smallest possible $c$ is known in terms of the dimension and volume of the polytope \cite{HPPS}. In this paper, we prove the following: \begin{thm} \label{thm:main} For every integral polytope $P$, there is a positive integer $c$ such that for all $c' \ge c$, $c'P$ admits a unimodular triangulation. \end{thm} The result may be a bit surprising, as there are known to be polytopes $P$ and integers $c$ such that $cP$ has a unimodular triangulation but $(c+1)P$ does not. We do not provide an explicit value for the $c$ in Theorem~\ref{thm:main}, but we note that the upper bound should be doubly exponential in the dimension and volume of $P$ using the arguments from \cite{HPPS}. We also note that the unimodular triangulation can be made regular, but to keep the paper simpler we do not prove this. The idea of the proof is as follows. The argument from \cite{KKMS} in fact shows that if $cP$ has a unimodular triangulation, then $c'P$ has a unimodular triangulation for any $c'$ a multiple of $c$. For our result, we prove the following: There exist relatively prime positive integers $a$ and $b$ such that for any nonnegative integers $r$ and $s$, $(ra + sb)P$ has a unimodular triangulation. In order to prove this, we extend the results of \cite{ALT18} to \emph{mixed subdivisions}, which can be thought of as coupled subdivisions of two polytopes. This ends up being more difficult than one might expect. While the constructions from \cite{ALT18} have natural mixed analogues, these natural analogues do not lead to a proof of the theorem; see the discussion at the beginning of Section~\ref{sec:main}. Therefore, we have to create some more complicated analogues as well as make some new constructions. Another feature of this proof is that it makes heavy use of an idea we call \emph{canonical subdivisions}. Canonical subdivisions are present in a more specialized form in \cite{KKMS} and \cite{HPPS}, and are expanded in a very general way in \cite{ALT18}. In \cite{ALT18}, the idea was explained using the language of categories and functors. In this paper we rework these ideas in terms of posets and poset maps, which turns out to be more flexible and general for this situation. Essentially, a canonical subdivision is a rule to subdivide every polytope within a family of polytopes so that this rule is compatible with the operation of restricting to a face of a polytope. The importance of this is that canonical subdivisions can be used to further subdivide arbitrary polytopal complexes in a consistent way. This allows us to iteratively construct a unimodular triangulation through many intermediate canonical subdivisions. This idea is implicit in Knudsen et al.'s original proof of the KMW theorem, and in fact they proved that any polytopal \emph{complex} $X$ of integral polytopes has a constant $c$ such that $cX$ has a unimodular triangulation. Our main result also extends to polytopal complexes in this way, which is immediate from the proof. The cornerstone of our canonical subdivision method is Theorem~\ref{thm:confluence}, which may be of independent interest. This theorem gives conditions under which a family of non-canonical subdivisions can be used to recursively construct a canonical subdivision. This theorem is important because our desired canonical subdivision is very complicated and canonicity is difficult to check. The theorem allows us to instead construct simpler non-canonical subdivisions, after which the criteria of the theorem are straightforward to check. To demonstrate Theorem~\ref{thm:confluence}, we have also included in Section~\ref{sec:canonicalexamples} some examples of well-known subdivisions in the literature that can be constructed using these methods. Finally, we would like to mention a few open problems. Despite the method of proof used in this paper, it is unknown whether $c_1P$ and $c_2P$ having unimodular triangulations implies that $(c_1+c_2)P$ has one as well. In addition, it is unknown whether for every dimension $d$ there exists an integer $c_d$ such that $c_d P$ has a unimodular triangulation for every $d$-dimensional polytope $P$. Finally, the question of whether specific classes of polytopes have unimodular triangulations has attracted significant attention. Classes of interest include smooth polytopes, matroid polytopes, and parallelepipeds. The paper is organized as follows. In Section~\ref{sec:prelim} we introduce the language of polytopes, posets, and canonical subdivisions. This section is very abstract, but the author believes the initial investment makes the main argument much easier to follow. In Section~\ref{sec:Cayleypolytopes}, we introduce Cayley polytopes which are the main building blocks of our constructions. In Section~\ref{sec:boxpoints} we introduce box points, which allow us to modify triangulations to produce triangulations with simplices of smaller volume. Our main constructions and proof are in Section~\ref{sec:main}. \section{Preliminaries} \label{sec:prelim} \subsection{Polytopes} \label{sec:polytopes} Throughout the paper, we work in $\mathbb R^d$ with some fixed $d$ unless otherwise specified. In this paper, a \emph{polytope} is a nonempty convex hull of finitely many points in $\mathbb R^d$. Given a polytope $P$ and a linear functional $\phi \in (\mathbb R^d)^\ast$, the \emph{face} of $P$ with respect to $\phi$ is the set of all points in $P$ at which $\phi$ reaches its maximum on $P$. We \textbf{do not} consider the empty set to be a face. A \emph{simplex} is the convex hull of a nonempty, affinely independent set of points. For any polytope $P \subset \mathbb R^d$, we define $V(P)$ to be the real span of the set $\{ u -v : u,v \text{ are vertices of } P \}$. We say that polytopes $P_1$, \dots, $P_n$ are \emph{independent} if $V(P_1)$, \dots, $V(P_n)$ are linearly independent vector subspaces, i.e. $\dim(V(P_1) + \dots + V(P_n)) = \dim V(P_1) + \dots + \dim V(P_n)$. A polytope of the form $\sum_{j=1}^n S_j$ where $S_1$, \dots, $S_n$ are independent simplices is called a \emph{polysimplex} or \emph{product of simplices}. In this paper, a \emph{lattice} is an additive subgroup of $\mathbb Z^d$. We define the \emph{normalized index}, or just \emph{index}, of a lattice $L$ to be the group index $[ \Span_{\mathbb R}(L) \cap \mathbb Z^d, L ]$. This index is always finite. We denote the index by $\ind(L)$. An \emph{integral polytope} is a polytope whose vertices have integer coordinates. Given an integral polytope $P$, we define $L(P)$ to be the lattice generated over $\mathbb Z$ by the set $\{ u -v : u,v \text{ are vertices of } P \}$. We define $N(P)$ to be the lattice $V(P) \cap \mathbb Z^d$. The \emph{index} of $P$ is the normalized index of $L(P)$, which equals $[N(P), L(P)]$. A \emph{unimodular simplex} is an integral simplex of index 1. For $d$-dimensional simplices in $\mathbb R^d$ this is equivalent to the definition given in the introduction; the current definition extends this to simplices with dimension lower than the ambient space. An \emph{ordered polytope} is a polytope along with a total ordering on its vertices. An \emph{ordered face}, or \emph{face}, of an ordered polytope is a face of the underlying polytope along with the vertex order induced by the original polytope. Any translation or positive dilation of an ordered polytope is also an ordered polytope, with the obvious ordering. \subsection{Posets and subdivisions} \subsubsection{Relative posets} Recall that a \emph{poset} is a set $\mathcal A$ along with a binary relation $\le_{\mathcal A}$ on $\mathcal A$ which is reflexive, antisymmetric, and transitive. We will always denote a poset by its set of elements, and if there is no risk of confusion we will use the symbol ``$\le$'' in place of $\le_{\mathcal A}$. Given two posets $\mathcal A$ and $\mathcal B$, a \emph{poset map} is a function $f : \mathcal A \to \mathcal B$ such that $f(x) \le_{\mathcal B} f(y)$ whenever $x \le_{\mathcal A} y$. A \emph{poset isomorphism} is a poset map which has an inverse which is a poset map. Let $\mathcal B$ be a poset. We define a \emph{$\mathcal B$-poset} to be a pair $(\mathcal A, p)$ where $\mathcal A$ is a poset and $p : \mathcal A \to \mathcal B$ is a poset map. If there is no risk of confusion, we denote $(\mathcal A, p)$ by just $\mathcal A$. Clearly, if $(\mathcal A, p)$ is a $\mathcal B$-poset and $(\mathcal A', p')$ is an $\mathcal A$-poset, then $(\mathcal A', p \circ p')$ is a $\mathcal B$-poset. Given two $\mathcal B$-posets $(\mathcal A, p)$ and $(\mathcal A', p')$, we define a \emph{poset map over $\mathcal B$} to be a poset map $f : \mathcal A \to \mathcal A'$ such that $p = p' \circ f$. Given a poset $\mathcal A$ and a subset $X \subset \mathcal A$, we let $\langle X \rangle_{\mathcal A}$ or $\langle X \rangle$ denote the set $X$ along with all members of $\mathcal A$ below $X$ in $\mathcal A$. If $x$ is a single element of $\mathcal A$, we use $\langle x \rangle$ as shorthand for $\langle \{x\} \rangle$. We let $\Max_{\mathcal A} X$ or $\Max X$ denote the maximal elements of $X$ with respect to $\le_{\mathcal A}$. \subsubsection{The poset of polytopes and subdivisions} Let $\mathcal P$ be the poset whose elements are all polytopes in $\mathbb R^d$, and with the partial order $F \le_{\mathcal P} P$ if $F$ is a face of $P$. The main class of posets we will work with in this paper are $\mathcal P$-posets. \begin{exam} Trivially, $(\mathcal P, \id)$ is a $\mathcal P$-poset, where $\id$ is the identity map. \end{exam} \begin{exam} Let $\mathcal O$ be the poset whose elements are ordered polytopes and with the partial order $F \le_{\mathcal O} P$ if $F$ is an ordered face of $P$. Then $(\mathcal O, \Forget)$, where $\Forget$ is the map $\mathcal O \to \mathcal P$ which forgets the vertex ordering, is a $\mathcal P$-poset. \end{exam} \begin{exam} \label{exam:pointsets} Let $\mathcal S$ be the poset whose elements are totally ordered finite subsets of $\mathbb R^d$, with the partial order $B \le_{\mathcal S} A$ if $B = A \cap F$, where $F$ is any face of $\conv(A)$, and the order on $B$ is the order induced by $A$. Then $(\mathcal S, \conv)$ is a $\mathcal P$-poset. \end{exam} A \emph{polytopal complex}, or \emph{$\mathcal P$-complex}, is a finite subset $X$ of $\mathcal P$ satisfying the following. \begin{enumerate}[label=(\alph*)] \item If $F$, $P \in \mathcal P$ such that $P \in X$ and $F \le P$, then $F \in X$. \item If $P$, $Q \in X$ are different, then the relative interiors of $P$ and $Q$ are disjoint. \end{enumerate} We define the \emph{support} of a polytopal complex $X$ to be \[ \abs{X} := \bigcup_{x \in X} x. \] If $\abs{X}$ is a polytope $Q$, we say that $X$ is a \emph{subdivision} of $Q$. A \emph{triangulation} is a subdivision all of whose elements are simplices. More generally, let $(\mathcal A, p)$ be a $\mathcal P$-poset. An \emph{$(\mathcal A, p)$-complex} or \emph{$\mathcal A$-complex} is subset $X$ of $\mathcal A$ such that $p(X)$ is a polytopal complex and the map $p : X \to p(X)$ is a poset isomorphism from the subposet of $\mathcal A$ induced on $X$ to the subposet of $\mathcal P$ induced on $p(X)$. A \emph{subcomplex} of $X$ is a subset of $X$ which is also an $\mathcal A$-complex. The \emph{support} of $X$ is defined to $\abs{p(X)}$. We denote the support by simply $\abs{X}$. If the support is a polytope $Q$, we say that $X$ is a \emph{$(\mathcal A,p)$-subdivision} or \emph{$\mathcal A$-subdivision} of $Q$. For any $\mathcal P$-poset $(\mathcal A,p)$, let $\Subd(\mathcal A)$ be the poset whose elements are $\mathcal A$-subdivisions of polytopes, and with partial order $X' \le_{\Subd(\mathcal A)} X$ if $X'$ is a subcomplex of $X$ such that $\abs{X'}$ is a face of $\abs{X}$. Then the map $\abs{\cdot} : \Subd(\mathcal A) \to \mathcal P$ which sends $X$ to $\abs{X}$ is a poset map, and $(\Subd(\mathcal A), \abs{\cdot})$ is a $\mathcal P$-poset. The following proposition is easy to verify. \begin{prop} \label{prop:mapofsubdivisions} Let $(\mathcal A, p)$ and $(\mathcal A', p')$ be $\mathcal P$-posets and let $f : \mathcal A \to \mathcal A'$ be a poset map over $\mathcal P$. Then the map $f : \Subd(\mathcal A) \to \Subd(\mathcal A')$ which sends $X$ to $f(X)$ is a poset map over $\mathcal P$. \end{prop} \subsubsection{Trivial subdivisions and perfect posets} For any polytope $P$, we define the \emph{trivial subdivision} of $P$ to be $\triv(P) := \langle P \rangle_{\mathcal P}$. Let $(\mathcal A, p)$ be a $\mathcal P$-poset. If, for all $x \in \mathcal A$, the set $\langle x \rangle_{\mathcal A}$ is an $\mathcal A$-complex, then we call $(\mathcal A,p)$ a \emph{perfect} $\mathcal P$-poset. In this situation we define $\triv_{\mathcal A}(x) := \langle x \rangle_{\mathcal A}$. We necessarily have that $p(\triv_{\mathcal A}(x))$ is the trivial subdivision of $p(x)$. Every example we have given so far is a perfect $\mathcal P$-poset. In particular, $(\Subd(\mathcal A), \abs{\cdot})$ is a perfect $\mathcal P$-poset for any $\mathcal P$-poset $\mathcal A$. \subsection{Mixed subdivisions} \label{sec:mixed} Let $n$ be a positive integer. Define $n\mathcal P$ to be the poset whose set of elements is \[ \underbrace{\mathcal P \times \dots \times \mathcal P}_{n \text{ times}} \] and with $(F_1, \dots, F_n) \le_{n \mathcal P} (P_1, \dots, P_n)$ if and only if $F_1$, \dots, $F_n$ are faces of $P_1$, \dots, $P_n$, respectively, such that $F_1+\dots+F_n$ is a face of $P_1+\dots+P_n$. Equivalently, $(F_1, \dots, F_n) \le_{n \mathcal P} (P_1, \dots, P_n)$ if and only if there exists a linear functional such that $F_1$, \dots, $F_n$ are the faces of $P_1$, \dots, $P_n$, respectively, with respect to this linear functional. In the future, whenever we write $F \le P$ for two $n$-tuples of polytopes $F$, $P$, we mean that $F \le_{n \mathcal P} P$. Let $\plus : n\mathcal P \to \mathcal P$ be the map sending $(P_1,\dots,P_n)$ to $P_1+\dots+P_n$. Then by the definition of $n\mathcal P$ this is a poset map. Furthermore, $(n\mathcal P, \plus)$ is a perfect $\mathcal P$-poset. Let $X$ be an $n\mathcal P$-complex. Define the \emph{$n$-support} of $X$ to be \[ \abs{X}_n := (Q_1,\dots,Q_n) \] where \[ Q_k = \bigcup_{(P_1,\dots,P_n) \in X} P_k. \] It is well-known \cite{San05} that if $\abs{X}$ is a polytope, then $Q_1$, \dots, $Q_n$ are polytopes and $\abs{X} = Q_1 + \dots + Q_n$. The set $X$ is known as a \emph{mixed subdivision} of $(Q_1,\dots,Q_n)$. In addition, the map $\abs{\cdot}_n : \Subd(n\mathcal P) \to n\mathcal P$ which sends $X$ to $\abs{X}_n$ is a poset map over $\mathcal P$. In particular, it is a poset map, so $(\Subd(n\mathcal P), \abs{\cdot}_n)$ is an $n\mathcal P$-poset. Now suppose $(\mathcal A, p)$ is an $n\mathcal P$-poset. Then $(\mathcal A, \plus \circ p)$ is a $\mathcal P$-poset. For the rest of the paper, whenever we define an $n \mathcal P$-poset $(\mathcal A, p)$, we will also implicitly treat $\mathcal A$ as a $\mathcal P$-poset as above. In particular, we can define $\Subd(\mathcal A)$ as the poset of $(\mathcal A, \plus \circ p)$-subdivisions. Let $(\mathcal A, p)$ be an $n\mathcal P$-poset, and let $X$ be an $\mathcal A$-complex. Then $p(X)$ is an $n\mathcal P$ complex. We call $\abs{p(X)}_n$ the \emph{$n$-support} of $X$, and for convenience we denote it by $\abs{X}_n$. Suppose additionally that $X$ is an $\mathcal A$-subdivision. By Proposition~\ref{prop:mapofsubdivisions}, the map $p : \Subd(\mathcal A) \to \Subd(n\mathcal P)$ given by $Y \mapsto p(Y)$, is a poset map over $\mathcal P$. As noted previously, $\abs{\cdot}_n : \Subd(n\mathcal P) \to n\mathcal P$ is also a poset map over $\mathcal P$. Hence, the map $\abs{\cdot}_n : \Subd(\mathcal A) \to n\mathcal P$ given by $X \mapsto \abs{p(X)}_n = \abs{X}_n$ is a poset map over $\mathcal P$. It follows that $\abs{X} = \Sum \abs{X}_n$. In addition, $(\Subd(\mathcal A), \abs{\cdot}_n)$ is an $n\mathcal P$-poset. We note the following fact for later. \begin{prop}\label{prop:nsupporttrivial} Let $(\mathcal A,p)$ be an $n \mathcal P$-poset and let $X$ be an $\mathcal A$-subdivision. If $X$ has a single maximal element $x$, then $\abs{X}_n = p(x)$. \end{prop} \begin{proof} If $X$ has a single maximal element $x$, then $p(X)$ has a single maximal element $p(x)$. Therefore, \begin{align*} \abs{X}_n &= \left( \bigcup_{(P_1,\dots,P_n)\in p(X)} P_k \right)_{k=1}^n \\ &= \left( p(x)_k \right)_{k=1}^n \\ &= p(x). \end{align*} \end{proof} \subsection{Canonical subdivisions} Let $(\mathcal A, p)$, $(\mathcal A', p')$ be $n\mathcal P$-posets. We define a \emph{canonical subdivsion} over $n \mathcal P$ to be any poset map over $n \mathcal P$ from $(\mathcal A,p)$ to $(\Subd(\mathcal A'),\abs{\cdot}_n)$. The importance of these maps is the following proposition. \begin{prop} \label{prop:canonicalrefinement} Let $\Sigma : \mathcal A \to \Subd(\mathcal A')$ be a canonical subdivision over $n\mathcal P$. For $X$ an $\mathcal A$-complex, define \[ \Sigma^\ast(X) := \bigcup_{x \in X} \Sigma(x). \] Then $\Sigma^\ast(X)$ is an $\mathcal A'$-complex with $n$-support $\abs{X}_n$. Moreover, the map $\Sigma^\ast : \Subd(\mathcal A) \to \Subd(\mathcal A')$ given by $X \mapsto \Sigma^\ast(X)$ is a canonical subdivision over $n\mathcal P$. \end{prop} Before proving this, we set some notation. Let $(\mathcal A, p)$, $(\mathcal A', p')$ be $n\mathcal P$-posets. Let $\mathcal A_0$ be any subset of $\mathcal A$. A \emph{subdivision} over $n\mathcal P$ is any map of sets $\Sigma : \mathcal A_0 \to \Subd(\mathcal A')$ such that $p(x) = \abs{\Sigma(x)}_n$ for all $x \in \mathcal A_0$. A canonical subdivision is therefore a subdivision $\Sigma : \mathcal A \to \Subd(\mathcal A')$ which is a poset map. Let $\Sigma : \mathcal A_0 \to \Subd(\mathcal A')$ be a subdivision over $n\mathcal P$, and let $x \in \mathcal A_0$ and $y \in \mathcal A$ with $y \le x$. The \emph{restriction} of $\Sigma$ from $x$ to $y$, denoted by $\Sigma(x|y)$, is the unique $Y \in \Subd(\mathcal A')$ such that $Y \le \Sigma(x)$ and $\abs{Y}_n = p(y)$. Equivalently, this is the unique $Y \in \Subd(\mathcal A')$ such that $Y \le \Sigma(x)$ and $\abs{Y} = (\Sum \circ p)(y)$. We note the following: \begin{prop} \label{prop:canonicalcriterion} A subdivision $\Sigma : \mathcal A \to \Subd(\mathcal A')$ over $n\mathcal P$ is a poset map (and hence a canonical subdivision over $n\mathcal P$) if and only if $\Sigma(x|y) = \Sigma(y)$ for all $x$, $y \in \mathcal A$ such that $y \le x$. \end{prop} \begin{proof} If $\Sigma$ is a poset map and $x$, $y \in \mathcal A$ such that $y \le x$, then $\Sigma(y) \le \Sigma(x)$. Since $\abs{\Sigma(y)}_n = p(y)$ by the definition of a subdivision over $n\mathcal P$, we must have $\Sigma(x|y) = \Sigma(y)$ by the definition of $\Sigma(x|y)$. Conversely, if $\Sigma(x|y) = \Sigma(y)$ for all $x$, $y \in \mathcal A$ such that $y \le x$, then in particular $\Sigma(y) = \Sigma(x|y) \le \Sigma(x)$, so $\Sigma$ is a poset map. \end{proof} \begin{proof}[Proof of Proposition~\ref{prop:canonicalrefinement}] Let $x$, $y \in X$ such that $(\Sum \circ p)(x)$ and $(\Sum \circ p)(y)$ share a face. Let $z$ be the element of $X$ such that $p(z)$ is this face, so $z \le x$ and $z \le y$. Since $\Sigma$ is a canonical subdivision, by Proposition~\ref{prop:canonicalcriterion} we have $\Sigma(x|z) = \Sigma(y|z) = \Sigma(z)$. Thus, the polytopal complexes $(\Sum \circ p')(\Sigma(x))$ and $(\Sum \circ p')(\Sigma(y))$ have the subdivision $(\Sum \circ p')(\Sigma(z))$ as their common intersection. This implies that $\Sigma^\ast(X)$ is indeed an $\mathcal A'$-complex. We next show that $\abs{\Sigma^\ast(X)}_n = \abs{X}_n$. We have \[ \abs{\Sigma^\ast(X)}_n = (Q_1,\dots,Q_n) \] where \begin{align*} Q_k &= \bigcup_{(P_1,\dots,P_n) \in p ((\Sigma^\ast(X))} P_k \\ &= \bigcup_{x \in X} \bigcup_{(P_1,\dots,P_n) \in p (\Sigma(x))} P_k \\ &= \bigcup_{x \in X} (k^\text{th} \text{ entry of } \abs{\Sigma(x)}_n) \\ &= \bigcup_{x \in X} (k^\text{th} \text{ entry of } p(x)) \\ &= \bigcup_{(P_1,\dots,P_n) \in p(X)} P_k \\ &= k^\text{th} \text{ entry of } \abs{X}_n \end{align*} where we have used in the fourth line that $p(x) = \abs{\Sigma(x)}_n$ since $\Sigma$ is a poset map over $n\mathcal P$. Hence $\abs{\Sigma^\ast(X)}_n = \abs{X}_n$, as desired. It follows that we have a map $\Sigma^\ast : \Subd(\mathcal A) \to \Subd(\mathcal A')$ that is a subdivision over $n \mathcal P$. To show that it is canonical, we need to show it is a poset map. Let $X$, $Y \in \Subd(\mathcal A)$ such that $Y \le X$. Thus $\abs{Y}$ is a face of $\abs{X}$, and hence $\abs{\Sigma^\ast(Y)}$ is a face of $\abs{\Sigma^\ast(X)}$ since $\abs{\Sigma^\ast(X)} = \abs{X}$ for all $X \in \Subd(\mathcal A)$. Moreover, since $Y$ is a subcomplex of $X$, it is easy to see from the definition of $\Sigma$ that $\Sigma(Y)$ is a subcomplex of $\Sigma(X)$. Thus $\Sigma^\ast(Y) \le \Sigma^\ast(X)$, as desired. \end{proof} \begin{exam} For any perfect $n \mathcal P$-poset $(\mathcal A,p)$, recall the trivial subdivision $\triv_{\mathcal A} : \mathcal A \to \Subd(\mathcal A)$ given by $\triv_{\mathcal A}(x) = \langle x \rangle_{\mathcal A}$. By Proposition~\ref{prop:nsupporttrivial}, we have $\abs{\triv_{\mathcal A}(x)}_n = p(x)$, hence $\triv_{\mathcal A}$ is a subdivision over $n \mathcal P$. It is also clear that if $x$, $y \in \mathcal A$ with $y \le x$ then $\triv_{\mathcal A}(x|y) = \triv_{\mathcal A}(y)$, so $\triv_{\mathcal A}$ is a canonical subdivision over $n \mathcal P$. \end{exam} \subsection{Confluent subdivisions} \label{sec:confluent} In this paper we will construct very complicated subdivisions recursively from smaller, simpler subdivisions. The purpose of this section is to give a systematic way to prove that the final subdivisions are well-defined and canonical. The key idea is the notion of confluence and Newman's lemma. Let $(\mathcal A, p)$ be an $n\mathcal P$-poset. Let $\{\mathcal A_\alpha\}$ be a (possibly infinite) collection of subsets of $\mathcal A$. For each $\alpha$, let $\sigma_\alpha : \mathcal A_\alpha \to \mathcal \Subd(\mathcal A)$ be a subdivision over $n\mathcal P$. Let $X$ be any finite subset of $\mathcal A$. Suppose $x$ is an element of $X$ and $x \in \mathcal A_\alpha$ for some $\alpha$. We define a \emph{$\sigma_\alpha$-move} on $X$ at $x$ to be the act of transforming $X$ into the set \[ X' := X \setminus \{x\} \cup \Max_{\mathcal A}\sigma_\alpha(x). \] We write this move as $X \xrightarrow{x,\sigma_\alpha} X'$, or simply $X \to X'$ if we do not need to specify $(x,\sigma_\alpha)$. We call a move $X \to X'$ \emph{non-trivial} if $X \neq X'$. Given another finite set $Y \subset \mathcal A$, we write $X \xrightarrow{\ast} Y$ if there exists a sequence of moves $X \to X_1 \to X_2 \to \dots \to Y$. We allow this sequence to contain only one term; in other words, we always have $X \xrightarrow{\ast} X$. Suppose that $X$, $Y \subset \mathcal A$ are finite. We say that $X$ and $Y$ are \emph{joinable} if there exists $Z \subset \mathcal A$ such that $X \xrightarrow{\ast} Z$ and $Y \xrightarrow{\ast} Z$. We say that the family of subdivisions $\{\sigma_\alpha\}$ is \emph{locally confluent} if for any finite $X \subset \mathcal A$ and any two moves $X \to Y_1$, $X \to Y_2$, we have that $Y_1$ and $Y_2$ are joinable. Note that this is equivalent to saying that for all $x \in \mathcal A$ and $\alpha$, $\beta$ such that $x \in \mathcal A_\alpha \cap \mathcal A_\beta$, we have that $\Max \sigma_\alpha(x)$ and $\Max \sigma_\beta(y)$ are joinable. We call a set $X \subset \mathcal A$ \emph{terminal} if there are no non-trivial moves from $X$. We call an element $x \in \mathcal A$ \emph{terminal} if $\{x\}$ is terminal. An element $x$ is terminal if and only if for every $\alpha$ such that $x \in \mathcal A_\alpha$, we have $x \in \sigma_\alpha(x)$. (This is because $\sigma_\alpha(x)$ is an $\mathcal A$-subdivision of $(\Sum \circ p)(x)$, so if $x \in \sigma_\alpha(x)$, then $x$ must be the only maximal element of $\sigma_\alpha(x)$.) From this, it follows that a set is terminal if and only if all its elements are terminal. We say that $\{\sigma_\alpha\}$ is \emph{terminating} if there are is no infinite sequence $X_1 \to X_2 \to X_3 \to \dots$ of non-trivial moves. Finally, we say that $\{\sigma_\alpha\}$ is \emph{facially compatible} if the following two properties hold: \begin{itemize} \item If $x$, $y \in \mathcal A$ such that $y \le x$ and $x \in \mathcal A_\alpha$, then $\Max\sigma_\alpha(x|y)$, $\{y\}$ are joinable. \item If $x$, $y \in \mathcal A$ such that $y \le x$ and $x$ is terminal, then $y$ is terminal. \end{itemize} Our main result is the following. \begin{thm} \label{thm:confluence} Let $(\mathcal A, p)$ be a perfect $n\mathcal P$-poset and let $\{\sigma_\alpha : \mathcal A_\alpha \to \Subd(\mathcal A)\}$ be a family of locally confluent, facially compatible, and terminating subdivisions. Then for any $x \in \mathcal A$, there is a unique terminal set $S(x) \subset \mathcal A$ such that $\{x\} \xrightarrow{\ast} S(x)$. Moreover, $\Sigma(x) := \langle S(x) \rangle_{\mathcal A}$ is an $\mathcal A$-subdivision, and the map $\Sigma : \mathcal A \to \Subd(\mathcal A)$ given by $x \mapsto \Sigma(x)$ is a canonical subdivision over $n \mathcal P$. \end{thm} \begin{proof} The fact that $S(x)$ exists and is unique follows directly from Newman's diamond lemma, which states that any locally confluent and terminating binary relation is globally confluent \cite{New42}. (In this case, the binary relation is non-trivial moves ``$\to$''.) It remains to show that $\Sigma(x)$ is an $\mathcal A$-subdivision, and the map $\Sigma : \mathcal A \to \Subd(\mathcal A)$ is a canonical subdivision over $n \mathcal P$. Define a binary relation $\prec$ on $\mathcal A$ by $y \prec x$ if there is a non-trivial move $\{x\} \to Y$ such that $y \in Y$. We prove the following: {\bf Claim:} $\prec$ is a well-founded relation on $\mathcal A$. \begin{proof}[Proof of claim] Suppose that $x_1 \succ x_2 \succ x_3 \succ \dots$ is an infinite descending sequence. Define a sequence of non-trivial moves $X_1 \to X_2 \to \dots$, inductively as follows. Define $X_1 = \{x_1\}$. Next fix $k \ge 2$, and assume by induction that we have constructed $X_1 \to \dots \to X_{k-1}$ such that $x_j \in X_j$ for all $1 \le j \le k-1$. By definition of $\prec$, there exists a non-trivial move $x_{k-1} \xrightarrow{(x_{k-1},\sigma_\alpha)} Y_k$ such that $x_k \in Y_k$. Define $X_k$ by $X_{k-1} \xrightarrow{(x_{k-1},\sigma_\alpha)} X_k = X_{k-1} \setminus \{x_{k-1}\} \cup Y_k$. Since the move $x_{k-1} \to Y_k$ is non-trivial, we have $x_{k-1} \notin Y_k$. Thus $x_{k-1} \notin X_k$, so $X_{k-1} \neq X_k$. Hence we have a non-trivial move $X_{k-1} \to X_k$ with $x_k \in X_k$, completing the induction. However, this contradicts the terminating property, proving the claim. \end{proof} We now prove by induction on $\prec$ that \begin{enumerate}[label=(\alph*)] \item $\Sigma(x)$ is an $\mathcal A$-subdivision \item $\abs{\Sigma(x)}_n = p(x)$ \item $\Sigma(x|y) = \Sigma(y)$ for all $y \le x$. \end{enumerate} By Proposition~\ref{prop:canonicalcriterion}, this will complete the proof. For the base case, assume that $x$ is terminal. Thus $S(x) = \{x\}$, so $\Sigma(x) = \langle x \rangle_{\mathcal A} = \triv_{\mathcal A}(x)$, since $\mathcal A$ is perfect. Hence $\Sigma(x)$ is an $\mathcal A$-subdivision, proving (a). Proposition~\ref{prop:nsupporttrivial} implies (b). Suppose $y \le x$. By facial compatibility, $y$ is also terminal, so $\Sigma(y) = \triv_{\mathcal A}(y)$. Since $\triv_{\mathcal A}(x|y) = \triv_{\mathcal A}(y)$, (c) is proved. For the inductive step, assume that $x$ is not terminal. Let $\{x\} \xrightarrow{(x,\sigma_\alpha)} Y$ be a non-trivial move. In particular, we have $x \notin Y$. Then $S(x) = \bigcup_{y \in Y} S(y)$, and hence $\Sigma(x) = \bigcup_{y \in Y} \Sigma(y)$. By induction, we have proved (a)-(c) for all $y \in Y$. Thus, the proof of Proposition~\ref{prop:canonicalrefinement} implies (a) and (b) for $x$. For (c), let $y \le x$. By facial compatibility, we have that $\Max\sigma_\alpha(x|y)$ and $\{y\}$ are joinable. It follows that \[ S(y) = \bigcup_{z \in \Max\sigma_\alpha(x|y)} S(z) \] and hence \begin{equation} \label{eq:Sigmay} \Sigma(y) = \bigcup_{z \in \Max\sigma_\alpha(x|y)} \Sigma(z). \end{equation} For every $z \in \Max\sigma_\alpha(x|y)$, there is $z' \in \Max \sigma_\alpha(x) = Y$ such that $z \le z'$. Since $z' \neq x$ by assumption that $\{x\} \to Y$ is non-trivial, we have $z' \prec x$, and therefore by induction $\Sigma(z) = \Sigma(z'|z)$. Hence $\Sigma(y)$ is a union of elements of the form $\Sigma(z'|z)$ with $z' \in Y$. Since $\Sigma(x) = \bigcup_{z' \in Y} \Sigma(z')$ and $\Sigma(z'|z)$ is a subcomplex of $\Sigma(z')$, it follows that $\Sigma(y)$ is a subcomplex of $\Sigma(x)$. On the other hand, the formula \eqref{eq:Sigmay} implies the support of $\Sigma(y)$ is $\abs{y}$. So we must have $\Sigma(x|y) = \Sigma(y)$, completing the proof. \end{proof} \subsection{Examples of canonical subdivisions} \label{sec:canonicalexamples} In this section we give two examples of well-known subdivisions which can be expressed as canonical subdivisions in our sense. This section can be skipped without logically affecting the main proof, but the ideas may be useful for understanding the later arguments. \subsubsection{Pulling triangulations} Let $(\mathcal S, \conv)$ be the $\mathcal P$-poset from Example~\ref{exam:pointsets}. Let $A \in \mathcal S$. A \emph{covector} of $A$ is a point $x \in A$ such that $\dim(A \setminus \{x\}) < \dim(A)$. (Here, $\dim(A)$ denotes the dimension of the smallest affine subspace containing $A$.) Every element of $A$ is a covector if and only if $A$ is affinely independent, i.e. $A$ is the set of vertices of a simplex. Let $\mathcal S^\ast$ be the set of elements of $\mathcal S$ which are not affinely independent. We define a subdivision $\pull : \mathcal S^\ast \to \Subd(\mathcal S)$ as follows. Let $A \in \mathcal S^\ast$, and let $x$ be the smallest element of $A$ (according to the order on $A$) which is not a covector. We define $\pull(A)$ to be the set of all $B$ and $x \cup B$ such that $B \le_{\mathcal S} A$ and $x \notin B$. Then $\pull(A)$ is an $\mathcal S$-subdivision of $\abs{A}$, so we have a subdivision $\pull : \mathcal S^\ast \to \Subd(\mathcal S)$ over $\mathcal P$. We now claim that the family $\{\pull\}$ consisting of a single subdivision is locally confluent, terminating, and facially compatible. Local confluence is trivial since the family has only one subdivision. If $A \in \mathcal S^\ast$ and $B \in \Max \pull(A)$, then $B \subsetneq A$, which proves termination. Finally, suppose $A \in \mathcal S^\ast$ and $B \le_{\mathcal S} A$. If $x \notin B$, then $\pull(A|B) = \triv_{\mathcal S}(B)$. If $x \in B$, then we can check that \[ \pull(A|B) = \begin{dcases*} \pull(B) & if $x$ is not a covector of $B$ \\ \triv_{\mathcal S}(B) & otherwise. \end{dcases*} \] In all cases, $\Max\pull(A|B)$ and $\{B\}$ are joinable, proving the first condition of facial compatibility. For the second property, we note that $A$ is terminal if and only if $A \notin \mathcal S^\ast$, i.e. $A$ is affinely independent. Clearly if this holds for $A$ then it holds for $B$, completing the proof. Thus, by Theorem~\ref{thm:confluence}, we have a canonical subdivision $\Pull : \mathcal S \to \Subd(\mathcal S)$ such that if $A \in \mathcal S$ and $B \in \Pull(A)$, then $B$ is terminal, i.e. affinely independent. This subdivision is known in the literature as the \emph{pulling triangulation}. \subsubsection{Dicing} Let $\mathscr H$ be a finite set of hyperplanes in $\mathbb R^d$. For each $H \in \mathscr H$, let $\mathcal P_H$ be set of polytopes in $\mathbb R^d$ whose relative interior intersects $H$. We define a subdivision $\dice_H : \mathcal P_H \to \Subd(\mathcal P)$ as follows. Let $H_1$, $H_2$ be the two closed half-spaces of $\mathbb R^d$ cut out by $H$. For $P \in \mathcal P_H$, we define \[ \dice_H(P) = \triv(P \cap H_1) \cup \triv(P \cap H_2) \] Then $\dice_H(P)$ is a polytopal subdivision of $P$ and $\dice_H : \mathcal P_H \to \Subd(\mathcal P)$ is a subdivision over $\mathcal P$. We claim that the family $\{ \dice_H \}_{H \in \mathscr H}$ is locally confluent, terminating, and facially compatible. We start with terminating. If $A \in \mathcal P_H$ and $B \in \Max\dice(A)$, then $B \notin \mathcal P_H$. Since $\mathscr H$ is finite, this implies $\{ \dice_H \}_{H \in \mathscr H}$ is terminating. We next show facial compatibility. Let $A \in \mathcal P_H$ and let $B \le_{\mathcal P} A$. If $B \in \mathcal P_H$, then $\dice_H(A|B) = \dice_H(B)$. Otherwise, $\dice_H(A|B) = \triv(B)$. Either way, $\Max \dice_H(A|B)$ and $\{B\}$ are joinable, proving the first condition of facial compatibility. Next, note that $A$ is terminal if and only if its relative interior does not intersect any $H \in \mathscr H$. If this holds for $A$ then it clearly holds for $B$, proving facial compatibility. Finally, we prove local confluence. Suppose $P \in \mathcal P_{G} \cap \mathcal P_{H}$ for distinct $G$, $H \in \mathscr H$. Consider $\Max \dice_G(P)$ and $\Max \dice_H(P)$. If we apply a $\dice_H$-move to both elements of $\Max \dice_G(P)$, then we obtain the same result as when we apply $\dice_G$-move to both elements of $\Max \dice_H(P)$. Thus $\Max \dice_G(P)$ and $\Max \dice_H(P)$ are joinable, so $\{ \dice_H \}_{H \in \mathscr H}$ is locally confluent. Thus, by Theorem~\ref{thm:confluence}, we have a canonical subdivision $\Dice : \mathcal P \to \Subd(\mathcal P)$ such that for all $P \in \mathcal P$ and $Q \in \Dice(P)$, the relative interior of $Q$ does not intersect any $H \in \mathscr H$. This is of course the subdivision obtained by intersecting $P$ with each of the closed regions of $\mathbb R^d$ cut out by $\mathscr H$. \section{Cayley polytopes} \label{sec:Cayleypolytopes} \subsection{Notation} Before proceeding, we set some notation regarding tuples and matrices. Let $a = (a_1, \dots, a_m)$ be an $m$-tuple (entry type unspecified). We allow $m=0$, in which case the tuple has no entries. We define $\abs{a} := m$. For $I \subset [m]$, we use $a_I$ to denote the tuple $(a_{i_1}, \dots, a_{i_k})$ where $I = \{i_1,\dots,i_k\}$ and $i_1 < \dots < i_k$. We use $a_{\setminus j}$ to denote $a_{[m] \setminus \{j\}}$. Similarly, let $a = (a_{ij})_{i,j=1}^{m,n}$ be an $m \times n$ matrix and $I \subset [m]$, $J \subset [n]$. We allow $m=0$ or $n=0$, in which case the matrix has no entries. We use $a_{I \times \bullet}$ to denote the matrix obtained by restricting $a$ to the rows indexed by $I$, preserving the order of the rows. We similarly define $a_{\bullet \times J}$ and $a_{I \times J}$. We use $a_{\setminus i \times \bullet}$ to denote $a_{I \setminus \{i\} \times \bullet}$, and similarly define $a_{\bullet \times \setminus j}$ and $a_{\setminus i \times \setminus j}$. \subsection{Cayley sums} \label{sec:Cayley} Let $P = (P_1,\dots,P_m)$ be an $m$-tuple of polytopes in $\mathbb R^d$, where $m \ge 1$. We say that $P$ is in \emph{Cayley position} if there exists a linear map $\pi: \mathbb R^d \to \mathbb R^d$ such that $\pi(P_i)$ is a point for all $i$ and the sequence of points $(\pi(P_i))_{i \in [m]}$ is affinely independent. In this situation, we define the \emph{Cayley sum} $\Cay(P)$ to be the convex hull of the entries of $P$. The faces of a Cayley sum $\Cay(P)$ are precisely Cayley sums of the form $\Cay(F_I)$, where $I$ is a nonempty subset of $[m]$ and $F = (F_1, \dots, F_m)$ where $F \le P$. (Recall that $F \le P$ means that there exists a linear functional such that $F_1$, \dots, $F_m$ are the faces of $P_1$, \dots, $P_m$, respectively, with respect to this linear functional.) Let $S = (S_1,\dots,S_n)$ be an $n$-tuple of independent simplices (recall the definition of independence from Section~\ref{sec:polytopes}), and let $a = (a_{ij})_{i,j=1}^{m,n}$ be an $m \times n$ matrix of nonnegative integers, where $m \ge 1$ and $n \ge 0$. We consider Cayley sums of the form $\Cay(P)$, where $P = (P_1,\dots,P_m)$ is an $m$-tuple of polytopes in Cayley position, and for each $i \in [m]$ we have \[ P_i = p_i + \sum_{j = 1}^n a_{ij} S_j \] for some $p_i \in \mathbb R^d$. The faces of $\Cay(P)$ are as follows. Let $S' = (S'_1,\dots,S'_n)$ be any $n$-tuple of polytopes such that $S'_j \le S_j$ for all $j$. Since the entries of $S$ are independent, this is equivalent to $S' \le S$. Let $P' = (P_1',\dots,P_m')$ be the tuple with \[ P_i' = p_i + \sum_{j=1}^n a_{ij} S_j'. \] Then $P' \le P$. Thus, for any nonempty $I \subset [m]$, $\Cay(P'_I)$ is a face of $\Cay(P)$. All faces of $\Cay(P)$ arise this way. \subsection{The poset $\mathcal C$} Let $\mathcal C_0$ be the poset defined as follows. Its elements are all tuples $(p,S,a)$ where \begin{itemize} \item $p = (p_1,\dots,p_m)$ is a tuple of points in $\mathbb R^d$, for some positive integer $m$. \item $S = (S_1,\dots,S_n)$ is a tuple of independent ordered integral simplices in $\mathbb R^d$, for some nonnegative integer $n$. \item $a = (a_{ij})_{i,j=1}^{m,n}$ is an $m \times n$ matrix of nonnegative integers. \item The polytopes $P_1$, \dots, $P_m$ are in Cayley position, where \[ P_i := p_i + \sum_{j = 1}^n a_{ij} S_j. \] \end{itemize} We equip $\mathcal C_0$ with the partial order $\le _{\mathcal C_0}$, where \[ (p_I,S',a_{I \times \bullet}) \le_{\mathcal C_0} (p,S,a) \] if $I$ is a nonempty subset of $[\abs{p}]$ and $S' \le S$. It is easy to see that this is a partial order. Let $\Cay : \mathcal C_0 \to \mathcal P$ be the map defined by \[ \Cay(p,S,a) = \Cay \left(p_i + \sum_{j = 1}^n a_{ij} S_j \right)_{i \in [\abs{p}]} \] By the discussion from the previous subsection, this is a poset map. Thus $(\mathcal C_0, \Cay)$ is a $\mathcal P$-poset. We now define an equivalence relation $\sim$ on $\mathcal C_0$ as follows. If $j \in [\abs{S}]$ is such that $S_j$ is a point or $a_{ij} = 0$ for all $i \in [\abs{p}]$, then we set \[ (p,S,a) \sim ((p_i+a_{ij}S_j)_{i \in [\abs{p}]},S_{\setminus j},a_{\bullet \times \setminus j}). \] We define $\mathcal C$ to be $\mathcal C_0 / \sim$. For $A$, $B \in \mathcal C$, we let $A \le_{\mathcal C} B$ if there are representatives $A_0$, $B_0$ of $A$ and $B$, respectively, in $\mathcal C_0$ such that $A_0 \le_{\mathcal C_0} B_0$. It is straightforward to check that this defines a partial order on $\mathcal C$ (for example, using standard form defined in the next subsection). Furthermore, we have that $\Cay : \mathcal C_0 \to \mathcal P$ is constant on equivalence classes of $\sim$. Thus there is a well-defined map $\Cay : \mathcal C \to \mathcal P$, this is a poset map, and $(\mathcal C, \Cay)$ is a $\mathcal P$-poset. If there is no risk of confusion, we will abuse notation and denote elements in $\mathcal C$ using their representatives in $\mathcal C_0$. \subsubsection{Standard form and $L(A)$} Given an object $A \in \mathcal C$, there is a unique representative $(p,S,a)$ of $A$ in $\mathcal C_0$ such that for all $j$, the $j$-th column of $a$ is not all zeroes. We call this the \emph{standard form} of $A$. It is easy to check that if $A_0$ is the standard form of $A \in \mathcal C$, then $B \le_{\mathcal C} A$ if and only if $B$ has a representative $B_0 \in \mathcal C_0$ such that $B_0 \le_{\mathcal C_0} A_0$. (In fact, this is true if $A_0$ is any representative of $A$, which can be proved from the previous sentence.) Using standard form, it is not hard to see that if $A \in \mathcal C$ and we have two different elements $B$, $C \in \mathcal C$ such that $B \le_{\mathcal C} A$ and $C \le_{\mathcal C} A$, then $\Cay(B)$ and $\Cay(C)$ are different faces of $\Cay(A)$. Moreover, for every face $F$ of $\Cay(A)$, there is a face $B$ of $A$ such that $\Cay(B) = F$. It follows that $\mathcal C$ is a perfect $\mathcal P$-poset. Let $A \in \mathcal C$ with standard form $(p,S,a)$. For $j \in [\abs{S}]$, let $v_j$ be the first vertex of $S_j$. Define \[ S_0(A) := \conv \left( p_i + \sum_{j=1}^{\abs{S}} a_{ij} v_j \right)_{i \in [\abs{p}]}. \] In other words, $S_0(A)$ is the face of $\Cay(A)$ whose vertices are the first vertices of each Cayley summand of $\Cay(A)$. Note that $S_0(A)$, $S_1$, \dots, $S_{\abs{S}}$ are independent simplices. We define \begin{align*} L(A) &:= L(S_0(A) + S_1 + \dots + S_{\abs{S}}) \\ &= L(S_0(A)) \oplus L(S_1) \oplus \dots \oplus L(S_{\abs{S}}). \end{align*} \subsection{$\gamma_T$ subdivisions} \label{sec:T} \subsubsection{Definition of $\gamma_T$} Let $T$ be an ordered integral simplex of dimension at least 1. Let $\mathcal C_T$ be the set of elements of $\mathcal C$ whose standard form $(p,S,a)$ has the property that $T$ is an entry of $S$. In this section we construct a subdivision $\gamma_T : \mathcal C_T \to \Subd(\mathcal C)$ over $\mathcal P$ such that the family $\{\gamma_T\}$ of all such subdivisions satisfies the conditions of Theorem~\ref{thm:confluence}. These subdivisions will be a main building block for future subdivisions. Let $A \in \mathcal C_T$ with standard form $(p,S,a)$. Let $j$ be the unique number such that $S_j = T$, and let $i$ be the smallest number such that $a_{ij} = \max_{i'j} a_{i'j}$. Let $v$ be the first vertex of $T$, and let $f$ be the facet of $T$ opposite $v$. Note that for any positive integer $c$, the dilated simplex $cT$ can be written as the union of the two polytopes \[ v + (c-1)T, \quad \Cay((c-1)f, cf). \] (Here, we are using the ordinary Cayley sum as defined in Section~\ref{sec:Cayley}.) This can be extended to a $\mathcal C$-subdivision of $\Cay(A)$ as follows. Recall that the standard form of $A$ is $(p,S,a)$. Let $m := \abs{p}$ and $n := \abs{S}$. We define \begin{equation} \label{eq:A'A''} \begin{aligned} A' &= (p',S,a') \in \mathcal C \\ A'' &= (p'', F, a'') \in \mathcal C \end{aligned} \end{equation} where \begin{itemize} \item $p'$ is the $m$-tuple obtained by replacing the $i$-th entry of $p$ with $p_{i} + v$. \item $a'$ is the $m \times n$ matrix obtained by subtracting 1 from the $(i,j)$ entry of $a$. \item $p''$ is the $(m+1)$-tuple obtained by inserting $p_{i} + v$ directly before the $i$-th entry of $p$. \item $F$ is the $n$-tuple obtained by replacing the $j$-th entry of $S$ with $f$. \item $a''$ is the $(m+1) \times n$ matrix obtained by inserting the $i$-th row of $a'$ directly above the $i$-th row of $a$. \end{itemize} We define \[ \gamma_T(A) := \triv_{\mathcal C}(A') \cup \triv_{\mathcal C}(A''). \] Then $\gamma_T(A)$ is a $\mathcal C$-subdivision of $\Cay(A)$. If $a_{ij} = 1$ and $a_{i'j} = 0$ for all $i' \neq i$, then this subdivision has a unique maximal element $A''$. Otherwise, the maximal elements are $A'$ and $A''$. \subsubsection{Properties of $\gamma_T$} We first prove the following: \begin{prop} \label{prop:LgammaT} Let $A \in \mathcal C_T$ and $B \in \Max \gamma_T(A)$. Then $L(A) = L(B)$. \end{prop} \begin{proof} We have either $B = A'$ or $B = A''$. In the former case, we have $S_0(A) = S_0(B)$ and $A$ and $B$ have the same second entry, so $L(A) = L(B)$ as desired. Assume $B = A''$. Without loss of generality, assume $i = j = 1$. For convenience, we will reuse the variables $i$ and $j$ in the below proof. Let $w$ be the second vertex of $T$. We have \begin{align*} S_0(B) &= \conv \left( \left\{ p_i + a_{i1} w + \sum_{j=2}^{\abs{S}} a_{ij} v_j \right\}_{i \in [\abs{p}]} \cup \right. \\ &\qquad\qquad \left. \left\{ p_{1} + v + (a_{11}-1)w + \sum_{j=2}^{\abs{S}} a_{1j}v_j \right\} \right) \\ &= \conv \left( \left\{ p_i + a_{i1} (w-v) + \sum_{j=1}^{\abs{S}} a_{ij} v_j \right\}_{i \in [\abs{p}]} \cup \right. \\ &\qquad\qquad \left. \left\{ p_{1} + (a_{11}-1)(w-v) + \sum_{j=1}^{\abs{S}} a_{1j}v_j \right\} \right). \end{align*} Subtracting the last entry in the above convex hull from the first entry, we get $w-v$. Hence $w-v \in L(S_0(B))$. It is then not hard to see from the above expression that $L(S_0(B)) = L(S_0(A)) \oplus \mathbb Z\langle w-v \rangle$. Thus, \begin{align*} L(B) &= L(S_0(A)) \oplus \mathbb Z\langle w-v \rangle \oplus L(f) \oplus L(S_2) \oplus \dots \oplus L(S_{\abs{S}}) \\ &= L(S_0(A)) \oplus L(T) \oplus L(S_2) \oplus \dots \oplus L(S_{\abs{S}}) \\ &= L(A) \end{align*} as desired. \end{proof} \begin{prop} \label{prop:facegammaT} Let $A \in \mathcal C_T$ and suppose $\mathcal B \in \mathcal C$ such that $B \le_{\mathcal C} A$. Then either $\gamma_T(A|B) = \gamma_{T'}(B)$ for some $T'$ or $\gamma_T(A|B) = \triv_{\mathcal C}(B)$. \end{prop} \begin{proof} Let the standard form of $A$ be $(p,S,a)$. Let $B = (p_I,S',a_{I \times \bullet})$ with $S' \le S$ and $I$ a nonempty subset of $[\abs{p}]$. Assume that $T = S_j$, and let $i$ be the smallest number such that $a_{ij} = \max_{i'} a_{i'j}$. If $i \in I$ and $S_j'$ contains the first vertex of $T$ and another vertex, then $\gamma_T(A|B) = \gamma_{S_j'}(B)$. Otherwise, we have $\gamma_T(A|B) = \triv_{\mathcal C}(B)$. \end{proof} We now prove the following. \begin{prop} \label{prop:confgammaT} The family of subdivisions $\{ \gamma_T \}$, where $T$ ranges over all ordered integral simplices of dimension at least 1, is locally confluent, terminating, and facially compatible. \end{prop} \begin{proof} We first prove local confluence. Let $A \in \mathcal C$ with standard form $(p,S,a)$, and let $T_1$, $T_2$ be distinct entries of $S$. We need to show that $\Max \gamma_{T_1}(A)$ and $\Max \gamma_{T_2}(A)$ are joinable. Without loss of generality, assume $T_1 = S_1$ and $T_2 = S_2$. Let $i_1$, $i_2$ be the smallest numbers such that $a_{i_1 1} = \max_i a_{i1}$ and $a_{i_22} = \max_i a_{i2}$, respectively. First suppose that $i_1 \neq i_2$. Then applying a $\gamma_{T_1}$-move on $\Max \gamma_{T_2}(A)$ at each of its elements yields the same result as applying a $\gamma_{T_2}$-move on $\Max \gamma_{T_1}(A)$ at each of its elements. Hence $\Max \gamma_{T_1}(A)$ and $\Max \gamma_{T_2}(A)$ are joinable, as desired. Now assume $i_1 = i_2$. Let $A_1'$, $A_1'' \in \gamma_{T_1}(A)$ be as in \eqref{eq:A'A''} and define $A_2'$, $A_2'' \in \gamma_{T_2}(A)$ analogously. Consider the following sequence of moves starting from $\gamma_{T_1}(A)$. First, if $A_1' \in \Max(\gamma_{T_1}(A))$, then apply a $\gamma_{T_2}$ move at $A_1'$. Next, apply a $\gamma_{T_2}$ move at $A_1''$. Since $A_1''$ has at least two entries greater than 0 in the second column, $\gamma_{T_2}(A_1'')$ has two maximal elements $(A_1'')'$, $(A_1'')''$, defined analogously to \eqref{eq:A'A''}. Finally, apply a $\gamma_{T_2}$ move at $(A_1'')'$. If we do an analogous sequence of moves starting from $\gamma_{T_2}(A)$ (with $\gamma_{T_1}$ moves replacing $\gamma_{T_2}$ moves), then we obtain the same set as we did with the above sequence. Hence, $\gamma_{T_1}(A)$ and $\gamma_{T_2}(A)$ are joinable, as desired. We next show that $\{ \gamma_T \}$ is terminating. Let $A \in \mathcal C_T$ with standard form $(p,S,a)$. Then if $B \in \mathcal C_T$ with standard from $(p',S',a')$, then either $\sum \dim S'_j < \sum \dim S_j$, or $S = S'$ and $\sum_{i,j} a'_{ij} < \sum_{i,j} a_{ij}$. Since the quantities $\sum \dim S_j$ and $\sum_{i,j} a_{ij}$ are both nonnegative integers, it follows that $\{ \gamma_T \}$ is terminating. Finally, we show that $\{ \gamma_T \}$ is facially compatible. Let $A \in \mathcal C_T$ and suppose $\mathcal B \in \mathcal C$ such that $B \le_{\mathcal C} A$. Proposition~\ref{prop:facegammaT} implies $\Max \gamma_T(A|B)$ and $\{B\}$ are joinable, as desired. Next, suppose $A \in \mathcal C$. From the definition of $\gamma_T$, $A$ is terminal with respect to $\{ \gamma_T \}$ if and only if $A \notin \mathcal C_T$ for any $T$. Thus, if $(p,S,a)$ is the standard form of $A$, then $A$ is terminal if and only if $\abs{S} = 0$. Clearly if this holds for $A$, it holds for any $B \le A$. Hence $\{ \gamma_T \}$ is facially compatible, completing the proof. \end{proof} From the previous two propositions we immediately have the following from Theorem~\ref{thm:confluence}. \begin{thm} \label{thm:Gamma} There is a canonical subdivision $\Gamma : \mathcal C \to \Subd(\mathcal C)$ over $\mathcal P$ such that for any $A \in \mathcal C$ and any $B \in \Max \Gamma(A)$, where $(p,S,a)$ is the standard form of $B$, we have $\abs{S} = 0$ and $L(A) = L(B)$. \end{thm} When restricted to elements $(p,S,a)$ of $\mathcal C$ where $\abs{p} = 1$ and $\abs{S} = 1$, $\Gamma$ is known as the ``canonical triangulation'' of a dilated ordered simplex \cite{HPPS}. \section{Box points and index-lowering} \label{sec:boxpoints} \subsection{Box points} In this section we describe the standard way of reducing the indices of integral polytopes, first defined in \cite{KKMS} as \emph{Waterman points}. We use the terminology of \cite{HPPS} and call them \emph{box points}. Let $S = (S_1,\dots,S_n)$ be a tuple of $n$ independent integral simplices in $\mathbb R^d$. Consider the lattices \begin{align*} L(S) &:= L(S_1) + \dots + L(S_n) \\ N(S) &:= N(S_1) + \dots + N(S_n). \end{align*} We define a \emph{box point} of $S$ to be any nonzero element of the quotient group $G(S) := N(S) / L(S)$. Let $v_1$, \dots, $v_n$ be any vertices of $S_1$, \dots, $S_n$, respectively. For any box point $x$ of $S$, there is a unique tuple $(c_1,\dots,c_n)$ of nonnegative integers such that the polysimplex $\sum_{j=1}^n c_j(S_j-v_j)$ contains exactly one representative of $x$ in $N(S)$.\footnote{Explicitly, $(c_1,\dots,c_n)$ can be described as follows. For each $1 \le j \le n$, let $\{u_1,\dots,u_{k_j}\}$ be the vertices of $S_j$ other than $v_j$, and define $e_j^i := u_i - v_j$. Then the $e_j^i$ form a basis for $L(S)$. Every $x \in G(S)$ has a unique representative in $N(S)$ of the form $\sum_{i,j} x_j^i e_j^i$, where each $0 \le x_j^i < 1$. Then $c_j = \lceil \sum_i x_j^i \rceil$.} The tuple $(c_1,\dots,c_n)$ does not depend on the choice of $v_1$, \dots, $v_n$. We denote the tuple $(c_1,\dots,c_n)$ by $c(S,x)$. We have $0 \le c_j \le d$ for all $j$. We define the \emph{support} of $c$ to be $\supp c := \{ j \in [n] : c_j \neq 0 \}$. Let $S$ be as above, and let $S' \le S$. Then we have a natural inclusion $G(S') \hookrightarrow G(S)$. Thus, any box point of $S'$ can be regarded as a box point of $S$. Suppose $x$ is a box point of $S'$, and thus a box point of $S$. Then $c(S',x) = c(S,x)$. Now, let $L$ be a $d$-dimensional sublattice of $\mathbb Z^d$. Let $x \in \mathbb Z^d / L$ be nonzero. For any tuple $S = (S_1,\dots,S_n)$ of independent integral simplices in $\mathbb Z^d$, we say that $x$ is a \emph{box point} of $S$ if there is $S' \le S$ such that $L(S') = L \cap N(S')$ and $N(S')$ contains a representative of $x$. In this situation, clearly $x$ can be naturally identified with a unique nonzero element of $G(S')$, and hence it can be identified with a unique nonzero element of $G(S)$. Let $x$ be as above. If $A \in \mathcal C$ has standard form $(p,S,a)$, we say that $x$ is a box point of $A$ if $x$ is a box point of $S$. \subsection{A proof of the KMW theorem} \label{sec:KMW} We now use box points to give a short proof of the KMW theorem. The fundamental construction is the same as in the original proof \cite{KKMS}, but the argument structure is simplified due to the use of canonical subdivisions. This section is optional and is not needed in the proof of the main theorem of the paper. Let $S$ be an ordered integral simplex. Let $x$ be a box point of $S$. Let $c$ be the single entry of $c(S,x)$. Let $x_0$ be the unique representative of $x + cv$ in $cS$, where $v$ is the first vertex of $S$. Let $\mathfrak F = \mathfrak F(S)$ be the set of all facets $F$ of $S$ such that $cF$ does not contain $x_0$. Consider the collection of simplices \[ \{ \conv(x_0,cF) : F \in \mathfrak F \}. \] These simplices are the maximal simplices of a triangulation of $cS$, called the \emph{stellar subdivision} of $cS$ at $x_0$. We can make this stellar subdivision a $\mathcal C$-subdivision by considering all elements in $\mathcal C$ with standard form \begin{equation} \label{eq:starsimplex} \left( (x_0,0), (F), \begin{pmatrix} 0 \\ c \end{pmatrix} \right) \end{equation} where $F \in \mathfrak F$. These elements map via $\Cay$ to the maximal elements of the above stellar subdivision. Thus, they are the maximal elements of a $\mathcal C$-subdivision whose image under $\Cay$ is the stellar subdivision. Crucially, for any $F \in \mathfrak F$, the lattice distance $h'$ between $cF$ and $x_0$ is strictly smaller than the lattice distance $h$ between $F$ and the opposite vertex of $S$. Thus, if $A \in \mathcal C$ is of the form \eqref{eq:starsimplex}, we have \[ \ind L(A) = h' \ind L(F) < h \ind L(F) = \ind L(S) \] Now consider the polytope $NcS$, where $N$ is a positive integer. Consider the collection of polytopes \begin{multline*} \{ \{ \conv(x_0+(N-1)cF,NcF) : F \in \mathfrak F \} \\ \cup \{ \conv((N-1)x_0+cF,(N-2)x_0+2cF) : F \in \mathfrak F \} \cup \dotsb \\ \cup \{ \conv(Nx_0,(N-1)x_0+cF) : F \in \mathfrak F \} \} \end{multline*} These are the maximal elements of a polytopal subdivision of $NcS$. We can make this a $\mathcal C$-subdivision by considering all elements in $\mathcal C$ with standard form \begin{equation} \label{eq:concentricpolytope} \left( \left( rx_0, (r-1)x_0 \right),(F), \begin{pmatrix} (N-r)c \\ (N-r+1)c \end{pmatrix} \right) \end{equation} where $F \in \mathfrak F$ and $1 \le r \le N$ is an integer. These are the maximal elements of a $\mathcal C$-subdivision whose image under $\Cay$ is the previous polytopal subdivision. For any $A \in \mathcal C$ of the form \eqref{eq:concentricpolytope}, we have $\ind L(A) < \ind L(S)$. We can now proceed with proving the KMW theorem. Let $L \subset \mathbb Z^d$ be a $d$-dimensional lattice, and fix some nonzero element $x \in \mathbb Z^d/L$. Let $\mathcal C_x := \mathcal C_x^\circ \cup \mathcal C_x^\bullet$, where \begin{itemize} \item $\mathcal C_x^\circ$ is the set of all elements in $\mathcal C$ which do not have $x$ as a box point. \item $\mathcal C_x^\bullet$ is the set of all elements in $\mathcal C$ whose standard form $(p,S,a)$ is such that $x$ is a box point of $S$, $\abs{p} = 1$, $\abs{S} = 1$, and the entry of $a$ is $d!$. \end{itemize} We let $\mathcal C_x$ inherit a poset structure and poset map $\Cay : \mathcal C_x \to \mathcal P$ from $\mathcal C$. We can check that $(\mathcal C_x, \Cay)$ is a perfect $\mathcal P$-poset. Let $\gamma_T^\circ$ be the restriction of $\gamma_T$ to $\mathcal C_x^\circ \cap \mathcal C_T$. It is easy to see that the image of $\gamma_T^\circ$ is contained in $\Subd(\mathcal C_x^\circ)$. Thus we have a well-defined subdivision $\gamma_T^\circ : \mathcal C_x^\circ \cap \mathcal C_T \to \Subd(\mathcal C_x)$ over $\mathcal P$. Now, define a subdivision $\stell_x : \mathcal C_x^\bullet \to \Subd(\mathcal C_x)$ as follows. Let $A \in \mathcal C_x^\bullet$ with standard form $(p,S,a)$. We will abuse notation and identify $p$ and $S$ with their single entries. Let $c := c(S,x)$ and $x_0$ be the unique representative of $x +cv$ in $S$, where $v$ is the first vertex of $S$. Using \eqref{eq:concentricpolytope}, we define $\stell_x(A)$ to be the $\mathcal C$-subdivision whose maximal elements are \[ \left( \left( p+rx_0, p+(r-1)x_0 \right),(F), \begin{pmatrix} (N-r)c \\ (N-r+1)c \end{pmatrix} \right) \] where $N = d!/c$, $1 \le r \le N$, and $F \in \mathfrak F(S)$. As in \eqref{eq:concentricpolytope}, this is a $\mathcal C$-subdivision of $\Cay(A)$, and since $x$ is not a box point of $F$ by definition of $\mathfrak F$, this is a $\mathcal C_x$-subdivision, as claimed. We leave the following as an exercise: \begin{prop} The family of subdivisions $\{\gamma_T^\circ\}_T \cup \{\stell_x\}$ is locally confluent, terminating, and facially compatible (as $\mathcal C_x$-subdivisions). \end{prop} From this and the previous discussions we have the following. \begin{thm} \label{thm:Gammax} There is a canonical subdivision $\Gamma_x : \mathcal C_x \to \Subd(\mathcal C_x)$ over $\mathcal P$ such that for all $A \in \mathcal C_x$, we have the following: \begin{enumerate}[label=(\alph*)] \item If $A \in \mathcal C_x^\circ$, then $\Gamma_x(A) = \Gamma(A)$. \item If $A \in \mathcal C_x^\bullet$ and $B \in \Max(\Gamma_x(A))$, where $(p,S,a)$ is the standard form of $B$, then $\abs{S} = 0$ and $\ind(L(B)) < \ind(L(A))$. \end{enumerate} \end{thm} We can now give a proof of Theorem~\ref{thm:KMW}. \begin{proof}[Proof of Theorem~\ref{thm:KMW}] For any triangulation $X$, we will consider the collection of lattices \[ \mathscr L(X) := \{ L(S) : S \in \Max X \}. \] Start with a $d$-dimensional triangulation $X$ in $\mathbb R^d$. Dilate the triangulation by $d!$ to form $d!X$. We can view $d!X$ as a $\mathcal C$-subdivision by replacing each simplex $d!S \in d!X$ by the element $((0), (S), (d!)) \in \mathcal C$. Call this $\mathcal C$-subdivision $Y$. Now suppose there is a $d$-dimensional simplex $S \in X$ such that $\ind L(S) > 1$. Choose any nonzero $x \in \mathbb Z^d / L(S)$. Note that $Y \in \Subd(\mathcal C_x)$. Let $X' = \Cay \Gamma_x^\ast(Y)$, where $\Gamma_x$ is from Theorem~\ref{thm:Gammax} and the $\ast$ construction is from Proposition~\ref{prop:canonicalrefinement}. For every $T' \in X'$, we have that $T'$ is a simplex, and either \begin{itemize} \item $\ind L(T') = \ind L(T)$ for some $T \in X$ where $x$ is not a box point of $T$, or \item $\ind L(T') < \ind L(T)$ for some $T \in X$ where $x$ is a box point of $T$. \end{itemize} Hence, comparing $\mathscr L(X)$ and $\mathscr L(X')$, we have that $\mathscr L(X')$ is obtained by replacing at least one lattice of $\mathscr L(X)$ with lattices of lower index, while keeping the other lattices the same. Thus, if we repeat the above process on $X'$, and so on, we will eventually obtain a triangulation $Z$ of $(d!)^N X$, for some $N$, such that $\mathscr L(Z)$ contains only the lattice $\mathbb Z^d$. Letting $X$ be any triangulation of an integral polytope $P$, we obtain the theorem. \end{proof} \subsection{The structure of box points} \label{sec:structboxpoints} In this section we investigate the distribution of representatives of box points in dilated polysimplices. Let $S = (S_1,\dots,S_n)$ be a tuple of ordered independent integral simplices and let $x$ be a box point of $S$. Let $c = c(S,x)$. By definition, the polysimplex $\sum_{j=1}^n c_j S_j$ contains a unique representative of $x + \sum_{j=1}^n c_j v_j$, where $v_j$ is the first vertex of $S_j$. Let $x_0$ be this representative. We call $x_0$ the \emph{focus} of $(S,x)$. Let $\mathfrak F = \mathfrak F(S,x)$ be the set of all tuples $F = (F_1,\dots,F_n)$ such that $F \le S$, $\sum F_j$ is a facet of $\sum S_j$, and $x$ is not a box point of $F$. Thus, if $F \in \mathfrak F$, the point $x_0$ is not contained in the facet $\sum_{j=1}^n c_j F_j$ of $\sum_{j=1}^n c_j S_j$. Moreover, the lattice distance from $x_0$ to $\sum_{j=1}^n c_j F_j$ is smaller than the lattice height of $\sum_{j=1}^n S_j$ with respect to the facet $\sum_{j=1}^n F_j$. As in Section~\ref{sec:KMW}, the elements of the form \[ \left((x_0,0),F, \begin{pmatrix} 0 & \cdots & 0 \\ c_1 & \cdots & c_n \end{pmatrix} \right) \] where $F \in \mathfrak F$ are the maximal elements of a $\mathcal C$-subdivision of $\sum_{j=1}^n c_j S_j$. If $B$ is such an element, then $\ind L(B) < \ind (L(S_1) + \dots + L(S_n))$. Now, consider the polysimplex $P = \sum_{j=1}^n a_j S_j$, where the $a_j$ are integers satisfying $a_j \ge c_j$ for all $j$. Since \[ P = \sum_{j=1}^n c_j S_j + \sum_{j=1}^n (a_j - c_j) S_j\] it follows that $P$ contains the polytope $P' := x_0 + \sum_{j=1}^n (a_j - c_j) S_j$. Moreover, if $F \in \mathfrak F$, then the lattice distance between the facet $x_0 + \sum_{j=1}^n (a_j - c_j) F_j$ of $P'$ and the facet $\sum_{j=1}^n a_j F_j$ of $P$ is equal to lattice distance from $x_0$ to $\sum_{j=1}^n c_j F_j$. Consider the elements of $\mathcal C$ of the form \[ \left( (x_0,0),F, \begin{pmatrix} a_1-c_1 & \cdots & a_n-c_n \\ a_1 & \cdots & a_n \end{pmatrix} \right). \] where $F$ ranges over $\mathfrak F$. These elements are the maximal elements of a $\mathcal C$-complex whose support is $\overline{P \setminus P'}$, the closure of $P \setminus P'$. Note that if there is some $j \in \supp c$ such that $a_j = c_j$, then $P'$ has smaller dimension than $P$, and so in this case this complex is a $\mathcal C$-subdivision of $P$. If $B \in \mathcal C$ is an element of the above form, then $\ind L(B) < \ind (L(S_1) + \dots + L(S_n))$. If $a_j \ge 2c_j$ for all $j$, then repeating the above argument on $P'$, we have that the elements of $\mathcal C$ of the form \[ \left( (2x_0,x_0),F, \begin{pmatrix} a_1-2c_1 & \cdots & a_n-2c_n \\ a_1-c_1 & \cdots & a_n-c_n \end{pmatrix} \right). \] are the maximal elements of a $\mathcal C$-complex whose support if $\overline{P' \setminus P''}$, where $P'' := 2x_0 + \sum_{j=1}^n (a_j-2c_j) F_j$. In general, if $a_j \ge Nc_j$ for some positive integer $N$, then the elements of $\mathcal C$ of the form \[ \left( \left(rx_0,(r-1)x_0 \right),F,\begin{pmatrix} a_1-rc_1 & \cdots & a_n-rc_n \\ a_1-(r-1)c_1 & \cdots & a_n-(r-1)c_n \end{pmatrix} \right), \] where $r = 1$, \dots, $N$ and $F \in \mathfrak F$, are the maximal elements of a $\mathcal C$-complex whose support is $\overline{P \setminus P^N}$, where $P^N := Nx_0 + \sum_{j=1}^n (a_j - Nc_j) F_j$. If in addition $a_j = Nc_j$ for some $j \in \supp c$, then this complex is a $\mathcal C$-subdivision of $P$. As before, if $B \in \mathcal C$ is a maximal element of this complex, then $\ind L(B) < \ind (L(S_1) + \dots + L(S_n))$. \subsection{$\kappa_{T,x}$-subdivisions} \label{sec:kappaTx} We will need one final geometric construction for the main proof. Like the previous constructions in this section, this construction uses box points to lower the indices of lattices, but in a way more analogous to $\gamma_T$. This is where our constructions begin to fundamentally diverge from the original proof of the KMW theorem. Let $L \subset \mathbb Z^d$ be a $d$-dimensional lattice and let $x \in \mathbb Z^d / L$ be nonzero. Let $T$ be an ordered integral simplex with dimension at least one. Let $\mathcal K_{T,x}$ be the set of elements of $\mathcal C$ whose standard form $(p,S,a)$ satisfies the following: \begin{itemize} \item $T$ is an entry of $S$, say the $j_0$-th entry. \item $x$ is a box point of $S$. \item $a_{ij} \ge c(S,x)_{j}$ for all $i$ and all $j$. \item $a_{ij_0} > c(S,x)_{j_0}$ for some $i$. \end{itemize} We now construct a subdivision $\kappa_{T,x} : \mathcal K_{T,x} \to \Subd(\mathcal C)$ over $\mathcal P$. Let $A \in \mathcal K_{T,x}$ with standard form $(p,S,a)$. Let $c = c(S,x)$. Let $j_0$ be the unique number such that $S_{j_0} = T$. Let $i_0$ be the smallest number such that $a_{i_0j_0} = \max_{i} a_{ij_0}$. Hence $a_{i_0j_0} > c_j$. To make the notation easier to read, we will assume from now on that $i_0 = j_0 = 1$; the below construction can be easily adjusted to allow for other values of $i_0$ and $j_0$. Let $v$ be the first vertex of $T$, and let $f$ be the facet of $T$ opposite $v$. Let $m = \abs{p}$ and $n = \abs{S}$. Let $F = (F_1,\dots,F_n)$ where $F_1 = f$ and $F_i = S_i$ for all $i > 1$. Let $A' = (p',S,a')$ and $A'' = (p'',F,a'')$ be as in Section~\ref{sec:T}, using $i = i_0 = 1$ and $j = j_0 = 1$. We consider two cases: \textbf{Case 1:} $x$ is a box point of $F$. In this case, we define $\kappa_{T,x}(A) = \gamma_T(A)$, as defined Section~\ref{sec:T}. \textbf{Case 2:} $x$ is not a box point of $F$. Let $\mathfrak F = \mathfrak F(S,x)$ be as in Section~\ref{sec:structboxpoints}, so $F \in \mathfrak F$. Let $x_0$ be the focus of $(S,x)$. Recall from Section~\ref{sec:structboxpoints} that the polytope $x_0 + \sum_{j=1}^n (a_{1j}-c_j)F_j$ is contained in $\sum_{j=1}^n a_{1j} S_j$. Moreover, the lattice distance between $x_0 + \sum_{j=1}^n (a_{1j}-c_j)F_j$ and $\sum_{j=1}^n a_{1j}F_j$ is less than the lattice height of $\sum_{j=1}^n S_j$ with respect to its facet $\sum_{j=1}^n F_j$. Hence, $x_0 + \sum_{j=1}^n (a_{1j}-c_j)F_j$ is contained in the polytope \[ \conv \left( v + (a_{11}-1)F_1 + \sum_{j=2}^n a_{1j} F_j, \sum_{j=1}^n a_{1j}F_j \right). \] Consider the tuple $(p''', F, a''')$, where $p'''$ is obtained from $p$ by replacing the first entry with $p_1 + x_0$, and $a'''$ is obtained from $a$ by replacing the first row with $(a_{1j}-c_j)_{j=1}^n$. It follows from the above discussion that $\Cay(p''',F,a''') \subset \Cay(A'')$. The polytope $\Cay(p''',F,a''')$ is parallel to the facets $\Cay(p,F,a)$ and $\Cay(p',F,a')$ of $\Cay(A'')$. We now define \begin{equation} \label{eq:sharpflat} \begin{aligned} A^\sharp &= (p^\sharp,F,a^\sharp) \in \mathcal C \\ A^\flat &= (p^\flat,F,a^\flat) \in \mathcal C \end{aligned} \end{equation} where \begin{itemize} \item $p^\sharp$ is the $(m+1)$-tuple obtained by inserting $p_1 + x_0$ directly before the $1$st entry of $p$. \item $a^\sharp$ is the $(m+1) \times n$ matrix obtained by inserting the row $(a_{1j}-c_j)_{j=1}^n$ directly above the $1$st row of $a$. \item $p^\flat$ is the $(m+1)$-tuple obtained by inserting $p_1 + x_0$ directly before the $1$st entry of $p'$. \item $a^\flat$ is the $(m+1) \times n$ matrix obtained by inserting the row $(a_{1j}-c_j)_{j=1}^n$ directly above the $1$st row of $a'$. \end{itemize} We have that \begin{align*} \Cay(A^\sharp) &= \conv( \Cay(p''',F,a'''), \Cay(p,F,a)) \\ \Cay(A^\flat) &= \conv( \Cay(p''',F,a'''), \Cay(p',F,a')). \end{align*} Now, let $\mathfrak G$ be the set of all tuples $G = (G_1,\dots,G_n)$ such that $G \le F$, $\sum_{j=1}^n G_j$ is a facet of $\sum_{j=1}^n F_j$, and $G \le F'$ for some $F' \in \mathfrak F$ with $F' \neq F$. Let \begin{equation} \label{eq:AG} A^G = (p^G, G, a^G) \in \mathcal C \end{equation} where \begin{itemize} \item $p^G$ is the $(m+2)$-tuple obtained by inserting $p_1+x_0$ directly before the 1st entry of $p''$. \item $a^G$ is the $(m+2)\times n$ matrix obtained by inserting the row $(a_{1j}-c_j)_{j=1}^n$ directly above the $1$st row of $a''$. \end{itemize} Then the elements $A^\sharp$, $A^\flat$, and $A^G$ over all $G \in \mathfrak G$ are the maximal elements of a $\mathcal C$-subdivision of $\Cay(A'')$. Thus, the elements $A'$, $A^\sharp$, $A^\flat$, and $A^G$ over all $G \in \mathfrak G$ are the maximal elements of a $\mathcal C$-subdivision of $\Cay(A)$. We define $\kappa_{T,x}(A)$ to be this subdivision. We prove the following two properties of $\kappa_{T,x}$. \begin{prop} \label{prop:kappaTx} Let $A \in \mathcal K_{T,x}$ and $B \in \Max \kappa_{T,x}(A)$. We have the following: \begin{itemize} \item If $x$ is a box point of $B$, then $L(A) = L(B)$. \item If $x$ is not a box point of $B$, then $\ind(L(B)) < \ind(L(A))$. \end{itemize} \end{prop} \begin{proof} If we are in Case 1, then $x$ is a box point of $B$ and $L(A) = L(B)$ by Proposition~\ref{prop:LgammaT}. Assume we are in Case 2. If $B = A'$, then $x$ is a box point of $B$ and $L(A) = L(B)$ by Proposition~\ref{prop:LgammaT}. So assume $B = A^\sharp$, $A^\flat$, or $A^G$ for some $G \in \mathfrak G$. Thus $x$ is not a box point of $B$. Let $F$, $\mathfrak F$ and $x_0$ be as above. For each $E \in \mathfrak F$, let $h_E$ be the lattice height of $\sum S_j$ with respect to its facet $\sum E_j$. Let $h_E'$ be the lattice distance between $x_0$ and $\sum E_j$. Hence $0 < h_E' < h_E$ for all $E \in \mathfrak F$. If $B = A^\sharp$, then we have \[ \ind(L(B)) = h_F' \ind(L(F)) < h_F \ind(L(F)) = \ind(L(A)). \] If $B = A^\flat$, then we have \[ \ind(L(B)) = (h_F-h_F') \ind(L(F)) < h_F \ind(L(F)) = \ind(L(A)). \] Finally, suppose $B = A^G$ for some $G \in \mathfrak G$. Let $E \in \mathfrak F$ such that $G \le E$ and $E \neq F$. Then \begin{multline*} \ind(L(B)) = h_E' \ind(L(p'',G,a'')) = h_E' \ind(L(E)) \\ < h_E \ind(L(E)) = \ind(L(A)). \end{multline*} In all cases $\ind(L(B)) < \ind(L(A))$, completing the proof. \end{proof} \begin{prop} \label{prop:facekappaTx} Let $A \in \mathcal K_{T,x}$ and suppose $\mathcal B \in \mathcal C$ such that $B \le_{\mathcal C} A$. Then one of the following hold: \begin{itemize} \item $\kappa_{T,x}(A|B) = \kappa_{T',x}(B)$ for some $T'$ \item $\kappa_{T,x}(A|B) = \gamma_{T'}(B)$ for some $T'$ \item $\kappa_{T,x}(A|B) = \triv_{\mathcal D}(B)$. \end{itemize} \end{prop} \begin{proof} Let the standard form of $A$ be $(p,S,a)$. Let $B = (p_I,S',a_{I \times \bullet})$ with $S' \le S$ and $I$ a nonempty subset of $[\abs{p}]$. Assume that $T = S_j$, and let $i$ be the smallest number such that $a_{ij} = \max_{i'} a_{i'j}$. If $i \in I$ and $S_j'$ contains the first vertex of $T$ and another vertex, then \[ \kappa_{T,x}(A|B) = \begin{dcases*} \kappa_{S_j',x}(B) & if $x$ is a box point of $S'$ \\ \gamma_{S_j'}(B) & if $x$ is not a box point of $S'$ \end{dcases*} \] Otherwise, we have $\gamma_T(A|B) = \triv_{\mathcal C}(B)$, as desired. \end{proof} \section{Main proof} \label{sec:main} Let $d$ be a positive integer and let $c$ be any integer greater than or equal to $d!+d$. In this section we prove the following. \begin{thm} \label{thm:main2} For any $d$-dimensional integral polytope $P$ in $\mathbb R^d$, there exists a positive integer $N$ such that for all nonnegative integers $r$ and $s$, $(rc^N + s(d!)^N)P$ has a unimodular triangulation. \end{thm} Taking $c$ to be relatively prime to $d!$, this theorem will imply Theorem~\ref{thm:main}, as desired. The overall strategy of the proof is similar to the proof of the KMW theorem in Section~\ref{sec:KMW}. We will fix a box point $x$, and then construct two types of subdivisions: one that preserves lattices that do not have $x$ as a box point, and one that replaces lattices which do have $x$ as a box point with lattices of smaller index. There are two main differences between the current proof and that of the KMW theorem. The first is that we work with intermediate dilations by $c$, rather than just $d!$. To deal with this we will use the subdivisions $\kappa_{T,x}$ defined earlier. The second difference is that we work over $2\mathcal P$. Roughly, this raises the following problem: We might produce an element $(P,Q) \in 2 \mathcal P$ where $P$ and $Q$ are independent and $\ind(P+Q) > \ind(P)\ind(Q)$. Such an element can never have a mixed subdivision into elements of index 1, even after dilation. Thus, we must avoid such elements. The way this is done is to work within a carefully defined $2\mathcal P$-poset $\mathcal D$ which always avoids such elements, and construct subdivisions which always remain in $\mathcal D$. These restrictions are the reason why this proof is much more complicated than the proof of the KMW theorem or the arguments in \cite{ALT18}. \subsection{The poset $\mathcal D$} \label{sec:posetD} \subsubsection{The poset $\mathcal M$} We first define a poset $\mathcal M_0$ as follows. Its elements are all elements \[ (p,S,a) \times (q,S,b) \times k \in \mathcal C_0 \times \mathcal C_0 \times \mathbb Z, \] satisfying the following properties. \begin{enumerate}[label=(\Roman*)] \item $\abs{p} = \abs{q}$. \item $0 \le k \le \abs{S}$. \item $p_i-q_i$ is constant over all $i$ \item \label{samedifference} For each $j \le k$, we have $a_{ij} = b_{ij}$ for all $i$. \item For each $j > k$, $a_{ij} = 1$ for all $i$, and either $b_{ij} = 0$ for all $i$ or $b_{ij} = 1$ for all $i$. \end{enumerate} We define an equivalence relation on $\mathcal M_0$ as follows. If $j$ is such that $j \le k$ and either $S_j$ is a point or $a_{ij} = 0$ for all $i$, we set \begin{multline*} (p,S,a) \times (q,S,b) \times k \sim \\ ((p_i+a_{ij}S_j)_{i \in [\abs{p}]}, S_{\setminus j}, a_{\bullet \times \setminus j}) \times ((q_i+b_{ij}S_j)_{i \in [\abs{q}]}, S_{\setminus j}, b_{\bullet \times \setminus j}) \times (k-1). \end{multline*} Also, if $j$ is such that $j > k$ and $S_j$ is a point, then we set \begin{multline*} (p,S,a) \times (q,S,b) \times k \sim \\ ((p_i+a_{ij}S_j)_{i \in [\abs{p}]}, S_{\setminus j}, a_{\bullet \times \setminus j}) \times ((q_i+b_{ij}S_j)_{i \in [\abs{q}]}, S_{\setminus j}, b_{\bullet \times \setminus j}) \times k. \end{multline*} (Note that if $j > k$, then by definition $a_{ij} \neq 0$ for any $i$.) We define $\mathcal M := \mathcal M_0 / \sim$. We define the \emph{standard form} of an element $A \in \mathcal M$ to be the unique representative $(p,S,a) \times (q,S,b) \times k \in \mathcal M_0$ of $A$ such that $(p,S,a)$ is in standard form. For $A$, $B \in \mathcal M$ we let $B \le_{\mathcal M} A$ if the standard form of $A$ is $(p,S,a) \times (q,S,b) \times k$ and $B$ has a representative of the form \[ (p_I,S',a_{I \times \bullet}) \times (q_I,S',b_{I \times \bullet}) \times k \] where $I$ is a nonempty subset of $[\abs{p}]$ and $S' \le S$. Then $\le_{\mathcal M}$ is a partial order on $\mathcal M$. Let $\Cay : \mathcal M \to 2\mathcal P$ be the map $\Cay(U \times V \times k) = (\Cay(U),\Cay(V))$. This is a well-defined poset map. Hence $(\mathcal M,\Cay)$ is a $2\mathcal P$-poset. Furthermore, $(\mathcal M,\Sum \circ \Cay)$ is a perfect $\mathcal P$-poset. \subsubsection{The poset $\mathcal N$} We similarly define a poset $\mathcal N$ as follows. Its elements are all elements \[ (p,S,a) \times (q,S,b) \times k \in \mathcal C_0 \times \mathcal C_0 \times \mathbb Z, \] satisfying the following properties. \begin{enumerate}[label=(\Roman*), resume] \item $\abs{q} = 1$. \item $0 \le k \le \abs{S}$. \item $a_{ij} \ge b_{1j}$ for all $i$, $j$. \item For each $j > k$, $a_{ij} = 1$ for all $i$, and $b_{1j} = 0$ or 1. \end{enumerate} We define equivalence relation $\sim$ on $\mathcal N_0$ analogously to the previous section and define $\mathcal N = \mathcal N_0 / \sim$. As before, we define the \emph{standard form} of an element $A \in \mathcal N$ to be the unique representative $(p,S,a) \times (q,S,b) \times k \in \mathcal N_0$ of $A$ such that $(p,S,a)$ is in standard form. For $A$, $B \in \mathcal N$ we let $B \le_{\mathcal N} A$ if the standard form of $A$ is $(p,S,a) \times (q,S,b) \times k$ and $B$ has a representative of the form \[ (p_I,S',a_{I \times \bullet}) \times (q,S',b) \times k \] where $I$ is a nonempty subset of $[\abs{p}]$ and $S' \le S$. Then $\le_{\mathcal N}$ is a partial order on $\mathcal N$. As before, let $\Cay : \mathcal N \to 2\mathcal P$ be the map $\Cay(U \times V \times k) = (\Cay(U),\Cay(V))$. This is a well-defined poset map. Hence $(\mathcal N,\Cay)$ is a $2\mathcal P$-poset. Furthermore, $(\mathcal N,\Sum \circ \Cay)$ is a perfect $\mathcal P$-poset. \subsubsection{The poset $\mathcal D$} The set $\mathcal M \cap \mathcal N$ is a poset ideal of both $\mathcal M$ and $\mathcal N$; specifically, it is the set of elements $(p,S,a) \times (q,S,b) \times k$ in $\mathcal M$ or $\mathcal N$ such that $\abs{p} = 1$ and \ref{samedifference} holds. Moreover, if $A$, $B \in \mathcal M \cap \mathcal N$, then $A \le_{\mathcal M} B$ if and only if $A \le_{\mathcal N} B$. Hence, we can define a poset $\mathcal D$ on the set $\mathcal M \cup \mathcal N$ where $A \le_{\mathcal D} B$ if and only if $A \le_{\mathcal M} B$ or $A \le_{\mathcal N} B$. As above, we define a map $\Cay : \mathcal D \to 2 \mathcal P$ by $\Cay(U \times V \times k) = (\Cay(U),\Cay(V))$. Then $(\mathcal D,\Cay)$ is a $2\mathcal P$-poset, and $(\mathcal D,\Sum \circ \Cay)$ is a perfect $\mathcal P$-poset. For $A \in \mathcal D$ with standard form $(p,S,a) \times (q,S,b) \times k$, we define $L(A) := L(p,S,a)$. We say that $x$ is a box point of $A$ if $x$ is a box point of $S_{[k]}$. \subsection{$\mu_T$, $\nu_T$, and $\epsilon$ subdivisions} In this section we define three subdivisions on subposets of $\mathcal D$. These subdivisions play the role of $\gamma_T$ in that they preserve lattices. \subsubsection{$\mu_T$ subdivisions} Let $T$ be an ordered integral simplex of dimension at least 1. Let $\mathcal M_T$ be the set of elements of $\mathcal M$ whose standard form $(p,S,a) \times (q,S,b) \times k$ has the property that $T$ is an entry of $S_{[k]}$. We construct a subdivision as follows $\mu_T : \mathcal M_T \to \Subd(\mathcal D)$. Let $A \in \mathcal M_T$ with standard form $U \times V \times k = (p,S,a) \times (q,S,b) \times k$. Assume $T$ is the $j$-th entry of $S$, where $j \le k$. We have that $U \in \mathcal C_T$, where $\mathcal C_T$ is as in Section~\ref{sec:T}. Also, by \ref{samedifference}, we have that the $j$-th column of $b$ is not all 0, so $V \in \mathcal C_T$. Following the construction of $\gamma_T(U)$ and $\gamma_T(V)$, let $U'$, $U''$, $V'$, and $V''$ be as in \eqref{eq:A'A''}. Then $U' \times V' \times k$, $U'' \times V'' \times k \in \mathcal M$. We define \[ \mu_T(A) = \triv_{\mathcal D}(U' \times V' \times k) \cup \triv_{\mathcal D}(U'' \times V'' \times k). \] Then $\mu_T(A)$ is a $\mathcal D$-subdivision with 2-support $\Cay(U) \times \Cay(V) = \Cay(A)$. Hence, $\mu_T : \mathcal M_T \to \Subd(\mathcal D)$ is a subdivision over $2 \mathcal P$. From Proposition~\ref{prop:LgammaT}, we have the following. \begin{prop} \label{prop:LmuT} Let $A \in \mathcal M_T$ and $B \in \Max \mu_T(A)$. Then $L(A) = L(B)$. \end{prop} From the proof of Proposition~\ref{prop:facegammaT}, we have the following. \begin{prop} \label{prop:facemuT} Let $A \in \mathcal M_T$ and suppose $\mathcal B \in \mathcal D$ such that $B \le_{\mathcal D} A$. Then either $\mu_T(A|B) = \mu_{T'}(B)$ for some $T'$ or $\mu_T(A|B) = \triv_{\mathcal D}(B)$. \end{prop} \subsubsection{$\nu_T$ subdivisions} Let $T$ be an ordered integral simplex of dimension at least 1. Let $\mathcal N_T$ be the set of elements of $\mathcal N$ whose standard form $(p,S,a) \times (q,S,b) \times k$ has the property that \begin{enumerate}[label=(\alph*)] \item $T$ is an entry of $S_{[k]}$, say the $j$-th entry. \item There is some $i$ such that $a_{ij} > b_{1j}$. \end{enumerate} We construct a subdivision as follows $\nu_T : \mathcal N_T \to \Subd(\mathcal D)$. Let $A \in \mathcal N_T$ with standard form $U \times V \times k = (p,S,a) \times (q,S,b) \times k$. Following the construction of $\gamma_T(U)$, let $U'$, $U''$ be as in \eqref{eq:A'A''}. Let $V_F := (q,F,b)$, where $F$ is as in \eqref{eq:A'A''}. Then $U' \times V \times k$, $U'' \times V_F \times k \in \mathcal N$. We define \[ \nu_T = \triv_{\mathcal D}(U' \times V \times k) \cup \triv_{\mathcal D}(U'' \times V_F \times k) \] Then $\nu_T(A)$ is a $\mathcal D$-subdivision with 2-support $\Cay(U) \times \Cay(V) = \Cay(A)$. Hence, $\nu_T : \mathcal N_T \to \Subd(\mathcal D)$ is a subdivision over $2 \mathcal P$. From Proposition~\ref{prop:LgammaT}, we have the following. \begin{prop} \label{prop:LnuT} Let $A \in \mathcal N_T$ and $B \in \Max \nu_T(A)$. Then $L(A) = L(B)$. \end{prop} From the proof of Proposition~\ref{prop:facegammaT}, we have the following. \begin{prop} \label{prop:facenuT} Let $A \in \mathcal N_T$ and suppose $\mathcal B \in \mathcal D$ such that $B \le_{\mathcal D} A$. Then either $\nu_T(A|B) = \nu_{T'}(B)$ for some $T'$ or $\nu_T(A|B) = \triv_{\mathcal D}(B)$. \end{prop} \subsubsection{$\epsilon$ subdivisions} \label{sec:epsilon} Finally, let $\mathcal E$ be the set of elements $A \in \mathcal D$ whose standard form $(p,S,a) \times (q,S,b) \times k$ satisfies the following: $\abs{p} > 1$, and one of the following hold: \begin{enumerate}[label=(\roman*)] \item $A \in \mathcal M$, and $k = 0$. \item $A \in \mathcal N$, and for all $j \le k$ we have $a_{ij} = b_{ij}$ for all $i$. \end{enumerate} Note that in either case, all the rows of $a$ are equal to each other and all the rows of $b$ are equal to each other. We construct a subdivision $\epsilon : \mathcal E \to \mathcal D$ as follows. Let $A \in \mathcal E$ with standard form $(p,S,a) \times (q,S,b) \times k$. Note that the entries of $p$ are affinely independent, because by definition of $\mathcal C_0$ the polytopes $p_i + \sum_{j=1}^n a_{ij}S_j$ are in Cayley position and $\sum_{j=1}^n a_{ij}S_j$ is constant over all $i$. Let $T$ be the ordered simplex with vertices given by the entries of $p$, in that order. Given a tuple $c$ and an object $d$, let $c \cup d$ denote the tuple obtained by concatenating $d$ to the end of $c$. Let $a_1$ denote the first row of $a$ and define $b_1$ similarly. If $A \in \mathcal M$, we define \[ A^T := ( (0), S \cup T, a_1 \cup 1 ) \times ((q_1-p_1), S \cup T, b_1 \cup 1) \times k \in \mathcal D. \] If $A \in \mathcal N$, we define \begin{equation} A^T := ( (0), S \cup T, a_1 \cup 1 ) \times (q, S \cup T, b \cup 0) \times k \in \mathcal D. \end{equation} Either way, we have $\Cay(A^T) = \Cay(A)$. We define $\epsilon(A) = \triv_{\mathcal D}(A^T)$. We have that $\epsilon : \mathcal E \to \mathcal D$ is a subdivision over $2 \mathcal P$. We have the following two properties of $\epsilon$. \begin{prop} \label{prop:Lepsilon} Let $A \in \mathcal E$ and $B \in \Max \epsilon(A)$. Then $L(A) = L(B)$. \end{prop} \begin{proof} This follows from the observation that $T$ is a translation of $S_0(p,S,a)$. \end{proof} From the proof of Proposition~\ref{prop:facegammaT}, we have the following. \begin{prop} \label{prop:faceepsilon} Let $A \in \mathcal E$ and suppose $\mathcal B \in \mathcal D$ such that $B \le_{\mathcal D} A$. Then either $\epsilon(A|B) = \epsilon(B)$ or $\epsilon(A|B) = \triv_{\mathcal D}(B)$. \end{prop} \begin{proof} Let the standard form of $A$ be $(p,S,a) \times (q,S,b) \times k$ and let $B = U \times V \times W$, where $U = (p_I,S',a_{I \times \bullet})$ for some nonempty $I \subseteq [\abs{p}]$ and $S' \le S$. If $\abs{I} > 1$, then $B \in \mathcal E$ and $\epsilon(A|B) = \epsilon(B)$. Otherwise, $\epsilon(A|B) = \triv_{\mathcal D}(B)$. \end{proof} \subsection{$\tau_x$, $\sigma_x$, and $\rho_x$ subdivisions} The final subdivisions we will construct are analogues of the $\stell_x$ and $\kappa_{T,x}$ subdivisions from earlier. These will allow us to lower indices of lattices. The order they are presented here is ``backwards'', in the sense that in practice, one would apply $\rho_x$ subdivisions first, then $\sigma_x$ subdivisions, then $\tau_x$. Throughout this section, we fix $L$ a $d$-dimensional lattice in $\mathbb R^d$ and some nonzero $x \in \mathbb Z^d / L$. \subsubsection{$\tau_x$ subdivisions} Let $\mathcal T_x$ be the set of elements of $\mathcal D$ whose standard form $(p,S,a) \times (q,S,b) \times k$ satisfies the following: \begin{enumerate}[label=(\alph*)] \item $x$ is a box point of $S_{[k]}$. \item $\abs{p} = 1$. \item For all $j \le k$, $a_{1j} = b_{1j} = d!$. \end{enumerate} We construct a subdivision $\tau_x : \mathcal T_x \to \Subd(\mathcal D)$ as follows. Let $A \in \mathcal T_x$ with standard form $(p,S,a) \times (q,S,b) \times k$. Let $n = \abs{S}$. Let $c$ be the $n$-tuple of integers where \[ c_j = \begin{dcases*} c(S_{[k]},x)_j & for $1 \le j \le k$ \\ 0 & for $k+1 \le j \le n$ \end{dcases*} \] Now, let $x_0$ be the focus of $(S,x)$, and let $\mathfrak F = \mathfrak F(S,x)$. Let $N = d!/\max_j c_j$. For all $F \in \mathfrak F$ and $r = 1$, \dots, $N$, define \begin{align*} U_{F,r} &:= \left( \left(p_1+rx_0,p_1+(r-1)x_0 \right),F,\right. \\ & \qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad \left. \begin{pmatrix} a_{11}-rc_1 & \cdots & a_{1n}-rc_n \\ a_{11}-(r-1)c_1 & \cdots & a_{1n}-(r-1)c_n \end{pmatrix} \right) \\ V_{F,r} &:= \left( \left(q_1+rx_0,q_1+(r-1)x_0 \right),F, \right. \\ & \qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad \left. \begin{pmatrix} b_{11}-rc_1 & \cdots & b_{1n}-rc_n \\ b_{11}-(r-1)c_1 & \cdots & b_{1n}-(r-1)c_n \end{pmatrix} \right). \end{align*} We have that $U_{F,r} \times V_{F,r} \times k \in \mathcal M$. From Section~\ref{sec:structboxpoints}, the set of all $U_{F,r} \times V_{F,r} \times k$ over $F \in \mathfrak F$ and $r \in [N+1]$ is the set of maximal elements of a $\mathcal D$-subdivision with 2-support $\Cay(p,S,a) \times \Cay(q,S,b) = \Cay(A)$. We define $\tau_x(A)$ to be this $\mathcal D$-subdivision. Hence, we have constructed a subdivision $\tau_x :\mathcal T_x \to \Subd(\mathcal D)$ over $2 \mathcal P$. From Section~\ref{sec:structboxpoints}, we have the following: \begin{prop} \label{prop:Ltaux} Let $A \in \mathcal T_x$ and $B \in \Max \tau_x(A)$. Then $L(B) < L(A)$. \end{prop} We also have the following. \begin{prop} \label{prop:facetaux} Let $A \in \mathcal T_x$ and suppose $\mathcal B \in \mathcal D$ such that $B \le_{\mathcal D} A$. Then either $\tau_x(A|B) = \tau_x(B)$ or $\tau_x(A|B) = \triv_{\mathcal D}(B)$. \end{prop} \begin{proof} If $x$ is a box point of $B$, then $\tau_x(A|B) = \tau_x(B)$. Otherwise, $\tau_x(A|B) = \triv_{\mathcal D}(B)$. \end{proof} \subsubsection{$\sigma_x$ subdivisions} Let $\mathcal S_x$ be the set of elements of $\mathcal D$ whose standard form $(p,S,a) \times (q,S,b) \times k$ satisfies the following: \begin{enumerate}[label=(\alph*)] \item $x$ is a box point of $S_{[k]}$. \item $(p,S,a) \times (q,S,b) \times k \in \mathcal N$. \item For all $j \le k$, $b_{1j} = 0$ or $d!$. \item For all $i$, either $a_{ij} = b_{1j}$ for all $j \le k$ or $a_{ij} = b_{1j} + c(S_{[k]},x)_j$ for all $j \le k$. \item There is some $i$ such that $a_{ij} = b_{1j} + c(S_{[k]},x)_j$ for all $j \le k$. \end{enumerate} We construct a subdivision $\sigma_x : \mathcal S_x \to \Subd(\mathcal D)$ as follows. Let $A \in \mathcal S_x$ with standard form $U \times V \times k = (p,S,a) \times (q,S,b) \times k$. Let $m = \abs{p}$ and $n = \abs{S}$. Let $c$ be the $n$-tuple of integers where $c_j = c(S_{[k]},x)_j$ for $j \le k$ and $c_j = 0$ for $j > k$, as before. Let $x_0$ be the focus of $(S,x)$, and let $\mathfrak F = \mathfrak F(S,x)$. Let $i$ be the smallest number such that $a_{ij} = b_{1j} + c_j$ for all $j \le k$. For all $F \in \mathfrak F$, we define \[ A^{\sharp,F} := U^{\sharp,F} \times V_F \times k \in \mathcal N \] where $V_F = (q,F,b)$, and $U^{\sharp,F} = (p^{\sharp},F,a^{\sharp})$ is defined analogously to $A^\sharp$ in \eqref{eq:sharpflat}; in other words, \begin{itemize} \item $p^{\sharp}$ is the $(m+1)$-tuple obtained by inserting $p_1 + x_0$ directly before the $i$th entry of $p$. \item $a^{\sharp}$ is the $(m+1) \times n$ matrix obtained by inserting the row $(a_{ij}-c_j)_{j=1}^n$ directly above the $i$th row of $a$. \end{itemize} In addition, we define \[ A^\star := (p''', S, a''') \times V \times k \in \mathcal N \] where $p'''$ and $a'''$ are defined right before \eqref{eq:sharpflat}; that is, \begin{itemize} \item $p'''$ is the $m$-tuple obtained by replacing the $i$th entry of $p$ with $p_1 + x_0$. \item $a'''$ is the $m \times n$ matrix obtained by replacing the $i$th row with $(a_{ij}-c_j)_{j=1}^n$. \end{itemize} From the discussion in Section~\ref{sec:structboxpoints}, $\{A^\star\} \cup \{A^{\sharp,F} : F \in \mathfrak F\}$ is the set of maximal elements of a $\mathcal D$-subdivision with 2-support $\Cay(U) \times \Cay(V) = \Cay(A)$. We define $\sigma_x(A)$ to be this $\mathcal D$-subdivision. Hence, we have constructed a subdivision $\sigma_x :\mathcal S_x \to \Subd(\mathcal D)$ over $2 \mathcal P$. Using similar arguments as the previous section, we have the following: \begin{prop} \label{prop:Lsigmax} Let $A \in \mathcal S_x$ and $B \in \Max \sigma_x(A)$. Then $L(B) < L(A)$. \end{prop} \begin{prop} \label{prop:facesigmax} Let $A \in \mathcal S_x$ and suppose $\mathcal B \in \mathcal D$ such that $B \le_{\mathcal D} A$. Then either $\sigma_x(A|B) = \sigma_x(B)$ or $\sigma_x(A|B) = \triv_{\mathcal D}(B)$. \end{prop} \subsubsection{$\rho_x$ subdivisions.} Let $\mathcal R_x$ be the set of elements of $\mathcal D$ whose standard from $(p,S,a) \times (q,S,b) \times k$ satisfies the following. \begin{enumerate}[label=(\alph*)] \item $x$ is a box point of $S$. \item $(p,S,a) \times (q,S,b) \times k \in \mathcal N$. \item For all $j \le k$, either $b_{1j} = 0$ or $b_{1j} = d!$. \item For all $i$ and all $j \le k$, we have $a_{ij} \ge b_{1j} + c(S,x)_j$. \item There exists some $i$ and some $j \le k$ satisfying $a_{ij} > b_{1j} + c(S,x)_j$. \end{enumerate} We construct a subdivision $\rho_x : \mathcal R_x \to \Subd(\mathcal D)$ as follows. Let $A \in \mathcal R_x$ with standard form $U \times V \times k = (p,S,a) \times (q,S,b) \times k$. Let $j \le k$ be the smallest number such that there exists $i$ satisfying $a_{ij} > b_{1j} + c(S,x)_j$. Let $T = S_j$. Let $f$ be the facet of $T$ opposite the first vertex of $T$, and let $F$ be the tuple obtained from $S$ by replacing $T$ with $f$. We consider two cases, in parallel to Section~\ref{sec:kappaTx}. \textbf{Case 1:} $x$ is a box point of $F$. In this case, we define $\rho_x(A) = \nu_T(A)$. \textbf{Case 2:} $x$ is not a box point of $F$. Following the construction of $\kappa_{T,x}$, define $U'$ as in \eqref{eq:A'A''}, and define $U^\sharp$, $U^\flat$ as in \eqref{eq:sharpflat}. In addition, define $\mathfrak G$ as in Section~\ref{sec:kappaTx}, and for each $G \in \mathfrak G$ define $U^G$ as in \eqref{eq:AG}. For any $R \le S$, let $V_R := (q,R,b)$. Then the set of elements \[ \{ U' \times V \times k, U^\sharp \times V_F \times k, U^\flat \times V_F \times k \} \cup \{ U^G \times V_G \times k : G \in \mathfrak G \} \subset \mathcal N \] is the set of maximal elements of a $\mathcal D$-subdivision with 2-support $\Cay(A)$. We define $\rho_x(A)$ to be this $\mathcal D$-subdivision. Hence we have defined a subdivision $\rho_x : \mathcal R_x \to \Subd(\mathcal D)$ over $2 \mathcal P$. From Proposition~\ref{prop:kappaTx}, we have the following. \begin{prop} \label{prop:Lrhox} Let $A \in \mathcal R_x$ and $B \in \Max \rho_x(A)$. We have the following. \begin{itemize} \item If $x$ is a box point of $B$, then $L(A) = L(B)$. \item If $x$ is not a box point of $B$, then $\ind(L(B)) < \ind(L(A))$. \end{itemize} \end{prop} Finally, we have the following, with an analogous proof to Proposition~\ref{prop:facekappaTx}. \begin{prop} \label{prop:facerhox} Let $A \in \mathcal R_x$ and suppose $\mathcal B \in \mathcal D$ such that $B \le_{\mathcal D} A$. Then one of the following hold: \begin{itemize} \item $\rho_x(A|B) = \rho_x(B)$ \item $\rho_x(A|B) = \nu_T(B)$ for some $T$ \item $\rho_x(A|B) = \triv_{\mathcal D}(B)$. \end{itemize} \end{prop} \subsubsection{The set $\mathcal D^\bullet_x$} Let $\mathcal D^\bullet_x$ be the set of elements of $\mathcal D$ whose standard form $(p,S,a) \times (q,S,b) \times k$ satisfies the following: \begin{enumerate}[label=(\alph*)] \item $x$ is a box point of $S_{[k]}$. \item $(p,S,a) \times (q,S,b) \times k \in \mathcal N$. \item For all $j \le k$, $b_{1j} = 0$ or $d!$. \item For all $i$, either $a_{ij} = b_{1j}$ for all $j \le k$ or $a_{ij} \ge b_{1j} + c(S_{[k]},x)_j$ for all $j \le k$. \end{enumerate} Observe that the sets $\mathcal R_x$, $\mathcal S_x$, $\mathcal T_x$, and $\mathcal E \cap \mathcal D^\bullet_x$ are pairwise disjoint and partition $\mathcal D^\bullet_x$. \subsection{Proof of Theorem~\ref{thm:main2}} With our constructions completed, we are now ready to prove Theorem~\ref{thm:main2}. The proof mirrors our proof of the KMW theorem from Section~\ref{sec:KMW}. Let $L$ be a $d$-dimensional lattice in $\mathbb R^d$, and fix some nonzero element $x \in \mathbb Z^d / L$. Define \[ \mathcal D_x := \mathcal D_x^\circ \cup \mathcal D_x^\bullet, \] where $\mathcal D_x^\circ$ is the set of all elements of $\mathcal D$ which do not have $x$ as a box point. We let $\mathcal D_x$ inherit a poset structure and poset map $\Cay : \mathcal D_x \to 2\mathcal P$ from $\mathcal D$. We can check that $(\mathcal D_x,\Cay)$ is a perfect $\mathcal P$-poset. Let $\mu_T^\circ$ and $\nu_T^\circ$ be the restrictions of $\mu_T$ and $\nu_T$, respectively, to \begin{align*} \mathcal M_T^\circ &:= \mathcal M_T \cap \mathcal D_x^\circ \\ \mathcal N_T^\circ &:= \mathcal N_T \cap \mathcal D_x^\circ \end{align*} respectively. Let $\epsilon_x$ be the restriction of $\epsilon$ to $\mathcal E_x := \mathcal E \cap \mathcal D_x$. We can check that for each of the subdivisions $\mu_T^\circ$, $\nu_T^\circ$, $\epsilon_x$, $\rho_x$, $\sigma_x$, and $\tau_x$, the output is always a $\mathcal D_x$-subdivision. Thus we have well defined subdivisions \begin{align*} \mu_T^\circ &: \mathcal M_T^\circ \to \Subd(\mathcal D_x) \\ \nu_T^\circ &: \mathcal N_T^\circ \to \Subd(\mathcal D_x) \\ \epsilon_x &: \mathcal E_x \to \Subd(\mathcal D_x) \\ \rho_x &: \mathcal R_x \to \Subd(\mathcal D_x) \\ \sigma_x &: \mathcal S_x \to \Subd(\mathcal D_x) \\ \tau_x &: \mathcal T_x \to \Subd(\mathcal D_x) \end{align*} over $2\mathcal P$. We now have the following. \begin{prop} \label{prop:S} The family of subdivisions \[ \mathscr S_x := \{\mu_T^\circ\}_T \cup \{\nu_T^\circ\}_T \cup \{\epsilon_x\} \cup \{\rho_x\} \cup \{\sigma_x\} \cup \{\tau_x\} \] is locally confluent, terminating, and facially compatible (as $\mathcal D_x$-subdivisions). \end{prop} \begin{proof} We first check local confluence. Note that for any simplices $T_1$, $T_2$, the sets $\mathcal M_{T_1}^\circ$, $\mathcal N_{T_2}^\circ$, $\mathcal E_x$, $\mathcal R_x$, $\mathcal S_x$, and $\mathcal T_x$ are pairwise disjoint. Thus, if $A \in \mathcal D_x$ and there are two distinct moves from $\{A\}$, then these moves must be either $\mu_{T_1}$, $\mu_{T_2}$ for some $T_1$, $T_2$ or $\nu_{T_1}$, $\nu_{T_2}$ for some $T_1$, $T_2$. The same argument from Proposition~\ref{prop:confgammaT} shows that the results of these moves are joinable. We next show the terminating property. Let $A \in \mathcal D_x$ with standard form $(p,S,a) \times (q,S,b) \times k$, and suppose we have a subdivision $f(A)$ where $f \in \mathscr S_x$. Let $B \in \Max f(A)$ with standard form $(p',S',a') \times (q',S',b') k'$. Then one of the following holds: \begin{enumerate} \item $\sum_{j \le k'} \dim S'_j < \sum_{j \le k} \dim S_j$. \item The above inequality is equality, and \[ \sum_{\substack{i \\ j \le k'}} a'_{ij} < \sum_{\substack{i \\ j \le k}} a_{ij}. \] \item The above two inequalities are equality, and $\abs{p'} < \abs{p}$. (This can only possibly occur if $f = \epsilon$.) \end{enumerate} It follows that $\mathscr S_x$ is terminating. Finally, we note that Propositions~\ref{prop:facemuT}, \ref{prop:facenuT}, \ref{prop:faceepsilon}, \ref{prop:facetaux}, \ref{prop:facesigmax}, and \ref{prop:facerhox} imply the first criterion of facial compatibility. For the second criterion, let $A \in \mathcal D_x$ with standard form $(p,S,a) \times (q,S,b) \times k$. We have that $A$ is terminal if and only if $A$ is not in any of the sets $\mathcal M^\circ$, $\mathcal N^\circ$, $\mathcal E_x$, $\mathcal R_x$, $\mathcal S_x$, $\mathcal T_x$. This occurs if and only if $\abs{p} = 1$ and $k = 0$. Clearly if this property holds for $A$, then it holds for any $B \le_{\mathcal D} A$. Hence $\mathscr S_x$ is facially compatible. \end{proof} From this we can conclude the following. \begin{thm} \label{thm:Deltax} There is a canonical subdivision $\Delta_x : \mathcal D_x \to \Subd(\mathcal D_x)$ over $2\mathcal P$ such that for all $A \in \mathcal D_x$ and $B \in \Max(\Delta_x(A))$, where $(p,S,a) \times (q,S,b) \times k$ is the standard form of $B$, we have the following: \begin{enumerate}[label=(\alph*)] \item $\abs{p} = 1$ and $k = 0$. \item If $A \in \mathcal D_x^\circ$, then $L(B) = L(A)$. \item If $A \in \mathcal D_x^\bullet$, then $L(B) < L(A)$. \end{enumerate} \end{thm} \begin{proof} Construct $\Delta_x$ from $\mathscr S_x$ and Theorem~\ref{thm:confluence}. Then, (a) follows from the proof of Proposition~\ref{prop:S}. (b) follows from Propositions~\ref{prop:LmuT}, \ref{prop:LnuT}, and \ref{prop:Lepsilon}. For (c), note that if $B$ satsifies (a), then $x$ is not a box point of $B$. Thus, (c) follows from Propositions~\ref{prop:LmuT}, \ref{prop:LnuT}, \ref{prop:Lepsilon}, \ref{prop:Ltaux}, \ref{prop:Lsigmax}, and \ref{prop:Lrhox}. \end{proof} We are now ready for the final proof. As in Theorem~\ref{thm:main2}, let $d$ be the dimension of the space we are working in and let $c$ be an integer $ \ge d!+d$. \begin{proof}[Proof of Theorem~\ref{thm:main2}] For any $\mathcal D$-complex $X$ with $\dim \abs{X} = d$, we consider the collection of lattices \[ \mathscr L(X) := \{ L(A) : A \in \Max X \}. \] Start with a $\mathcal D$-subdivision $X$ with $\dim \abs{X} = d$ and 2-support $(P,Q)$. Assume that every element of $X$ is terminal with respect to $\mathscr S_x$. Consider a map $\theta : X \to \mathcal D$ defined as follows: If $A \in X$ has standard form $(p,S,a) \times (q,S,b) \times 0$, then \[ \theta(A) := (cp, S, ca) \times (d!q, S, d!b) \times \abs{S}. \] Then $\theta(X)$ is a $\mathcal D$-subdivision with 2-support $(cP, d!Q)$. Choose some lattice $L \in \mathscr L(X)$ and some nonzero $x \in \mathbb R^d / L$. Note that $\theta(X)$ is a $\mathcal D_x$-complex; indeed, for all $j \le \abs{S}$, we have $b_{1j} = 0$ or $d!$ and $a_{1j} = c \ge d! + d \ge b_{1j} + d$. Let $X' := \Delta_x^\ast(\theta(X))$, where $\Delta_x$ is from Theorem~\ref{thm:Deltax} and the $\ast$ construction is from Proposition~\ref{prop:canonicalrefinement}. Then $X'$ is a $\mathcal D$-subdivision with 2-support $(cP, d!Q)$ where every element is terminal with respect to $\mathscr S_x$. Comparing $\mathscr L(X)$ and $\mathscr L(X')$, by Theorem~\ref{thm:Deltax}, we have that $\mathscr L(X')$ is obtained by replacing at least one lattice of $\mathscr L(X)$ with lattices of lower index, while keeping the other lattices the same. Thus, if we repeat the above process on $X'$ instead of $X$, and so on, we will eventually obtain a $\mathcal D$-subdivision $Y$ with 2-support $(c^N P, (d!)^N Q)$ for some $N$, such that every element of $Y$ is terminal and $\mathscr L(Y) = \{ \mathbb Z^d \}$. Let $r$, $s$ be nonnegative integers. Consider the map $\omega_{r,s} : Y \to \mathcal C$ given by \[ \omega((p,S,a) \times (q,S,b) \times 0) = (rp+sq,S,ra+sb). \] Then $\omega_{r,s}(Y)$ is a $\mathcal C$-subdivision with support $rc^N P + s(d!)^N Q$. Let \[ Z := \Gamma^\ast(\omega_{r,s}(Y)), \] where $\Gamma$ is from Theorem~\ref{thm:Gamma}. Then $Z$ is a $\mathcal C$-subdivision with support $rc^N P + s(d!)^N Q$ such that for all $(p,S,a) \in Z$ in standard form, we have $\abs{S} = 0$ and $L(p,S,a) = \mathbb Z^d$. Thus $\Cay(Z)$ is a unimodular triangulation of $rc^N P + s(d!)^N Q$. Now, let $P$ be a $d$-dimensional integral polytope in $\mathbb R^d$, and let $X_0$ be any triangulation of $P$ into integral simplices. Let \[ X := \{ ((0),(T),(1)) \times ((0),(T),(1)) \times 0 \in \mathcal D : T \in X_0 \} \] Then $X$ is $\mathcal D$-subdivision with 2-support $(P,P)$, all of whose elements are terminal. Hence, applying the above argument to $X$ gives the result. \end{proof}
{ "timestamp": "2021-12-10T02:06:53", "yymm": "2112", "arxiv_id": "2112.04654", "language": "en", "url": "https://arxiv.org/abs/2112.04654", "abstract": "An integral polytope is a polytope whose vertices have integer coordinates. A unimodular triangulation of an integral polytope in $\\mathbb{R}^d$ is a triangulation in which all simplices are integral with volume $1/d!$. A classic result of Knudsen, Mumford, and Waterman states that for every integral polytope $P$, there exists a positive integer $c$ such that $cP$ has a unimodular triangulation. We strengthen this result by showing that for every integral polytope $P$, there exists $c$ such that for every positive integer $c' \\ge c$, $c'P$ admits a unimodular triangulation. This answers a longstanding question in the area.", "subjects": "Combinatorics (math.CO)", "title": "Unimodular triangulations of sufficiently large dilations", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.98866824479921, "lm_q2_score": 0.8104789086703224, "lm_q1q2_score": 0.8012947600818668 }
https://arxiv.org/abs/1603.00677
Karhunen-Loeve expansions of Levy processes
Karhunen-Loeve expansions (KLE) of stochastic processes are important tools in mathematics, the sciences, economics, and engineering. However, the KLE is primarily useful for those processes for which we can identify the necessary components, i.e., a set of basis functions, and the distribution of an associated set of stochastic coefficients. Our ability to derive these components explicitly is limited to a handful processes. In this paper we derive all the necessary elements to implement the KLE for a square-integrable Levy process. We show that the eigenfunctions are sine functions, identical to those found in the expansion of a Wiener process. Further, we show that stochastic coefficients have a jointly infinitely divisible distribution, and we derive the generating triple of the first d coefficients. We also show, that, in contrast to the case of the Wiener process, the coefficients are not independent unless the process has no jumps. Despite this, we develop a series representation of the coefficients which allows for simulation of any process with a strictly positive Levy density. We implement our theoretical results by simulating the KLE of a variance gamma process.
\section{Introduction} Fourier series are powerful tools in mathematics and many other fields. The Karhunen-Lo{\`e}ve theorem (KLT) allows us to create generalized Fourier series from stochastic processes in an, in some sense, optimal way. Arguably the most famous application of the KLT is to derive the classic sine series expansion of a Wiener process $W$ on $[0,1]$. Specifically, \begin{align}\label{eq:wklt} W_t = \sqrt{2}\sum_{k \geq 1}Z_k \frac{\sin\left(\pi(k-\frac{1}{2}t)\right)}{\pi\left(k - \frac{1}{2}\right)} \end{align} where convergence of the series is in $L^2(\Omega,{\mathbb P})$ and uniform in $t \in [0,1]$, and the $\{Z_k\}_{k\geq 1}$ are i.i.d. standard normal random variables. The main result of this paper is to show that a square integrable L\'evy process admits a similar representation as a series of sine functions; the key difference is that the stochastic coefficients are no longer normal nor independent.\\ \\ The KLT applies much more generally and is thus an important tool in many fields. For example, we see applications of the KLT and Principal Component Analysis, its discrete time counterpart, in physics and engineering \cite{Ghanesh,Phoon}, \cite[Chapter 10]{deep}, in signal and image processing \cite{unser1998wavelets}, \cite[Chapter 1]{sigbook}, in finance and economics \cite{Benko,cont2002dynamics,gun1} and other areas. For interesting recent theses on the KLT from three different points of view see also \cite{shijin} (probability and time series), \cite[Chapter 7]{Luo} (stochastic partial differential equations), and \cite{wang} (statistics).\\ \\ Deriving the Karhunen--L{\`o}eve expansion (KLE) of the type \eqref{eq:wklt} for a square integrable stochastic process $X$ on $[a,b]$ requires two steps: first, one must solve a Fredholm integral equation to obtain the basis functions $\{e_k\}_{k\geq1}$ (c.f. the sine functions in Equation \ref{eq:wklt}). Second, one must identify the distribution of the stochastic coefficients \begin{align}\label{eq:stocoef} Z_k := \int_a^bX_te_k(t){\textnormal d} t,\quad k\in \mathbb{N}. \end{align} In general, obtaining both the basis functions and the distribution of the stochastic coefficients is not an easy task, and we have full knowledge in only a few specific cases. Besides the Wiener process, the Brownian Bridge process, the Anderson--Darling process, and spherical fractional Brownian Motion (see \cite{Istas20061578} for the latter) are some examples. For further examples with derivation see \cite[Chapter 1]{shijin}. Non-Gaussian processes pose an additional challenge and the problem of deriving the KLE is usually left to numerical means (see e.g., \cite{Phoon}). \\ \\ In this paper we derive all the elements of the KLE for a square integrable L\'{e}vy process on the interval $[0,T]$. The result is timely since in many of the fields mentioned above, especially in finance, but recently also in the area of image/signal processing (see e.g., \cite{bouya}), L\'{e}vy models are becoming increasingly popular. In Section \ref{sec:klt} we show that the basis functions are sine functions, identical to those in \eqref{eq:wklt}, and that the first $d$ stochastic coefficients are jointly distributed like an infinitely divisible (ID) random vector. We identify the generating triple of this vector from which it follows that the coefficients are independent only when the process has no jumps, i.e., when the process is a scaled Wiener process with drift. Although simulating dependent multivariate random variables from a characteristic function is generally difficult, in Section \ref{sec:shotnoise} we derive a shot-noise (series) representation for \begin{align}\label{eq:vecstoco} Z^{(d)} := (Z_1,Z_2,\ldots,Z_d)^{\text{\textbf{T}}},\quad d \in \mathbb{N}, \end{align} for those processes which admit a strictly positive L\'{e}vy density. This result, in theory, allows us to simulate the truncated KLE for a large class of L\'{e}vy models. We conclude by generating some paths of a $d$-term KLE approximation of a variance gamma process.\\ \\ To begin, we recall the necessary facts from the theory of L\'{e}vy processes and ID random vectors. \section{Facts from the theory of L\'{e}vy processes}\label{sec:introlev} The L\'{e}vy-Khintchine theorem states that every $d$-dimensional ID random vector $\xi$ has a Fourier transform of the form \begin{align*} {\mathbb E}[e^{\i\langle \mathbf{z},\xi\rangle}] = e^{-\Psi(\mathbf{z})},\quad \mathbf{z} \in {\mathbb R}^d, \end{align*} where \begin{align}\label{eq:char} \Psi(\mathbf{z}) = \frac{1}{2}\mathbf{z}^{\textbf{T}}Q\mathbf{z} - \i\langle \mathbf{a}, \mathbf{z} \rangle - \int_{{\mathbb R}^d\backslash\{\mathbf{0}\}}e^{\i\langle \mathbf{z}, \mathbf{x} \rangle} - 1 - \i\langle \mathbf{z},\mathbf{x} \rangle h(\mathbf{x})\nu({\textnormal d} \mathbf{x}), \end{align} and where $\mathbf{a} \in {\mathbb R}^d$, $Q$ is a positive semi-definite matrix, and $\nu({\textnormal d} \mathbf{x})$ is a measure on ${\mathbb R}^d\backslash\{\mathbf{0}\}$ satisfying \begin{align}\label{eq:int_cond} \int_{{\mathbb R}^d\backslash\{\mathbf{0}\}}\min(1,\vert \mathbf{x} \vert^2)\nu({\textnormal d} \mathbf{x} ) < \infty. \end{align} The function $h$ is known as the cut-off function; in general, we need such a function to ensure convergence of the integral. An important fact is that up to a choice of $h$, the generating triple $(\mathbf{a},Q,\nu)$ uniquely identifies the distribution of $\xi$. The L\'{e}vy-Khintchine theorem for L\'{e}vy processes gives us an analogously powerful result, specifically, for any $d$-dimensional L\'{e}vy process $X$ we have \begin{align*} {\mathbb E}[e^{i\langle \mathbf{z},X_t\rangle}] = e^{-t\Psi(\mathbf{z})},\quad \mathbf{z} \in {\mathbb R}^d,\,t \geq 0, \end{align*} where $\Psi$ is as in \eqref{eq:char} and $X$ is uniquely determined, up to identity in distribution, by the triple $(\mathbf{a},Q,\nu)$. Following convention, we will refer to the function $\Psi$ as the characteristic exponent of $\xi$ (resp. $X$) and will write $\Psi_{\xi}$ (resp. $\Psi_{X}$) if there is the potential for ambiguity. In one dimension we will write $(a,\sigma^2,\nu)$ for the generating triple; the measure $\nu$ will always be referred to as the L\'{e}vy measure. When $\nu({\textnormal d} x) = \pi(x){\textnormal d} x$ for some density function $\pi$, we will write $(a,\sigma^2,\pi)$ and refer to $\pi$ as the L\'{e}vy density. If we wish to be specific regarding the cut-off function we will write $(\mathbf{a},Q,\nu)_{h\equiv \cdot}$ or $(a,\sigma^2,\nu)_{h\equiv\cdot}$ for the generating triples. \\ \\ \noindent In this article we will work primarily with one dimensional L\'{e}vy processes having zero mean and finite second moment; by this we mean that ${\mathbb E}[X_t] = 0$ and ${\mathbb E}[X_t^2] < \infty$ for every $t \geq 0$. We will denote the set of all such L\'{e}vy processes by $\mathcal{K}$. One may show that the later condition implies that $\Psi$ is twice differentiable. Thus, when we work with a process $X \in \mathcal{K}$, we can express the variance of $X_t$ as \begin{align*} \textnormal{Var}(X_t) = {\mathbb E}[X_t^{2}] = \Psi''(0)t, \end{align*} and the covariance of $X_t$ and $X_s$ as \begin{align*} \textnormal{Cov}(X_s,X_t) = {\mathbb E}[X_sX_t] = \Psi''(0)\min(s,t). \end{align*} For notational convenience we will set $\alpha := \Psi''(0)$. \\ \\ \noindent The existence of moments for both L\'{e}vy processes and ID random vectors can be equivalently expressed in terms of the L\'{e}vy measure. An ID random vector $\xi$ or L\'{e}vy process $X$ with associated L\'{e}vy measure $\nu$ has a finite second moment (meaning the component-wise moments) if, and only if, \begin{align}\label{eq:conditionA} \int_{\vert \mathbf{x}\vert > 1} \vert \mathbf{x}\vert^2\nu({\textnormal d} \mathbf{x}) < \infty \tag{Condition A}. \end{align} We will denote the class of ID random vectors with zero first moment and finite second moment by $\mathcal{C}$. The subset of $\mathcal{C}$ which also satisfies \begin{align}\label{eq:conditionB} \int_{\vert \mathbf{x}\vert \leq 1} \vert \mathbf{x}\vert\nu({\textnormal d} \mathbf{x}) < \infty. \tag{Condition B} \end{align} will be denoted $\mathcal{CB}$ and $\mathcal{KB}$ will denote the analogous subset of $\mathcal{K}$. We remark that any $\xi \in \mathcal{C}$ (resp. $X \in \mathcal{K}$) necessarily has a representation of the form $(\mathbf{0},Q,\nu)_{h\equiv 1}$ (resp. $(0,\sigma^2,\nu)_{h\equiv 1}$). Additionally, any $d$-dimensional $\xi \in \mathcal{CB}$ necessarily has representation $(\mathbf{a},Q,\nu)_{h\equiv 0}$ where $\mathbf{a}$ has entries \begin{align*} -\int_{\mathbb{R}^d\backslash\{\mathbf{0}\}}P_k(\mathbf{x})\nu({\textnormal d} \mathbf{x}),\quad k \in \{1,2,\ldots d\} \end{align*} and $P_k$ is the projection onto the $k$-th component. Analogously, if $X \in \mathcal{KB}$ then we have representation $(a,\sigma^2,\nu)_{h\equiv 0}$ where $a = -\int_{\mathbb{R}\backslash\{0\}}x\nu({\textnormal d} x)$. \section{The Karhunen--Lo\`{e}ve theorem}\label{sec:klt} Given a real valued continuous time stochastic process $X$ defined on an interval $[a,b]$ and an orthonormal basis $\{\phi_k\}_{k \geq 1}$ for $L^2([a,b])$ we might try to express $X$ as a generalized Fourier series \begin{align}\label{eq:eigen} X_t = \sum_{k=1}^\infty Y_k\phi_k(t),\quad\text{ where }\quad Y_k := \int_a^bX_t\phi_k(t){\textnormal d} t. \end{align} In this section, our chosen basis will be derived from the eigenfunctions corresponding to the non-zero eigenvalues $\{\lambda_k\}_{k \geq 1}$ of the integral operator $K:L^{2}([a,b]) \rightarrow L^{2}([a,b])$, \begin{align*} (Kf)(s):=\int_a^b\textnormal{Cov}(X_{s},X_t)f(t){\textnormal d} t. \end{align*} When the covariance satisfies a continuity condition it is known (see for example \cite{Ghanesh} Section 2.3.3) that the normalized set of eigenfunctions $\{e_k\}_{k \geq 1}$ of $K$ is countable and forms a basis for $L^{2}([a,b])$. When we choose this basis in \eqref{eq:eigen} we adopt the special notation $\{Z_k\}_{d \geq 1}$ for the stochastic coefficients. In this case, the expansion is optimal in a number of ways. Specifically, we have: \begin{theorem}[The Karhunen-Lo\`{e}ve Theorem]\label{theo:klt} Let $X$ be a real valued continuous time stochastic process on $[a,b]$ such that $0 \leq a \leq b < \infty$ and let ${\mathbb E}[X_t] = 0$ and ${\mathbb E}[X^2_t] < \infty$ for each $t \in [a,b]$. Further, suppose $\textnormal{Cov}(X_s,X_t)$ is continuous on $[a,b]\times[a,b]$. \begin{enumerate}[(i)] \item Then, \begin{align*} {\mathbb E}\left[\left( X_t - \sum_{k=1}^{d}Z_ke_k(t)\right)^2\right] \rightarrow 0,\quad\text{ as }\quad d \rightarrow \infty \end{align*} uniformly for $t \in [a,b]$. Additionally, the random variables $\{Z_k\}_{k \geq 1}$ are uncorrelated and satisfy ${\mathbb E}[Z_k] = 0$ and ${\mathbb E}[Z_k^2] = \lambda_k$. \item For any other basis $\{\phi_k\}_{k \geq 1}$ of $L^2([a,b])$, with corresponding stochastic coefficients $\{Y_k\}_{k \geq 1}$, and any $d \in \mathbb{N}$, we have \begin{align*} \int_{a}^{b}{\mathbb E}\left[\left(\varepsilon_d(t)\right)^2\right]{\textnormal d} t \leq \int_{a}^{b}{\mathbb E}\left[\left(\tilde\varepsilon_d(t)\right)^2\right]{\textnormal d} t, \end{align*} where $\varepsilon_d$ and $\tilde\varepsilon_d$ are the remainders $\varepsilon_d(t) :=\sum_{d+1}^{\infty}Z_ke_k(t)$ and $\tilde\varepsilon_d(t) :=\sum_{d+1}^{\infty}Y_k\phi_k(t)$. \end{enumerate} \end{theorem} \noindent Going forward we assume the order of the eigenvalues, eigenfunctions, and the stochastic coefficients is determined according to $\lambda_1 \geq \lambda_2 \geq \lambda_3, \ldots$. \\ \\ \noindent According to Ghanem and Spanos \cite{Ghanesh} the Karhunen-Lo\`{e}ve theorem was proposed independently by Karhunen \cite{Karhun}, Lo\`{e}ve \cite{love}, and Kac and Siegert \cite{Katchy}. Modern proofs of the first part of the theorem can be found in \cite{Ashy} and \cite{Ghanesh} and the second part -- the optimality of the truncated approximation -- is also proven in \cite{Ghanesh}. A concise and readable overview of this theory is given in \cite[Chapter 7.1]{Luo}.\\ \\ \noindent We see that although the KLT is quite general, it is best applied in practice when can determine the three components necessary for a Karhunen-Lo\'{e}ve expansion: the eigenfunctions $\{e_k\}_{k \geq 1}$; the eigenvalues $\{\lambda_k\}_{k \geq 1}$; and the distribution of the stochastic coefficients $\{Z_k\}_{k \geq 1}$. If we wish to use the KLE for simulation then we need even more: We also need to know how to simulate the random vector $Z^{(d)} = (Z_1,Z_2,\ldots,Z_d)$ which, in general, has uncorrelated but not necessarily independent components. \\ \\ For Gaussian processes, the second obstacle is removed, since one can show that the $\{Z_k\}_{k \geq 1}$ are again Gaussian, and therefore independent. There are, of course, many ways to simulate a vector of independent Gaussian random variables. For a process $X \in \mathcal{K}$, the matter is slightly more complicated as we establish in Theorem \ref{theo:main1}. However, since the covariance function of a process $X \in \mathcal{K}$ differs from that of a Wiener process only by the scaling factor $\alpha$, the method for determining the eigenfunctions and the eigenvalues for a L\'{e}vy process is identical to that employed for a Wiener process. Therefore, we omit the proof of the following proposition, and direct the reader to \cite[pg. 41]{Ashy} where the proof for the Wiener process is given. \begin{proposition}\label{prop:joe} The eigenvalues and associated eigenfunctions of the operator $K$ defined on $L^{2}([0,T])$ with respect to $X \in \mathcal{K}$ are given by \begin{align}\label{eq:efuncval} \lambda_k= \frac{\alpha T^2}{\pi^2\left(k - \frac{1}{2}\right)^2},\quad\text{ and }\quad e_k(t)= \sqrt{\frac{2}{T}}\sin\left(\frac{\pi}{T}\left(k-\frac{1}{2}\right)t\right),\quad k\in \mathbb{N},\,t\in[0,T]. \end{align} \end{proposition} \noindent A nice consequence of Proposition \ref{prop:joe} and Theorem \ref{theo:klt} is that it allows us to estimate the amount of total variance $v(T) := \int_0^T\textnormal{Var}(X_t){\textnormal d} t = \int_0^T{\mathbb E}[X^2_t]{\textnormal d} t = \alpha T^2/2$ we capture when we represent our process by a truncated KLE. Using the orthogonality of the $\{e_k\}_{k \geq 1}$, and the fact that ${\mathbb E}[Z_k^2] = \lambda_k$ for each $k$, it is straightforward to show that the total variance satisfies $v(T) = \sum_{k \geq 1}\lambda_k$. Therefore, the total variance explained by a $d$-term approximation is \begin{align*} \frac{\sum_{k=1}^{d}\lambda_k}{v(T)} = \frac{2}{\pi^2}\sum_{k=1}^{d}\frac{1}{\left(k-\frac{1}{2}\right)^2}. \end{align*} By simply computing the quantity on the right we find that the first 2, 5 and 21 terms already explain $90\%$, $95\%$, and $99\%$ of the total variance of the process. Additionally, we see that this estimate holds for all $X \in \mathcal{K}$ independently of $\alpha$ or $T$.\\ \\ \noindent The following lemma is the important first step in identifying the joint distribution of the stochastic coefficients of the KLE for $X \in \mathcal{K}$. The reader should note, however, that the lemma applies to more general L\'{e}vy processes, and is not just restricted to the set $\mathcal{K}$. \begin{lemma}\label{lem:bert} Let $X$ be a L\'evy process and let $\{f_k\}_{k = 1}^d$ be a collection of functions which are in $L^1([0,T])$. Then the vector $\mathbf{\xi}$ consisting of elements \begin{align*} \xi_k = \int_0^{T}X_tf_k(s){\textnormal d} s, \quad k \in \{1,2,\ldots, d\}, \end{align*} has an ID distribution with characteristic exponent \begin{align}\label{eq:bert} \Psi_{\mathbf{\xi}}(\mathbf{z}) = \int_0^T\Psi_X\left(\langle \mathbf{z}, \mathbf{u}(t) \rangle\right){\textnormal d} t,\quad \mathbf{z} \in {\mathbb R}^d, \end{align} where $\mathbf{u}:[0,T]\rightarrow {\mathbb R}^{d}$ is the function with $k$-th component $u_{k}(t) :=\int_t^T f_{k}(s){\textnormal d} s$, $k \in \{1,2,\ldots,d\}$. \end{lemma} \begin{remark} A similar identity to \eqref{eq:bert} is known, see pg. 128 in \cite{handbook}. In the proof of Lemma \ref{lem:bert}, we borrow some ideas from there. Since the proof is rather lengthy we relegate it to the Appendix. \end{remark} \noindent With Lemma \ref{lem:bert} and Proposition \ref{prop:joe} in hand, we come to our first main result. In the following theorem we identify the generating triple of the vector $Z^{(d)}$ containing the first $d$ stochastic coefficients of the KLE for a process $X \in \mathcal{K}$. Although it follows that $Z^{(d)}$ has dependent entries (see Corollary \ref{cor:main2}), Theorem \ref{theo:main1}, and in particular the form of the L\'{e}vy measure $\Pi$, will also be the key to simulating $Z^{(d)}$. Going forward we use the notation $\mathcal{B}_{S}$ for the Borel sigma algebra on the topological space $S$. \begin{theorem}\label{theo:main1} If $X \in \mathcal{K}$ with generating triple $(0,\sigma^2,\nu)_{h \equiv 1}$ then $Z^{(d)} \in \mathcal{C}$ with generating triple \\ $(\mathbf{0},\mathcal{Q},\Pi)_{h\equiv 1}$ where $\mathcal{Q}$ is a diagonal $d\times d$ matrix with entries \begin{align}\label{eq:themat} q_{k,k} := \frac{\sigma^2}{2}\frac{T^2}{\pi^2\left(k - \frac{1}{2}\right)^2},\quad k \in \{1,2,\ldots,d\}, \end{align} and $\Pi$ is the measure, \begin{align}\label{eq:levymeas} \Pi(B) := \int_{{\mathbb R}\backslash\{0\}\times[0,T]}{\mathbb I}(f(\mathbf{v}) \in B)(\nu\times\lambda)({\textnormal d} \mathbf{v}),\quad B \in \mathbb{B}_{\mathbb{R}^d\backslash\{\mathbf{0}\}}, \end{align} where $\lambda$ is the Lebesgue measure on $[0,T]$ and $f:{\mathbb R}\times[0,T] \rightarrow {\mathbb R}^d$ is the function \begin{align}\label{eq:theff} (x,t) \mapsto \frac{\sqrt{2T}x}{\pi}\left( \frac{\cos\left(\frac{\pi}{T}\left(1 - \frac{1}{2}\right) t \right)}{\left(1 - \frac{1}{2}\right)},\frac{\cos\left(\frac{\pi}{T}\left(2 - \frac{1}{2}\right) t \right)}{\left(2 - \frac{1}{2}\right)},\ldots,\frac{\cos\left(\frac{\pi}{T}\left(d - \frac{1}{2}\right) t \right)}{\left(d - \frac{1}{2}\right)}\right)^{\textnormal{\textbf{T}}}. \end{align} \end{theorem} \begin{proof} We substitute the formula for the characteristic exponent (Formula \ref{eq:char} with $a=0$ and $h\equiv 1$) and the eigenfunctions (Formula \ref{eq:efuncval}) into \eqref{eq:bert} and carry out the integration. Then \eqref{eq:themat} follows from the fact that \begin{align*} u_k(t) = \int_t^Te_k(s){\textnormal d} s = \sqrt{\frac{2}{T}}\int_t^T\sin\left(\frac{\pi}{T}\left(k - \frac{1}{2}\right)s\right){\textnormal d} s = \sqrt{2T}\frac{\cos\left(\frac{\pi}{T}\left(k - \frac{1}{2}\right)t\right)}{\pi(k-\frac{1}{2})}, \quad k \in \mathbb{N} \end{align*} and that the $\{u_k\}_{k \geq 1}$ are therefore also orthogonal on $[0,T]$. \\ \\ \noindent Next we note that $f$ is a continuous function from ${\mathbb R}\backslash\{0\}\times[0,T]$ to ${\mathbb R}^d$ and is therefore $\left(\mathcal{B}_{{\mathbb R}\backslash\{0\}\times[0,T]},\mathcal{B}_{{\mathbb R}^d\backslash\{\mathbf{0}\}}\right)$ measurable. Therefore, $\Pi$ is nothing other than the push forward measure obtained from $(\nu\times\lambda)$ and $f$; in particular, it is a well-defined measure on $\mathcal{B}_{{\mathbb R}^d\backslash\{\mathbf{0}\}}$. It is also a L\'{e}vy measure that satisfies \ref{eq:conditionA} since \begin{align}\label{eq:prolev} \int_{\vert \mathbf{x} \vert > 1}\vert \mathbf{x} \vert^2\Pi(d \mathbf{x}) \leq \int_{{\mathbb R}^{d}\backslash\{\mathbf{0}\}}\vert \mathbf{x}\vert^2\Pi({\textnormal d} \mathbf{x}) = \frac{2T}{\pi^2}\int_0^T\left(\sum_{k=1}^du_k^2(t)\right){\textnormal d} t\int_{{\mathbb R}\backslash\{0\}}x^2\nu(d x) < \infty, \end{align} where the final inequality follows from the fact that $X \in \mathcal{K}$. Applying Fubini's theorem and a change of variables, i.e., \begin{align*} \int_0^T\int_{{\mathbb R}\backslash\{0\}}e^{\i x\langle \mathbf{z}, \mathbf{u}(t) \rangle} - 1 - \i x\langle \mathbf{z},\mathbf{u}(t) \rangle \nu({\textnormal d} x){\textnormal d} t &= \int_{{\mathbb R}\backslash\{0\}\times[0,T]}e^{\i \langle \mathbf{z}, f(\mathbf{v}) \rangle} - 1 - \i \langle \mathbf{z},f(\mathbf{\mathbf{v}}) \rangle (\nu\times\lambda)({\textnormal d} \mathbf{v})\\ &= \int_{{\mathbb R}^d\backslash\{\mathbf{0}\}}e^{\i\langle \mathbf{z}, \mathbf{x} \rangle} - 1 - \i\langle \mathbf{z},\mathbf{x} \rangle \Pi({\textnormal d} \mathbf{x}), \end{align*} concludes the proof of infinite divisibility. Finally, noting that \begin{align*} {\mathbb E}[Z_k] = {\mathbb E}\left[\int_0^TX_te_k(t){\textnormal d} t\right] = \int_0^T{\mathbb E}[X_t]e_k(t){\textnormal d} t = 0,\quad k \in \{1,2,\ldots,d\}, \end{align*} shows that $Z^{(d)} \in \mathcal{C}$. \end{proof} \begin{remark} Note, that if we set $\sigma=1$, $\nu \equiv 0$, and $T=1$ we may easily recover the KLE of the Wiener process, i.e., \eqref{eq:wklt}, from Theorem \ref{theo:main1}. \end{remark} \noindent We gather some fairly obvious but important consequences of Theorem \ref{theo:main1} in the following corollary. \begin{corollary}\label{cor:main1} Suppose $X \in \mathcal{K}$, then: \begin{enumerate}[(i)] \item $X \in \mathcal{KB}$ with generating triple $(a,\sigma^2,\nu)_{h\equiv 0}$ if, and only if, $Z^{(d)} \in \mathcal{CB}$ with generating triple $(\mathbf{a},\mathcal{Q},\Pi)_{h\equiv 0}$, where $\mathcal{Q}$ and $\Pi$ are as defined in \eqref{eq:themat} and \eqref{eq:levymeas} and $\mathbf{a}$ is the vector with entries \begin{align}\label{eq:drift} a_k := a\frac{(-1)^{k+1}\sqrt{2}T^{\frac{3}{2}}}{\pi^2\left(k - \frac{1}{2}\right)^2},\quad k \in \{1,2,\ldots,d\}. \end{align} \item $X$ has finite L\'{e}vy measure $\nu$ if, and only if, $Z^{(d)}$ has finite L\'{e}vy measure $\Pi$. \end{enumerate} \end{corollary} \begin{proof}\ \\ \emph{(i)} Since \begin{align*} \int_{\mathbb{R}^d\backslash\{\mathbf{0}\}}\vert \mathbf{x} \vert \Pi({\textnormal d} \mathbf{x}) = \frac{\sqrt{2T}}{\pi}\int_{0}^{T}\left\vert\sum_{k=1}^{d}u_k(t)\right\vert{\textnormal d} t\int_{\mathbb{R}\backslash\{0\}}\vert x \vert \nu({\textnormal d} x) \end{align*} and \ref{eq:conditionA} is satisfied by both $\nu$ and $\Pi$ it follows that \ref{eq:conditionB} is satisfied for $\nu$ if, and only if, it is satisfied for $\Pi$. Formula \ref{eq:drift} then follows from the fact that \begin{align*} -\int_{\mathbb{R}^d\backslash\{\mathbf{0}\}}P_k(\mathbf{x}) \Pi({\textnormal d} \mathbf{x}) = -\int_{\mathbb{R}\backslash\{0\}}x\nu({\textnormal d} x)\frac{\sqrt{2T}}{\pi}\int_0^T\frac{\cos\left(\frac{\pi}{T}\left(k - \frac{1}{2}\right) t \right)}{\left(k - \frac{1}{2}\right)}{\textnormal d} t = a\frac{(-1)^{k+1}\sqrt{2}T^{\frac{3}{2}}}{\pi^2\left(k - \frac{1}{2}\right)^2}. \end{align*} \noindent \emph{(ii)} Straightforward from the definition of $\Pi$ in Theorem \ref{theo:main1}. \end{proof} \noindent Also intuitively obvious, but slightly more difficult to establish rigorously, is the fact that the entries of $Z^{(d)}$ are dependent unless $\nu \equiv 0$. \begin{corollary}\label{cor:main2} If $X \in \mathcal{K}$ then $Z^{(d)}$ has independent entries if, and only if, $\nu$ is the zero measure. \end{corollary} \noindent To prove Corollary \ref{cor:main2} we use the fact that a $d$-dimensional ID random vector with generating triple $(\mathbf{a},Q,\nu)$ has independent entries if, and only if, $\nu$ is supported on the union of the coordinate axes and $Q$ is diagonal (see E 12.10 on page 67 in \cite{sato}). For this purpose we define, for a vector $\mathbf{x} = (x_1,x_2,\ldots,x_d)^{\textbf{T}} \in {\mathbb R}^d$ such that $\,x_k > 0,\,k\in\{1,2,\ldots,d\}$, the sets \begin{align*} \mathcal{I^+}(\mathbf{x}) := \Pi_{k=1}^{d}(x_k,\infty),\quad\text{and}\quad\mathcal{I^-}(\mathbf{x}) := \Pi_{k=1}^{d}(-\infty,-x_k), \end{align*} where we caution the reader that the symbol $\Pi$ indicates the Cartesian product and not the L\'{e}vy measure of $Z^{(d)}$. \\ \\ \noindent In the proof below, and throughout the remainder of the paper, $f$ will always refer to the function defined in \eqref{eq:theff}, and $f_k$ to the $k$-th coordinate of $f$. \begin{proof}[Proof of Corollary \ref{cor:main2}]\ \\ \noindent ($\Leftarrow$) The assumption $\nu \equiv 0$ implies our process is a scaled Wiener process in which case it is well established that $Z^{(d)}$ has independent entries. Alternatively, this follows directly the fact that the matrix $\mathcal{Q}$ in Theorem \ref{theo:main1} is diagonal.\\ \\ \noindent ($\Rightarrow$) We assume that $\nu$ is not identically zero and show that there exists $\mathbf{x}$ such that either $\Pi(\mathcal{I}^+(\mathbf{x}))$ or $\Pi(\mathcal{I}^-(\mathbf{x}))$ is strictly greater than zero.\\ \\ \noindent Since $\nu(\mathbb{R}\backslash\{0\}) > 0$ there must exist $\delta > 0$ such that one of $\nu((-\infty,-\delta))$ and $\nu((\delta,\infty))$ is strictly greater than zero; we will initially assume the latter. We observe that for $d \in \mathbb{N}$, $d \geq 2$, the zeros of the function $h_d:[0,T]\rightarrow{\mathbb R}$ defined by \begin{align*} t \mapsto \frac{\cos\left(\frac{\pi}{T}\left(d - \frac{1}{2}\right) t \right)}{\left(d - \frac{1}{2}\right)}, \end{align*} occur at points $\{nT/(2d-1)\}_{n=1}^{2d -1}$, and therefore the smallest zero is $t_d := T/(2d-1)$. From the fact that the cosine function is positive and decreasing on $[0,\pi/2]$ we may conclude that \begin{align*} \frac{\cos\left(\frac{\pi}{T}\left(k - \frac{1}{2}\right) t \right)}{\left(k - \frac{1}{2}\right)} > \epsilon,\quad k\in\{1,2,\ldots,d\},\quad t \in [0,t_d/2], \end{align*} where $\epsilon = h_d(t_d/2) > 0$. Now, let $\mathbf{x}$ be the vector with entries $x_k = \delta\epsilon\sqrt{2T}/\pi$ for $k \in \{1,2,\ldots,d\}$. Then, \begin{align*} (\delta,\infty) \times [0,t_d/2] \subset f^{-1}\left(\mathcal{I}^+\left(\mathbf{x}\right)\right), \end{align*} since for $(x,t)\in (\delta,\infty) \times [0,t_d/2]$ we have \begin{align*} f_k(x,t) = \frac{\sqrt{2T}}{\pi}x\frac{\cos\left(\frac{\pi}{T}\left(k - \frac{1}{2}\right) t \right)}{\left(k - \frac{1}{2}\right)} > \delta\epsilon\frac{\sqrt{2T}}{\pi} = x_k,\quad k \in \{1,2,\ldots,d\}. \end{align*} But then, \begin{align} \Pi(\mathcal{I}^+(\mathbf{x})) \geq \nu((\delta,\infty))\times\lambda([0,t_d/2]) > 0. \end{align} If we had initially assumed that $\nu((-\infty,-\delta)) > 0$ we would have reached the same conclusion by using the interval $(-\infty,-\delta)$ and $\mathcal{I}^-(\mathbf{x})$. We conclude that $\Pi$ is not supported on the union of the coordinate axes, and so $Z^{(d)}$ does not have independent entries. \end{proof} \section{Shot-noise representation of $Z^{(d)}$}\label{sec:shotnoise} Although we have characterized the distribution of our stochastic coefficients $Z^{(d)}$ we are faced with the problem of simulating a random vector with dependent entries with only the knowledge of the characteristic function. In general, this seems to be a difficult problem, even generating random variables from the characteristic function is not straightforward (see for example \cite{fromchar}). In our case, thanks to Theorem \ref{theo:main1} we know that $Z^{(d)}$ is infinitely divisible and that the L\'{e}vy measure $\Pi$ has a special disintegrated form. This will help us build the connection with the so-called shot-noise representation of our vector $Z^{(d)}$. The goal is to represent $Z^{(d)}$ as an almost surely convergent series of random vectors. \\ \\ \noindent To explain this theory -- nicely developed and explained in \cite{rosy,theoapp} -- we assume that we have two random sequences $\{V_i\}_{i \geq 1}$ and $\{\Gamma_i\}_{i \geq 1}$ which are independent of each other and defined on a common probability space. We assume that each ${\Gamma_i}$ is distributed like a sum of $i$ independent exponential random variables with mean 1, and that the $\{V_i\}_{i \geq 1}$ take values in a measurable space $D$, and are i.i.d. with common distribution $F$. Further, we assume we have a measurable function $H:(0,\infty)\times D \rightarrow {\mathbb R}^{d}$ which we use to define the random sum \begin{align}\label{eq:shot} S_n:= \sum_{i = 1}^{n}H(\Gamma_i,V_i),\quad n \in \mathbb{N}, \end{align} and the measure \begin{align}\label{eq:meas} \mu(B) := \int_0^{\infty}\int_{D}{\mathbb I}(H(r,v) \in B)F({\textnormal d} v){\textnormal d} r,\quad B \in B_{\mathbb{R}^d\backslash\{\mathbf{0}\}}. \end{align} The function $C:(0,\infty)\rightarrow {\mathbb R}^{d}$ is defined by \begin{align}\label{eq:funA} C_k(s) := \int_0^{s}\int_DP_k(H(r,v))F({\textnormal d} v){\textnormal d} r,\quad k \in \{1,2,\ldots,d\}, \end{align} where, as before, $P_k$ is the projection onto the $k$-th component. The connection between \eqref{eq:shot} and ID random vectors is then explained in the following theorem whose results can be obtained by restricting Theorems 3.1, 3.2, and 3.4 in \cite{rosy} from a general Banach space setting to ${\mathbb R}^d$. \begin{theorem}[Theorems 3.1, 3.2, and 3.4 in \cite{rosy}]\label{theo:ros} Suppose $\mu$ is a L\'evy measure, then: \begin{enumerate}[(i)] \item If \ref{eq:conditionB} holds then $S_n$ converges almost surely to an ID random vector with generating triple $(\mathbf{0},\mathbf{0},\mu)_{h\equiv 0}$ as $n \rightarrow \infty$. \item If \ref{eq:conditionA} holds, and for each $v \in S$ the function $r \rightarrow \vert H(r,v) \vert$ is non increasing, then \begin{align}\label{eq:compsum} M_n := S_n - C(n), \quad n\in \mathbb{N} \end{align} converges almost surely to an ID random vector with generating triple $(\mathbf{0},\mathbf{0},\mu)_{h\equiv 1}$. \end{enumerate} \end{theorem} \noindent The name ``shot-noise representation" comes from the idea that $\vert H \vert$ can be interpreted as a model for the volume of the noise of a shot $V_i$ that occurred $\Gamma_i$ seconds ago. If $\vert H \vert $ is non increasing in the first variable, as we assume in case (ii) in Theorem \ref{theo:ros}, then the volume decreases as the elapsed time grows. The series $\lim_{n\rightarrow \infty}S_n$ can be interpreted as the total noise at the present time of all previous shots.\\ \\ \noindent The goal is to show that for any process in $\mathcal{K}$ whose L\'{e}vy measure admits a strictly positive density $\pi$, the vector $Z^{(d)}$ has a shot-noise representation of the form \eqref{eq:shot} or \eqref{eq:compsum}. To simplify notation we make some elementary but necessary observations/assumptions: First, we assume that $X$ has no Gaussian component $\sigma^2$. There is no loss of generality to this assumption, since if $X$ does have a Gaussian component then $Z^{(d)}$ changes by the addition of a vector of independent Gaussian random variables. This poses no issue from a simulation standpoint. Second, from \eqref{eq:char} we see that any L\'{e}vy process $X$ with representation $(0,0,\pi)_{h\equiv j}$, $j \in \{0,1\}$ can be decomposed into the difference of two independent L\'{e}vy processes, each having only positive jumps. Indeed, splitting the integral and making a change of variable $x \mapsto -x$ gives \begin{align}\label{eq:split} \Psi_{X}(z) &= -\int_{\mathbb{R}\backslash\{0\}} e^{\i zx} - 1 - \i zxj\pi(x){\textnormal d} x \nonumber \\ &= -\int_{0}^{\infty} e^{\i zx} - 1 - \i zxj\pi(x){\textnormal d} x -\int_{0}^{\infty}e^{\i z(-x)} - 1 - \i z(-x)j\pi(-x){\textnormal d} x \nonumber \\ &= \Psi_{X^+}(z) + \Psi_{-X^-}(z) \end{align} where $X^+$ (resp. $X^-$) has L\'{e}vy density $\pi(\cdot)$ (resp. $\pi(-\cdot)$) restricted to $(0,\infty)$. In light of this observation, the results of Theorem \ref{theo:main2} are limited to L\'{e}vy processes with positive jumps. It should be understood that for a general process we can obtain $Z^{(d)}$ by simulating $Z^{(d)}_+$ and $Z^{(d)}_-$ -- corresponding to $X^+$ and $X^-$ respectively -- and then subtracting the second from the first to obtain a realization of $Z^{(d)}$.\\ \\ \noindent Last, for a L\'{e}vy process with positive jumps and strictly positive L\'{e}vy density $\pi$, we define the function \begin{align}\label{eq:thefung} g(x) := \int_x^{\infty}\pi(s){\textnormal d} s. \end{align} which is just the tail integral of the L\'{e}vy measure. We see that $g$ is strictly monotonically decreasing to zero, and so admits a strictly monotonically decreasing inverse $g^{-1}$ on the domain $(0,g(0))$. \begin{theorem}\label{theo:main2} Let $\pi$ be a strictly positive L\'{e}vy density on $(0,\infty)$ and identically zero elsewhere. \begin{enumerate}[(i)] \item If $X \in \mathcal{KB}$ with generating triple $(a,0,\pi)_{h\equiv 0}$, then $Z^{(d)}$ has a shot noise representation \begin{align}\label{eq:shotnoise} Z^{(d)} \,{\buildrel d \over =}\ \mathbf{a} + \sum_{i \geq 1}H(\Gamma_i,U_i) \end{align} where $f$ and $\mathbf{a}$ are defined in \eqref{eq:theff} and \eqref{eq:drift} respectively, $\{U_i\}_{i \geq 1}$ is an i.i.d. sequence of uniform random variables on $[0,1]$, and \begin{align}\label{eq:H} H(r,v) := f(g^{-1}(r/T){\mathbb I}( 0 < r < g(0)),Tv). \end{align} \item If $X \in \mathcal{K}$ with generating triple $(0,0,\pi)_{h\equiv 1}$, then $Z^{(d)}$ has a shot noise representation \begin{align}\label{eq:const1} Z^{(d)} \,{\buildrel d \over =}\ \lim_{n\rightarrow\infty}\sum_{i = 1}^{n}H(\Gamma_i,U_i) - C(n), \end{align} where $H$ and $\{U_i\}_{i\geq 1}$ are as in Part $(i)$ and $C$ is defined as in \eqref{eq:funA}. \end{enumerate} \end{theorem} \begin{proof} Rewriting \eqref{eq:levymeas} to suit our assumptions and making a change of variables $t = Tv$ gives, for any $B \in \mathcal{B}_{\mathbb{R}^d\backslash\{\mathbf{0}\}}$ \begin{align*} \Pi(B) &= \int_0^T\int_0^{\infty}{\mathbb I}(f(x,t) \in B)\pi(x){\textnormal d} x{\textnormal d} t = \int_0^1\int_0^{\infty}{\mathbb I}(f(x,Tv) \in B)T\pi(x){\textnormal d} x{\textnormal d} v. \end{align*} Making a further change of variables $r = Tg(x)$ gives \begin{align*} \Pi(B) &= \int_0^1\int_0^{g(0)}{\mathbb I}(f(g^{-1}(r/T),Tv) \in B){\textnormal d} r{\textnormal d} v. \end{align*} Since $0 \notin B$, so that ${\mathbb I}(0 \in B) = 0$, we may conclude that \begin{align*} \Pi(B) = \int_0^{\infty}\int_{0}^1{\mathbb I}(f(g^{-1}(r/T){\mathbb I}( 0 < r < g(0)),Tv) \in B){\textnormal d} v{\textnormal d} r. \end{align*} From the definition of the function $f$ (Formula \ref{eq:theff}), and that of $g^{-1}$, it is clear that \begin{align}\label{eq:myh} (r,v) \mapsto f(g^{-1}(r/T){\mathbb I}( 0 < r < g(0)),Tv) \end{align} is measurable and non increasing in absolute value for any fixed $v$. Therefore, we can identify \eqref{eq:myh} with the function $H$, the uniform distribution on $[0,1]$ with $F$, and $\Pi$ with $\mu$. The results then follow by applying the results of Theorems \ref{theo:main1} and \ref{theo:ros} and Corollary \ref{cor:main1}. \end{proof} \noindent Going forward we will write simply $H(r,v) = f(g^{-1}(r/T),Tv)$ where it is understood that $g^{-1}$ vanishes outside the interval $(0,g(0))$. \subsubsection*{Discussion} There are two fairly obvious difficulties with the series representations of Theorem \ref{theo:main2}. The first -- this a common problem for all series representations of ID random variables when the L\'evy measure is not finite -- is that we have to truncate the series when $g(0) =\infty$ (equivalently $\nu(\mathbb{R}\backslash\{0\}) = \infty$). Besides the fact that in these cases our method fails to be exact, computation time may become an issue if the series converge too slowly. The second issue is that $g^{-1}$ is generally not known in closed form. Thus, in order to apply the method we will need a function $g$ that is amenable to accurate and fast numerical inversion. In the survey \cite{rosy} Rosi{\'n}ski reviews several methods, which depend on various properties of the L\'evy measure (for example, absolute continuity with respect to a probability distribution), that avoid this inversion. In a subsequent paper \cite{rosinski2007tempering} he develops special methods for the family of tempered $\alpha$-stable distributions that also do not require inversion of the tail of the L\'evy measure. We have made no attempt to adapt these techniques here, as the fall outside of the scope of this paper. However, this seems to be a promising area for further research.\\ \\ A nice feature of simulating a $d$-dimensional KLE of a L\'{e}vy process $X \in \mathcal{K}$ via Theorem \ref{theo:main2} is that we may increase the dimension incrementally. That is, having simulated a path of the $d$-term KLE approximation of $X$, \begin{align}\label{eq:partsum} S_{t}^{(d)} := \sum_{k=1}^{d}Z_ke_k(t),\quad t\in[0,T], \end{align} we may derive a path of $S^{(d+1)}$ directly from $S^{(d)}$ as opposed to starting a fresh simulation. We observe that a realization $z_k$ of $Z_k$ can be computed individually once we have the realizations $\{\gamma_i,u_i\}_{i\geq 1}$ of $\{\Gamma_i,\,U_i\}_{i\geq 1}$. Specifically, \begin{align*} z_k = a_k + \sum_{i \geq 1}\frac{\sqrt{2T}g^{-1}(\gamma_i/T)}{\pi}\frac{\cos\left(\pi\left(k - \frac{1}{2}\right)u_i\right)}{\left(k-\frac{1}{2}\right)}, \end{align*} when \ref{eq:conditionB} holds, with an analogous expression when it does not. Thus, if $s^{(d)}_t$ is our realization of $S^{(d)}_t$ we get a realization of $S^{(d+1)}_t$ via $s^{(d+1)}_t = s^{(d)}_t + z_{d+1}e_{d+1}(t)$.\\ \\ \noindent It is also worthwhile to compare the series representations for L\'{e}vy processes found in \cite{rosy} and the proposed method. As an example, suppose we have a subordinator $X$ with a strictly positive L\'{e}vy density $\pi$. Then, it is also true that \begin{align}\label{eq:regseries} \{X_t:t\in[0,T]\} \,{\buildrel d \over =}\ \left\{\sum_{i\geq 1}g^{-1}(\Gamma_i/T){\mathbb I}(TU_i < t):t\in[0,T]\right\}. \end{align} The key difference between the approaches, is that the series in \eqref{eq:regseries} depends on $t$, whereas the series representation of $Z^{(d)}$ is independent of $t$. Therefore, in \eqref{eq:regseries} we have to recalculate the series for each $t$, adding those summands for which $U_iT < t$. Of course, the random variables $\{\Gamma_i, U_i\}_{i\geq 1}$ need to be generated only once. On the other hand, while we have to simulate $Z^{(d)}$ only once for all $t$, each summand requires the evaluation of $d$ cosine functions, and for each $t$ we have to evaluate $d$ sine functions when we form the KLE. However, since there is no more randomness once we have generated $Z^{(d)}$ the second computation can be done in advance. \subsubsection*{Example} Consider the Variance Gamma (VG) process which was first introduced in \cite{initvg} and has since become a popular model in finance. The process can be constructed as the difference of two independent Gamma processes, i.e., processes with L\'{e}vy measures of the form \begin{align}\label{eq:gammasub} \nu({\textnormal d} x) = c\frac{e^{-\rho x}}{x}{\textnormal d} x,\quad x > 0, \end{align} where $c,\,\rho > 0$. For this example we use a Gamma process $X^+$ with parameters $c=1$ and $\rho=1$ and subtract a Gamma process $X^-$ with parameters $c=1$ and $\rho = 2$ to yield a VG process $X$. Assuming no Gaussian component or additional linear drift, it can be shown (see Proposition 4.2 in \cite{contan}) that the characteristic exponent of $X$ is then \begin{align*} \Psi_X(z) = -\left(\int_{0}^{\infty}(e^{\i zx} -1)\frac{e^{-x}}{x}{\textnormal d} x + \int_0^{\infty}(e^{-\i zx} -1)\frac{e^{-2x}}{x}{\textnormal d} x \right) = \log\left(1-\i z\right) + \log\left(1 + \frac{\i z}{2}\right). \end{align*} We observe that $X^+,\,X^{-} \notin \mathcal{K}$ since \begin{align*} {\mathbb E}[X^+_t] = \i t\Psi'_{X^+}(0) = t \neq 0\quad\text{ and }\quad{\mathbb E}[X^-_t] = \i t\Psi'_{X^-}(0) = \frac{t}{2} \neq 0. \end{align*} However, this is not a problem, since we can always construct processes $\tilde{X}^+,\,\tilde{X}^- \in \mathcal{K}$ by subtracting $t$ and $t/2$ from $X^+$ and $X^-$ respectively. We then generate the KLE of $\tilde{X}^+$ and add back $t$ to the result, and apply the analogous procedure for $X^{-}$. This is true generally as well, i.e., for a square integrable L\'{e}vy process with expectation ${\mathbb E}[X_t] = \i t\Psi_X'(0) \neq 0$ we can always construct a process $\tilde{X} \in \mathcal{K}$ by simply subtracting the expectation $\i t\Psi_X'(0)$.\\ \\ \noindent From \eqref{eq:gammasub} we see that the function $g$ will have the form \begin{align*} g(x) = c\int_x^{\infty}\frac{e^{-\rho s}}{s}{\textnormal d} s = cE_1(\rho x), \end{align*} where $E_1(x) := \int_x^{\infty}s^{-1}e^{-s}{\textnormal d} s$ is the exponential integral function. Therefore, \begin{align*} g^{-1}(T^{-1}r) = \frac{1}{\rho}E_1^{-1}\left(\frac{r}{Tc}\right). \end{align*} There are many routines available to compute $E_1$; we choose a Fortran implementation to create a lookup table for $E^{-1}_1$ with domain $[6.226\times10^{-22},45.47]$. We discretize this domain into $200000$ unevenly spaced points, such that the distance between two adjacent points is no more than 0.00231. Then we use polynomial interpolation between points.\\ \\ \noindent When simulating $Z^{(d)}_+$ we truncate the series \eqref{eq:shotnoise} when $(Tc)^{-1}\Gamma_i > 45.47$; at this point we have $g^{-1}(T^{-1}\Gamma_i) < \rho^{-1}10^{-19}$. Using the fact that the $\{\Gamma_i\}_{i\geq 1}$ are nothing other than the waiting times of a Poisson process with intensity one, we estimate that we need to generate on average $45Tc$ random variables to simulate $Z^{(d)}_+$ and similarly for $Z^{(d)}_-$. We remark that for the chosen process both the decay and computation of $g^{-1}$ are manageable.\\ \\ \noindent We simulate sample paths of $S^{(d)}$ for $d \in \{5,10,15,20,25,100,3000\}$ using the described approach. We also compute a Monte Carlo (MC) approximation of the expectation of $X$ by averaging over $10^6$ sample paths of the $d$-term approximation. Some sample paths and the results of the MC simulation are depicted in Figure \ref{fig1}, where the colors black, grey, red, green, blue, cyan, and magenta correspond to $d$ equal to 5, 10, 15, 20, 25, 100, and 3000 respectively.\\ \\ \noindent In Figure \ref{figa} we show the sample paths resulting from a simulation of $S^{(d)}$. We notice that the numerical results correspond with the discussion of Section \ref{sec:klt}: the large movements of the sample path are already captured by the 5-term approximation. We also notice peaks resulting from rapid oscillations before the bigger ``jumps" in the higher term approximations. This behaviour is magnified for the 3000-term approximation in Figure \ref{figb}. In classical Fourier analysis this is referred to as the Gibbs phenomenon; the solution in that setting is to replace the partial sums by Ces\`{a}ro sums. We can employ the same technique here, replacing $S^{(d)}$ with $C^{(d)}$, which is defined by \begin{align*} C_t^{(d)} :=\frac{1}{d}\sum_{k=1}^dS_t^{(k)}. \end{align*} It is relatively straightforward to show that $C^{(d)}$ converges to $X$ in the same manner as $S^{(d)}$ (as described in Theorem \ref{theo:klt} (i)). In Figure \ref{figc} we show the effect of replacing $S^{(d)}$ with $C^{(d)}$ on all sample paths, and in Figure \ref{figd} we show the $C^{(3000)}$ approximation -- now the Gibbs phenomenon is no longer apparent. \\ \\ In Figure \ref{fige} we show the MC simulation of $E[S^{(5)}_t]$ (black +) plotted together with $E[X_t] = t/2$ (green $\circ$). We see the 5-term KLE already gives a very good approximation. In Figure \ref{figf} we also show the errors $E[S^{(d)}_t] - E[X_t]$ for $d=5$ (black +), $d=25$ (blue $\circ$), and $d=3000$ (magenta $\square$). Again we have agreement with the discussion in Section \ref{sec:klt}: little is gained in our MC approximation of $E[X_t]$ by choosing a KLE with more than 25 terms. Recall that a KLE with 25 terms already captures more than 99\% of the total variance of the given process. \begin{figure} \centering \subfloat[] {\label{figa}\includegraphics[height =5.5cm]{all}} \subfloat[] {\label{figb}\includegraphics[height =5.5cm]{close}}\\ \subfloat[] {\label{figc}\includegraphics[height =5.5cm]{ces_all}} \subfloat[] {\label{figd}\includegraphics[height =5.5cm]{ces_close}}\\ \subfloat[] {\label{fige}\includegraphics[height =5.5cm]{exp_5}} \subfloat[] {\label{figf}\includegraphics[height =5.4cm]{exp_err}} \caption{(a) KLE sample paths (b) Example of Gibbs phenomenon (c) KLE with Ces\`{a}ro sums \\(d) Mitigated Gibbs phen. (e) ${\mathbb E}[X_t]=t/2$ and MC sim. of ${\mathbb E}[S^{(5)}_t]$ (f) MC Err. ${\mathbb E}[S^{(d)}_t] - t/2$} \label{fig1} \end{figure} \section*{Author's acknowledgements} My work is supported by the Austrian Science Fund (FWF) under the project F5508-N26, which is part of the Special Research Program ``Quasi-Monte Carlo Methods: Theory and Applications". I would like to thank Jean Bertoin for explaining his results in \cite{handbook} to me. This helped me extend identity \eqref{eq:bert} of Lemma \ref{lem:bert} from $C^1$ functions to $L^1$ functions. Further I would like to thank Alexey Kuznetsov and Gunther Leobacher for reading a draft of this paper and offering helpful suggestions. \newpage \begin{appendices} \section{Additional proof}\label{app:A} \begin{proof}[Proof of Lemma \ref{lem:bert}] We give a proof for continuously differentiable $\{f_k\}_{k = 1}^d$ first and then prove the general case. Accordingly, we fix $\mathbf{z} \in {\mathbb R}^d$, a collection of continuously differentiable $\{f_k\}_{k = 1}^d$ defined on $[0,T]$, and a L\'{e}vy process $X$ with state space ${\mathbb R}$. Instead of proving identity \eqref{eq:bert} directly for $X$ we will prove that \begin{align}\label{eq:withb} \Psi_{\xi^{(b)}}(\mathbf{z}) = \int_0^T\Psi_{X^{(b)}}\left(\langle \mathbf{z}, \mathbf{u}(t) \rangle\right){\textnormal d} t = b\int_0^T\Psi_X\left(\langle \mathbf{z}, \mathbf{u}(t) \rangle\right){\textnormal d} t,\quad \mathbf{z} \in {\mathbb R}^d,\quad b > 0, \end{align} where $X^{(b)}$ is the process defined by $X^{(b)}_t := X_{bt}$ and $\xi^{(b)}$ is the vector with entries \begin{align*} \xi^{(b)}_k := \int_{0}^{T}X_{bt}f_k(t){\textnormal d} t,\quad k \in \{1,2,\ldots, k\}. \end{align*} It is clear that $X^{(b)}$ is a L\'{e}vy process, that $\Psi_{X^{(b)}} = b\Psi_X$, and that \eqref{eq:bert} corresponds to the special case $b=1$. We focus on this more general result because it will lead directly to a proof of infinite divisibility. We begin by defining \begin{align}\label{eq:arr} {R^{(k)}_N} := \frac{T}{N}\sum_{n=0}^{N-1}f_k\left(\frac{(n+1)T}{N}\right)X_{\frac{b(n+1)T}{N}},\quad k \in \{1,2,\ldots, k\},\, N \in \mathbb{N}, \end{align} which are $N$-point, right-endpoint Riemann sum approximations of the random variables $\xi^{(b)}_k$. By the usual telescoping sum technique for L\'evy processes we can write \begin{align*} X_{\frac{b(n+1)T}{N}} &= \left(X_{\frac{b(n+1)T}{N}} - X_{\frac{bnT}{N}}\right) + \left(X_{\frac{bnT}{N}} - X_{\frac{b(n-1)T}{N}}\right) + \ldots + \left(X_{\frac{b2T}{N}} - X_{\frac{bT}{N}}\right) + X_{\frac{bT}{N}} \\ &\,{\buildrel d \over =}\ X^{(1)} + X^{(2)} + \ldots +X^{(n+1)}, \end{align*} where the random variables $X^{(i)}$ are independent and each distributed like $X_{bT/N}$. This allows us to rearrange the sum $R^{(k)}_N$ according to the random variables $X^{(i)}$, gathering together those with the same index. Therefore, we have \begin{align*} R^{(k)}_N \,{\buildrel d \over =}\ \sum_{n=0}^{N-1}X^{(n+1)}\left(\frac{T}{N}\sum_{j=n}^{N-1}f_k\left(\frac{(j+1)T}{N}\right)\right). \end{align*} We notice that the term in brackets on the right-hand side is a $(N - n)$-point, right-endpoint Riemann sum approximation for the integral of $f_k$ over the interval $[nT/N,T]$. Let us therefore define \begin{align}\label{eq:esstee} t^{(k)}_{n,N} := \frac{T}{N}\sum_{j=n}^{N-1}f_k\left(\frac{(j+1)T}{N}\right), \quad \text{ and } \quad s^{(k)}_{n,N} := \int_{\frac{nT}{N}}^{T}f_k(s){\textnormal d} s, \end{align} as well as the $d$-dimensional vectors $\mathbf{t}_{n,N}$ and $\mathbf{s}_{n,N}$ consisting of entries $t^{(k)}_{n,N}$ and $s^{(k)}_{n,N}$ respectively. We observe that \begin{align}\label{eq:convlev} {\mathbb E}[\exp(\i\langle \mathbf{z}, \xi^{(b)}\rangle] =\lim_{N \rightarrow \infty}{\mathbb E}\left[\exp\left(\sum_{n=0}^{N-1}\i X^{(n+1)}\langle \mathbf{z}, \mathbf{t}_{n,N}\rangle\right)\right] =\lim_{N \rightarrow \infty}\exp\left(-\frac{bT}{N}\sum_{n=0}^{N-1}\Psi_X(\langle \mathbf{z}, \mathbf{t}_{n,\,N}) \rangle\right), \end{align} where we have used the dominated convergence theorem to obtain the first equality, and the independence of the $X^{(i)}$ to obtain the final equality. Further, we get \begin{align*} \exp\left(-\int_0^T\Psi_{X^{(b)}}\left(\langle \mathbf{z}, \mathbf{u}(t) \rangle \right){\textnormal d} t\right) = \lim_{N \rightarrow \infty}\exp\left(-\frac{bT}{N}\sum_{n=0}^{N-1}\Psi_X(\langle \mathbf{z}, \mathbf{s}_{n,\,N}) \rangle \right), \end{align*} by using the left-endpoint Riemann sums. We note that $\vert \langle \mathbf{z}, \mathbf{t}_{n,\,N} \rangle - \langle \mathbf{z}, \mathbf{s}_{n,\,N} \rangle \vert \rightarrow 0$ uniformly in $n$ since \begin{align}\label{eq:bnd} \vert \langle \mathbf{z}, \mathbf{t}_{n,\,N} \rangle - \langle \mathbf{z}, \mathbf{s}_{n,\,N} \rangle\vert \leq \sum_{k=1}^{d}\vert z_k \vert\left \vert t^{(k)}_{n,N} - s^{(k)}_{n,N} \right \vert \leq \frac{d T^2}{N} \max_{1\leq k \leq d}\left\{\vert z_k\vert\sup_{x \in [0,T]}\vert f'_k(x) \vert \right\}, \end{align} where the last estimate follows from the well-known error bound $((c-a)^2\sup_{x \in [a,c]}\vert g'(x) \vert)/N$ for the absolute difference between an $N$-point, right end-point Riemann sum and the integral of a $C^1$ function $g$ over $[a,c]$. Then, by the continuity of $\Psi_X$, for any $\epsilon > 0$ we may choose an appropriately large $N$ such that \begin{align*} \left \vert \frac{1}{N}\sum_{n=0}^{N-1}\psi_X(\langle \mathbf{z}, \mathbf{t}_{n,\,N} \rangle ) - \frac{1}{N}\sum_{n=0}^{N-1}\psi_X(\langle \mathbf{z}, \mathbf{s}_{n,\,N} \rangle )\right\vert &\leq \frac{1}{N}\sum_{n=0}^{N-1}\vert\psi_X(\langle \mathbf{z}, \mathbf{t}_{n,\,N} \rangle ) - \psi_X(\langle \mathbf{z}, \mathbf{t}_{n,\,N} \rangle )\vert \leq \epsilon. \end{align*} This proves \eqref{eq:withb} and therefore also \eqref{eq:bert} for $C^{1}$ functions. \\ \\ To establish the infinite divisibility of $\xi$ we note that \eqref{eq:withb} shows that $\Psi_{\xi^{(b)}} = b\Psi_{\xi^{(1)}} = b\Psi_{\xi}$ and that $e^{-b\Psi_{\xi}}$ is therefore a positive definite function for every $b$ since it is the characteristic function of the random vector $\xi^{(b)}$. Positive definiteness follows from Bochner's Theorem (see for example Theorem 2.13 in \cite{levtype}). Also, we clearly have $\Psi_{\xi}(\mathbf{0}) = 0$ since $\Psi_{X}(0) = 0$. By Theorem 2.15 in \cite{levtype} these two points combined show that $\Psi_{\xi}$ is the characteristic exponent of an ID probability distribution, and hence $\xi$ is an ID random vector.\\ \\ Now one can extend the lemma to $L^{1}$ functions $\{f_k\}_{k = 1}^d$ by exploiting the density of $C^{1}([0,T])$ in $L^{1}([0,T])$. In particular, for each $k$ we can find a sequence of $C^{1}$ functions $\{f_{n,k}\}_{n\geq 1}$ which converges in $L^1$ to $f_k$. Then, \begin{align*} \vert u_k(t) - u_{n,k}(t) \vert = \left\vert \int_t^T f_k(t){\textnormal d} t - \int_t^T f_{n,k}(t){\textnormal d} t\right\vert \leq \int_0^T\left\vert f_k(t) - f_{n,k}(t)\right\vert{\textnormal d} t \end{align*} showing that $u_{n,k} \rightarrow u_k$ uniformly in $t$. This shows that for each $\mathbf{z}$ the functions $\{\Psi_{X}(\langle \mathbf{z},\mathbf{u}_n(\cdot)\rangle)\}_{n\geq 1}$, with $\mathbf{u}_n := (u_{n,1},\cdots,u_{n,d})^{\textnormal{\textbf{T}}}$, are uniformly bounded on $[0,T]$, so that the dominated convergence theorem applies and we have \begin{align}\label{eq:rhs} \lim_{n \rightarrow \infty}\exp\left(-\int_0^T\Psi_X\left(\langle \mathbf{z}, \mathbf{u}_n(t) \rangle\right){\textnormal d} t\right) = \exp\left(-\int_0^T\Psi_X\left(\langle \mathbf{z}, \mathbf{u}(t) \rangle\right){\textnormal d} t\right). \end{align} On the other hand, $X$ is a.s. bounded on $[0,T]$, so that \begin{align*} \lim_{n\rightarrow\infty}\vert \xi_k - \xi_{n,k} \vert = \lim_{n\rightarrow\infty}\left\vert \int_0^{T}X_tf_{k}(t){\textnormal d} t - \int_0^{T}X_t f_{n,k}(t){\textnormal d} t \right \vert \leq\left(\sup_{t\in[0,T]}\vert X_t \vert\right) \lim_{n\rightarrow\infty}\int_0^{T}\vert f_{k}(t) - f_{n,k}(t) \vert{\textnormal d} t = 0, \end{align*} a.s.. Therefore $\Xi_n := (\xi_{n,1},\cdots,\xi_{n,d})^{\textnormal{\textbf{T}}}$ converges a.s. and consequently also in distribution to $\xi$. Together with \eqref{eq:rhs}, this implies that for each $\mathbf{z}$ \begin{align}\label{eq:final} \lim_{n\rightarrow\infty}{\mathbb E}[e^{i\langle \mathbf{z},\Xi_n \rangle}] = {\mathbb E}[e^{i\langle \mathbf{z},\xi \rangle}] = \exp\left(-\int_0^T\Psi_X\left(\langle \mathbf{z}, \mathbf{u}(t) \rangle\right){\textnormal d} t\right). \end{align} Therefore, \eqref{eq:bert} is also proven for functions in $L^1$. Since each $\Xi_n$ has an ID distribution Lemma 3.1.6 in \cite{Messer} guarantees that $\xi$ is also an ID random vector. \end{proof} \end{appendices} \bibliographystyle{plain}
{ "timestamp": "2016-03-03T02:09:04", "yymm": "1603", "arxiv_id": "1603.00677", "language": "en", "url": "https://arxiv.org/abs/1603.00677", "abstract": "Karhunen-Loeve expansions (KLE) of stochastic processes are important tools in mathematics, the sciences, economics, and engineering. However, the KLE is primarily useful for those processes for which we can identify the necessary components, i.e., a set of basis functions, and the distribution of an associated set of stochastic coefficients. Our ability to derive these components explicitly is limited to a handful processes. In this paper we derive all the necessary elements to implement the KLE for a square-integrable Levy process. We show that the eigenfunctions are sine functions, identical to those found in the expansion of a Wiener process. Further, we show that stochastic coefficients have a jointly infinitely divisible distribution, and we derive the generating triple of the first d coefficients. We also show, that, in contrast to the case of the Wiener process, the coefficients are not independent unless the process has no jumps. Despite this, we develop a series representation of the coefficients which allows for simulation of any process with a strictly positive Levy density. We implement our theoretical results by simulating the KLE of a variance gamma process.", "subjects": "Probability (math.PR)", "title": "Karhunen-Loeve expansions of Levy processes", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9886682488058381, "lm_q2_score": 0.8104789040926008, "lm_q1q2_score": 0.8012947588033065 }
https://arxiv.org/abs/2008.06577
Cycles of a given length in tournaments
We study the asymptotic behavior of the maximum number of directed cycles of a given length in a tournament: let $c(\ell)$ be the limit of the ratio of the maximum number of cycles of length $\ell$ in an $n$-vertex tournament and the expected number of cycles of length $\ell$ in the random $n$-vertex tournament, when $n$ tends to infinity. It is well-known that $c(3)=1$ and $c(4)=4/3$. We show that $c(\ell)=1$ if and only if $\ell$ is not divisible by four, which settles a conjecture of Bartley and Day. If $\ell$ is divisible by four, we show that $1+2\cdot\left(2/\pi\right)^{\ell}\le c(\ell)\le 1+\left(2/\pi+o(1)\right)^{\ell}$ and determine the value $c(\ell)$ exactly for $\ell = 8$. We also give a full description of the asymptotic structure of tournaments with the maximum number of cycles of length $\ell$ when $\ell$ is not divisible by four or $\ell\in\{4,8\}$.
\section{Introduction} \label{sec:intro} In this paper, we address one of the most natural extremal problems concerning tournaments: \emph{What is the maximum number of cycles of a given length that can be contained in an $n$-vertex tournament?} The cases of cycles of length three and four are well-understood. An $n$-vertex tournament has at most $\frac{n(n^2-1)}{24}$ cycles of length three (cyclic triangles) if $n$ is odd, and at most $\frac{n(n^2-4)}{24}$ if $n$ is even; both bounds are the best possible. This result can be traced back to 1940 to the work of Kendall and Babington Smith~\cite{KenB40} and of Szele~\cite{Sze43}, also see~\cite{Moo15}, and it is well-known that the number of cycles of length three is determined by the degree sequence of a tournament~\cite{Goo59}. Beineke and Harary~\cite{BeiH65} and Colombo~\cite{Col64} proved in the 1960s the best possible bounds on the number of cycles of length four: $\frac{n(n^2-1)(n-3)}{48}$ when $n$ is odd and $\frac{n(n^2-4)(n-3)}{48}$ when $n$ is even. The asymptotics of the case of cycles of length five was determined only recently by Komarov and Mackey~\cite{KomM17} who showed that the number of cycles of length five is asymptotically maximized if and only if the tournament is almost regular, i.e., the behavior in this case is completely analogous to that of cycles of length three; exact results on cycles of length five for regular tournaments and tournaments of odd order were obtained by Savchenko in~\cite{Sav16,SavXX}. In this paper, we employ algebraic techniques to provide asymptotically optimal results on the maximum number of cycles for each cycle length not divisible by four, and determine the limit behavior in the case of cycle lengths divisible by four. To state our results precisely, we need to fix some notation. Let $C(n,\ell)$ be the maximum number of cycles of length $\ell$ in an $n$-vertex tournament. We compare this quantity to the expected number of cycles of length $\ell$ in a random $n$-vertex tournament, which is $R(n,\ell)=\frac{(\ell-1)!}{2^{\ell}}\binom{n}{\ell}$, and define \[c(\ell)=\lim_{n\to\infty}\frac{C(n,\ell)}{R(n,\ell)}.\] In particular, the results that we have mentioned earlier imply that $c(3)=c(5)=1$ and $c(4)=4/3$. Bartley and Day conjectured the following. \begin{conjecture}[{Bartley~\cite[Conjecture 104]{Bar18} and Day~\cite[Conjecture 40]{Day17}}] \label{conj:equiv} For $\ell\ge 3$, it holds that $c(\ell)=1$ if and only if $\ell$ is not divisible by four. \end{conjecture} We remark that the statement of Conjecture~\ref{conj:equiv} has been proven for \emph{regular} tournaments by Savchenko~\cite{Sav16}, also see~\cite{Sav17}, and also jointly by Bartley and Day~\cite{Day17,Bar18}, and for $\ell\le 8$ by Bartley~\cite[Theorem 109]{Bar18}. Since it is known that $c(\ell)>1$ for all $\ell$ divisible by four, the following theorem, which is implied by Theorems~\ref{thm:ck1} and~\ref{thm:ck2}, settles the conjecture. \begin{theorem} \label{thm:ck1+2} Let $\ell\ge 3$. If $\ell$ is not divisble by four, then $c(\ell)=1$. \end{theorem} If $\ell$ is divisible by four, we establish an asymptotically tight upper bound (Theorem~\ref{thm:ck4}): \[1+2\cdot\left(2/\pi\right)^{\ell}\le c(\ell)\le 1+\left(2/\pi+o(1)\right)^{\ell}\] Our asymptotic result on $c(\ell)$ for $\ell$ divisible by four provides a strong evidence for the following conjecture on the value of $c(\ell)$ for such $\ell$ (we remark that Conjecture~\ref{conj:div4} is stated in~\cite{Day17} by giving an extremal construction, which we mention at the end of Section~\ref{sec:prelim}); the conjecture is an extension of an earlier problem posed by Savchenko~\cite{Sav16} for regular tournaments. \begin{conjecture}[{Bartley~\cite[Conjecture 106]{Bar18} and Day~\cite[Conjecture 45]{Day17}}] \label{conj:div4} If $\ell$ is divisible by four, then \[c(\ell)=1+2\cdot\sum_{i=1}^{\infty}\left(\frac{2}{(2i-1)\pi}\right)^{\ell}.\] \end{conjecture} Our asymptotic result agrees with the conjecture on the dominant term of the sum. We also show that $c(8)=332/315$ (Theorem~\ref{thm:c8}) and classify the extremal constructions (Theorem~\ref{thm:c48}); note that the value of $c(8)$ is the one given in Conjecture~\ref{conj:div4}. In addition to the results on the value of $c(\ell)$, we have also been able to determine the asymptotic structure of extremal tournaments when $\ell$ is not divisible by four. If $\ell$ is odd, then tournaments achieving the maximum number of cycles of length $\ell$ are exactly those that are almost regular (Theorem~\ref{thm:ck1}), and if $\ell$ is even but not divisible by four, then a tournament achieves the maximum number of cycles of length $\ell$ if and only if it is quasirandom (Theorem~\ref{thm:ck2}). In particular, maximizing the density of cycles of length $4k+2$ is a \emph{quasirandom-forcing} property, i.e., a property that a tournament has if and only if it is quasirandom. We remark that in the induced setting, in addition to the density of transitive tournaments with four or more vertices, which is known to be quasirandom-forcing, see~\cite{CorR17} and \cite[Exercise 10.44]{Lov93}, there is only one additional tournament such that its density is quasirandom-forcing~\cite{BucLSS21,CorPS19,HanKKMPSV19}, which is the unique $5$-vertex strongly connected tournament with diameter four. \section{Preliminaries} \label{sec:prelim} In this section, we fix the notation used throughout the paper and present analytic and algebraic tools needed for our arguments. The set of the first $n$ positive integers is denoted by $[n]$. A \emph{tournament} is an orientation of a complete graph, and the \emph{random tournament} is an orientation of a complete graph where each edge is directed with probability $1/2$ in each of the two possible directions independently of the other edges. Let $C(T,\ell)$ be the ratio of the number of cycles of length $\ell$ in an $n$-vertex tournament $T$ and the expected number of cycles of length $\ell$ in the random $n$-vertex tournament. In particular, the maximum value of $C(T,\ell)$, where the maximum is taken over all $n$-vertex tournaments, is $C(n,\ell)/R(n,\ell)$. The \emph{adjacency matrix} $A$ of a tournament $T$ is the zero-one matrix with rows and columns indexed by vertices of $T$ such that $A_{ij}=1$ iff $T$ contains an edge from the $i$-th vertex to the $j$-th vertex. The \emph{tournament matrix} of an $n$-vertex tournament $T$ is the matrix obtained from the adjacency matrix of $T$ by setting its diagonal entries to be equal to $1/2$ and then dividing each entry of the matrix by $n$. We say that a real square matrix $A$ of order $n$ is \emph{skew-symmetric} if $A=-A^T$, i.e., $A_{ij}=-A_{ji}$ for all $i,j\in [n]$, and $A$ is \emph{complementary} if $A$ is non-negative and $A_{ij}+A_{ji}=1/n$ for all $i,j\in [n]$. In particular, the tournament matrix of a tournament is complementary. Finally, if $A$ is an $n\times n$ matrix, then its \emph{Frobenius norm}, which is denoted by $\|A\|_F$, is \[\|A\|_F=\sqrt{\sum_{i,j\in [n]}A_{ij}^2}.\] We recall that $\|Av\|\le\|A\|_F\cdot\|v\|$ for every vector $v\in\RR^n$. Since the trace of the $\ell$-th power of the adjacency matrix of $T$ is the number of closed walks of length $\ell$, we obtain the following; note that we state the next proposition for tournament matrices rather than adjacency matrices. \begin{proposition} \label{prop:eigen} Let $A$ be the tournament matrix of an $n$-vertex tournament $T$, $\lambda_1,\ldots,\lambda_n$ be its eigenvalues, and $\ell\ge 3$ an integer. It holds that \[C(T,\ell)=\frac{2^{\ell}}{n^\ell}\cdot\sum_{i=1}^n \lambda_i^{\ell}+O(n^{-1}).\] \end{proposition} We next recall some basic properties of tournament matrices, and more generally complementary matrices, used in~\cite{ChaGKN19}; we remark that similar results were also used earlier by Brauer and Gentry~\cite{BraG68}. \begin{proposition} \label{prop:matrix} Let $A$ be a complementary matrix. There is a positive real number $\rho$ such that $\rho$ is a real eigenvalue of $A$, and the absolute value of each eigenvalue of $A$ is at most $\rho$. In addition, each eigenvalue of $A$ has non-negative real part, and the sum of the eigenvalues is equal to $1/2$. \end{proposition} \subsection{Tournament limits} \label{subsec:limit} We now define tournament limits, which are analogous to graph limits described in detail in the monograph by Lov\'asz~\cite{Lov12} and which were used earlier in~\cite{ChaGKN19,Tho18,ZhaZ20}, also see~\cite{DiaJ08,LovS10i} for related concepts. Most of the results translate readily from the setting of graphs to that of tournaments. However, the results on the convergence of spectra seem to be an exception as we point out further. A \emph{tournamenton} is a measurable function $W:[0,1]^2\to [0,1]$ such that $W(x,y)+W(y,x)=1$ for all $(x,y)\in[0,1]^2$. The \emph{density} of a tournament $T$ in a tournament $T_0$, which is denoted by $d(T,T_0)$, is the probability that a uniformly randomly chosen subset of $|T|$ vertices of $T_0$ induces a tournament isomorphic to $T$; if $|T|>|T_0|$, we set $d(T,T_0)=0$. The \emph{density} of an $n$-vertex tournament $T$ in a tournamenton $W$ is defined as \[d(T,W)=\frac{|T|!}{|\Aut(T)|!}\int_{x_1,\ldots,x_n\in [0,1]}\prod_{i\to j}W(x_i,x_j)\;\mathrm{d} x_1\cdots x_n,\] where the product is taken over all $i$ and $j$ such that the $i$-th vertex is joined by an edge to the $j$-th vertex. Two tournamentons $W$ and $W'$ are \emph{weakly isomorphic} if $d(T,W)=d(T,W')$ for every tournament $T$. We say that a sequence $(T_n)_{n\in\NN}$ of tournaments is \emph{convergent} if $|T_n|$ tends to infinity and the sequence $d(T,T_n)$ converges for every tournament $T$. A tournamenton $W$ is a \emph{limit} of a convergent sequence $(T_n)_{n\in\NN}$ of tournaments if $d(T,W)$ is equal to the limit of $d(T,T_n)$ for every tournament $T$. For example, the tournamenton equal to $1/2$ everywhere is the limit of the sequence of random $n$-vertex tournaments with probability one. A tournamenton $W$ is called \emph{regular} if \[\int_{[0,1]}W(x,y)\;\mathrm{d} y=1/2\] for almost every $x\in [0,1]$; such tournamentons are limits of tournaments where the in-degrees and out-degrees of most of the vertices are asymptotically equal to half of the total number of vertices. An analogous line of arguments as in the graph case yields that every convergent sequence of tournaments has a limit and every tournamenton is a limit of a convergent sequence of tournaments. In particular, a tournamenton $W$ is a limit of $W$-random tournaments that we next define. An $n$-vertex \emph{$W$-random tournament} is obtained as follows: sample $n$ points $x_1,\ldots,x_n$ uniformly and independently in $[0,1]$ and orient the edge between the $i$-th and $j$-th vertex from the $i$-th vertex to the $j$-th vertex with probability $W(x_i,x_j)$. Note that the expected density of a tournament $T$ in a $W$-random tournament is equal to $d(T,W)$. We next introduce the quantity $C(W,\ell)$, which is the limit analogue of $C(T,\ell)$ defined earlier: \[C(W,\ell)=2^{\ell}\int_{x_1,\ldots,x_\ell\in [0,1]}W(x_1,x_2)W(x_2,x_3)\cdots W(x_{\ell-1},x_{\ell})W(x_{\ell},x_1)\;\mathrm{d} x_1\cdots x_{\ell}.\] It follows that $c(\ell)$ is the maximum of $C(W,\ell)$ where the maximum is taken over all tournamentons $W$ (and it can be shown that the maximum is indeed attained). Note that the expected value of $C(T,\ell)$ for an $n$-vertex $W$-random tournament $T$, $n\ge\ell$, is equal to $C(W,\ell)$. If $W$ and $W'$ are two tournamentons, then the \emph{cut distance} between $W$ and $W'$, which is denoted by $\cut{W}{W'}$, is defined as \[\cut{W}{W'}=\sup_{X,Y\subseteq [0,1]}\left|\int_{X\times Y}W(x,y)-W'(x,y)\;\mathrm{d} x\;\mathrm{d} y\right|,\] where the supremum is taken over all measurable subsets $X$ and $Y$ of $[0,1]$. As in the graph case, it can be shown that \begin{equation} \left|d(T,W)-d(T,W')\right|\le |T|^2\cut{W}{W'}\label{eq:cut} \end{equation} for every tournament $T$ and all tournamentons $W$ and $W'$. Every complementary matrix $A$ of order $k$ can be associated with a tournamenton $W$ as follows: the interval $[0,1]$ is split into $k$ disjoint measurable sets $U_1,\ldots,U_k$ each of measure $1/k$ and $W(x,y)=k\cdot A_{ij}$ if $x\in U_i$ and $y\in U_j$; tournamentons that can be obtained in this way are called \emph{step tournamentons}. A \emph{step approximation} of a tournamenton $W$ is a complementary matrix $A$ such that there exists a partition of the interval $[0,1]$ to $k$ disjoint measurable sets $U_1,\ldots,U_k$ each of measure $1/k$ that \[A_{ij}=k\int_{U_i\times U_j}W(x,y)\;\mathrm{d} x\;\mathrm{d} y\] for every $i,j\in [k]$. The step tournamenton associated with a step approximation $A$ and the sets $U_1,\ldots,U_k$ used to define $A$ is denoted by $W[A]$. We say that a sequence of step approximations $(A_n)_{n\in\NN}$ of a tournamenton $W$ is \emph{convergent} if $\cut{W}{W[A_n]}$ converges to zero. The Regularity Lemma yields that for every $\varepsilon>0$ and every tournamenton $W$, there exists a step approximation $A$ of the tournamenton $W$ such that $\cut{W}{W[A]}\le\varepsilon$. In particular, every tournamenton has a convergent sequence of step approximations and we obtain using \eqref{eq:cut} the following. \begin{proposition} \label{prop:cut} If $W$ is a tournamenton and $(A_n)_{n\in\NN}$ is a convergent sequence of its step approximations, then \[d(T,W)=\lim_{n\to\infty}d(T,W[A_n])\] for every tournament $T$. In particular, it holds that \[C(W,\ell)=2^\ell\lim_{n\to\infty}\Trace A_n^\ell\] for every $\ell\ge 3$. \end{proposition} We finish this subsection by describing a tournamenton that is believed to be extremal for Conjecture~\ref{conj:div4} for every $\ell$ divisible by four. For $x,y\in[0,1]$, define $W_C(x,x)=1/2$, $W_C(x,y)=1$ if $y\in (x-1,x-1/2)\cup (x,x+1/2]$, and $W_C(x,y)=0$ otherwise. The tournamenton $W_C$, which we refer to as the carousel tournamenton, is depicted in Figure~\ref{fig:carouselle}, and it is the limit of the following tournaments described in~\cite{Day17} in relation to Conjecture~\ref{conj:div4}: take vertices $0,\ldots,2n$ and join a vertex $i$ to the vertices $i+1,\ldots,i+n$ (computations modulo $2n+1$). These tournaments are called \emph{carousel tournaments}. The value of $C(W_C,\ell)$ for every positive integer $\ell$ divisible by four is equal to the value of $c(\ell)$ given in Conjecture~\ref{conj:div4}. \begin{figure} \begin{center} \epsfysize 3cm \epsfbox{tourn-ck-7.mps} \hskip 2cm \epsfbox{tourn-ck-1.mps} \end{center} \caption{The tournament matrix of the $9$-vertex carousel tournament and the carousel tournamenton. The origin of the coordinate system is in the top left corner, the $x$-axis is vertical and the $y$-axis is horizontal. The black color represents the value $1$ and the white color the value $0$.} \label{fig:carouselle} \end{figure} \subsection{Spectral properties of tournaments and their limits} \label{subsec:spectrum} A tournamenton $W$ can be viewed as a linear operator from $L_2[0,1]$ to $L_2[0,1]$ defined as \[(Wf)(x)=\int_{[0,1]}W(x,y)f(y)\;\mathrm{d} y.\] Since this operator is compact (as all Hilbert-Schmidt integral operators are), its spectrum $\sigma(W)$ is either finite or countably infinite, the only accumulation point of $\sigma(W)$ can be zero, and every non-zero element of $\sigma(W)$ is an eigenvalue of $W$. Moreover, for every non-zero $\lambda\in\sigma(W)$, there exists $k_{\lambda}\in\NN$ such that the kernels of $(W-\lambda)^{k_{\lambda}}$ and $(W-\lambda)^{k_{\lambda}+1}$ are the same and their dimension is finite. For example, the spectrum of the carousel tournamenton $W_C$ defined at the end of Subsection~\ref{subsec:limit} consists of $1/2$, $\pm\unit/((2k-1)\pi)$ for $k\in\NN$, and $0$. It is plausible that if $(T_n)_{n\in\NN}$ is a convergent sequence of tournaments and $W$ is a limit tournamenton, then the normalized spectra of the tournament matrices of $T_n$ converge to the spectrum of $W$ in the sense used in the graph setting in~\cite[Section 6]{BorCLSV12}, also see~\cite[Chapter 11]{Lov12}. However, the equality between the density of cycles of length $\ell$ and the trace of $W^\ell$ for $\ell\ge 3$ (note that $W^\ell$ is a trace-class operator for $\ell\ge 2$), which forms the core of the argument in~\cite[Section 6]{BorCLSV12} and is straightforward in the case of graphons, is not obvious in the case of tournamentons. While we establish a close analogy of this equality as \eqref{eq:limitpower} in Proposition~\ref{prop:limitset}, it does not seem to be strong enough to give results completely analogous to those on the convergence of spectra of graph limits. To proceed with our exposition, we need to define a notion of convergence for multisets of complex numbers. If $z$ is a complex number, we write $N_{\varepsilon}(z)$ for the set of all complex numbers $z'$ with $|z-z'|\le\varepsilon$. We say that a sequence $(X_i)_{i\in\NN}$ of multisets of complex numbers \emph{converges as multisets} to a multiset $X$ if the following holds: \begin{itemize} \item if $z$ is an element of $X$ with a finite multiplicity, then there exists $\varepsilon_0>0$ such that for every $\varepsilon\in(0,\varepsilon_0)$, there exists $i_0$ such that both $|X_i\cap N_{\varepsilon}(z)|$ and $|X\cap N_{\varepsilon}(z)|$ are finite and equal for every $i\ge i_0$, and \item if $z$ is an element of $X$ with infinite multiplicity, then for every $\varepsilon>0$ and every $k\in\NN$, there exists $i_0$ such that $|X_i\cap N_{\varepsilon}(z)|\ge k$ for every $i\ge i_0$, \end{itemize} We say that a sequence $(A_n)_{n\in\NN}$ of step approximations of a tournamenton $W$ is \emph{strongly convergent} if it is convergent and the spectra of $A_n$ converge as multisets. Later, we will show that every convergent sequence of step approximations is also strongly convergent but we treat the two notions as distinct until we establish their equivalence. We summarize properties of the limit multiset of a strongly convergent sequence of step approximations in the next proposition. \begin{proposition} \label{prop:limitset} Let $W$ be a tournamenton and let $(A_n)_{n\in\NN}$ be a strongly convergent sequence of step approximations of $W$. The limit multiset $X$ of the spectra of $A_n$ satisfies the following. \begin{itemize} \item The set $X$ contains zero and its multiplicity is infinite. \item Every non-zero element of $X$ has finite multiplicity. \item Every element of $X$ has a non-negative real part. \item The real parts of the elements of $X$ sum to at most $1/2$ (taking their multiplicities into account). \item If $x$ is an element of $X$ that is not real, then $X$ contains the complex conjugate of $x$ and the multiplicities of $x$ and its complex conjugate are the same. \item If $X$ contains a non-zero element, then it contains a positive real $\rho$ such that the absolute value of all elements of $X$ is at most $\rho$. \item The sum of the $\ell$-th powers of the elements of $X$ is absolutely convergent for every $\ell\ge 2$. \item It holds that \begin{equation} C(W,\ell)=2^{\ell}\cdot\sum_{x\in X}x^\ell\label{eq:limitpower} \end{equation} for every $\ell\ge 3$. \end{itemize} \end{proposition} \begin{proof} Let $X_n$ be the spectrum of $A_n$. By Proposition~\ref{prop:matrix}, all the elements of $X_n$ have non-negative real parts and the sum of their real parts is $1/2$. Hence, all the elements of $X$ have non-negative real parts and their real parts sum to at most $1/2$. Since the matrix $A_n$ is real, every non-real eigenvalue of $A_n$ comes in a pair with a complex conjugate eigenvalue of the same multiplicity. Hence, the pairs of complex conjugate non-real elements of $X$ must have the same multiplicity. Let $\rho_n$ be the largest real eigenvalue of $A_n$. Note that the absolute value of all the elements of $X_n$ is at most $\rho_n$ by Proposition~\ref{prop:matrix}. The sequence $\rho_n$ converges (because the sets $X_n$ converge as multisets) and its limit $\rho$ belongs to $X$. If $\rho=0$, then $X$ has no non-zero elements. If $\rho>0$, then the absolute value of all elements of $X$ is at most $\rho$. We next show that the sum of the $\ell$-th powers of the elements of $X$ is absolutely convergent for every $\ell\ge 2$. Let $X^+$ and $X^-$ be the multiset of the non-zero elements $x$ of $X$ such that $\Real x\ge|x|/2$ and $\Real x\le|x|/2$, respectively. Note that $\Real x^2 \le - |x^2|/2 < 0$ for every $x\in X^-$. First observe that \begin{equation} \sum_{x\in X^+}|x|\le 2\sum_{x\in X^+}\Real x\le 1.\label{eq:X+} \end{equation} Since $\Real x^2\le (\Real x)^2$ for every $x\in X_n$ and the sum of real parts of the elements of $X_n$ is $1/2$, it follows that \[\sum_{x\in X_n,\Real x^2>0}\Real x^2\le\frac{1}{4}.\] Since the trace of $A_n^2$ is non-negative, we obtain that \[-\sum_{x\in X_n,\Real x^2<0}\Real x^2\le\frac{1}{4}.\] In particular, it holds that \[ -\sum_{x\in X^-}\Real x^2\le\frac{1}{4}. \] As $-\Real x^2\ge |x^2|/2$ for every $x\in X^-$, we obtain that \begin{equation} \sum_{x\in X^-}|x|^2\le -2\sum_{x\in X^-}\Real x^2\le\frac{1}{2}.\label{eq:X-} \end{equation} The inequalities \eqref{eq:X+} and \eqref{eq:X-} yield that all elements of the multiset $X$ except for $0$ have finite multiplicity and the sum of the $\ell$-th powers of the elements of $X$ is absolutely convergent for every $\ell\ge 2$. Since the sizes of the multisets $X_n$ tend to infinity and $|x|\le 1/2$ for all their elements, the multiset $X$ must contain an element with infinite multiplicity; since such an element can be only $0$, the multiset $X$ contains $0$ with infinite multiplicity. It remains to establish \eqref{eq:limitpower}. Fix $\varepsilon\in (0,1)$. Similarly to the previous paragraph, define $X_n^+$ to be the multiset of the elements $x$ of $X_n$ such that $\Real x\ge|x|/2$ and $X_n^-$ to be the multiset of the elements $x$ of $X_n$ such that $\Real x\le|x|/2$. Along the lines leading to \eqref{eq:X+} and \eqref{eq:X-}, we obtain that \[\sum_{x\in X_n^+}|x|^2\le\frac{1}{2}\sum_{x\in X_n^+}|x|\le\frac{1}{2}\qquad\mbox{and}\qquad\sum_{x\in X_n^-}|x|^2\le\frac{1}{2}.\] Hence, we obtain for $\ell \geq 3$ that \begin{equation} \left|\Trace A_n^\ell-\sum_{x\in X_n,|x|>\varepsilon}x^\ell\right| \le\sum_{x\in X_n,|x|\le\varepsilon}|x|^\ell \le\varepsilon\sum_{x\in X_n,|x|\le\varepsilon}|x|^2\le\varepsilon.\label{eq:Xneps} \end{equation} Similarly, we obtain that \begin{equation} \left|\sum_{x\in X,|x|\le\varepsilon}x^\ell\right| \le\sum_{x\in X,|x|\le\varepsilon}|x|^\ell \le\varepsilon\sum_{x\in X,|x|\le\varepsilon}|x|^2\le\varepsilon.\label{eq:Xeps} \end{equation} Consequently, it follows from \eqref{eq:Xneps} and \eqref{eq:Xeps} that \begin{equation} \lim_{n\to\infty}\left|\Trace A_n^\ell-\sum_{x\in X_n}x^\ell\right|\le 2\varepsilon.\label{eq:Xnlim} \end{equation} Since the estimate \eqref{eq:Xnlim} holds for every $\varepsilon\in (0,1)$, It follows that \[\lim_{n\to\infty}\Trace A_n^\ell=\sum_{x\in X_n}x^\ell.\] The identity \eqref{eq:limitpower} now follows from Proposition~\ref{prop:cut}. \end{proof} By compactness, every convergent sequence of step approximations of $W$ has a strongly convergent subsequence. The limit multisets of two such strongly convergent subsequences must be the same since no two different multisets can be absolutely convergent and satisfy \eqref{eq:limitpower} for every $\ell\ge 3$. Hence, every convergent sequence of step approximations of $W$ is also strongly convergent, which we state as a corollary. \begin{corollary} \label{cor:strongconv} Every convergent sequence $(A_n)_{n\in\NN}$ of step approximations of a tournamenton $W$ is strongly convergent. \end{corollary} Corollary~\ref{cor:strongconv} implies that the limit multiset of every convergent sequence of step approximations of a tournamenton $W$ is the same; we will write $\wsigma(W)$ for this limit multiset after removing $0$. Proposition~\ref{prop:limitset} now yields that \begin{equation} C(W,\ell)=2^{\ell}\cdot\sum_{\lambda\in\wsigma(W)}\lambda^\ell\label{eq:lambda} \end{equation} for every $\ell\ge 3$. We conclude with two propositions relating structural properties of a tournamenton $W$ and $\wsigma(W)$. \begin{proposition} \label{prop:reg} A tournamenton $W$ is regular if and only if $1/2\in\wsigma(W)$. \end{proposition} \begin{proof} Fix a tournamenton $W$. Let $(A_n)_{n\in\NN}$ be a convergent sequence of step approximations of $W$. Let $k_n$ be the order of $A_n$, $\rho_n$ the largest real eigenvalue of $A_n$, and $v_n$ the corresponding eigenvector with norm one. Further let $\JJ_n$ be the $k_n\times k_n$ matrix with all entries equal to $k_n^{-1}$ and $j_n$ the $k_n$-dimensional vector with all entries equal to $k_n^{-1/2}$. Note that $j_n=\JJ_n j_n$ and $1$ is the only non-zero eigenvalue of $\JJ_n$. Finally, let $j_0$ be the function $[0,1]\to [0,1]$ such that $j_0(x)=1$ for all $x\in [0,1]$. If $W$ is a regular tournamenton, then $A_nj_n=j_n/2$. It follows that $1/2$ is an eigenvalue of $A_n$ for every $n\in\NN$ and so $1/2\in\wsigma(W)$. We next assume that $1/2\in\wsigma(W)$ and show that the tournamenton $W$ is regular. Since the matrix $A_n$ cannot have a real eigenvalue larger than $1/2$ by Proposition~\ref{prop:matrix}, it follows that the values of $\rho_n$ converge to $1/2$. Observe that \[v_n^T\JJ_nv_n=v_n^T(A_n^T+A_n)v_n=v_n^TA_n^Tv_n+v_n^TA_nv_n=2\rho_n.\] It follows that $<v_n|j_n>^2=2\rho_n$, which implies that \[\|v_n-j_n\|=2-2<v_n|j_n>=2-2\sqrt{2\rho_n}.\] Since $\|A_n\|_F\le 1$, we obtain that \begin{align*} \|j_n-2A_nj_n\| &\le\|j_n-v_n\|+\|v_n-2A_nv_n\|+2\|A_nv_n-A_nj_n\|\\ &\le 3\|j_n-v_n\|+1-2\rho_n = 7-2\rho_n-6\sqrt{2\rho_n}. \end{align*} It follows that \[\lim_{n\to\infty}\|j_n-2A_nj_n\|=0.\] Since the step tournamentons $W[A_n]$ converge to the tournamenton $W$ in the cut distance, we obtain that $\|j_0-2Wj_0\|_2=0$ where $Wj_0$ is the function resulting from applying the linear operator given by $W$ to the function $j_0$. It follows that the tournamenton $W$ is regular. \end{proof} \begin{proposition} \label{prop:qrand} A tournamenton $W$ is equal to $1/2$ almost everywhere if and only if $\wsigma(W)=\{1/2\}$. \end{proposition} \begin{proof} Fix a tournamenton $W$. Let $(A_n)_{n\in\NN}$ be a convergent sequence of step approximations of $W$, and let $k_n$ be the order of $A_n$. Further let $\JJ_n$ be the $k_n\times k_n$ matrix with all entries equal to $(2k_n)^{-1}$ and $j_n$ the $k_n$-dimensional vector with all entries equal to $k_n^{-1/2}$. Note that the definition of the matrix $\JJ_n$ differs from that in the proof of Proposition~\ref{prop:reg}. In particular, it holds that $\JJ_nj_n=j_n/2$. If the tournamenton $W$ is equal to $1/2$ almost everywhere, then $A_n=\JJ_n$ for every $n\in\NN$ and the only non-zero eigenvalue of $A_n$ is $1/2$. It follows that $\wsigma(W)=\{1/2\}$. We next assume that $\wsigma(W)=\{1/2\}$ and show that $W$ is equal to $1/2$ almost everywhere. Since $1/2\in\wsigma(W)$, the tournamenton $W$ is regular and it follows that $A_nj_n=j_n/2$. Define $B_n=A_n-\JJ_n$ and observe that $B_n$ is a skew-symmetric matrix and that $B_nj_n$ is the zero vector, i.e., the vector $j_n$ belongs to the kernel of~$B_n$. Hence, the non-zero eigenvalues of $A_n$ are $1/2$ and the non-zero eigenvalues of~$B_n$, which are square roots of the (real and negative) eigenvalues of the symmetric matrix $B_n^2$; in particular, they all are purely imaginary. Suppose that $W$ is not equal to $1/2$ almost everywhere, in particular, the cut distance of $W$ and the tournamenton equal to $1/2$ everywhere is positive. Since the tournamentons $W[A_n]$ converge to $W$ in the cut distance, there exists a sequence of vectors $v_n\in\RR^{k_n}$ with $\|v_n\|=1$ and a real $\delta>0$ such that \[\lim_{n\to\infty}\|B_nv_n\|\ge\delta.\] In particular, \[\lim_{n\to\infty} |v_n^TB_n^2v_n|\ge\delta^2,\] which implies that the smallest eigenvalue $B_n^2$ is less than $-\delta^2$. Hence, every $B_n$ has a purely imaginary eigenvalue with absolute value at least $\delta$. However, this is impossible since $\wsigma(W)=\{1/2\}$. We conclude that $W$ is equal to $1/2$ almost everywhere. \end{proof} \section{Cycles of length not divisible by four} \label{sec:ck12} In this section, we compute $c(\ell)$ when $\ell$ is not divisible by four and we characterize tournamentons that are extremal. The proofs of both Theorem~\ref{thm:ck1} and~\ref{thm:ck2} are based on the analysis of spectra of linear operators associated with tournamentons, however, the arguments apply the same in the setting of tournament matrices. \begin{theorem} \label{thm:ck1} If $\ell\ge 3$ is odd, then $C(W,\ell)\le 1$ for every tournamenton $W$, and equality holds if and only if $W$ is regular. In particular, $c(\ell)=1$. \end{theorem} \begin{proof} Fix $\ell\ge 3$ and a tournamenton $W$. Let $\rho$ be the largest positive real contained in $\wsigma(W)$. We start with establishing the following. If $z$ is a complex number with $|z|\le\rho$ and $\Real z\ge 0$, then \begin{equation} \Real z^{\ell}\le \ell \rho^{\ell-1} \Real z, \label{eq:ck1} \end{equation} and equality holds if and only if $\Real z=0$. Consider such $z$ and let $\alpha$ be such that $\Real z=|z|\cdot\cos\alpha$ and $\Imag z=|z|\cdot\sin\alpha$. If $\Real z=0$, the estimate \eqref{eq:ck1} holds with equality. Hence, we can assume that $\alpha\in [0,\pi/2)$ (considering the complex conjugate of $z$ if needed). If $\ell$ is one modulo four, we set $\beta=\pi/2-\alpha$ and obtain the following: \[ \Real z^{\ell}=|z|^{\ell}\cos\ell\alpha =|z|^{\ell}\sin\ell\beta <\ell|z|^{\ell}\sin\beta =\ell|z|^{\ell}\cos\alpha =\ell|z|^{\ell-1}\Real z. \] If $\ell$ is three modulo four, we set $\beta=-\pi/2+\alpha$ and obtain the following: \[ \Real z^{\ell}=|z|^{\ell}\cos\ell\alpha =|z|^{\ell}\sin\ell\beta <-\ell|z|^{\ell}\sin\beta =\ell|z|^{\ell}\cos\alpha =\ell|z|^{\ell-1}\Real z. \] We now obtain the estimate \eqref{eq:ck1} using $|z|\le\rho$. We next bound the sum of the $\ell$-th powers of the elements of $\wsigma(W)$ by applying~\eqref{eq:ck1} to every element of $\wsigma(W)$ except for $\rho$. Note that we treat $\wsigma(W)$ as a multiset, i.e., if the multiplicity of $\rho$ in $\wsigma(W)$ is larger than one, then $\wsigma(W)\setminus\{\rho\}$ contains~$\rho$. \begin{equation} \sum_{\lambda\in\wsigma(W)}\lambda^{\ell} =\sum_{\lambda\in\wsigma(W)}\Real\lambda^{\ell} \le\rho^\ell+\;\; \sum_{\mathclap{\lambda\in\wsigma(W)\setminus\{\rho\}}}\;\; \ell \rho^{\ell-1}\Real\lambda \le\left(\rho+\;\; \sum_{\mathclap{\lambda\in\wsigma(W)\setminus\{\rho\}}}\;\; \Real\lambda\right)^\ell \label{eq:ck1s} \end{equation} Since the sum of the real parts of the elements of $\wsigma(W)$ is at most $1/2$ by Proposition~\ref{prop:limitset}, we obtain that \[\sum_{\lambda\in\wsigma(W)}\lambda^{\ell}\le\frac{1}{2^{\ell}}.\] The identity \eqref{eq:lambda} now yields that $C(W,\ell)\le 1$ and since the choice of $W$ was arbitrary, it follows that $c(\ell)\le 1$. Moreover, if $C(W,\ell)=1$, the sum of real parts of the elements of $\wsigma(W)$ is $1/2$ and the estimate \eqref{eq:ck1} holds with equality for every element of $\wsigma(W)\setminus\{\rho\}$. In particular, the real part of every element of $\wsigma(W)\setminus\{\rho\}$ is zero. Hence, if $C(W,\ell)=1$, then $\rho=1/2$ and the tournamenton $W$ is regular by Proposition~\ref{prop:reg}. \end{proof} We next focus on the case when $\ell$ is even but not divisible by four. \begin{theorem} \label{thm:ck2} If $\ell\ge 6$ is even but not divisible by four, then $C(W,\ell)\le 1$ for every tournamenton $W$ and equality holds if and only if $W$ is equal to $1/2$ almost everywhere. In particular, $c(\ell)=1$. \end{theorem} \begin{proof} Fix $\ell\ge 6$ and a tournamenton $W$, and let $\rho$ be the largest positive real contained in $\wsigma(W)$. We start with establishing the following. If $z$ is a complex number with $|z|\le\rho$ and $\Real z\ge 0$, then \begin{equation} \Real z^{\ell}\le \ell \rho^{\ell-1} \Real z, \label{eq:ck2} \end{equation} and equality holds only if $z=0$. Consider such $z\not=0$ and let $\alpha$ be such that $\Real z=|z|\cdot\cos\alpha$ and $\Imag z=|z|\cdot\sin\alpha$. By symmetry, we can assume that $\alpha\in [0,\pi/2]$. Let $\beta=\pi/2-\alpha$. We first show that \begin{equation} -\cos\ell\beta<\ell\sin\beta.\label{eq:ck2sin} \end{equation} If $0\le\beta<\frac{\pi}{2\ell}$, then $\cos\ell\beta>0$, and the inequality in \eqref{eq:ck2sin} holds since its left side is negative while the right side is non-negative. If $\beta\ge\frac{\pi}{2\ell}$, then $\sin\beta>\frac{1}{\ell}$, and the inequality in \eqref{eq:ck2sin} holds since its right side is larger than one. We now apply \eqref{eq:ck2sin} as follows. \[ \Real z^{\ell}=|z|^{\ell}\cos\ell\alpha =-|z|^{\ell}\cos\ell\beta <\ell|z|^{\ell}\sin\beta =\ell|z|^{\ell}\cos\alpha =\ell|z|^{\ell-1}\Real z. \] We now obtain the estimate \eqref{eq:ck2} using $|z|\le\rho$. Similarly to the proof of Theorem~\ref{thm:ck2}, we bound the sum of the $\ell$-th powers of the elements of $\wsigma(W)$ using \eqref{eq:ck2} as follows. \[ \sum_{\lambda\in\wsigma(W)}\lambda^{\ell} =\sum_{\lambda\in\wsigma(W)}\Real\lambda^{\ell} \le\rho^\ell+\sum_{\lambda\in\wsigma(W)\setminus\{\rho\}}\ell \rho^{\ell-1}\Real\lambda \le\left(\rho+\sum_{\lambda\in\wsigma(W)\setminus\{\rho\}}\Real\lambda\right)^\ell \] Note that the inequality is strict unless $\rho$ is the only non-zero element of $\wsigma(W)$. Since the sum of the real parts of the elements of $\wsigma(W)$ is at most $1/2$, we obtain using \eqref{eq:lambda} that \[C(W,\ell)=2^\ell\sum_{\lambda\in\wsigma(W)}\lambda^{\ell}\le 1.\] Since the choice of $W$ was arbitrary, it follows that $c(\ell)\le 1$. Moreover, if $C(W,\ell)=1$, then $\rho$ is the only element of $\wsigma(W)$ and $\rho=1/2$. Consequently, if $C(W,\ell)=1$, then $W$ is equal to $1/2$ almost everywhere by Proposition~\ref{prop:qrand}. \end{proof} \section{Cycles of length divisible by four} \label{sec:ck4} The proof of the main result of this section requires bounding the spectral radius of skew-symmetric matrices with all entries between $-1$ and $+1$. To get the tight bound on the spectral radius of such matrices, we need the following auxiliary lemma. \begin{lemma} \label{lm:sumsq} Let $s_1,\ldots,s_k$ be positive reals. Further let $x_1,\ldots,x_k$ be non-negative reals such that $x_1\ge x_2\ge \cdots \ge x_k\ge 0$ and \[ \begin{array}{ccccccccccccccccc} x_1 & & & & & & & \le & s_1 & + & s_2 & + & s_3 & + & \cdots & + & s_k\\ x_1 & + & x_2 & & & & & \le & s_1 & + & 2s_2 & + & 2s_3 & + & \cdots & + & 2s_k\\ x_1 & + & x_2 & + & x_3 & & & \le & s_1 & + & 2s_2 & + & 3s_3 & + & \cdots & + & 3s_k\\ \vdots && \vdots && & \ddots && & \vdots && \vdots & & \vdots & & & & \vdots\\ x_1 & + & x_2 & + & \cdots & + & x_k & \le & s_1 & + & 2s_2 & + & 3s_3 & + & \cdots & + & ks_k, \end{array} \] i.e., the $m$-th inequality, $m\in [k]$ is \[\sum_{i=1}^m x_i \le \sum_{i=1}^m \min\{i,m\}\cdot s_i.\] It then holds that \[\sum_{i=1}^k x_i^2\le\sum_{i=1}^k\left(\sum_{j=i}^k s_j\right)^2\] and equality holds if and only if $x_i=s_i+s_{i+1}+\cdots+s_k$ for all $i\in [k]$. \end{lemma} \begin{proof} Let $S_i$, $i\in [k]$, be the right side of the $i$-th inequality listed in the statement of the lemma, and set $S_0=0$. Observe that \[S_1-S_0>S_2-S_1>\cdots>S_k-S_{k-1}=s_k.\] Since the set of reals $x_1\ge x_2\ge \cdots \ge x_k\ge 0$ that satisfy the $k$ inequalities $x_1+\cdots+x_i\le S_i$, $i\in [k]$, is compact, there exists a $k$-tuple $x_1,\ldots,x_k$ that maximizes the sum $x_1^2+\cdots+x_k^2$ subject to $x_1\ge x_2\ge \cdots \ge x_k\ge 0$ and the $k$ inequalities. Fix such a $k$-tuple. Observe that $x_k\ge s_k>0$. We first establish that $x_1>x_2>\cdots>x_k$. Suppose that there exists $i$ such that $x_i=x_{i+1}$. Choose the smallest such $i$, and let $i'$ be the largest index such that $x_{i'}=x_i$. Next suppose that $x_1+\cdots+x_j=S_j$ for some $j\in\{i,\ldots,i'-1\}$. Since $x_1+\cdots+x_{j-1}\le S_{j-1}$, it follows that $x_j\ge S_j-S_{j-1}$. We next obtain using that $x_j=x_{j+1}$ the following: \[x_1+\cdots+x_j+x_{j+1}=x_1+\cdots+x_j+x_j\ge S_j+(S_j-S_{j-1})>S_j+(S_{j+1}-S_j)=S_{j+1},\] which violates the $(j+1)$-th inequality listed in the statement of the lemma. Hence, it holds that $x_1+\cdots+x_j<S_j$ for every $j=i,\ldots,i'-1$. Choose $\varepsilon>0$ such that \begin{itemize} \item $x_1+\cdots+x_j+\varepsilon\le S_j$ for every $j=i,\ldots,i'-1$, \item if $i\ge 2$, then $x_i+\varepsilon\le x_{i-1}$, \item if $i'\le k-1$, then $x_{i'}-\varepsilon\ge x_{i'+1}$, and \item if $i'=k$, then $x_{i'}-\varepsilon\ge 0$. \end{itemize} Consider $x'_1,\ldots,x'_k$ such that \[x_j=\begin{cases} x_j+\varepsilon & \mbox{if $j=i$,} \\ x_j-\varepsilon & \mbox{if $j=i'$, and} \\ x_j & \mbox{otherwise.} \end{cases}\] Observe that $x'_1\ge x'_2\ge\cdots\ge x'_k\ge 0$ and $x'_1+\cdots+x'_j\le S_j$ for every $j\in [k]$. Since the sum of the squares of $x'_1,\ldots,x'_k$ is larger than the sum of the squares of $x_1,\ldots,x_k$, we obtain that the $k$-tuple $x_1,\ldots,x_k$ does not maximize the sum of the squares subject to $x_1\ge x_2\ge \cdots \ge x_k\ge 0$ and the $k$ inequalities listed in the statement of the lemma. This contradicts the choice of $x_1,\ldots,x_k$. Hence, we have established that $x_1>x_2>\cdots>x_k$. We next show that $x_1+\cdots+x_i=S_i$ for every $i\in [k]$. Suppose the opposite, that $x_1+\cdots+x_i<S_i$ for some $i$, and choose $\varepsilon>0$ such that \begin{itemize} \item $x_1+\cdots+x_i+\varepsilon\le S_i$, \item if $i\ge 2$, then $x_i+\varepsilon\le x_{i-1}$, \item if $i\le k-2$, then $x_{i+1}-\varepsilon\ge x_{i+2}$, and \item if $i=k-1$, then $x_{i+1}-\varepsilon\ge 0$. \end{itemize} Consider $x'_1,\ldots,x'_k$ such that \[x_j=\begin{cases} x_j+\varepsilon & \mbox{if $j=i$,} \\ x_j-\varepsilon & \mbox{if $j=i+1$, and} \\ x_j & \mbox{otherwise.} \end{cases}\] Observe that $x'_1\ge x'_2\ge\cdots\ge x'_k\ge 0$ and $x'_1+\cdots+x'_j\le S_j$ for every $j\in [k]$. Since the sum of the squares of $x'_1,\ldots,x'_k$ is larger than the sum of the squares of $x_1,\ldots,x_k$, we obtain that the $k$-tuple $x_1,\ldots,x_k$ does not maximize the sum of the squares subject to $x_1\ge x_2\ge \cdots \ge x_k\ge 0$ and the $k$ inequalities listed in the statement of the lemma. Hence, we conclude that $x_1+\cdots+x_i=S_i$ for every $i\in [k]$, which implies that $x_i=S_i-S_{i-1}$ for every $i\in [k]$. Since it holds that $S_i-S_{i-1}=s_i+\cdots+s_k$, the statement of the lemma now follows. \end{proof} We next bound the spectral radius of skew-symmetric matrices with entries between $-1$ and $+1$. For $n\in\NN$, define $D_n$ to be the skew-symmetric matrix with all entries above the diagonal equal to $+1$ and all entries below the diagonal equal to $-1$. The next lemma asserts that the matrix $D_n$ has the largest possible spectral radius among all skew-symmetric matrices with entries between $-1$ and~$+1$. \begin{lemma} \label{lm:antisym} For every $n\in\NN$, the spectral radius of a skew-symmetric matrix $A\in [-1,1]^{n\times n}$ is at most the spectral radius of $D_n$. \end{lemma} \begin{proof} We fix $n$ and write $D$ for $D_n$ throughout the proof. We establish that for every vector $v\in\RR^n$ with $\|v\|=1$, there exists a vector $w$ that can be obtained from $v$ by permuting the entries of $v$ and changing their signs such that $\|Av\|\le\|Dw\|$. This would imply that the spectral radius of $A$ does not exceed that of $D$. First observe that it is enough to prove the inequality for vectors $v\in\RR^n$ with non-zero entries (if the statement fails for a vector $v$, it also fails for any unit vector obtained by any small perturbation of $v$). Next observe that if $A'$ is obtained by changing the signs of all entries in the $i$-th row and the $i$-th column and $v'$ is obtained from $v$ by changing the sign of its $i$-th entry, then $(Av)_i=-(A'v')_i$ and $(Av)_j=(A'v')_j$ for all $j\not=i$; in particular, $\|Av\|=\|A'v'\|$. Hence, we will assume without loss of generality that all entries of $v$ are positive. By permuting rows and columns of $A$ symmetrically and applying the same permutation to $v$, we can assume that $(Av)_1\ge (Av)_2\ge \cdots \ge (Av)_n$. Let $k$ be the largest index such that $(Av)_k\ge 0$, and let $w$ be the vector obtained from $v$ by permuting its first $k$ entries and its remaining $n-k$ entries separately in a way that $w_1\le w_2\le\dots\le w_k$ and $w_{k+1}\ge w_{k+2}\ge\dots\ge w_n$. We will show that \begin{equation} \sum_{i=1}^k(Av)_i^2\le\sum_{i=1}^k(Dw)_i^2\quad\mbox{and}\quad\sum_{i=k+1}^n(Av)_i^2\le\sum_{i=k+1}^n(Dw)_i^2.\label{eq:antisym1} \end{equation} The arguments for the two cases are symmetric and so we focus on establishing that \begin{equation} \sum_{i=1}^k(Av)_i^2\le\sum_{i=1}^k(Dw)_i^2.\label{eq:antisym2} \end{equation} Observe that the following holds for every $m\in [k]$ (we use that $A_{ij}+A_{ji}=0$ for all $i,j\in [m]$): \begin{align*} \sum_{i=1}^m (Av)_i & = \sum_{i=1}^m\sum_{j=1}^n A_{ij}v_j \\ & \le \sum_{1\le i<j\le m}\max\{v_i-v_j,v_j-v_i\}+\sum_{i=1}^m\sum_{j=m+1}^n v_j \\ & = \sum_{1\le i<j\le m} \left(v_i+v_j-2\min\{v_i,v_j\}\right)+m\sum_{j=m+1}^n v_j \\ & = -\sum_{1\le i,j\le m} \min\{v_i,v_j\}+m\sum_{j=1}^n v_j \\ & \le -\sum_{1\le i,j\le m} \min\{w_i,w_j\}+m\sum_{j=1}^n w_j \\ & = \sum_{1\le i<j\le m} \left(w_j-w_i\right)+m\sum_{j=m+1}^n w_j \\ & = \sum_{i=1}^m\sum_{j=1}^n D_{ij}w_j = \sum_{i=1}^m (Dw)_i. \end{align*} Let $k'$ be the largest index such that $k'\le k$ and $(Dw)_{k'}>0$. We choose $\varepsilon>0$ and apply Lemma~\ref{lm:sumsq} with the following parameters: $x_i=(Av)_i$, $i\in [k]$, $s_i=(Dw)_i-(Dw)_{i+1}$ for $i\in [k'-1]$, $s_{k'}=(Dw)_{k'}$ and $s_i=\varepsilon$, $i\in [k]\setminus [k']$. Observe that $x_1,\ldots,x_{k}$ and $s_1,\ldots,s_{k}$ satisfy the assumptions of Lemma~\ref{lm:sumsq}. Hence, Lemma~\ref{lm:sumsq} implies that \[\sum_{i=1}^{k} x_i^2\le\sum_{i=1}^{k'}\left((Dw)_i+(k-k')\varepsilon\right)^2+(k-k')\varepsilon^2.\] Since this inequality holds for every $\varepsilon>0$, we obtain that \[\sum_{i=1}^{k} x_i^2\le\sum_{i=1}^{k'}(Dw)_i^2.\] This establishes \eqref{eq:antisym2}. The other inequality in \eqref{eq:antisym1} can be proven analogously. Hence, we conclude that $\|Av\|\le\|Dw\|$ as desired. \end{proof} We next use Lemma~\ref{lm:antisym} to bound elements of $\wsigma(W)$ of a regular tournamenton~$W$. \begin{lemma} \label{lm:ck4} Let $W$ be a tournamenton. If $1/2$ is contained in $\wsigma(W)$, then each other element of $\wsigma(W)$ has absolute value at most $1/\pi$. \end{lemma} \begin{proof} Let $A_n$, $n\in\NN$, be a convergent sequence of step approximations of $W$, and let $k_n$ be the order of $A_n$. Since $W$ is regular by Proposition~\ref{prop:reg}, the sum of each row of $A_n$ is $1/2$. This yields that $1/2$ is an eigenvalue of $A_n$ and the associated eigenvector is $(1,\ldots,1)$. Let $J_{k_n}$ be the square matrix of order $k_n$ with all entries equal to $1$ and let $B_n=J_{k_n}-2k_n\cdot A_n$. Note that the matrix $B_n$ is skew-symmetric and all its entries are between $-1$ and $+1$. Also note that the vector $(1,\ldots,1)$ is an eigenvector of $A_n$ associated with the eigenvalue $1/2$, and it is also an eigenvector of $B_n$ associated with the eigenvalue $0$. Since $B_n$ is skew-symmetric, all its non-zero eigenvalues are purely imaginary and there exists an orthonormal basis of~$\CC^{k_n}$ formed by eigenvectors of the matrix $B_n$ (see the beginning of Section~\ref{sec:c8} for a more detailed exposition on properties of skew-symmetric matrices). This means that we can assume that every eigenvector of $B_n$ associated with a non-zero eigenvalue is orthogonal to the vector $(1,\ldots,1)$ in the space $\CC^{k_n}$. Since $(1,\ldots,1)$ is an eigenvector of $B_n$ associated with the eigenvalue $0$, every eigenvector of~$B_n$ associated with a non-zero eigenvalue of $B_n$ is also an eigenvector of $A_n$. Thus, if $\lambda$ is an eigenvalue of $A_n$, then either $\lambda$ is equal to $1/2$ (for the vector $(1,\ldots,1)$) or $-2k_n\lambda$ is an eigenvalue of~$B_n$. Since the spectral radius of $B_n$ is at most the spectral radius of the matrix~$D_{k_n}$ by Lemma~\ref{lm:antisym} and the spectral radiuses of the matrices $D_{k_n}$ divided by $k_n$ converge to $2/\pi$, it follows that the limit of the maximum absolute value of an eigenvalue of $A_n$ different from $1/2$ is at most $1/\pi$. The statement of the lemma now follows. \end{proof} We are now ready to asymptotically determine $c(\ell)$ for $\ell$ divisible by four. Since the multiset $\wsigma(W_C)$ for the carousel tournamenton $W_C$, which we described at the end of Section~\ref{sec:prelim}, consists of $1/2$ and $\pm\unit/((2i-1)\pi)$ for $i\in\NN$, we obtain using \eqref{eq:lambda} that \[c(\ell)\ge 1+2\cdot\sum_{i=1}^{\infty}\left(\frac{2}{(2i-1)\pi}\right)^{\ell}\] for every $\ell$ divisible by four. The next theorem of this section provides an asymptotically matching upper bound. \begin{theorem} \label{thm:ck4} For every $\varepsilon>0$, there exists $\ell_0$ such that the following holds for every $\ell\ge\ell_0$ divisible by four: \[c(\ell)\le 1+\left(\frac{2}{\pi}+\varepsilon\right)^{\ell}.\] \end{theorem} \begin{proof} Let $W_k$ be a tournamenton that maximizes the sum of the $(4k)$-th powers of $\wsigma(W_k)$. Suppose that the statement of the theorem is false, i.e., there exists $\varepsilon>0$ and a sequence $(k_i)_{i\in\NN}$ such that \begin{equation}\label{eq:ck4c} C(W_{k_i},4k_i)>1+\left(\frac{2}{\pi}+\varepsilon\right)^{4k_i} \end{equation} for every $i\in\NN$. Without loss of generality, we may assume that the sequence $(W_{k_i})_{i\in\NN}$ is convergent in the cut distance and let $W$ be the tournamenton that is its limit. Since partitions of $[0,1]$ corresponding to fine enough step approximations of $W_{k_i}$ yield step approximations close to $W$ in the cut distance and the step approximations of $W_{k_i}$ and $W$ are also close in the cut distance, the sets $\left(\wsigma(W_{k_i})\cup\{0^\infty\}\right)_{i\in\NN}$ converge to $\wsigma(W)\cup\{0^\infty\}$ as multisets. Let $\rho$ be the largest positive real contained in $\wsigma(W)$. If $\rho$ is smaller than $1/2$, then it holds that \[\lim_{i\to\infty}C(W_{k_i},4k_i)=0.\] Since this contradict the choice of $(k_i)_{i\in\NN}$, we can assume that $\rho=1/2$, i.e., $\wsigma(W)$ contains $1/2$. Hence, all other non-zero elements of $\wsigma(W)$ are purely imaginary and the absolute value of every such element at most $1/\pi$ by Lemma~\ref{lm:ck4}. Let $m$ be the number elements of $\wsigma(W)$ with the absolute value $1/\pi$ (note that $m$ can be zero). We can assume that $\varepsilon$ is small enough that the absolute value of any element of $\wsigma(W)$ with absolute value smaller than $1/\pi$ is at most $1/\pi-\varepsilon$. It follows that \begin{equation} \lim_{i\to\infty}\frac{C(W_{k_i},4k_i)-2^{4k_i}\sum_{\lambda\in\wsigma(W_{k_i}),|\lambda|\ge 1/\pi-\varepsilon/2}\lambda^{4k_i}}{(2/\pi)^{4k_i}}=0.\label{eq:ck4a} \end{equation} The choice of $m$ implies that there exists $i_0$ such that \begin{equation} \sum_{\lambda\in\wsigma(W_{k_i}),|\lambda|\ge 1/\pi-\varepsilon/2}\lambda^{4k_i}\le\frac{1}{2^{4k_i}}+m\left(\frac{1}{\pi}+\frac{\varepsilon}{4}\right)^{4k_i}\label{eq:ck4b} \end{equation} for every $i\ge i_0$. Using \eqref{eq:ck4c}, \eqref{eq:ck4a} and \eqref{eq:ck4b} we get a contradiction. \end{proof} \section{Cycles of length eight} \label{sec:c8} In this section, we present our results on cycles of length eight. We also present the analogous arguments in the (simpler) case of cycles of length four to make the exposition more accessible, although the presented results for cycles of length four have been previously proven. To be able to present our arguments, we need to recall some results on matrices and particularly on skew-symmetric matrices. If $A$ is a square matrix of order $n$ and $X\subseteq [n]$, then $A[X]$ is the square matrix formed by entries in the rows and the columns indexed by the elements of $X$. Throughout this section, $J_n$ denotes the square matrix of order $n$ with all entries equal to $1$. Recall that $D_n$ is the skew-symmetric $n\times n$ matrix with all entries above the diagonal equal to $+1$ and all entries below the diagonal equal to $-1$. We say that two skew-symmetric matrices are \emph{sign-equivalent} if one can be obtained from the other by permuting the rows and columns symmetrically and multiplying some of the rows and the symmetric set of columns by $-1$. It is well-known that for every real skew-symmetric matrix~$A$, there exists an orthogonal (real) matrix $Q$ (i.e. a square matrix $Q$ such that $Q^TQ$ is the identity matrix) such that the matrix $Q^TAQ$ is a block diagonal matrix with two kinds of blocks: blocks of size two of the form $\begin{pmatrix} 0 & a \\ -a & 0 \end{pmatrix}$ and blocks of size one equal to the zero matrix. In particular, there exists an orthogonal basis formed by eigenvectors of the matrix $A^2$ and the values $-a^2$ from the blocks of the matrix $Q^TAQ$ are the eigenvalues of $A^2$ (each with multiplicity two). In addition, if $v$ and $v'$ are two rows of $Q$ associated with the single block of $Q^TAQ$ with a value $a$, i.e., $av=Av'$ and $-av'=Av$, the complex vectors $v+\unit v'$ and $v-\unit v'$ are eigenvectors of $A$ associated with the eigenvalues $a\unit$ and $-a\unit$, respectively, and the vectors $v+\unit v'$ and $v-\unit v'$ are orthogonal in the space $\CC^n$. We apply the just reviewed results on skew-symmetric matrices to get the following upper bound the trace of the fourth and eight powers of the sum of the all-one matrix and a skew-symmetric matrix. \begin{lemma} \label{lm:midterms} Let $B$ be a skew-symmetric matrix of order $n$ with entries between $-1$ and $+1$. It holds that \begin{align*} \Trace (J_n+B)^4 &=\Trace J_n^4+\Trace B^4-4n||Bj||^2 \quad\mbox{and}\\ \Trace (J_n+B)^8 &\le\Trace J_n^8+\Trace B^8-2n^5||Bj||^2, \end{align*} where $j$ is the vector with all entries equal to one. In particular, it holds that \[\Trace (J_n+B)^4\le\Trace J_n^4+\Trace B^4\quad\mbox{and}\quad \Trace (J_n+B)^8\le\Trace J_n^8+\Trace B^8,\] and equality holds if and only if the sum of each row of $B$ is zero. \end{lemma} \begin{proof} We start with the trace of the $4$-th power of $J_n+B$. We obtain the following by expanding $(J_n+B)^4$ and using that $\Trace XY=\Trace YX$: \[\Trace (J_n+B)^4=\Trace J_n^4+4\Trace J_n^3B+4\Trace J_n^2B^2+2\Trace J_nBJ_nB+4\Trace J_nB^3+\Trace B^4.\] Since $B$ is skew-symmetric, any odd power of $B$ is also skew-symmetric. In particular, $J_nB^{2k-1}J_n$ is the zero matrix and $\Trace J_n B^{2k-1}=0$ for every $k\in\NN$. It follows that \begin{equation} \Trace (J_n+B)^4=\Trace J_n^4+4\Trace J_n^2B^2+\Trace B^4.\label{eq:mid1} \end{equation} Let $Q$ be the orthogonal matrix such that $Q^TBQ$ has the block structure described before the statement of this lemma, let $k$ be the number of blocks of size two, and let $a_1,\ldots,a_k$ be the (non-zero) numbers associated with these blocks. Since the trace of $B^2$ is equal to $2(a_1^2+\cdots+a_k^2)$, it follows that \begin{equation} a_1^2+\cdots+a_k^2\le \frac{n^2}{2}.\label{eq:mida} \end{equation} Further, let $q_i$ and $q'_i$ be the two rows of $Q$ corresponding to the block with $a_i$, $i\in [k]$, and let $\alpha_i\in [0,\pi/2]$ be the angle between the vector $j$ and the plane generated by $q_i$ and $q'_i$, $i\in [k]$. Note that the $(n-2k)$-dimensional subspace orthogonal to the space generated by $q_1,\ldots,q_k$ and $q'_1,\ldots,q'_k$ is the kernel of $B$ (as it is generated by the rows of $Q$ corresponding to the blocks of size one). Since the rows of $Q$ form an orthogonal basis, it follows that \begin{equation} \sum_{i=1}^k\cos^2\alpha_i\le 1.\label{eq:mid2} \end{equation} Observe that the following identities hold. \begin{align*} \Trace J_n^4 & = n^4 \\ \Trace J_n^2B^2 & = -n^2\sum_{i=1}^k a_i^2\cos^2\alpha_i \\ \Trace B^4 & = 2\sum_{i=1}^k a_i^4 \end{align*} In particular, the second term in \eqref{eq:mid1} is non-positive and equal to $-4n||Bj||^2$. Hence, $\Trace (J_n+B)^4\le \Trace J_n^4+\Trace B^4$ and equality holds if and only if the vector $j$ is in the kernel of $B$. The latter holds if and only if the sum of each row of $B$ is zero. This establishes the statement of the lemma concerning the trace of the $4$-th power of $J_n+B$. We next analyze the trace of the $8$-th power of $J_n+B$. As in the case of the $4$-th power the trace of some terms in the expansion of $(J_n+B)^8$ is zero, and we obtain the following. \begin{align} \Trace (J_n+B)^8 & = \Trace J_n^8+8\Trace J_n^6B^2+8\Trace J_n^4B^4+8\Trace J_n^3B^2J_nB^2\nonumber\\ & + 4\Trace J_n^2B^2J_n^2B^2+8\Trace J_n^2B^6+8\Trace J_nB^2J_nB^4+\Trace B^8\label{eq:mid3} \end{align} As in the previous case, we can express some of the terms in \eqref{eq:mid3} using $a_i$ and $\alpha_i$, $i\in [k]$. \begin{align*} \Trace J_n^8 & = n^8 \\ \Trace J_n^6B^2 & = -n^6\sum_{i=1}^k a_i^2\cos^2\alpha_i \\ \Trace J_n^4B^4 & = n^4\sum_{i=1}^k a_i^4\cos^2\alpha_i \\ \Trace J_n^3B^2J_nB^2 & = \Trace J_n^2B^2J_n^2B^2 = n^4\left(\sum_{i=1}^k a_i^2\cos^2\alpha_i\right)^2 \\ \Trace J_n^2B^6 & = -n^2\sum_{i=1}^k a_i^6\cos^2\alpha_i \\ \Trace J_nB^2J_nB^4 & = -n^2\left(\sum_{i=1}^k a_i^2\cos^2\alpha_i\right)\left(\sum_{i=1}^k a_i^4\cos^2\alpha_i\right)\\ \Trace B^8 & = 2\sum_{i=1}^k a_i^8 \end{align*} We derive using \eqref{eq:mid2} that \begin{align} &2\Trace J_n^6B^2+8\Trace J_n^3B^2J_nB^2+8\Trace J_nB^2J_nB^4\nonumber\\ &=-2n^2\left(\sum_{i=1}^k a_i^2\cos^2\alpha_i\right)\left(n^4-4n^2\sum_{i=1}^k a_i^2\cos^2\alpha_i+4\sum_{i=1}^k a_i^4\cos^2\alpha_i\right)\nonumber\\ &\le-2n^2\left(\sum_{i=1}^k a_i^2\cos^2\alpha_i\right)\left(n^4\sum_{i=1}^k\cos^2\alpha_i-4n^2\sum_{i=1}^k a_i^2\cos^2\alpha_i+4\sum_{i=1}^k a_i^4\cos^2\alpha_i\right)\nonumber\\ &=-2n^2\left(\sum_{i=1}^k a_i^2\cos^2\alpha_i\right)\left(\sum_{i=1}^k\left(n^2-2a_i^2\right)^2\cos^2\alpha_i\right)\le 0.\label{eq:mid4+} \end{align} Using \eqref{eq:mida} and \eqref{eq:mid2}, we obtain that \[\sum_{i=1}^k a_i^2\cos^2\alpha_i\le\frac{n^2}{2},\] which yields that \begin{align} 2\Trace J_n^6B^2+4\Trace J_n^2B^2J_n^2B^2&=\nonumber\\ 2n^4\left(\sum_{i=1}^k a_i^2\cos^2\alpha_i\right)\left(-n^2+2\sum_{i=1}^k a_i^2\cos^2\alpha_i\right)&\le 0.\label{eq:mid4} \end{align} We finally bound a portion of the second term, and the third and sixth terms in \eqref{eq:mid3} as follows. \begin{align} 2\Trace J_n^6B^2+8\Trace J_n^4B^4+8\Trace J_n^2B^6&=\nonumber\\ -2n^2\sum_{i=1}^k\left(n^4-4n^2a_i^2+4a_i^4\right)a_i^2\cos^2\alpha_i&=\nonumber\\ -2n^2\sum_{i=1}^k\left(n^2-2a_i^2\right)^2a_i^2\cos^2\alpha_i&\le 0\label{eq:mid5} \end{align} Using \eqref{eq:mid4+}, \eqref{eq:mid4} and \eqref{eq:mid5}, we obtain the following estimate on $\Trace (J_n+B)^8$ using the expansion in \eqref{eq:mid3}. \begin{equation} \Trace (J_n+B)^8 \le \Trace J_n^8+\Trace B^8+2\Trace J_n^6B^2\label{eq:mid6} \end{equation} Since it holds that $\Trace J_n^6B^2=-n^5||Bj||^2\le 0$, the inequality from the statement of the lemma now follows. Hence, $\Trace (J_n+B)^8\le \Trace J_n^8+\Trace B^8$. Moreover, since the inequalities \eqref{eq:mid4+}, \eqref{eq:mid4} and \eqref{eq:mid5} hold with equality if $\cos\alpha_i=0$ for every $i\in [k]$, which holds if $j$ is in the kernel of $B$, we can conclude that $\Trace (J_n+B)^8=\Trace J_n^8+\Trace B^8$ if and only if the sum of each row of $B$ is zero. \end{proof} The following two lemmas will be important to establish an upper bound on the trace of the last term in the upper bound given in Lemma~\ref{lm:midterms}. To state the lemmas, we need the following definition. For a square matrix $A$ of order $n$ we define the \emph{cyclic index} of $A$ as \[\Cycl A=\sum_{\pi\in S_n}\prod_{i=1}^nA_{\pi(i)\pi(i+1)},\] where the computation with indices is modulo $n$, i.e., $\pi(n+1)=\pi(1)$. \begin{lemma} \label{lm:trace4} Let $B$ be a skew-symmetric matrix of order $4$ such that each off-diagonal entry of $B$ is $+1$ or $-1$. The cyclic index of $B$ is at most the cyclic index of $D_4$ and equality holds if and only if $B$ is sign-equivalent to $D_4$. \end{lemma} \begin{proof} Since the cyclic index of sign-equivalent matrices is the same, we may assume without loss of generality that the first row of $B$ contains $+1$ only. Hence, we need to consider the following two matrices (after a permutation of rows and columns): \[ \begin{pmatrix} 0 & +1 & +1 & +1 \\ -1 & 0 & +1 & +1 \\ -1 & -1 & 0 & +1 \\ -1 & -1 & -1 & 0 \\ \end{pmatrix} \quad\mbox{and}\quad \begin{pmatrix} 0 & +1 & +1 & +1 \\ -1 & 0 & +1 & -1 \\ -1 & -1 & 0 & +1 \\ -1 & +1 & -1 & 0 \\ \end{pmatrix} \] The cyclic index of the left matrix is $8$ and the cyclic index of the right matrix is $-24$. The statement of the lemma follows. \end{proof} To state the next lemma, we need to introduce another skew-symmetric matrix. \[D'_8= \begin{pmatrix} 0 & +1 & +1 & +1 & +1 & +1 & +1 & +1 \\ -1 & 0 & +1 & -1 & +1 & +1 & +1 & +1 \\ -1 & -1 & 0 & -1 & +1 & +1 & +1 & +1 \\ -1 & +1 & +1 & 0 & -1 & -1 & -1 & -1 \\ -1 & -1 & -1 & +1 & 0 & +1 & +1 & -1 \\ -1 & -1 & -1 & +1 & -1 & 0 & +1 & +1 \\ -1 & -1 & -1 & +1 & -1 & -1 & 0 & +1 \\ -1 & -1 & -1 & +1 & +1 & -1 & -1 & 0 \end{pmatrix} \] \begin{lemma} \label{lm:trace8} Let $B$ be a skew-symmetric matrix of order $8$ such that each off-diagonal entry of $B$ is $+1$ or $-1$. The cyclic index of $B$ is at most the cycle index of $D_8$ and equality holds if and only if $B$ is sign-equivalent to $D_8$ or to $D'_8$. \end{lemma} The proof of Lemma~\ref{lm:trace8} proceeds by a computer assisted inspection of all skew-symmetric $8 \times 8$ matrices where the off-diagonal entries in the first row are $+1$, those in the first column are $-1$, and all other off-diagonal entries are either $+1$ or $-1$. In an independent way, we have prepared a C program and a C++ program to verify Lemma~\ref{lm:trace8}, i.e., to check that the cyclic index of every skew-symmetric $8 \times 8$ matrix of the above form is at most $2\,176$ and the equality holds if and only if the matrix is sign-equivalent to $D_8$ or to $D'_8$; the code of the C program and its output are available as ancillary files on arXiv. We are now ready to compute the value of $c(8)$. \begin{theorem} \label{thm:c8} It holds that $c(4)=4/3$ and $c(8)=332/315$. \end{theorem} \begin{proof} Fix $\ell\in\{4,8\}$ and let $A$ be the tournament matrix of an $n$-vertex tournament $T$. Proposition~\ref{prop:eigen} yields that \[C(T,\ell)=\frac{2^\ell}{n^\ell}\Trace A^\ell+O(n^{-1}).\] Let $B=J_n-2A$ and note that $A=\frac{J_n+B}{2}$. By Lemma~\ref{lm:midterms}, we obtain that \[\Trace A^\ell\le\frac{1}{2^\ell}\left(\Trace J_n^\ell+\Trace B^\ell\right).\] The trace of $B^\ell$ can be combinatorially interpreted as the sum taken over all closed walks with length $\ell$ in $T$ where the sum contains $+1$ for every such walk with an even number of forward edges and $-1$ for every such walk with an odd number of forward edges. Such walks that are not cycles contribute to the sum only $O(n^{\ell-1})$ and those that are cycles can be counted as cyclic indices of the square submatrices with rows and columns indexed by the vertices of the cycle. Hence, we obtain the following. \[\Trace B^\ell=\sum_{X\in\binom{[n]}{\ell}}\Cycl B[X]+O(n^{\ell-1}).\] By Lemmas~\ref{lm:trace4} and~\ref{lm:trace8}, it holds $\Cycl B[X]\le\Cycl D_\ell$ for every $X\in\binom{[n]}{\ell}$, which yields that \[\Trace B^\ell\le\Trace D_n^\ell+O(n^{\ell-1}).\] Hence, we obtain that \[C(T,\ell)\le\frac{1}{n^\ell}\left(n^\ell+\Trace D_n^\ell+O(n^{\ell-1})\right).\] We next proceed separately for $\ell=4$ and $\ell=8$. Analyzing the spectrum of the matrix $D_n$ yields that \[\lim_{n\to\infty}\frac{\Trace D_n^4}{n^4}=2\sum_{i=1}^{\infty}\left(\frac{2}{(2i-1)\pi}\right)^4=\frac{1}{3},\] which implies that \[C(T,4)\le\frac{4}{3}+O(n^{-1}).\] Similarly, we obtain that \[\lim_{n\to\infty}\frac{\Trace D_n^8}{n^8}=2\sum_{i=1}^{\infty}\left(\frac{2}{(2i-1)\pi}\right)^8=\frac{17}{315},\] which implies that \[C(T,8)\le\frac{332}{315}+O(n^{-1}).\] Since it holds that $C(W_C,4)=4/3$ and $C(W_C,8)=332/315$ for the carousel tournamenton $W_C$, the statement of the theorem now follows. \end{proof} The methods used to prove Theorem~\ref{thm:c8} actually provide the characterization of extremal tournamentons. We fix some additional notation: $T^4$ is the $4$-vertex transitive tournament, $C^4$ is the unique $4$-vertex hamiltonian tournament, $L^4$ is the unique $4$-vertex non-transitive tournament with a sink, and $W^4$ is the unique $4$-vertex non-transitive tournament with a source. The four tournaments are depicted in Figure~\ref{fig:CLTW}. It is also interesting to note that the matrix $D'_8$ is sign-equivalent to the following matrix $D''_8$: \[D''_8= \begin{pmatrix} 0 & +1 & +1 & -1 & +1 & +1 & +1 & +1 \\ -1 & 0 & +1 & +1 & +1 & +1 & +1 & +1 \\ -1 & -1 & 0 & +1 & +1 & +1 & +1 & +1 \\ +1 & -1 & -1 & 0 & +1 & +1 & +1 & +1 \\ -1 & -1 & -1 & -1 & 0 & +1 & +1 & -1 \\ -1 & -1 & -1 & -1 & -1 & 0 & +1 & +1 \\ -1 & -1 & -1 & -1 & -1 & -1 & 0 & +1 \\ -1 & -1 & -1 & -1 & +1 & -1 & -1 & 0 \end{pmatrix} . \] The $8$-vertex tournament with the tournament matrix $\frac{J_8+D''_8}{2}$ is the tournament obtained from two copies of $C^4$ by adding edges directed from the first copy to the second; this tournament is depicted in Figure~\ref{fig:C4C4}. \begin{figure} \begin{center} \epsfbox{tourn-ck-4.mps} \hskip 1cm \epsfbox{tourn-ck-2.mps} \hskip 1cm \epsfbox{tourn-ck-5.mps} \hskip 1cm \epsfbox{tourn-ck-3.mps} \end{center} \caption{The tournaments $T^4$, $C^4$, $L^4$ and $W^4$.} \label{fig:CLTW} \end{figure} \begin{figure} \begin{center} \epsfbox{tourn-ck-6.mps} \end{center} \caption{The $8$-vertex tournament with the tournament matrix $\frac{J_8+D''_8}{2}$. The edges between the two copies of $C^4$ are drawn in gray to better display the structure of the tournament.} \label{fig:C4C4} \end{figure} \begin{theorem} \label{thm:c48} Let $W$ be a tournamenton. It holds that $C(W,4)=c(4)=4/3$ if and only if $W$ is weakly isomorphic to the carousel tournamenton $W_C$, and it holds that $C(W,8)=c(8)=332/315$ if and only if $W$ is weakly isomorphic to the carousel tournamenton $W_C$. \end{theorem} \begin{proof} If $W$ is weakly isomorphic to the the carousel tournamenton $W_C$, then $C(W,4)=C(W_C,4)=c(4)$ and $C(W,8)=C(W_C,8)=c(8)$. Hence, we focus on proving the converse implications and start with the one concerning cycles of length four. Let $W$ be a tournamenton such that $C(W,4)=c(4)$. Let $T_n$, $n\in\NN$, be an $n$-vertex $W$-random tournament, and let $B_n$ be the skew-symmetric matrix such that $\frac{J_n+B_n}{2}$ is the tournament matrix of $T_n$. Note that the limit of $C(T_n,4)$ is $C(W,4)$ with probability one. Similarly to the proof of Theorem~\ref{thm:c8}, we obtain using Lemma~\ref{lm:midterms} that \[C(T_n,4)\le 1+\frac{1}{n^4}\Trace D_n^4-\frac{4}{n^3}||B_nj||^2+O(n^{-1}),\] where $j$ is the vector with all entries equal to one. It follows that \[C(W,4)=\lim_{n\to\infty} C(T_n,4)\le 1+\lim_{n\to\infty}\frac{1}{n^4}\Trace D_n^4=c(4).\] Furthermore, equality can hold only if \begin{equation} \lim_{n\to\infty}\frac{\|B_nj\|^2}{n^3}=0\label{eq:Bnj} \end{equation} and the proportion of principal submatrices of $B_n$ of order four that are sign-equivalent to $D_4$ tends to $1$. The latter implies that $d(T^4,W)+d(C^4,W)=1$ and $d(L^4,W)+d(W^4,W)=0$, i.e., the only $4$-vertex tournaments with positive density in $W$ are $T^4$ and $C^4$. We conclude that all $4$-vertex subtournaments of $T_n$ are $T^4$ and $C^4$ (with probability one), in particular, the in-neighborhood and the out-neighborhood of every vertex of $T_n$ is transitive. If \eqref{eq:Bnj} holds, then $1/2\in\wsigma(W)$ and the tournamenton $W$ is regular by Proposition~\ref{prop:reg}. Hence, the in-degree of every vertex of $T_n$ is close to $n/2$ with high probability, which implies that $W$ is a limit of the carousel tournaments described at the end of Section~\ref{sec:prelim}. Since the tournamenton $W_C$ is also a limit of the carousel tournaments, the tournamentons $W$ and $W_C$ are weakly isomorphic. We now deal with the case of cycles of length eight. Let $W$ be a tournamenton such that $C(W,8)=c(8)$. As in the previous case, we conclude that $W$ is regular and the only $8$-vertex tournaments with positive density in $W$ are those whose tournament matrix $A$ satisfies that $2A-J_n$ is sign-equivalent to $D_8$ and $D'_8$. A~straightforward case analysis yields that every skew-symmetric matrix $B$ of order nine such that every principal submatrix of order eight of $B$ is sign-equivalent to $D_8$ or $D'_8$ satisfies that $B$ is sign-equivalent to $D_9$. It follows that the only $9$-vertex tournaments with positive density in $W$ are those whose tournament matrix $A$ satisfies that $2A-J_n$ is sign-equivalent to $D_9$. Consequently, the only $4$-vertex tournaments with positive density in $W$ are those whose tournament matrix $A$ satisfies that $2A-J_n$ is sign-equivalent to $D_4$, i.e., the tournaments $T^4$ and $C^4$. Analogously to the previous case, we now conclude that the tournamentons $W$ and $W_C$ are weakly isomorphic. \end{proof} \bibliographystyle{bibstyle}
{ "timestamp": "2022-07-25T02:10:38", "yymm": "2008", "arxiv_id": "2008.06577", "language": "en", "url": "https://arxiv.org/abs/2008.06577", "abstract": "We study the asymptotic behavior of the maximum number of directed cycles of a given length in a tournament: let $c(\\ell)$ be the limit of the ratio of the maximum number of cycles of length $\\ell$ in an $n$-vertex tournament and the expected number of cycles of length $\\ell$ in the random $n$-vertex tournament, when $n$ tends to infinity. It is well-known that $c(3)=1$ and $c(4)=4/3$. We show that $c(\\ell)=1$ if and only if $\\ell$ is not divisible by four, which settles a conjecture of Bartley and Day. If $\\ell$ is divisible by four, we show that $1+2\\cdot\\left(2/\\pi\\right)^{\\ell}\\le c(\\ell)\\le 1+\\left(2/\\pi+o(1)\\right)^{\\ell}$ and determine the value $c(\\ell)$ exactly for $\\ell = 8$. We also give a full description of the asymptotic structure of tournaments with the maximum number of cycles of length $\\ell$ when $\\ell$ is not divisible by four or $\\ell\\in\\{4,8\\}$.", "subjects": "Combinatorics (math.CO)", "title": "Cycles of a given length in tournaments", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9886682451330957, "lm_q2_score": 0.8104789018037399, "lm_q1q2_score": 0.8012947535637022 }
https://arxiv.org/abs/1201.6649
Discriminant coamoebas through homology
Understanding the complement of the coamoeba of a (reduced) A-discriminant is one approach to studying the monodromy of solutions to the corresponding system of A-hypergeometric differential equations. Nilsson and Passare described the structure of the coamoeba and its complement (a zonotope) when the reduced A-discriminant is a function of two variables. Their main result was that the coamoeba and zonotope form a cycle which is equal to the fundamental cycle of the torus, multiplied by the normalized volume of the set A of integer vectors. That proof only worked in dimension two. Here, we use simple ideas from topology to give a new proof of this result in dimension two, one which can be generalized to all dimensions.
\section*{Introduction} $A$-hypergeometric functions, which are solutions to $A$-hypergeometric systems of differential equations~\cite{GKZ89,GKZ94,SST}, enjoy two complimentary analytical formulae which together give an approach to studying the monodromy of the solutions~\cite{Beukers} at non-resonant parameters. One formula is as explicit power series whose convergence domains in ${\bf{C}}^{N+1}$ have an action of the group ${\bf{T}}^{N+1}$ of phases. These power series form a basis of solutions, with known local monodromy around loops from ${\bf{T}}^{N+1}$. Another formula is as $A$-hypergeometric Mellin-Barnes integrals~\cite{Nilsson} evaluated at phases $\theta\in{\bf{T}}^{N+1}$. When the Mellin-Barnes integrals give a basis of solutions, they may be used to glue together the local monodromy groups and determine a subgroup of the monodromy group, which may sometimes be the full monodromy group. Here, $A\subset{\bf{Z}}^n$ consists of $N{+}1$ integer vectors that generate ${\bf{Z}}^n$. Considering ${\bf{Z}}^n\subset{\bf{Z}}^{n+1}$ as the vectors with first coordinate 1, we regard $A$ as a collection of $N{+}1$ vectors in ${\bf{Z}}^{n+1}$. The $A$-discriminant is a multihomogeneous polynomial in $N{+}1$ variables with $n{+}1$ homogeneities corresponding to $A$. Removing these homogeneities gives the reduced $A$-discriminant, $D_B$, which is a hypersurface in ${\bf{C}}^d$ ($d:=N{-}n$) that depends upon a vector configuration $B\subset{\bf{Z}}^d$ Gale dual to $A$. This reduction corresponds to a homomorphism $\beta\colon({\bf{C}}^*)^{N+1}\to({\bf{C}}^*)^d$ and induces a corresponding map $\Arg(\beta)$ on phases. The Mellin-Barnes integrals at $\theta\in{\bf{T}}^{N+1}$ give a basis of solutions when $\Arg(\beta)(\theta)$ has a neighborhood in ${\bf{T}}^{d}$ with the property that no point of $D_B$ has a phase lying in that neighborhood~\cite{Nilsson}. By results in~\cite{Johansson,NS}, this means that $\Arg(\beta)(\theta)$ lies in the complement of the closure of the coamoeba ${\mathcal{A}}_B$ of $D_B$. When $d=2$, the closure of ${\mathcal{A}}_B$ and its complement were described in~\cite{NP10} as topological chains in ${\bf{T}}^2$ (induced from natural chains in its universal cover ${\bf{R}}^2$, where ${\bf{T}}^2=({\bf{R}}/2\pi{\bf{Z}})^2$). The closure of the coamoeba is an explicit chain depending on $B$. Its edges coincide with the edges of the zonotope $Z_B$ generated by $B$. The main result of~\cite{NP10} is the following theorem. \begin{itheorem}\label{T:NP} The sum of the coamoeba chain $\overline{{\mathcal{A}}_B}$ and the zonotope $Z_B$ forms a two-dimensional cycle in ${\bf{T}}^2$ that is equal to $n!\vol(A)$ times the fundamental cycle. \end{itheorem} Here, $n!\vol(A)$ is the normalized volume of the convex hull of $A$, which is the dimension of the space of solutions to the (non-resonant) $A$-hypergeometric system. The zonotope $Z_B$ gives points in the complement of ${\mathcal{A}}_B$, by Theorem~\ref{T:NP}. Its proof in~\cite{NP10} only works when $d=2$ and it is not clear how to generalize it to $d>2$. However, any such generalization would be important, for Mellin-Barnes integrals at a set of phases $\theta$ where $\Arg(\beta)(\theta)$ are distinct points of $Z_B$ with the same image in ${\bf{T}}^d$ are linearly independent. We give a proof of Theorem~\ref{T:NP} which explains the occurrence of the zonotope and can be generalized to higher dimensions. This proof uses the Horn-Kapranov parametrization of the $A$-discriminant~\cite{K91}, which implies that the discriminant coamoeba is the image of the coamoeba of a line $\ell_B$ in $\P^N$ under the map $\Arg(\beta)$. We construct a piecewise linear \demph{zonotope chain} in ${\bf{T}}^N$ (the quotient of ${\bf{T}}^{N+1}$ by the diagonal torus) which is a cone over the boundary of the coamoeba of $\ell_B$, and compute the homology class of the sum of the coamoeba and this zonotope chain. This gives a formula for the image of this cycle under $\Arg(\beta)$, which we show is $n!\vol(A)$ times the fundamental cycle of ${\bf{T}}^2$. Theorem~\ref{T:NP} follows as the map $\Arg(\beta)$ sends the coamoeba of $\ell$ to the coamoeba ${\mathcal{A}}_B$ of $D_B$ and sends the zonotope chain to $Z_B$. While for $A$-discriminants, the set $A$ consists of distinct integer vectors and consequently its Gale dual $B$ generates ${\bf{Z}}^2$ and has no two vectors parallel, we establish Theorem~\ref{T:NP} in the greater generality of any finite multiset $B$ of integer vectors in ${\bf{Z}}^2$ with sum ${\bf 0}$ that spans ${\bf{R}}^2$. This generality is useful in our primary application to hypergeometric systems, for example the classical systems of Appell~\cite{appell} and Lauricella~\cite{lauricella} may be expressed as $A$-hypergeometric systems with repeated vectors in the Gale dual $B$. In this setting, we replace the reduced $A$-discriminant by the Horn-Kapranov parametrization given by the vectors $B$, and study the coamoeba ${\mathcal{A}}_B$ of the image, which is also written $D_B$. The normalized volume $n!\vol(A)$ of the configuration $A$ is replaced by a quantity $d_B$ that depends upon the vectors in $B$. We collect some preliminaries in Section~\ref{S:one}. In Section~\ref{S:realline} we study the coamoeba of a line in $\P^N$ defined over the real numbers and define its associated zonotope chain. Our main result is a computation of the homology class of the cycle formed by these two chains. In Section~\ref{S:three} we show that under the map $\Arg(\beta)$ the coamoeba and zonotope chains map to the coamoeba ${\mathcal{A}}_B$ and the zonotope $Z_B$, and a simple application of the result in Section~\ref{S:realline} shows that the homology class of $\overline{{\mathcal{A}}_B}+Z_B$ is $d_B$ times the fundamental cycle of ${\bf{T}}^2$.\medskip \noindent{\bf Remark.} This approach to reduced $A$-discriminant coamoebas and their complements was developed during the Winter 2011 semester at the Institut Mittag-Leffler, with the main result obtained in August 2011, along with a sketch of a program to extend it to $d\geq 2$. With the tragic death of Mikael Passare on 15 September 2011, the task of completing this paper fell to the second author, and the program extending these results is being carried out in collaboration with Mounir Nisse. \section{Coamoebas and cohomology of tori}\label{S:one} Throughout $N$ will be an integer strictly greater than 1. Let $\defcolor{\P^N}$ be $N$-dimensional complex projective space, which will always have a preferred set of coordinates $[x_1:\dotsb:x_N:x_{N+1}]$ (up to reordering). Similarly, $\defcolor{{\bf{C}}^N}$, $\defcolor{({\bf{C}}^*)^N}$, $\defcolor{{\bf{R}}^N}$, and $\defcolor{{\bf{Z}}^N}$ are $N$-tuples of complex numbers, non-zero complex numbers, real numbers, and integers, all with corresponding preferred coordinates. We will write ${\bf e}_i$ for the $i$th basis vector in a corresponding ordered basis. The argument map ${\bf{C}}^*\ni z=re^{\sqrt{-1}\theta}\mapsto\theta\in{\bf{T}}:={\bf{R}}/2\pi{\bf{Z}}$ induces an argument map $\defcolor{\Arg}\colon({\bf{C}}^*)^N\to{\bf{T}}^N$. To a subvariety $X\subset\P^N$ (or ${\bf{C}}^N$ or $({\bf{C}}^*)^N$) we associate its \demph{coamoeba} $\defcolor{{\mathcal{A}}(X)}\subset{\bf{T}}^N$ which is the image of $X\cap({\bf{C}}^*)^N$ under $\Arg$. The closure of the coamoeba ${\mathcal{A}}(X)$ was studied in~\cite{Johansson,NS}. This closure contains ${\mathcal{A}}(X)$, together with all limits of arguments of unbounded sequences in $X\cap({\bf{C}}^*)^N$, which constitute the \demph{phase limit set of $X$, ${\mathcal{P}}^\infty(X)$}. The main result of~\cite{NS} (proven when $X$ is a complete intersection in~\cite{Johansson}) is that ${\mathcal{P}}^\infty(X)$ is the union of the coamoebas of all initial degenerations of $X\cap({\bf{C}}^*)^N$. Lines in ${\bf{C}}^3$ were studied in~\cite{NS}, and the arguments there imply some basic facts about coamoebas of lines. When $X=\defcolor{\ell}\subset{\bf{C}}^N$ is a line which is not parallel to a sum of coordinate directions (${\bf e}_{i_1}+\dotsb+{\bf e}_{i_s}$ for some subset $\{i_1,\dotsc,i_s\}$ of $\{1,\dotsc,N\}$), its coamoeba is two-dimensional and its phase limit set is a union of at most $N{+}1$ one-dimensional subtori of ${\bf{T}}^N$, one for each point of $\ell$ at infinity, whose directions are parallel to sums of coordinate directions. If $\ell'\subset{\bf{C}}^M$ ($M<N$) is the image of $\ell$ under a coordinate projection, then the coamoeba ${\mathcal{A}}(\ell')$ is the image of ${\mathcal{A}}(\ell)$ under the induced projection. If $\ell'$ is not parallel to a sum of coordinate directions, then the map $\overline{{\mathcal{A}}(\ell)}\to\overline{{\mathcal{A}}(\ell')}$ is an injection except for those components of the phase limit set which are collapsed to points. \medskip The integral cohomology of the compact torus ${\bf{T}}^N$ is the exterior algebra $\wedge^*{\bf{Z}}^N$. Under the natural identification of homology with the linear dual of cohomology (which is again $\wedge^*{\bf{Z}}^N$), we will write ${\bf e}_i$ for the fundamental $1$-cycle $[{\bf{T}}_i]$ of the coordinate circle ${\bf{T}}_i:= 0^{i-1}\times {\bf{T}}\times 0^{N-i}$ and ${\bf e}_i\wedge {\bf e}_j$ is the fundamental cycle $[{\bf{T}}_{i,j}]$ of the coordinate 2-torus ${\bf{T}}_{i,j}\simeq{\bf{T}}^2$ in the directions $i$ and $j$ with the implied orientation. Given a continuous map $\rho\colon{\bf{T}}^N\to{\bf{T}}^2$, the induced map in homology is $\rho_*\colon H_*({\bf{T}}^N,{\bf{Z}})\to H_*({\bf{T}}^2,{\bf{Z}})$ where $\rho_*({\bf e}_i)=[\rho({\bf{T}}_i)]$, where we interpret $[\rho({\bf{T}}_i)]$ as a cycle---the set of points in $\rho({\bf{T}}_i)$ over which $\rho$ has degree $n$ will appear in $[\rho({\bf{T}}_i)]$ with coefficient $n$. By the identification of $H_*({\bf{T}}^N,{\bf{Z}})$ with $\wedge^*{\bf{Z}}^N$, such a map is determined by its action on $H_1({\bf{T}}^N,{\bf{Z}})$, where it is an integer linear map ${\bf{Z}}^N\to{\bf{Z}}^2$. \section{The coamoeba and zonotope chains of a real line}\label{S:realline} We study the coamoeba ${\mathcal{A}}(\ell)$ of a line $\ell$ in $\P^N$ defined by real equations. Its closure $\overline{{\mathcal{A}}(\ell)}$ is a two-dimensional chain in ${\bf{T}}^N$ whose boundary consists of at most $N{+}1$ one-dimensional subtori parallel to sums of coordinate directions. We describe a piecewise linear two-dimensional chain---the \demph{zonotope chain} of $\ell$---which has the same boundary as the coamoeba, but with opposite orientation. The union of the coamoeba and the zonotope chain forms a cycle whose homology class we compute. The line $\ell$ has a parametrization \[ \Phi\ \colon\ \P^1\ni z\ \longmapsto\ [b_1(z)\,:\, b_2(z)\,:\, \dotsb\,:\,b_{N+1}(z)]\ \in\ \P^N\,, \] where $b_1,\dotsc,b_{N+1}$ are real linear forms with zeroes $\xi_1,\dotsc,\xi_{N+1}\in{\bf{R}}\P^1$. The formulation and statement of our results about the coamoeba of $\ell$ will be with respect to particular orderings of the forms $b_i$, which we now describe. \begin{definition}\label{D:conventions} Suppose that these zeroes are in a weakly increasing cyclic order on ${\bf{R}}\P^1$, \begin{equation}\label{Eq:A} \xi_1\ \leq\ \xi_2\ \leq\ \dotsb\ \leq\ \xi_{N+1}\,. \end{equation} Next, identify $\P^1\smallsetminus\{\xi_{N+1}\}$ with ${\bf{C}}$, so that $\xi_{N+1}$ is the point $\infty$ at infinity, and suppose that the distinct zeroes are \begin{equation}\label{Eq:B} \zeta_1\ <\ \zeta_2\ <\ \dotsb\ <\ \zeta_M\ <\ \zeta_{M{+}1}\ =\ \infty\,. \end{equation} (Note that $M\leq N$.) Let ${\bf{R}}={\bf{R}}\P^1\smallsetminus\{\infty\}$ and consider the forms $b_i$ as affine functions on ${\bf{R}}$. Fix a scaling of these functions so that $b_{N+1}=1$. On the interval $(-\infty,\zeta_1)$ the sign of each function $b_i$ is constant. Define $\defcolor{\sgn_i}\in\{\pm 1\}$ to be this sign. By~\eqref{Eq:A} and~\eqref{Eq:B}, there exist numbers $1=m_1<\dotsb<m_{M+1}<m_{M+2}=N{+}2$ such that $b_i(\zeta_j)=0$ if and only if $i\in[m_j,m_{j+1})$. We further suppose that on each of these intervals $[m_j,m_{j+1})$ the signs $\sgn_i$ are weakly ordered. Specifically, there are integers $n_1,\dotsc,n_{M{+}1}$ with $m_j< n_j \leq m_{j+1}$ such that one of the following holds \begin{eqnarray} \sgn_{m_j}=\sgn_{m_j+1}=\dotsb=\sgn_{n_j-1}=-1 &< &1=\sgn_{n_j}=\dotsb=\sgn_{m_{j+1}-1}\,,\makebox[.1in][l]{\qquad or}\label{Eq:inc}\\ \sgn_{m_j}=\sgn_{m_j+1}=\dotsb=\sgn_{n_j-1}=1 &> &-1=\sgn_{n_j}=\dotsb=\sgn_{m_{j+1}-1}\,,\label{Eq:dec} \end{eqnarray} for $j=1,\dotsc,M{+}1$. If $n_j=m_{j+1}$, then all the signs are the same; otherwise both signs occur. Since $b_{N+1}=1$, either~\eqref{Eq:inc} occurs with $n_{M+1}\leq N{+}1$ or~\eqref{Eq:dec} occurs with $n_{M+1}=N{+}1$. \hfill\includegraphics[height=10pt]{figures/QED.eps} \end{definition} The point $\Arg(b_1(z),\dotsc,b_N(z))\in{\bf{T}}^N$ is constant for $z$ in each interval of ${\bf{R}}^1\smallsetminus\{\zeta_1,\dotsc,\zeta_M\}$. Let $\defcolor{p_1}:=(\arg(\sgn_i)\mid i=1,\dotsc,N)$ be the point coming from the interval $(-\infty,\zeta_1)$, and for each $j=1,\dotsc,M$, let \defcolor{$p_{j+1}$} be the point coming from the interval $(\zeta_j,\zeta_{j+1})$. These $M{+}1$ points $p_1,\dotsc,p_{M+1}$ of ${\bf{T}}^N$ are the vertices of the coamoeba ${\mathcal{A}}(\ell)$ of $\ell$. To understand the rest of the coamoeba, note that when $M\geq 2$ the map $\Arg\circ\Phi$ is injective on $\P^1\smallsetminus{\bf{R}}\P^1$ (see~\cite[\S~2]{NS}). (When $M=1$, $\ell$ is parallel to a sum of coordinate directions and ${\mathcal{A}}(\ell)$ is a translate of the corresponding one-dimensional subtorus of ${\bf{T}}^N$.) It suffices to consider the image of the upper half plane, as the image of the lower half plane is obtained by multiplying by $-1$ (induced by complex conjugation). For the upper half plane, consider $\Arg\circ\Phi(z)$ for $z$ lying on a contour $C$ as shown in Figure~\ref{F:contour} \begin{figure}[htb] \begin{picture}(223,77) \put(0,10){\includegraphics{figures/contour.eps}} \put(40,0){$\zeta_1$} \put(85,0){$\zeta_2$} \put(185,0){$\zeta_M$} \put(228,18){${\bf{R}}$} \put(71,50){$C$} \end{picture} \caption{Contour in upper half plane} \label{F:contour} \end{figure} that contains semicircles of radius $\epsilon$ centered at each root $\zeta_j$ and a semicircle of radius $1/\epsilon$ centered at 0, but otherwise lies along the real axis, for $\epsilon$ a sufficiently small positive number. As $z$ moves along $C$, $\Arg\circ\Phi(z)$ takes on values $p_1,\dotsc,p_{M+1}$, for $z\in C\cap{\bf{R}}$. On the semicircular arc around $\zeta_j$, it traces a curve from $p_{j}$ to $p_{j+1}$ in which nearly every component is constant, except for those $i$ where $b_i(\zeta_j)=0$, each of which decreases by $\pi$. In the limit as $\epsilon\to 0$, this becomes the line segment between $p_{j}$ and $p_{j+1}$ with direction $-{\bf f}_j$, where \[ \defcolor{{\bf f}_j}\ :=\ \sum_{i\colon b_i(\xi_j)=0} {\bf e}_i \ =\ \sum_{i=m_j}^{m_{j+1}-1} {\bf e}_i\,, \] and where we set ${\bf e}_{N+1}:=-({\bf e}_1+\dotsb+{\bf e}_N)$. This is because we are really working in the torus for $\P^N$, which is the quotient ${\bf{T}}^{N+1}/\Delta({\bf{T}})$ of ${\bf{T}}^{N+1}$ modulo the diagonal torus, and ${\bf e}_i\in{\bf{T}}^{N+1}/\Delta({\bf{T}})$ is the image of the standard basis element in ${\bf{T}}^{N+1}$. Thus ${\bf e}_1+\dotsb+{\bf e}_{N+1}=0$. Along the arc near infinity, $\Arg\circ\Phi(z)$ approaches the line segment between $p_{M+1}$ and $p_1$ which has direction $-{\bf f}_{M+1}$, where \begin{equation}\label{Eq:bbf_M+1} {\bf f}_{M{+}1}\ =\ - \sum_{i\colon b_i(\infty)\neq0} {\bf e}_j\ =\ -({\bf f}_1+\dotsb+{\bf f}_M)\,. \end{equation} This polygonal path connecting $p_1,\dotsc,p_{M+1}$ in cyclic order forms the boundary of the image of the upper half plane under $\Arg\circ\Phi$, which is a two-dimensional membrane in ${\bf{T}}^N$. The boundary of the image of the lower half plane is also a piecewise linear path connecting $p_1,\dotsc,p_{M+1}$ in cyclic order, but the edge directions are ${\bf f}_1,\dotsc,{\bf f}_{M{+}1}$. \begin{example}\label{Ex:tc} Let $N=3$ and suppose that the affine functions $b_i$ are $z$, $1{-}2z$, $z{-}2$, and $1$. Then $M=N$, $\xi_i=\zeta_i$, $\zeta_1=0$, $\zeta_1=1/2$, $\zeta_2=2$, and ${\bf f}_i={\bf e}_i$. The vertices of ${\mathcal{A}}(\ell)$ are \[ p_1\ =\ (\pi,0,\pi)\,,\quad p_2\ =\ (0,0,\pi)\,,\quad p_3\ =\ (0,-\pi,\pi)\,,\quad \mbox{and}\quad p_4\ =\ (0,-\pi,0)\,. \] Figure~\ref{F:linecoamoeba} shows two views of ${\mathcal{A}}(\ell)$ in the fundamental domain $[-\pi,\pi]^3\subset{\bf{R}}^3$ of ${\bf{T}}^3$, where the opposite faces of the cube are identified to form ${\bf{T}}^3$. \hfill$\includegraphics[height=10pt]{figures/QED.eps}$ \end{example} \begin{figure}[htb] \begin{picture}(133,115) \put(0,0){\includegraphics[height=115pt]{figures/three_1.eps}} \put(35, 78){$p_1$} \put(53,104){$p_2$} \put(13,110){$p_3$} \put(10,62){$p_4$} \put(75,33){$p_1$} \put( 55, 7){$p_2$} \put(97, 2){$p_3$} \put( 97,50){$p_4$} \end{picture} \qquad\qquad \begin{picture}(142,115)(-23,0) \put(0,0){\includegraphics[height=115pt]{figures/three_2.eps}} \put( 43,104){$p_1$} \put( 32, 84){$p_2$} \put(110, 83){$p_3$} \put(108,86){\vector(-1,0){20}} \put(110, 50){$p_4$} \put(108,53){\vector(-1,0){20}} \put( 30, -2){$p_1$} \put( 36, 31){$p_2$} \put(-23, 28){$p_3$} \put(-12,30){\vector(1,0){20}} \put(-23, 61){$p_4$} \put(-12,63){\vector(1,0){20}} \end{picture} \caption{Two views of ${\mathcal{A}}(\ell)$} \label{F:linecoamoeba} \end{figure} \begin{example}\label{Ex:CoArepeat} We consider three examples when $N=3$ in which the affine functions have repeated zeroes. For the first, suppose that the affine functions $b_i$ are $-1{-}z,-1{-}z,2z$, and $2$. These have zeroes $-1\leq-1<0<\infty$ and the vertices of the coamoeba ${\mathcal{A}}(\ell)$ are \[ (0,0,\pi)\,,\quad (-\pi,-\pi,\pi)\,,\quad \mbox{and}\quad (-\pi,-\pi,0)\,. \] So ${\mathcal{A}}(\ell)$ consists of two triangles with edges parallel to ${\bf e}_1{+}{\bf e}_2$, ${\bf e}_3$, and ${\bf e}_1{+}{\bf e}_2{+}{\bf e}_3$. It lies in the plane $\theta_1=\theta_2$. For a second example, suppose that the affine functions $b_i$ are $\frac{1}{2}+z,\frac{1}{2}-z,-2$, and $1$. These have zeroes $-1,1,\infty$, and $\infty$. The vertices of the coamoeba ${\mathcal{A}}(\ell)$ are \[ (\pi,0,\pi)\,,\quad (0,0,\pi)\,,\quad \mbox{and}\quad (0,-\pi,\pi)\,. \] So ${\mathcal{A}}(\ell)$ consists of two triangles with edges parallel to ${\bf e}_1$, ${\bf e}_2$, and ${\bf e}_1{+}{\bf e}_2$. It lies in the plane $\theta_3=\pi$. Finally, suppose that the affine functions $b_i$ are $-z,1-z,2z-2$, and $1$. These have zeroes $0,1,1$, and $\infty$. The vertices of the coamoeba ${\mathcal{A}}(\ell)$ are \[ (0,0,\pi)\,,\quad (-\pi,0,\pi)\,,\quad \mbox{and}\quad (-\pi,-\pi,0)\,. \] So ${\mathcal{A}}(\ell)$ consists of two triangles with edges parallel to ${\bf e}_1$, ${\bf e}_2{+}{\bf e}_3$, and ${\bf e}_1{+}{\bf e}_2{+}{\bf e}_3$. It lies in the plane $\theta_3=\theta_2+\pi$. We display all three coamoebas in Figure~\ref{F:repeat}. \hfill$\includegraphics[height=10pt]{figures/QED.eps}$ \end{example} The \demph{coamoeba chain $\overline{{\mathcal{A}}(\ell)}$} of $\ell$ is the closure of the coamoeba of $\ell$ in which the image of each half plane (under $\Arg\circ\Phi(\cdot)$) is oriented so that its boundary is an oriented polygonal path connecting $p_1,\dotsc,p_{M+1},p_1$. On the upper half plane this agrees with the orientation induced by the parametrization $\P^1\smallsetminus{\bf{R}}\P^1\to{\mathcal{A}}(\ell)$, but it has the opposite orientation on the lower half plane. The boundary of $\overline{{\mathcal{A}}(\ell)}$ consists of $M{+}1$ circles in which $p_j$ and $p_{j+1}$ are antipodal points on the $j$th circle and both semicircles (each is the boundary of the image of a half plane) are oriented to point from $p_j$ to $p_{j+1}$. This coamoeba chain is not a closed chain, as it has nonempty oriented boundary, but there is a natural zonotope chain $Z(\ell)$ such that $\overline{{\mathcal{A}}(\ell)}+Z(\ell)$ is closed. Intuitively, $Z(\ell)$ is the cone over the boundary of $\overline{{\mathcal{A}}(\ell)}$ with vertex the origin $\defcolor{{\bf 0}}:=(0,\dotsc,0)$. Unfortunately, there is no notion of a cone in ${\bf{T}}^N$ and the zonotope chain may be more than just this cone. We instead define a chain in ${\bf{R}}^N$ as the cone over an oriented polygon $P(\ell)$ with vertex the origin and set $Z(\ell)$ to be the image of this chain in ${\bf{T}}^N$. \begin{definition}\label{Def:Zonotope_chain} Recall that the affine functions $b_1,\dotsc,b_N,b_{N{+}1}=1$ are ordered in the following way. Their zeroes are $\zeta_1<\dotsb<\zeta_M<\zeta_{M{+}1}=\infty$ and there are integers $1=m_1<\dotsb<m_{M+1}\leq N{+}1$ and $n_1,\dotsc,n_{M{+}1}$ with $m_j< n_j\leq m_{j+1}$ such that one of~\eqref{Eq:inc} or~\eqref{Eq:dec} holds, where $\sgn_i$ is the sign of $b_i$ on $(-\infty,\zeta_1)$. We had defined ${\bf f}_j:=\sum_{i=m_j}^{m_{j+1}-1}{\bf e}_i$. We will need the following vectors \[ \defcolor{{\bf g}_j}\ :=\ \sum_{i=m_j}^{n_j-1}{\bf e}_i \qquad\mbox{and}\qquad \defcolor{{\bf h}_j}\ :=\ \sum_{i=m_j}^{m_{j+1}-1}\sgn_i {\bf e}_i\ =\ \sgn_{m_j}(2{\bf g}_j-{\bf f}_j)\ \,. \] We first define a sequence of points $\widetilde{p}\hspace{1.2pt}_1,{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_1,\dotsc,\widetilde{p}\hspace{1.2pt}_{2M+2},{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_{2M+2}\in(\pi{\bf{Z}})^N$ with the property that $\widetilde{p}\hspace{1.2pt}_i,{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_i,\widetilde{p}\hspace{1.2pt}_{M{+}1{+}i},$ and ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_{M{+}1{+}i}$ all map to $p_i\in{\bf{T}}^N$. To begin, set $\widetilde{p}\hspace{1.2pt}_1$ to be the unique point in $\{0,\pi\}^N\subset{\bf{R}}^N$ which maps to $p_1\in{\bf{T}}^N$, \begin{equation}\label{Eq:ptilde} \widetilde{p}\hspace{1.2pt}_{1,i}\ =\ \arg(\sgn_i)\ =\ \left\{ \begin{array}{rcl}\pi&\ &\mbox{if }\sgn_i=-1\\ 0&&\mbox{if }\sgn_i=1\end{array}\right.\ . \end{equation} For each $j=1,\dotsc,M{+}1$, set $\defcolor{\widetilde{p}\hspace{1.2pt}_{j+1}}:=\widetilde{p}\hspace{1.2pt}_j+\pi{\bf h}_j$. Since ${\bf h}_j=\sgn_{m_j}(2{\bf g}_j-{\bf f}_j)$, we have that $\widetilde{p}\hspace{1.2pt}_{j+1}$ maps to $p_{j+1}$, as $p_{j+1}=p_j-\pi{\bf f}_j\mod (2\pi{\bf{Z}})^N$. For the remainder of the points, if $n_j<m_{j+1}$, so that both signs occur, set $\defcolor{{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j} := \widetilde{p}\hspace{1.2pt}_j+2\pi\sgn_{m_j}{\bf g}_j$, and otherwise set $\defcolor{{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j} := \widetilde{p}\hspace{1.2pt}_j$. Observe that ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j$ maps to $p_j$ and that in every case, $\widetilde{p}\hspace{1.2pt}_{j+1}={\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j-\pi\sgn_{m_j}{\bf f}_j$. We claim that $\widetilde{p}\hspace{1.2pt}_{M{+}2}=-\widetilde{p}\hspace{1.2pt}_1$. Since $\widetilde{p}\hspace{1.2pt}_{M{+}2}=\widetilde{p}\hspace{1.2pt}_1+\pi({\bf h}_1+\dotsb+{\bf h}_{M{+}1})$, we need to show that $\pi({\bf h}_1+\dotsb+{\bf h}_{M{+}1})=-2\widetilde{p}\hspace{1.2pt}_1$. By definition, \[ {\bf h}_1+\dotsb+{\bf h}_{M+1}\ =\ \sum_{i=1}^{N+1} \sgn_i{\bf e}_i\,. \] We have $\sgn_{N+1}=1$ as $b_{N{+}1}=1$. Since we defined ${\bf e}_{N+1}$ to be $-({\bf e}_1+\dotsb+{\bf e}_N)$, we see that \[ {\bf h}_1+\dotsb+{\bf h}_{M+1}\ =\ \sum_{i=1}^N (\sgn_i-1){\bf e}_i\,. \] The $i$th component of this sum is $-2$ if $\sgn_i=-1$ and $0$ if $\sgn_i=1$. Since $\widetilde{p}\hspace{1.2pt}_{1,i}=\arg(\sgn_i)$, this proves the claim. Finally, for each $M{+}2\leq j\leq 2M{+}2$, set \[ \defcolor{\widetilde{p}\hspace{1.2pt}_j}\ :=\ -\widetilde{p}\hspace{1.2pt}_{j-(M{+}1)}\qquad\mbox{and}\qquad \defcolor{{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j}\ :=\ -\widetilde{p}\hspace{1.2pt}_{j-(M{+}1)}\,, \] and let $P(\ell)$ be the cyclically oriented path obtained by connecting \[ {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_{2M+2}\,,\,\widetilde{p}\hspace{1.2pt}_{2M+2}\,,\, {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_{2M+1}\,,\,\widetilde{p}\hspace{1.2pt}_{2M+1}\,,\, \dotsc, {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_2\,,\,\widetilde{p}\hspace{1.2pt}_2\,,\,{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_1\,,\, \widetilde{p}\hspace{1.2pt}_1 \] in cyclic order. The cone over $P(\ell)$ with vertex the origin is the union of possibly degenerate triangles of the form \[ \conv({\bf 0},\widetilde{p}\hspace{1.2pt}_{i+1},{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_{i}) \qquad\mbox{and}\qquad \conv({\bf 0},{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_{i},\widetilde{p}\hspace{1.2pt}_{i}) \qquad\mbox{for}\qquad i=2M{+}2,\dotsc,2,1\,, \] where $\widetilde{p}\hspace{1.2pt}_{2M+3}:=\widetilde{p}\hspace{1.2pt}_1$. Each triangle is oriented so its three vertices occur in positive order along its boundary. If a point $\widetilde{p}\hspace{1.2pt}_i$ or ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_i$ is ${\bf 0}$, then the triangles involving it degenerate into line segments, as do triangles $\conv({\bf 0},{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_{i},\widetilde{p}\hspace{1.2pt}_{i})$ when ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_{i}=\widetilde{p}\hspace{1.2pt}_{i}$. Let $\widetilde{Z(\ell)}$ be the union of these oriented triangles, which is a chain in ${\bf{R}}^N$. Define the \demph{zonotope chain $Z(\ell)$} to be the image in ${\bf{T}}^N$ of $\widetilde{Z(\ell)}$. \hfill$\includegraphics[height=10pt]{figures/QED.eps}$ \end{definition} \begin{example}\label{Ex:CoZrepeat} Figure~\ref{F:ZC} shows two views of the zonotope chain with the coamoeba chain of Figure~\ref{F:linecoamoeba}. \begin{figure}[htb] \begin{picture}(153,150) \put(0,0){\includegraphics[height=150pt]{figures/ZC_1.eps}} \put( 44,101){$\widetilde{p}\hspace{1.2pt}_1$} \put(69,134){$\widetilde{p}\hspace{1.2pt}_2$} \put(120,126){$\widetilde{p}\hspace{1.2pt}_3$} \put(127,64){$\widetilde{p}\hspace{1.2pt}_4$} \put(101, 46){$\widetilde{p}\hspace{1.2pt}_5$} \put(73,10){$\widetilde{p}\hspace{1.2pt}_6$} \put( 18, 32){$\widetilde{p}\hspace{1.2pt}_7$} \put(18,82){$\widetilde{p}\hspace{1.2pt}_8$} \put(70.5,60){${\bf 0}$} \end{picture} \qquad\qquad \begin{picture}(170,150)(-20,0) \put(0,0){\includegraphics[height=150pt]{figures/ZC_2.eps}} \put( 59,135){$\widetilde{p}\hspace{1.2pt}_1$} \put( 46,110){$\widetilde{p}\hspace{1.2pt}_2$} \put( 0,126){$\widetilde{p}\hspace{1.2pt}_3$} \put(-20, 78){$\widetilde{p}\hspace{1.2pt}_4$} \put(-8,82){\vector(1,0){18}} \put( 55, 9){$\widetilde{p}\hspace{1.2pt}_5$} \put( 65, 35){$\widetilde{p}\hspace{1.2pt}_6$} \put(113, 21){$\widetilde{p}\hspace{1.2pt}_7$} \put(133, 66){$\widetilde{p}\hspace{1.2pt}_8$} \put(131,69){\vector(-1,0){18}} \put(53,66){${\bf 0}$} \end{picture} \caption{Two views of the coamoeba and zonotope chains} \label{F:ZC} \end{figure} Now consider the zonotope chains for the three lines of Example~\ref{Ex:CoArepeat}. When $\ell$ is defined by $z\mapsto[-1-z,-1-z,2z,2]$, the points $\widetilde{p}\hspace{1.2pt}_1,\dotsc,{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_6$ (omitting repeated points) are \[ (0,0, \pi)\,,\ ( \pi, \pi, \pi)\,,\ ( \pi, \pi,0)\,,\ (0,0,-\pi)\,,\ (-\pi,-\pi,-\pi)\,,\quad\mbox{ and }\quad (-\pi,-\pi,0)\,. \] We display the coamoeba chain and the zonotope chain of $\ell$ at the left of Figure~\ref{F:repeat}. \begin{figure}[htb] \begin{picture}(102,150)(0,-20) \put(0,0){\includegraphics[height=110pt]{figures/equal.eps}} \put( 51, 95){$\widetilde{p}\hspace{1.2pt}_1$} \put(79, 72){$\widetilde{p}\hspace{1.2pt}_2$} \put( 79, 35){$\widetilde{p}\hspace{1.2pt}_3$} \put( 39, 10){$\widetilde{p}\hspace{1.2pt}_4$} \put( 11, 35){$\widetilde{p}\hspace{1.2pt}_5$} \put( 11, 70){$\widetilde{p}\hspace{1.2pt}_6$} \thicklines \put( 12,118){\White{\vector(2,-3){21}}} \put( 90, -8){\White{\vector(-2,3){21}}} \thinlines \put(0 ,121){${\mathcal{A}}(\ell)$}\put( 12,118){\vector(2,-3){20}} \put(80,-18){${\mathcal{A}}(\ell)$}\put( 90, -8){\vector(-2,3){20}} \end{picture} \qquad \begin{picture}(116,150)(-5,-20) \put(0,0){\includegraphics[height=110pt]{figures/parallelI.eps}} \put(12, 75){$\widetilde{p}\hspace{1.2pt}_1$} \put( 50,97){$\widetilde{p}\hspace{1.2pt}_2$} \put(79, 71){$\widetilde{p}\hspace{1.2pt}_3$} \put( 72,-5){${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_3$} \put(88, 30){$\widetilde{p}\hspace{1.2pt}_4$} \put( 6, 4){$\widetilde{p}\hspace{1.2pt}_5$} \put( 25,18){$\widetilde{p}\hspace{1.2pt}_6$} \put( 25,107){${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_6$} \put(-5,122){${\mathcal{A}}(\ell)$} \put( 28, -19){${\mathcal{A}}(\ell)$} \thicklines \put(16, 9){\White{\vector(4,1){36}}} \put(8,118){\White{\vector(1,-1){28}}} \put(42,-9){\White{\vector(1, 1){28}}} \thinlines \put(16,9){\vector(4,1){35}} \put(8,118){\vector(1,-1){27}} \put(42,-9){\vector(1, 1){27}} \end{picture} \qquad \begin{picture}(181,150)(-5,-20) \put(0,0){\includegraphics[height=110pt]{figures/parallelII.eps}} \put( 83, 97){$\widetilde{p}\hspace{1.2pt}_1$} \put(104, 89){$\widetilde{p}\hspace{1.2pt}_2$} \put(166, 101){${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_2$} \put(144, 47){$\widetilde{p}\hspace{1.2pt}_3$} \put( 81, 7){$\widetilde{p}\hspace{1.2pt}_4$} \put( 60, 19){$\widetilde{p}\hspace{1.2pt}_5$} \put( -5, 3){${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_5$} \put( 20, 64){$\widetilde{p}\hspace{1.2pt}_6$} \put( 0,120){${\mathcal{A}}(\ell)$} \put(145, -10){${\mathcal{A}}(\ell)$} \thicklines \put( 18,114){\White{\vector(2,-1){49}}} \put(153, 0){\White{\vector(-2,1){49}}} \thinlines \put( 18,114){\vector(2,-1){48}} \put(153, 0){\vector(-2,1){48}} \end{picture} \caption{Coamoeba and zonotope chains} \label{F:repeat} \end{figure} When $\ell$ is defined by $z\mapsto [\frac{1}{2}+z,\frac{1}{2}-z,-2,1]$, the points $\widetilde{p}\hspace{1.2pt}_1,\dotsc,{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_6$ are \begin{eqnarray*} &\widetilde{p}\hspace{1.2pt}_1\ =\ (\pi,0,\pi)\,,\ \widetilde{p}\hspace{1.2pt}_2\ =\ (0,0,\pi)\,,\ \widetilde{p}\hspace{1.2pt}_3\ =\ (0,\pi,\pi)\,,\ \Magenta{{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_3\ =\ (0,\pi,-\pi)}\,,&\\ &\widetilde{p}\hspace{1.2pt}_4\ =\ (-\pi,0,-\pi)\,,\ \widetilde{p}\hspace{1.2pt}_5\ =\ (0,0,-\pi)\,,\ \widetilde{p}\hspace{1.2pt}_6\ =\ (0,-\pi,-\pi)\,,\ \Magenta{\widetilde{p}\hspace{1.2pt}_6'\ =\ (0,-\pi,\pi)}\,.& \end{eqnarray*} We display the coamoeba and zonotope chains of $\ell$ in the middle of Figure~\ref{F:repeat}. When $\ell$ is defined by $z\mapsto [-z,1-z,2z-2,1]$, the points $\widetilde{p}\hspace{1.2pt}_1,\dotsc,{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_6$ are \begin{eqnarray*} &\widetilde{p}\hspace{1.2pt}_1\ =\ (0,0,\pi)\,,\ \widetilde{p}\hspace{1.2pt}_2\ =\ (\pi,0,\pi)\,,\ \Magenta{{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_2\ =\ (\pi,2\pi,\pi)}\,,\ \widetilde{p}\hspace{1.2pt}_3\ =\ (\pi,\pi,0)\,,&\\ &\widetilde{p}\hspace{1.2pt}_4\ =\ (0,0,-\pi)\,,\ \widetilde{p}\hspace{1.2pt}_5\ =\ (-\pi,0,-\pi)\,,\ \Magenta{{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_5\ =\ (-\pi,-2\pi,-\pi)}\,,\ \widetilde{p}\hspace{1.2pt}_6\ =\ (-\pi,-\pi,0)\,.& \end{eqnarray*} We display the coamoeba and zonotope chains of $\ell$ on the right of Figure~\ref{F:repeat}. \hfill$\includegraphics[height=10pt]{figures/QED.eps}$ \end{example} We state the main result of this section. \begin{theorem}\label{Th:homology} The sum, $\overline{{\mathcal{A}}(\ell)}+Z(\ell)$, of the coamoeba chain and the zonotope chain forms a cycle in ${\bf{T}}^N$ whose homology class is \[ [\overline{{\mathcal{A}}(\ell)}+Z(\ell)]\ =\ \sum_{\substack{1\leq i<j\leq N\\(\widetilde{p}\hspace{1.2pt}_{1,i},\widetilde{p}\hspace{1.2pt}_{1,j})=(0,\pi)}} {\bf e}_i\wedge{\bf e}_j\,. \] \end{theorem} \begin{example}\label{Ex:homology_class} For the line of Example~\ref{Ex:tc}, $\widetilde{p}\hspace{1.2pt}_1=(\pi,0,\pi)$, and the only entries $i<j$ with 0 at $i$ and $\pi$ at $j$ are $i=2$ and $j=3$, and so \[ [\overline{{\mathcal{A}}(\ell)}+Z(\ell)]\ =\ {\bf e}_2\wedge{\bf e}_3\,. \] For the first line of Example~\ref{Ex:CoArepeat}, $\widetilde{p}\hspace{1.2pt}_1=(0,0,\pi)$, and so \[ [\overline{{\mathcal{A}}(\ell)}+Z(\ell)]\ =\ {\bf e}_1\wedge{\bf e}_3 + {\bf e}_2\wedge{\bf e}_3\,. \] For the second line of Example~\ref{Ex:CoArepeat}, $\widetilde{p}\hspace{1.2pt}_1=(\pi,0,\pi)$, so that $[\overline{{\mathcal{A}}(\ell)}+Z(\ell)]={\bf e}_2\wedge{\bf e}_3$. For the third line of Example~\ref{Ex:CoArepeat}, $\widetilde{p}\hspace{1.2pt}_1=(0,0,\pi)$, and $[\overline{{\mathcal{A}}(\ell)}+Z(\ell)]={\bf e}_1\wedge{\bf e}_3 + {\bf e}_2\wedge{\bf e}_3$. These homology classes are apparent from Figures~\ref{F:ZC} and~\ref{F:repeat}. \hfill$\includegraphics[height=10pt]{figures/QED.eps}$ \end{example} \begin{example}\label{Ex:plane} Our proof of Theorem~\ref{Th:homology} rests on the case of $N=2$. Suppose first that $M=2$. Up to positive rescaling and translation in the domain ${\bf{R}}\P^1$, there are four lines. \[ \begin{picture}(65,62)(-7.5,-12) \put(0,0){\includegraphics{figures/lineI.eps}} \put(-4.5,-12){$[z:z{-}1:1]$} \end{picture} \qquad \begin{picture}(65,62)(-7.5,-12) \put(0,0){\includegraphics{figures/lineIV.eps}} \put(-4.5,-12){$[z:1{-}z:1]$} \end{picture} \qquad \begin{picture}(65,62)(-7.5,-12) \put(0,0){\includegraphics{figures/lineIII.eps}} \put(-8.5,-12){$[-z:1{-}z:1]$} \end{picture} \qquad \begin{picture}(65,62)(-7.5,-12) \put(0,0){\includegraphics{figures/lineII.eps}} \put(-8.5,-12){$[-z:z{-}1:1]$} \end{picture} \] For these, the initial point $p_1$ is $(\pi,\pi)$, $(\pi,0)$, $(0,0)$, and $(0,\pi)$, respectively. The four coamoeba chains are, in the fundamental domain $[-\pi,\pi]^2$, \[ \includegraphics{figures/2dcoamoebaI.eps} \qquad \includegraphics{figures/2dcoamoebaIV.eps} \qquad \includegraphics{figures/2dcoamoebaIII.eps} \qquad \includegraphics{figures/2dcoamoebaII.eps} \] and the corresponding zonotope chains are as follows. \[ \includegraphics{figures/2dzonotopeI.eps} \qquad \includegraphics{figures/2dzonotopeIV.eps} \qquad \includegraphics{figures/2dzonotopeIII.eps} \qquad \includegraphics{figures/2dzonotopeII.eps} \] For each, the sum $\overline{{\mathcal{A}}(\ell)}+Z(\ell)$ of chains is a cycle. This cycle is homologous to zero for the first three, and it forms the fundamental cycle ${\bf e}_1\wedge{\bf e}_2$ of ${\bf{T}}^2$ for the fourth. Now suppose that $M=1$. We may assume that $\xi_1=0$. Up to positive rescaling there are eight possibilities for the parametrization of $\ell$, \begin{eqnarray*} &[-z:-z:1]\,,\ [ z: z:1]\,,\ [-z: 1:1]\,,\ [ z: 1:1]\,,&\\ &[ z:-z:1]\,,\ [ z:-1:1]\,,\ [-z:-1:1]\,,\ [-z: z:1]\,.& \end{eqnarray*} For all of these, the coamoeba is one-dimensional. In the first four, the zonotope chain is one-dimensional. Table~\ref{T:paths} gives the parametrization, the vertices of the coamoeba of the upper half plane, and the path $P(\ell)=\widetilde{p}\hspace{1.2pt}_4,\widetilde{p}\hspace{1.2pt}_3,\widetilde{p}\hspace{1.2pt}_2,\widetilde{p}\hspace{1.2pt}_1$ for these four. \begin{table}[htb] \caption{Coamoeba and zonotope chains.}\label{T:paths} \begin{tabular}{|r||l|l|}\hline \multicolumn{1}{|c||}{$\ell$} &\multicolumn{1}{|c|}{${\mathcal{A}}(\ell)$}&\multicolumn{1}{|c|}{$P(\ell)$}\\\hline $[-z:-z:1]$&$(0,0)\,,\,(-\pi,-\pi)$&$(-\pi,-\pi)\,,\,(0,0)\,,\,(\pi,\pi)\,,\,(0,0)$\\\hline $[ z: z:1]$&$(\pi,\pi)\,,\,(0,0) $&$(0,0)\,,\,(-\pi,-\pi)\,,\,(0,0)\,,\,(\pi,\pi)$\\\hline $[-z: 1:1]$&$(0,0)\,,\,(-\pi,0) $&$(-\pi,0)\,,\,(0,0)\,,\,(\pi,0)\,,\,(0,0)$\\\hline $[ z: 1:1]$&$(\pi,0)\,,\,(0,0) $&$(0,0)\,,\,(-\pi,0)\,,\,(0,0)\,,\,(\pi,0)$\\\hline \end{tabular} \end{table} The remaining parametrizations are more interesting. When $\ell$ is given by $z\mapsto[z:-z:1]$, we have $p_1=(\pi,0)$ and $p_2=(0,-\pi)$, and $P(\ell)$ is \[ \widetilde{p}\hspace{1.2pt}_4=(0,-\pi)\,,\ {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_3=(\pi,0)\,,\ \widetilde{p}\hspace{1.2pt}_3=(-\pi,0)\,,\ \widetilde{p}\hspace{1.2pt}_2=(0,\pi)\,,\ {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_1=(-\pi,0)\,,\quad\mbox{and}\quad\widetilde{p}\hspace{1.2pt}_1=(\pi,0)\,, \] and the zonotope chain is shown on the left in Figure~\ref{F:M=1}. The path $\widetilde{p}\hspace{1.2pt}_4{-}{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_3{-}\widetilde{p}\hspace{1.2pt}_3{-}\widetilde{p}\hspace{1.2pt}_2{-}{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_1{-}\widetilde{p}\hspace{1.2pt}_1{-}\widetilde{p}\hspace{1.2pt}_4$ zig-zags over itself, once in each direction, and consequently each triangle is covered twice, once with each orientation, and therefore $[Z(\ell)]=0$ in homology. \begin{figure}[htb] \begin{picture}(90,89)(-23,-24) \put(0,0){\includegraphics{figures/2dzonotopeV.eps}} \put( 42,33){$\widetilde{p}\hspace{1.2pt}_1={\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_3$} \put(-23,15){${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_1=\widetilde{p}\hspace{1.2pt}_3$} \put(23,-9){$\widetilde{p}\hspace{1.2pt}_4$} \put(23,58){$\widetilde{p}\hspace{1.2pt}_2$} \put(0,-24){$[ z:-z:1]$} \end{picture} \qquad \begin{picture}(81,89)(-15,-24) \put(0,0){\includegraphics{figures/2dzonotopeVI.eps}} \put(-11,-7){$\widetilde{p}\hspace{1.2pt}_3$} \put(57,52){$\widetilde{p}\hspace{1.2pt}_1$} \put(21,-9){${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_2{\,=\,}\widetilde{p}\hspace{1.2pt}_4$} \put(-2,58){$\widetilde{p}\hspace{1.2pt}_2{\,=\,}{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_4$} \put(0,-24){$[ z:-1:1]$} \end{picture} \qquad \begin{picture}(81,89)(-15,-24) \put(0,0){\includegraphics{figures/2dzonotopeVII.eps}} \put(-11,55){${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_4$} \put(-11,-7){$\widetilde{p}\hspace{1.2pt}_4$} \put(22,-9){$\widetilde{p}\hspace{1.2pt}_3$} \put(57,-7){${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_2$} \put(57,52){$\widetilde{p}\hspace{1.2pt}_2$} \put(22,58){$\widetilde{p}\hspace{1.2pt}_1$} \put(0,-24){$[-z:-1:1]$} \end{picture} \qquad \begin{picture}(120,89)(-32,-24) \put(0,0){\includegraphics{figures/2dzonotopeVIII.eps}} \put(-11,30){$\widetilde{p}\hspace{1.2pt}_4$} \put(-36,-3){${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_3$} \put(23,-9){$\widetilde{p}\hspace{1.2pt}_3$} \put(56,19){$\widetilde{p}\hspace{1.2pt}_2$} \put(81,52){${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_1$} \put(22,58){$\widetilde{p}\hspace{1.2pt}_1$} \put(0,-24){$[-z: z:1]$} \end{picture} \caption{Four more zonotope chains.}\label{F:M=1} \end{figure} When $\ell$ is given by $[z:-1:1]$, we have $p_1=(\pi,\pi)$ and $p_2=(0,\pi)$, and $P(\ell)$ is \[ {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_4=(0,\pi)\,,\ \widetilde{p}\hspace{1.2pt}_4=(0,-\pi)\,,\ \widetilde{p}\hspace{1.2pt}_3=(-\pi,-\pi)\,, \ {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_2=(0,-\pi)\,,\ \widetilde{p}\hspace{1.2pt}_2=(0,\pi)\,,\quad\mbox{and}\quad\widetilde{p}\hspace{1.2pt}_1=(\pi,\pi)\,, \] and the zonotope chain is shown on the left center of Figure~\ref{F:M=1}. As before, each triangle is covered twice, once with each orientation, and therefore $[Z(\ell)]=0$ in homology. When $\ell$ is given by $[-z:-1:1]$, we have $p_1=(0,\pi)$ and $p_2=(-\pi,\pi)$, and $P(\ell)$ is \[ {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_4=(-\pi,\pi)\,,\ \widetilde{p}\hspace{1.2pt}_4=(-\pi,-\pi)\,,\ \widetilde{p}\hspace{1.2pt}_3=(0,-\pi)\,, {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_2=(\pi,-\pi)\,,\ \widetilde{p}\hspace{1.2pt}_2=(\pi,\pi)\,,\quad\mbox{and}\quad \widetilde{p}\hspace{1.2pt}_1=(0,\pi)\,, \] and the zonotope chain is shown on the right center of Figure~\ref{F:M=1}. The triangles $\conv({\bf 0},\widetilde{p}\hspace{1.2pt}_2,{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_2)$ and $\conv({\bf 0},\widetilde{p}\hspace{1.2pt}_4,{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_4)$ are shaded differently. The zonotope chain is equal to the fundamental cycle of ${\bf{T}}^2$, with the standard positive orientation. Thus $[Z(\ell)]={\bf e}_1\wedge{\bf e}_2$ in homology. Finally, when $\ell$ is given by $[-z: z:1]$, we have $p_1=(0,\pi)$ and $p_2=(-\pi,0)$, and $P(\ell)$ is \[ \widetilde{p}\hspace{1.2pt}_4=(-\pi,0)\,,\ {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_3=(-2\pi,-\pi)\,,\ \widetilde{p}\hspace{1.2pt}_3=(0,-\pi)\,, \widetilde{p}\hspace{1.2pt}_2=( \pi,0)\,,\ {\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_1=( 2\pi, \pi)\,,\quad\mbox{and}\quad \widetilde{p}\hspace{1.2pt}_1=(0,\pi)\,, \] and the zonotope chain is shown on the right of Figure~\ref{F:M=1}. Again, $[Z(\ell)]=[{\bf{T}}^2]$. Observe that ${\mathcal{A}}(\ell)+Z(\ell)$ forms a cycle which is homologous to zero unless $\widetilde{p}\hspace{1.2pt}_1=(0,\pi)$, in which case it equals the fundamental cycle ${\bf e}_1\wedge{\bf e}_2$ of ${\bf{T}}^2$. \hfill$\includegraphics[height=10pt]{figures/QED.eps}$ \end{example} \begin{proof}[Proof of Theorem~$\ref{Th:homology}$] We show that the two chains $\overline{{\mathcal{A}}(\ell)}$ and $Z(\ell)$ have the same boundary, but with opposite orientation, which implies that their sum is a cycle. We observed that the boundary of ${\mathcal{A}}(\ell)$ lies along the $M{+}1$ circles in which the $j$th contains $p_j$ and $p_{j+1}$ (with $p_{M+2}=p_1$) and has direction parallel to ${\bf f}_j$. On this $j$th circle the boundary of ${\mathcal{A}}(\ell)$ consists of the two semicircles oriented from $p_j$ to $p_{j+1}$. There are two types of edges forming the boundary of the zonotope cycle $Z(\ell)$. The first comes from the edges of $P(\ell)$ with direction $\pm{\bf f}_j$ connecting $\widetilde{p}\hspace{1.2pt}_{j+1}$ to ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j$ and $\widetilde{p}\hspace{1.2pt}_{M+1+j+1}$ to ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_{M+1+j}$, and the second comes from edges connecting ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j$ to $\widetilde{p}\hspace{1.2pt}_j$, when ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j\neq\widetilde{p}\hspace{1.2pt}_j$. The first type of edge gives a part of the boundary of $Z(\ell)$ which is equal to the boundary of ${\mathcal{A}}(\ell)$, but with opposite orientation. (The edges point from $p_{j+1}$ to $p_j$.) The edges of the second type come in pairs which cancel each other. Indeed, when $\widetilde{p}\hspace{1.2pt}_j\neq{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j$, then the edge from ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j$ to $\widetilde{p}\hspace{1.2pt}_j$ is the directed circle connecting $p_j$ with itself and having direction $\pm{\bf g}_j$, which is equal to, but opposite from, the edge connecting ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_{M+1+j}$ to $\widetilde{p}\hspace{1.2pt}_{M+1+j}$. Thus $\overline{{\mathcal{A}}(\ell)}+Z(\ell)$ forms a cycle in homology. We determine the homology class $[\overline{{\mathcal{A}}(\ell)}+Z(\ell)]$ by computing its pushforward to each two-dimensional coordinate projection of ${\bf{T}}^N$. Let $1\leq i<j\leq N$ be two coordinate directions and consider the projection onto the plane of the coordinates $i$ and $j$, which is a map ${\it pr}\colon{\bf{T}}^N\to{\bf{T}}^2$. The image of $\ell$ under ${\it pr}$ is parametrized by \begin{equation}\label{Eq:proj_param} z\ \longmapsto\ [b_i(z)\,:\,b_j(z)\,:\,b_{N+1}(z)]\,. \end{equation} If $b_i,b_j$, (and $b_{N{+}1}=1$) all vanish at $\xi_{N{+}1}=\infty$, then the image of $\ell$ under ${\it pr}$ is a point, and the image of $Z(\ell)$ is either a point or is one-dimensional, and so ${\it pr}_*[{\mathcal{A}}(\ell)+Z(\ell)]=0$. In this case $(\widetilde{p}\hspace{1.2pt}_{1,i},\widetilde{p}\hspace{1.2pt}_{1,j})$ is either $(0,0)$, $(\pi,0)$, or $(\pi,\pi)$, by~\eqref{Eq:inc} and~\eqref{Eq:ptilde}. Otherwise, the image of $\ell$ under the projection of $\P^N$ to the $(i,j)$-coordinate plane is the line $\ell'$ parameterized by~\eqref{Eq:proj_param}. It is immediate from the definitions that \[ {\it pr}(\overline{{\mathcal{A}}(\ell)})\ = \overline{{\mathcal{A}}(\ell')} \qquad\mbox{and}\qquad {\it pr}(Z(\ell))\ =\ Z(\ell')\,. \] When $b_i$ and $b_j$ have distinct (finite) zeroes, say $\zeta_a$ and $\zeta_b$, then ${\it pr}$ is injective on the interior of $\overline{{\mathcal{A}}(\ell)}$ and on the edges with directions $\pm{\bf f}_a$, $\pm{\bf f}_b$, and $\pm{\bf f}_{M+1}$ (sending them to edges with directions $\pm{\bf e}_1$, $\pm{\bf e}_2$, and $\pm({\bf e}_1{+}{\bf e}_2)$) and collapsing the others to points. In the other cases, ${\mathcal{A}}(\ell')$ is a circle. However, in all cases ${\it pr}$ is one-to-one over the interiors of each triangle in the image zonotope cycle $Z(\ell')$, collapsing the other triangles to line segments or to points. Thus \[ {\it pr}_*[\overline{{\mathcal{A}}(\ell)}+Z(\ell)]\ =\ [\overline{{\mathcal{A}}(\ell')}+Z(\ell')]\,. \] Since the last vertex of the path $P(\ell')$ is $(\widetilde{p}\hspace{1.2pt}_{1,i},\widetilde{p}\hspace{1.2pt}_{1,j})$, the theorem follows from the computation of Example~\ref{Ex:plane}. \end{proof} \section{Structure of discriminant coamoebas in dimension two}\label{S:three} Suppose now that $B\subset{\bf{Z}}^2$ is a multiset of $N{+}1$ vectors which span ${\bf{R}}^2$ and have sum ${\bf 0}=(0,0)$. We use $B=\{{\bf b}_1,\dotsc,{\bf b}_{N{+}1}\}$ to define a rational map ${\bf{C}}^2-\to{\bf{C}}^2$ \begin{equation}\label{Eq:HK} z\ \longmapsto\ \Big( \prod_{i=1}^{N+1} \langle {\bf b}_i,z\rangle^{{\bf b}_{i,1}}\,,\, \prod_{i=1}^{N+1} \langle {\bf b}_i,z\rangle^{{\bf b}_{i,2}}\Bigr)\,. \end{equation} Since $\sum_i {\bf b}_i={\bf 0}$, each coordinate is homogenous of degree $0$, and so~\eqref{Eq:HK} induces a rational map $\Psi_B\colon\P^1\to\P^2$ (where the image has distinguished coordinates). Define \defcolor{$D_B$} to be the image of this map~\eqref{Eq:HK}. When $B$ consists of distinct vectors that span ${\bf{Z}}^2$, then it is Gale dual to a set of vectors of the form $(1,{\bf a})$ for ${\bf a}\in A\subset{\bf{Z}}^{n+2}$. In this case,~\eqref{Eq:HK} is the Horn-Kapranov parametrization~\cite{K91} of the reduced $A$-discriminant. We use Theorem~\ref{Th:homology} to study the coamoeba \defcolor{${\mathcal{A}}_B$} of $D_B$ and its complement, for any multiset $B$. The results of Section~\ref{S:realline} are applicable because the map~\eqref{Eq:HK} factors, \[ \begin{array}{rcrcl} {\bf{C}}^2\ni z&\longmapsto&(\langle{\bf b}_1,z\rangle,\langle{\bf b}_2,z\rangle, \dotsc,\langle{\bf b}_{N+1},z\rangle)\in{\bf{C}}^{N+1}\\ &&{\bf{C}}^{N+1}\ni (x_1,x_2,\dotsc,x_{N+1})&\longmapsto& {\displaystyle\Big( \prod_{i=1}^{N+1} x_i^{{\bf b}_{i,1}}\,,\, \prod_{i=1}^{N+1} x_i^{{\bf b}_{i,2}}\Bigr)\in{\bf{C}}^2} \end{array} \] The first map, $\Phi_B$, is linear and the second, $\beta$, is a monomial map. They induce maps $\P^1\to\P^N-\to\P^2$, with the second a rational map. Let \defcolor{$\ell_B$} be the image of $\Phi_B$ in $\P^N$, which is a real line as in Section~\ref{S:realline}. The map $\Arg(\beta)$ is the homomorphism ${\bf{T}}^N\to{\bf{T}}^2$ induced by the linear map on the universal covers, (also written $\Arg(\beta)$), \[ \Arg(\beta)\ \colon\ {\bf{R}}^N\ \ni\ {\bf e}_i\ \longmapsto\ {\bf b}_i\ \in\ {\bf{R}}^2\,, \] and the following is immediate. \begin{lemma}\label{L:coamoeba_structure} The coamoeba ${\mathcal{A}}_B$ is the image of the coamoeba ${\mathcal{A}}(\ell_B)$ under the map $\Arg(\beta)$. \end{lemma} \begin{example}\label{Ex:rational_cubic} Let $B$ be the vector configuration $\{(1,0), (-2,1), (1,-2), (0,1)\}$. Observe that ${\bf b}_1+{\bf b}_2+{\bf b}_3+{\bf b}_4=0$ and $3{\bf b}_1+2{\bf b}_2+{\bf b}_3=0$, thus $B$ is Gale dual to the vector configuration $\{(1,3),(1,2),(1,1),(1,0)\}\subset\{1\}\times{\bf{Z}}$. So $A$ is simply $\{0,1,2,3\}$ if we identify ${\bf{Z}}$ with $\{1\}\times{\bf{Z}}$. We show these two configurations. \[ \begin{picture}(63,58)(-10,-4) \put(0,0){\includegraphics{figures/vectorsB.eps}} \put( 27,47){${\bf b}_4$} \put(47,26){${\bf b}_1$} \put(-3,50){${\bf b}_2$} \put(47,-4){${\bf b}_3$} \put(5,5){$B$} \end{picture} \qquad\qquad \begin{picture}(68,59)(-2,0) \put(0,30){\includegraphics{figures/vectorsA.eps}} \put(-1.5,23.4){$0$} \put(18.8,23.4){$1$} \put(38.8,23.4){$2$} \put(58.8,23.4){$3$} \put(29,44){$A$} \end{picture} \] Observe that the convex hull of $A$ has volume $d_B=3$. The map~\eqref{Eq:HK} becomes \[ (x,y)\ \longmapsto\ \Bigl( \frac{x(x-2y)}{(y-2x)^2}\,,\, \frac{y(y-2x)}{(x-2y)^2}\Bigr)\,, \] whose image is the curve below. \[ \begin{picture}(81,100) \put(0,0){\includegraphics[height=100pt]{figures/Adiscr.eps}} \put(21,76.5){$-2$} \put(78,29){$-2$} \end{picture} \] The line $\ell_B$ is the line of Example~\ref{Ex:tc} and so ${\mathcal{A}}_B$ is the image of the coamoeba of Figure~\ref{F:linecoamoeba} under the map \[ \Arg(\beta)\ \colon\ (\theta_1,\theta_2,\theta_3)\ \longmapsto\ (\theta_1{-}2\theta_2{+}\theta_3, \theta_2{-}2\theta_3)\,. \] We display this image below, first in the fundamental domain $[\pi,\pi]^2$ of ${\bf{T}}^2$, and then in universal cover ${\bf{R}}^2$ of ${\bf{T}}^2$ (each square is one fundamental domain). \begin{equation}\label{Eq:A-discr} \raisebox{-40pt}{\begin{picture}(60,90) \put(0,0){\includegraphics{figures/AdiscrCo.eps}} \put(23,2){${\mathcal{A}}_B$} \end{picture} \qquad\qquad \begin{picture}(165,90)(5,0) \put(37,0){\includegraphics{figures/AdiscrCoUC.eps}} \put(0,2){${\mathcal{A}}_B$} \put(20,5){\vector(1,0){92}} \put(11,13){\vector(1,2){31}} \put(84,46){$0$} \put(105,75){$Z_B$} \put(104,78){\vector(-2,-1){30}} \end{picture}} \end{equation} In the picture on the left, the darker shaded regions are where the argument map is two-to-one. The octagon on the right is the zonotope $Z_B$ generated by $B$ and it is the image of the zonotope chain of Figure~\ref{F:ZC} under the map $\Arg(\beta)$. Observe that the union of the coamoeba and the zonotope covers the fundamental domain $d_B=3$ times. \hfill$\includegraphics[height=10pt]{figures/QED.eps}$ \end{example} What we observe in this example is in fact quite general. We first use Lemma~\ref{L:coamoeba_structure} to describe the coamoeba ${\mathcal{A}}_B$ more explicitly, then study the zonotope $Z_B$ generated by $B$, before making an important definition and giving our proof of Theorem~\ref{T:NP}. The line $\ell_B$ is parametrized by the forms $z\mapsto\langle {\bf b}_i,z\rangle$, for $i=1,\dotsc,N{+}1$. Let $\xi_i\in{\bf{R}}\P^1$ be the zero of the $i$th form, and suppose these are in a weakly increasing cyclic order on ${\bf{R}}\P^1$, \[ \xi_1\ \leq\ \xi_2\ \leq\ \dotsb\ \leq\ \xi_{N+1}\,. \] Next, identify $\P^1\smallsetminus\{\xi_{N+1}\}$ with ${\bf{C}}$, so that $\xi_{N+1}$ is the point $\infty$ at infinity, and suppose that the distinct zeroes are \[ \zeta_1\ <\ \zeta_2\ <\ \dotsb\ <\ \zeta_M\ <\ \zeta_{M{+}1}\ =\ \infty\,. \] By the description of the coamoeba ${\mathcal{A}}(\ell_B)$ of Section~\ref{S:realline} and Lemma~\ref{L:coamoeba_structure}, we see that the coamoeba ${\mathcal{A}}_B$ is composed of two components, each bounded by polygonal paths that are the images of the boundary of ${\mathcal{A}}(\ell_B)$ under the map $\Arg(\beta)$. For each $j=1,\dotsc,M{+}1$, set \[ \defcolor{{\bf c}_j}\ :=\ \Arg(\beta)({\bf f}_j)\ =\ \sum_{i\colon \langle{\bf b}_i,\zeta_j\rangle=0} {\bf b}_i\,. \] The components of ${\mathcal{A}}_B$ correspond to the half planes of $\P^1$, and the boundary along each is the polygonal path with edges $\pm\pi{\bf c}_1,\dotsc,\pm\pi{\bf c}_{M+1}$ with the $+$ signs for the upper half plane and $-$ signs for the lower half plane. The complete description requires the following proposition, which is explained in~\cite[\S~2]{NP10}. \begin{proposition}\label{P:unrammified} Suppose that $M>1$. Then the composition \[ \P^1\smallsetminus\{\zeta_1,\dotsc,\zeta_{M+1}\} \xrightarrow{\,\Psi_B\,}D_B\xrightarrow{\,\Arg\,}{\mathcal{A}}_B\ \subset\ {\bf{T}}^2 \] is an immersion when restricted to $\P^1\smallsetminus{\bf{R}}\P^1$ (in fact it is locally a covering map). \end{proposition} The edges $\pm\pi{\bf c}_1,\dotsc,\pm\pi{\bf c}_{M+1}$ decompose ${\bf{T}}^2$ into polygonal regions. Over each polygonal region the map of Proposition~\ref{P:unrammified} has a constant number of preimages. This number of preimages equals the winding number of the polygonal path around that region. Then the pushforward $\Arg(\beta)_*(\overline{{\mathcal{A}}(\ell_B)})$ of the coamoeba chain of the line $\ell_B$ is the chain in ${\bf{T}}^2$ where the multiplicity of a region is this number of preimages/winding number. This equals the coamoeba chain of $D_B$. We will write \defcolor{$\overline{{\mathcal{A}}_B}$} for this chain $\Arg(\beta)_*(\overline{{\mathcal{A}}(\ell_B)})$, as our arguments use the pushforward. There is another natural chain we may define from the vector configuration $B$. Let $\defcolor{\overline{{\bf 0},\pi{\bf b}_i}}$ be the directed line segment in ${\bf{R}}^2$ connecting the origin to the endpoint of the vector $\pi{\bf b}_i$. Let $\defcolor{Z_B}\subset{\bf{R}}^2$ be the Minkowski sum of the line segments $\overline{{\bf 0},\pi{\bf b}_i}$ for ${\bf b}_i\in B$. This is a centrally symmetric zonotope as $\sum_i {\bf b}_i={\bf 0}$. We will also write $Z_B$ for its image in ${\bf{T}}^2$, considered now as a chain. For any ${\bf v}\in{\bf{R}}^2$, the points \[ q\ :=\ \sum_{\langle{\bf b}_i,{\bf v}\rangle>0} {\bf b}_i \qquad\mbox{and}\qquad q'\ :=\ \sum_{\langle{\bf b}_i,{\bf v}\rangle\geq0} {\bf b}_i \] are vertices of $Z_B$ which are extreme in the direction of ${\bf v}$. These differ only if the line ${\bf{R}}{\bf v}$ represents a zero $\zeta_j$ of one of the forms, and then the edge between them is $\pi{\bf d}_j$, where \begin{equation}\label{Eq:bdj} {\bf d}_j\ :=\ \sum_{i\colon \langle{\bf b}_i,{\bf v}\rangle=0} \sign(\langle{\bf b}_i,{\bf w}\rangle)\; {\bf b}_i,, \end{equation} where ${\bf w}$ is a vector such that $\langle -{\bf w},q\rangle>\langle -{\bf w},q'\rangle$ and $\sign(x)\in\{\pm1\}$ is the sign of the real number $x$. Thus ${\bf d}_j$ is the vector parallel to any ${\bf b}_i$ with $\langle{\bf b}_i,\zeta_j\rangle=0$ whose length is the sum of the lengths of these vectors and its direction is such that $\langle{\bf d}_j,{\bf w}\rangle>0$. Starting at a vertex of $Z_B$ and moving, say clockwise, the successive edge vectors will be the vectors $\{\pm\pi{\bf d}_1,\dotsc,\pm\pi{\bf d}_M,\pm\pi{\bf d}_{M+1}\}$ occuring in a cyclic clockwise order. This may be seen on the right in~\eqref{Eq:A-discr}, where $Z_B$ is the octagon. Its southeastern-most vertex is $\pi{\bf b}_1+\pi{\bf b}_3$ (corresponding to the vector ${\bf v}_1=-{\bf b}_2$, and the edges encountered from there in clockwise order are $-\pi{\bf b}_1,\pi{\bf b}_2,-\pi{\bf b}_3, \pi{\bf b}_4, \pi{\bf b}_1,-\pi{\bf b}_2, \pi{\bf b}_3,-\pi{\bf b}_4$. (Here, ${\bf d}_j={\bf b}_j$) Before giving our proof of Theorem~\ref{T:NP}, we make an important definition. Let $B=\{{\bf b}_1,\dotsc,{\bf b}_{N+1}\}$ be a multiset of vectors in ${\bf{Z}}^2$ that span ${\bf{R}}^2$ and whose sum is ${\bf 0}$. Write \demph{$\cone({\bf b}_i,{\bf b}_j)$} for the cone generated by the vectors ${\bf b}_i, {\bf b}_j$. Suppose that ${\bf v}$ is any vector in ${\bf{R}}^2$ not pointing in the direction of a vector in $B$, and set \begin{equation}\label{Eq:dBv} \defcolor{d_{B,{\bf v}}}\ :=\ \sum_{{\bf v}\in\cone({\bf b}_i,{\bf b}_j)} |{\bf b}_i\wedge {\bf b}_j|\ . \end{equation} Here $|{\bf b}_i\wedge {\bf b}_j|$ is the absolute value of the determinant of the matrix whose columns are the two vectors, which is the area of the parallelogram generated by ${\bf b}_i$ and ${\bf b}_j$. \begin{lemma}\label{L:independent} The sum~$\eqref{Eq:dBv}$ is independent of choice of ${\bf v}$. \end{lemma} \begin{proof} The rays generated by elements of $B$ divide ${\bf{R}}^2$ into regions. The sum~\eqref{Eq:dBv} depends only upon the region containing ${\bf v}$---it is a sum over all cones containing the given region. To show its independence of region, let ${\bf v},{\bf v}'$ lie in adjacent regions with $\defcolor{{\bf u}}$ a vector generating the ray separating the regions. Suppose that the vectors in $B$ are indexed so that ${\bf b}_\kappa,{\bf b}_{\kappa+1},\dotsc,{\bf b}_{\mu-1}$ are the vectors with direction $-{\bf u}$ and ${\bf b}_\mu,{\bf b}_{\mu+1},\dotsc,{\bf b}_\lambda$ are the vectors with direction ${\bf u}$. Then the sums for $d_{B,{\bf v}}$ and $d_{B,{\bf v}'}$ both include the sum over all cones whose relative interior contains ${\bf u}$, but have different terms involving cones with one generator among ${\bf b}_\mu,\dotsc,{\bf b}_\lambda$. All such cones appear, and up to a sign, the difference $d_{B,{\bf v}}-d_{B,{\bf v}'}$ is equal to \begin{multline*} \bigl({\bf b}_\mu +\dotsb+ {\bf b}_\lambda\bigr) \wedge \bigl({\bf b}_1 +\dotsb+ {\bf b}_{\kappa-1}\ +\ {\bf b}_{\lambda+1}+\dotsb+ {\bf b}_{N+1}\bigr)\\ \ =\ \bigl({\bf b}_\mu +\dotsb +{\bf b}_\lambda\bigr) \wedge ({\bf b}_1+\dotsb+{\bf b}_{N+1})\ =\ 0\,, \end{multline*} which proves the lemma. \end{proof} \begin{remark} The sum~\eqref{Eq:dBv} is known to coincide with the normalized volume of the convex hull of the vector configuration $A$ that is Gale dual to $B$ (see~\cite{DS02}), so Lemma~\ref{L:independent} also follows from this fact. We will henceforth write \demph{$d_B$} for this volume/sum. \hfill$\includegraphics[height=10pt]{figures/QED.eps}$ \end{remark} \begin{example} Consider the sum~\eqref{Eq:dBv} for the vector configuration $B$ of Example~\ref{Ex:rational_cubic}. There are four choices for the vector ${\bf v}$ as indicated below \begin{equation}\label{Eq:B-and-v} \raisebox{-46pt}{\begin{picture}(100,93)(-15,-3) \put(0,0){\includegraphics{figures/vectorsB-v.eps}} \put( 36,63){${\bf b}_4$} \put(62,37){${\bf b}_1$} \put(-13,57){${\bf b}_2$} \put(62, -3){${\bf b}_3$} \put( 36,12){${\bf v}_1$} \put(81,21){${\bf v}_2$} \put( 61,61){${\bf v}_3$} \put( 7,79){${\bf v}_4$} \end{picture}} \end{equation} The vector ${\bf v}_1$ lies only in $\cone({\bf b}_2,{\bf b}_3)$, and we have ${\bf b}_2\wedge {\bf b}_3=|\begin{smallmatrix}-2&1\\1&-2\end{smallmatrix}|=3$. The vector ${\bf v}_2$ lies in $\cone({\bf b}_3,{\bf b}_1)$ and $\cone({\bf b}_3,{\bf b}_4)$, and we have ${\bf b}_3\wedge {\bf b}_1+{\bf b}_3\wedge {\bf b}_4= |\begin{smallmatrix}1&-2\\1&0\end{smallmatrix}|+ |\begin{smallmatrix}1&-1\\0&1\end{smallmatrix}|=2+1=3$. Similarly, ${\bf v}_3$ lies in $\cone({\bf b}_3,{\bf b}_4)$, $\cone({\bf b}_1,{\bf b}_4)$, and $\cone({\bf b}_1,{\bf b}_2)$, and ${\bf b}_3\wedge {\bf b}_4+{\bf b}_1\wedge {\bf b}_4+{\bf b}_1\wedge {\bf b}_2=1+1+1=3$, and the calculation for ${\bf v}_4$ is the mirror-image of that for ${\bf v}_2$. In every case, $d_{B,{\bf v}_i}=3$, and so $d_B=3$. \hfill$\includegraphics[height=10pt]{figures/QED.eps}$ \end{example} \begin{theorem}\label{Th:Cycle} The sum, $\overline{{\mathcal{A}}_B}+Z_B$, of the coamoeba chain of $D_B$ and the $B$-zonotope chain is a cycle in ${\bf{T}}^2$ which equals $d_B[{\bf{T}}^2]$. \end{theorem} \begin{proof} We will show that $\Arg(\beta)_*[Z(\ell_B)]=[Z_B]$, which implies that \[ [\overline{{\mathcal{A}}_B}+Z_B]\ =\ \Arg(\beta)_*[\overline{{\mathcal{A}}(\ell_B)} + Z(\ell_B)] \] is a cycle, as $\Arg(\beta)_*[\overline{{\mathcal{A}}(\ell_B)}]=[\overline{{\mathcal{A}}_B}]$. Since $\Arg(\beta)_*({\bf e}_i\wedge{\bf e}_j)={\bf b}_i\wedge{\bf b}_j\cdot[{\bf{T}}^2]$, the formula of Theorem~\ref{Th:homology} will give us the homology class of $[\overline{{\mathcal{A}}_B}+Z_B]$. We will use~\eqref{Eq:dBv} and Lemma~\ref{L:independent} to show that it equals $d_B[{\bf{T}}^2]$. This will imply the theorem as we will show that there is an ordering of the vectors $B$ such that the map $\Arg(\beta)\colon Z(\ell_B)\to Z_B$ in the universal covers ${\bf{R}}^N\to{\bf{R}}^2$ is injective. Recall that $\xi_1,\dotsc,\xi_{N+1}$ are points of ${\bf{R}}\P^1$ with $\langle {\bf b}_i,\xi_i\rangle=0$ and $\zeta_1,\dotsc,\zeta_{M+1}$ are the distinct points among them. Let ${\bf 0}\neq\defcolor{{\bf v}}\in{\bf{R}}^2$ represent $\xi_{N+1}=\zeta_{M+1}$ (so that $\langle{\bf b}_{N+1},{\bf v}\rangle=0$) and choose $x\in{\bf{R}}^2$ to be a point with $\langle{\bf b}_{N+1},x\rangle=1$. Then $t\mapsto x+t{\bf v}$ gives a parametrization of ${\bf{R}}\P^1$ with $\infty=\zeta_{M+1}$, and identifies ${\bf{R}}$ with ${\bf{R}}\P^1\smallsetminus\{\infty\}$. To agree with Definition~\ref{D:conventions}, we suppose that the points of $B$ are ordered so that~\eqref{Eq:A} and~\eqref{Eq:B} hold. Thus there are integers $1=m_1<\dotsb<m_{M+1}<m_{M+2}=N{+}2$ such that \[ \langle {\bf b}_i,\zeta_j\rangle=0\ \Longleftrightarrow\ m_j\leq i<m_{j+1}\,. \] We further suppose that $B$ is ordered so that one of~\eqref{Eq:inc} or~\eqref{Eq:dec} holds for every $j=1,\dotsc,M{+}1$. Specifically, let ${\bf w}:=x+\tau {\bf v}$ for some fixed $\tau<\zeta_1$. Then there exist integers $n_1,\dotsc,n_{M+1}$ such that for each $j=1,\dotsc,M{+}1$ we have $m_j< n_j\leq m_{j+1}$ and either \begin{eqnarray*} \langle{\bf b}_{m_j},{\bf w}\rangle\,,\ \dotsc\,,\ \langle{\bf b}_{n_j-1},{\bf w}\rangle &<\ 0\ <& \langle{\bf b}_{n_j},{\bf w}\rangle\,,\ \dotsc\,,\ \langle{\bf b}_{m_{j+1}-1},{\bf w}\rangle\,, \makebox[.1in][l]{\qquad or}\\ \langle{\bf b}_{m_j},{\bf w}\rangle\,,\ \dotsc\,,\ \langle{\bf b}_{n_j-1},{\bf w}\rangle &>\ 0\ >& \langle{\bf b}_{n_j},{\bf w}\rangle\,,\ \dotsc\,,\ \langle{\bf b}_{m_{j+1}-1},{\bf w}\rangle\,. \end{eqnarray*} For $i=1,\dotsc,N{+}1$, let $\sgn_i\in\{\pm 1\}$ be the sign of $\langle{\bf b}_i,{\bf w}\rangle$. Note that $\sgn_{N+1}=1$. Define ${\bf f}_j,{\bf g}_j,{\bf h}_j$ as in Definition~\ref{D:conventions}, \[ {\bf f}_j\ :=\ \sum_{i=m_j}^{m_{j+1}-1} {\bf e}_i\,,\qquad {\bf g}_j\ :=\ \sum_{i=m_j}^{n_{j}-1} {\bf e}_i\,,\qquad\mbox{and}\qquad {\bf h}_j\ :=\ \sum_{i=m_j}^{m_{j+1}-1} \sgn_i{\bf e}_i\,. \] Consider now the following affine parametrization of $\ell_B\subset\P^N$, \[ \Phi_B\ \colon\ t\ \longmapsto\ [\langle{\bf b}_1,x+t{\bf v}\rangle\ \colon \dotsb\ \colon\ \langle{\bf b}_N,x+t{\bf v}\rangle\ \colon\ \langle{\bf b}_{N+1},x+t{\bf v}\rangle=1]\,. \] Let $\widetilde{p}\hspace{1.2pt}_1\in\{0,\pi\}^N\in{\bf{R}}^N$ be the point whose $i$th coordinate is $\arg(\sgn_i)$. Its image $p_1\in{\bf{T}}^N$ is the point on the coamoeba of $\ell_B$ coming from the real points $\Phi_B(-\infty,\zeta_1)$. We describe $\Arg(\beta)(Z(\ell_B))$ in the universal cover ${\bf{R}}^2$ of ${\bf{T}}^2$. For each $j=1,\dotsc,2M{+}2$, set $\defcolor{\widetilde{q}_j}:=\Arg(\beta)(\widetilde{p}\hspace{1.2pt}_j)$ and $\defcolor{{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j}:=\Arg(\beta)({\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j)$. Since \begin{equation}\label{Eq:def_p1tilde} \widetilde{p}\hspace{1.2pt}_{1,i}\ =\ \left\{ \begin{array}{rcl} \pi &&\mbox{if }\langle{\bf b}_i,{\bf w}\rangle<0\\ 0 &&\mbox{if }\langle{\bf b}_i,{\bf w}\rangle>0 \end{array}\right.\ , \end{equation} we have \[ \widetilde{q}_1\ =\ \pi\cdot\sum_{\langle{\bf b}_i,{\bf w}\rangle<0}{\bf b}_i\ , \] and so $\widetilde{q}_1$ is a vertex of $Z_B$ which is extreme in the direction of $-{\bf w}$. The zonotope chain $Z(\ell_B)$ is a union of the triangles \begin{equation}\label{Eq:ZC_triangles} \conv({\bf 0},\widetilde{p}\hspace{1.2pt}_{j+1},{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j) \qquad\mbox{and}\qquad \conv({\bf 0},{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j,\widetilde{p}\hspace{1.2pt}_j) \qquad\mbox{for}\ j=M{+}2,\dotsc,1\,, \end{equation} where the second is degenerate if $\widetilde{p}\hspace{1.2pt}_j={\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j$. Thus $\Arg(\beta)(Z(\ell_B))$ will be the union of the (possibly degenerate) triangles \begin{equation}\label{Eq:Zone_triangles} \conv({\bf 0},\widetilde{q}_{j+1},{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j) \qquad\mbox{and}\qquad \conv({\bf 0},{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j,\widetilde{q}_j) \qquad\mbox{for}\ j=M{+}2,\dotsc,1\,, \end{equation} For $j\leq M{+}1$, $\widetilde{p}\hspace{1.2pt}_{j+1}=\widetilde{p}\hspace{1.2pt}_j+\pi{\bf h}_j$, so \[ \widetilde{q}_{j+1}\ =\ \widetilde{q}_j+\pi\Arg(\beta)({\bf h}_j)\ =\ \widetilde{q}+\pi{\bf d}_j\,, \] which we see by~\eqref{Eq:bdj} (with the vector ${\bf w}=x+\tau{\bf v}$) and our definition of $\sgn_i$. If we fix the orientation so that ${\bf v}$ is clockwise of ${\bf b}_{N+1}$, then by our choice of ordering of the zeroes $\zeta_j$, the lines ${\bf{R}}{\bf d}_1, \dotsc,{\bf{R}}{\bf d}_{M+1}$ occur in clockwise order. Since $\langle{\bf d}_j,{\bf w}\rangle>0$ and $\widetilde{q}_1$ is extreme in the direction of $-{\bf w}$, the vectors $\pi{\bf d}_1,\dotsc,\pi{\bf d}_{M{+}1}$ will form the edges of the zonotope starting at $\widetilde{q}_1$ and moving clockwise. It follows from the discussion following~\eqref{Eq:bdj} that $\widetilde{q}_1,\dotsc,\widetilde{q}_{2M+2}$ form the vertices of the zonotope $Z_B$. This implies that no $\widetilde{q}_j$ coincides with the origin ${\bf 0}$. All that remains is to understand the two triangles~\eqref{Eq:Zone_triangles} for those $j$ when ${\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j\neq\widetilde{q}_j$. In this case, ${\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j=\widetilde{p}\hspace{1.2pt}_j+2\pi\sgn_{m_j}{\bf g}_j$, and so \[ {\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j\ =\ \widetilde{p}\hspace{1.2pt}_j\ +\ 2\pi\sgn_{m_j}\sum_{i=m_j}^{n_j-1} {\bf b}_i\ =\ \widetilde{p}\hspace{1.2pt}_j\ +\ 2\pi\sum_{i=m_j}^{n_j-1} \sgn_i{\bf b}_i\,. \] Since ${\bf b}_{m_j},\dotsc,{\bf b}_{m_{j+1}-1}$ are parallel, $\widetilde{q}_j,{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j$, and $\widetilde{q}_{j+1}$ are collinear. This implies that \[ \Arg(\beta)_* [ \conv({\bf 0},{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j,\widetilde{p}\hspace{1.2pt}_j) + \conv({\bf 0},\widetilde{p}\hspace{1.2pt}_{j+1},{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j)]\ =\ [\conv({\bf 0},\widetilde{q}_{j+1},\widetilde{q}_j)]\,, \] which shows that $\Arg(\beta)_*[Z(\ell_B)]=[Z_B]$. Indeed, if ${\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j$ lies between $\widetilde{q}_j$ and $\widetilde{q}_{j+1}$ then $\Arg(\beta)$ preserves the orientation of the triangles~\eqref{Eq:ZC_triangles} and is therefore injective over their images, whose union is $\conv({\bf 0},\widetilde{q}_{j+1},\widetilde{q}_j)$. Otherwise, the two triangles~\eqref{Eq:Zone_triangles} have opposite orientations and \[ \conv({\bf 0},{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j,\widetilde{q}_j)\ \supset\ \conv({\bf 0},\widetilde{q}_{j+1},{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j)\,, \] so that $\Arg(\beta)_* [ \conv({\bf 0},{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j,\widetilde{p}\hspace{1.2pt}_j) + \conv({\bf 0},\widetilde{p}\hspace{1.2pt}_{j+1},{\widetilde{p}\hspace{1.6pt}'\hspace{-3.4pt}}_j)]$ equals \[ [ \conv({\bf 0},{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j,\widetilde{q}_j)]-[ \conv({\bf 0},\widetilde{q}_{j+1},{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j)] \ =\ [\conv({\bf 0},\widetilde{q}_{j+1},\widetilde{q}_j)]\,. \] Theorem~\ref{Th:homology}, Equation~\eqref{Eq:def_p1tilde}, and $\Arg(\beta)_*({\bf e}_i\wedge {\bf e}_j)={\bf b}_i\wedge {\bf b}_j\cdot[{\bf{T}}^2]$, show that \[ \Arg(\beta)_*[\overline{{\mathcal{A}}(\ell_B)}+Z(\ell_B)]\ =\ [{\bf{T}}^2]\cdot \sum_{\substack{1\leq i<j\leq N\\\langle{\bf b}_i,{\bf w}\rangle>0>\langle{\bf b}_j,{\bf w}\rangle}} {\bf b}_i\wedge {\bf b}_j\,. \] We will show that this equals $d_B[{\bf{T}}^2]$. Observe that if ${\bf b}_i$ and ${\bf b}_j$ are parallel, then ${\bf b}_i\wedge{\bf b}_j=0$ and they do not contribute to the sum. We will consider the sum with the restriction that the vectors ${\bf b}_i$ and ${\bf b}_j$ are not parallel. Set $\defcolor{{\bf w}^\perp}:=-{\bf b}_{N+1}+{\bf w}/\langle{\bf w},{\bf w}\rangle$, which is orthogonal to ${\bf w}$. Suppose that ${\bf v}$ is clockwise of ${\bf b}_{N+1}$, as below. \[ \begin{picture}(93,104)(-13,-20) \put(-10,-10){\includegraphics{figures/bij.eps}} \put(-1,-12){${\bf b}_i$} \put(58,-13){${\bf b}_j$} \put(25,78){${\bf b}_{N+1}$} \put(61,33){${\bf v}$} \put(-13,32){${\bf w}$} \put(16,-18){${\bf w}^\perp$} \end{picture} \] By our choice of ${\bf w}$, the lines ${\bf{R}}{\bf w}^\perp,{\bf{R}}{\bf b}_1,\dotsc,{\bf{R}}{\bf b}_{N+1}$ occur in weak clockwise order with ${\bf{R}}{\bf w}^\perp$ distinct from the rest. Suppose now that $1\leq i<j\leq N$ where \begin{equation}\label{Eq:condition} \langle{\bf b}_i,{\bf w}\rangle\ >\ 0\ >\ \langle{\bf b}_j,{\bf w}\rangle\,, \end{equation} and ${\bf b}_i$ and ${\bf b}_j$ are not parallel. The cone spanned by ${\bf b}_i$ and ${\bf b}_j$ meets a half ray of ${\bf{R}}{\bf w}^\perp$, with ${\bf b}_i$ to the left of ${\bf{R}}{\bf w}^\perp$ and ${\bf b}_j$ to the right of ${\bf{R}}{\bf w}^\perp$, by~\eqref{Eq:condition}. Since ${\bf{R}}{\bf w}^\perp,{\bf{R}}{\bf b}_i$, and ${\bf{R}}{\bf b}_j$ occur in clockwise order, we must have that ${\bf w}^\perp\in\cone({\bf b}_i,{\bf b}_j)$, which shows that \[ \sum_{\substack{1\leq i<j\leq N\\\langle{\bf b}_i,{\bf w}\rangle<0<\langle{\bf b}_j,{\bf w}\rangle}} {\bf b}_i\wedge {\bf b}_j \ =\ \sum_{\substack{1\leq i<j\leq N\\{\bf w}^\perp\in\cone({\bf b}_i,{\bf b}_j)}} {\bf b}_i\wedge {\bf b}_j\ =\ d_{B,{\bf w}^\perp}\ =\ d_B\,. \] The sum equals $d_{B,{\bf w}^\perp}$ because if ${\bf b}_j$ is counter clockwise from ${\bf b}_i$ by~\eqref{Eq:condition} and the condition that ${\bf w}^\perp\in\cone({\bf b}_i,{\bf b}_j)$ with $i<j$. Thus ${\bf b}_i\wedge{\bf b}_j>0$. We complete the proof by noting that ${\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_j$ will lie between $\widetilde{q}_j$ and $\widetilde{q}_{j+1}$ if either $n_j=m_{j+1}$, so that ${\bf g}_j={\bf f}_j$, or if \[ \|{\bf g}_j\|\ =\ \|\sum_{i=m_j}^{n_j-1}{\bf b}_j\| \ =\ \sum_{i=m_j}^{n_j-1}\|{\bf b}_j\|\ \leq\ \sum_{i=n_j}^{m_{j+1}-1}\|{\bf b}_j\|\ =\ \|{\bf f}_j-{\bf g}_j\|\,, \] as ${\bf b}_{m_j},\dotsc,{\bf b}_{n_j-1}$ have the same direction which is opposite to the (common) direction of ${\bf b}_{n_j},\dotsc,{\bf b}_{m_{j+1}-1}$. If this does not occur for our given order, then we simply reverse the vectors ${\bf b}_{m_j},\dotsc,{\bf b}_{m_{j+1}-1}$, replacing ${\bf g}_j$ with ${\bf f}_j-{\bf g}_j$. \end{proof} \begin{example}\label{ex:parallel} The last point in the proof about the injectivity of \[ \Arg(\beta)\ \colon\ Z(\ell_B)\ \longrightarrow\ Z_B \] (and more generally the arguments when $B$ has parallel vectors) is geometrically subtle. We expose this subtlety in the following two examples. Suppose that $B$ consists of the vectors $(1,0),(0,1),(-2,-2)$, and $(1,1)$, \[ \includegraphics{figures/vectorsparallel.eps} \] When ${\bf v}=(1,-1)$ and $x=(\frac{1}{2},\frac{1}{2})$, then $\ell_B$ has the parametrization \begin{equation}\label{Eq:first_P} z\ \longmapsto\ [ \tfrac{1}{2}+z\;:\; \tfrac{1}{2}-z\;:\; -2\;:\; 1]\,, \end{equation} which is the second line in our running Examples~\ref{Ex:CoArepeat},~\ref{Ex:CoZrepeat}, and~\ref{Ex:homology_class}. In this case the image $\Arg(\beta)(Z(\ell_B))$ is shown on the left of Figure~\ref{F:images}. It is superimposed over a fundamental domain and dashed lines $\theta_1,\theta_2=n\pi$ for $n\in{\bf{Z}}$. The segments $\widetilde{q}_3,{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_2$ and $\widetilde{q}_6,{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_5$ are covered in both directions as $\Arg(\beta)(P(\ell_B))$ backtracks over these segments. In fact, the triangles \[ \conv({\bf 0},\widetilde{q}_3,{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_2)\qquad\mbox{and}\qquad \conv({\bf 0},\widetilde{q}_6,{\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_5) \] have orientation opposite of the other triangles. The medium shaded parts (near ${\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_2$ and ${\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_5$) are covered twice and the darker shaded parts near ${\bf 0}$ are covered thrice. \begin{figure}[htb] \begin{picture}(177,126)(-11,-11) \put(0,0){\includegraphics{figures/imageParallelI.eps}} \put( 71, 52){${\bf 0}$} \put( 23,-10){$\widetilde{q}_1$} \put( 48,-10){$\widetilde{q}_2$} \put(156,102){${\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_2$} \put(129, 67){$\widetilde{q}_3$} \put(129,107){$\widetilde{q}_4$} \put( 93,107){$\widetilde{q}_5$} \put(-11, -3){${\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_5$} \put( 16, 32){$\widetilde{q}_6$} \end{picture} \qquad \begin{picture}(177,126)(-11,-11) \put(0,0){\includegraphics{figures/imageParallelII.eps}} \put( 71, 54){${\bf 0}$} \put( 23,-10){$\widetilde{q}_1$} \put( 48,-10){$\widetilde{q}_2$} \put(104, 42){${\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_2$} \put(129, 67){$\widetilde{q}_3$} \put(129,107){$\widetilde{q}_4$} \put( 93,107){$\widetilde{q}_5$} \put( 39, 57){${\widetilde{q}\hspace{1.6pt}'\hspace{-3.4pt}}_5$} \put( 16, 32){$\widetilde{q}_6$} \thicklines \put(130,11){\White{\vector(-1,0){114}}} \put(130,11){\White{\vector(-1,0){115}}} \put(138,21){\White{\vector(0,1){78}}} \put(138,22){\White{\vector(0,1){78}}} \thinlines \put(130,11){\vector(-1,0){113}} \put(138,21){\vector(0,1){77}} \put(132,9){${\mathcal{A}}_B$} \end{picture} \caption{Images of $\Arg(\beta)(Z(\ell_B))$} \label{F:images} \end{figure} Now suppose that the vectors in $B$ are in the order $(1,0)$, $(1,1)$, $(-2,-2)$, and $(0,1)$, and $v=(-1,0)$ and $x=(0,1)$. Then $\ell_B$ is parametrized by \[ z\ \longmapsto\ [-z\;:\; 1-z \;:\; 2z-2 \;:\; 1]\,, \] In this case the image $\Arg(\beta)(Z(\ell_B))$ is equal to the zonotope $Z_B$, and is shown on the right of Figure~\ref{F:images}, together with the coamoeba ${\mathcal{A}}_B$. As explained in the proof of Theorem~\ref{Th:Cycle}, the image equals the zonotope because in the pair of parallel vectors $(1,1)$ and $(-2,-2)$, the shorter comes first in this case, while in the previous case, the shorter one came second. In both cases (which are just different parametrizations of the same line) $\Arg(\beta)_*[Z(\ell_B)]=[Z_B]$ as shown in the proof of Theorem~\ref{Th:Cycle}, and the coamoebas coincide. Furthermore, $[\overline{{\mathcal{A}}_B}+Z_B]=2[{\bf{T}}^2]$ for both, as $d_B=2$. \hfill\includegraphics[height=10pt]{figures/QED.eps} \end{example} \def$'${$'$} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2]{% \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2}
{ "timestamp": "2012-08-03T02:04:43", "yymm": "1201", "arxiv_id": "1201.6649", "language": "en", "url": "https://arxiv.org/abs/1201.6649", "abstract": "Understanding the complement of the coamoeba of a (reduced) A-discriminant is one approach to studying the monodromy of solutions to the corresponding system of A-hypergeometric differential equations. Nilsson and Passare described the structure of the coamoeba and its complement (a zonotope) when the reduced A-discriminant is a function of two variables. Their main result was that the coamoeba and zonotope form a cycle which is equal to the fundamental cycle of the torus, multiplied by the normalized volume of the set A of integer vectors. That proof only worked in dimension two. Here, we use simple ideas from topology to give a new proof of this result in dimension two, one which can be generalized to all dimensions.", "subjects": "Algebraic Geometry (math.AG)", "title": "Discriminant coamoebas through homology", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9857180673335565, "lm_q2_score": 0.8128673223709251, "lm_q1q2_score": 0.8012580060060713 }
https://arxiv.org/abs/2103.14350
The convergence of the Stochastic Gradient Descent (SGD) : a self-contained proof
We give here a proof of the convergence of the Stochastic Gradient Descent (SGD) in a self-contained manner.
\section{Introduction} The Stochastic Gradient Descent (SGD) or other algorithms derived from it are used extensively in Deep Learning, a branch of Machine Learning; but the proof of convergence is not always easy to find. The goal of this paper is to adapt various proofs from the literature in a simple format. In particular no claim of originality is made (see \cite{turinici2020convergence,ayadi2020stochastic,turinici2020xray,Turinici_2019 } for some of my recent research papers in this area); on the contrary please cite this work if you find it useful (arxiv or DOI: \text{10.5281/zenodo.4638695}). This proof can be used in any domain where a self-contained presentation is needed. \section{Recall of the general framework} Suppose $(\Omega,F,\mathbb{P})$ is a probability space, $L:\Omega \times \mathbb{R}^N \to \mathbb{R}$ a function depending on a random argument $\omega$ and a parameter $X$ (second argument) to be optimized. Denote \begin{equation} \mathcal{L}(X) = \mathbb{E}_\omega [ L(\omega,X)]. \end{equation} The goal of the SGD is to find a minimum of $\mathcal{L}$. It operates iteratively by taking at iteration $n$: \begin{itemize} \item a (deterministic) "learning rate" $\rho_n$ (schedule fixed {\it a priori} \item a random $\omega_n \in \Omega$ independent of any other previous random variables is drawn (following the law $\mathbb{P}$) \item and updating by the formula \begin{equation} X_{n+1} = X_n - \rho_n \nabla_x L(\omega_n,X_n). \end{equation} \end{itemize} \section{Hypothesis on $L$ and $\mathcal{L}$} In order to prove the convergence we need some hypothesis that are detailed below \begin{enumerate} \item The gradient of $L$ is bounded: \begin{equation} \exists B > 0 : \ \sup_\omega \| \nabla_X L(\omega,X) \|^2 \le B, \ \forall X \in \mathbb{R}^N. \label{eq:hypgradbounded} \end{equation} \item $\mathcal{L}$ is strongly convex: \begin{equation} \exists \mu > 0 : \ \mathcal{L}(Y) \ge \mathcal{L}(X) + \langle \nabla \mathcal{L}(X), Y-X \rangle + \frac{\mu}{2} \|X-Y\|^2, \ \forall X,Y \in \mathbb{R}^N. \label{eq:hypstrongconvex} \end{equation} Note that for $\mu=0$ this is just the usual convexity, i.e. the function is above its tangent. For general $\mu$ this tells that the function is even above a parabola centered in any $X$. For regular functions this means that the Hessian $D^2 \mathcal{L}$ of $\mathcal{L}$ satisfies $D^2 \mathcal{L} \ge \mu \cdot I_N$\footnote{Here $I_N$ is the $N\times N$ identity matrix.}. \end{enumerate} \section{A convergence result and its proof} We fill prove the following \begin{theorem} Suppose that each $L(\omega,\cdot)$ is differentiable (a.e. $\omega \in \Omega$)\footnote{This requirement can be largely weakened. For instance in the case of ReLU activation, which corresponds to the positive part $x\mapsto x_+$, one can employ any suitable sub-gradient of the $x_+$ function and in particular take at the non-regular point $x=0$ any value between $0$ and $1$.} and that $\mathcal{L}$ satisfies the hypothesis \eqref{eq:hypgradbounded} and \eqref{eq:hypstrongconvex}. Then \begin{enumerate} \item the function $\mathcal{L}$ has an unique minimum $X_*$; \label{item:uniqueness} \item \label{item:inequalitydn} For any $n \ge 0$ denote \begin{equation} d_n = \mathbb{E} \left[ \|X_{n}-X_*\|^2 \right]. \end{equation} Then \begin{equation} d_{n+1} \le (1-\rho_n \mu) d_n + \rho_n^2 B. \label{eq:dnplusone} \end{equation} \item \label{item:constant} For any $\epsilon > 0$ there exists a $\rho > 0$ such that if $\rho_n = \rho$ then \begin{equation} \limsup_{n\to \infty} \mathbb{E} \left[ \|X_{n+1}-X_*\|^2 \right] \le \epsilon. \label{eq:estimationvoisinage} \end{equation} \item \label{item:convergence} Take $\rho_n$ a sequence such that: \begin{equation} \rho_n \to 0 \text{ and } \sum_{n\ge 1} \rho_n = \infty. \label{eq:hyprhon} \end{equation} Then $d_n \to 0$, that is $\lim_{n \to \infty }X_n = X_*$, where the convergence is the $L^2$ convergence of random variables. \end{enumerate} \begin{proof} {\bf Item \ref{item:uniqueness}}: The existence and uniqueness of the optimum is guaranteed by the assumptions of strong convexity and smoothness of $\mathcal{L}$. \noindent {\bf Item \ref{item:inequalitydn}}: We have \begin{align} & \mathbb{E} \left[ \|X_{n+1}-X_*\|^2 \right] = \mathbb{E} \left[ \|X_{n}-X_* - \rho_n \nabla_x L(\omega_n,X_n)\|^2 \right] \nonumber \\&= \mathbb{E} \left[ \|X_{n}-X_*\|^2 \right] + \rho_n^2 \mathbb{E} \left[ \|\nabla_x L(\omega_n,X_n)\|^2 \right] - 2 \rho_n \mathbb{E} \left[ \langle X_{n}-X_*, \nabla_x L(\omega_n,X_n)\rangle \right]. \end{align} First we remark that\footnote{The formal justification is as follows: denote by $\mathcal{F}_n$ the sigma algebra generated by $X_1$, ..., $X_n$, $\omega_1$, ..., $\omega_{n-1}$. In particular $\omega_n$ is independent of $\mathcal{F}_n$. Recall now that for any random variables $U$ measurable with respect to $\mathcal{F}_n$ and $V$ independent of $\mathcal{F}_n$: $ \mathbb{E}[ g(U,V) | \mathcal{F}_n] = \int g(v,U) P_V(dv) $ and in particular $\mathbb{E}[ g(U,V)]= \mathbb{E}[\mathbb{E}[ g(U,V) | \mathcal{F}_n]]= \mathbb{E}[ \int g(v,U) P_V(dv)]$. } $$ \mathbb{E} \left[ \langle X_{n}-X_*, \nabla_x L(\omega_n,X_n)\rangle \right] = \mathbb{E} \left[ \langle X_{n}-X_*, \nabla \mathcal{L}(X_n)\rangle \right]. $$ But at its turn \begin{align} & \mathbb{E} \left[ \langle X_{n}-X_*, \nabla \mathcal{L}(X_n)\rangle \right] \ge \mathbb{E} \left[ \mathcal{L}(X_n) - \mathcal{L}(X_*) + \frac{\mu}{2} \|X_{n}-X_*\|^2 \right] \nonumber\\& \ge \frac{\mu}{2} \mathbb{E} [\|X_{n}-X_*\|^2], \end{align} the last inequality being guaranteed by the fact that $X_*$ is the minimum. Putting together all relations proved so far one obtains the relation \eqref{eq:dnplusone}. \noindent {\bf Item \ref{item:constant}}: When $\rho_n$ is constant equal to $\rho$ inequality \eqref{eq:dnplusone} is equivalent to $$d_{n+1} - \rho \frac{B}{\mu} \le (1-\rho \mu) (d_n - \rho \frac{B}{\mu}).$$ Since the function $x \mapsto x_+$ (the positive part) is increasing we obtain for $\rho < 1/\mu$: $$\left(d_{n+1} - \rho \frac{B}{\mu} \right)_+ \le (1-\rho \mu) \left(d_n - \rho \frac{B}{\mu}\right)_+,$$ and by iteration, for any $k\ge 1$: $$\left(d_{n+k} - \rho \frac{B}{\mu} \right)_+ \le (1-\rho \mu)^k \left(d_n - \rho \frac{B}{\mu}\right)_+.$$ Taking $k\to \infty$ we obtain $\limsup_k \left(d_{k} - \rho \frac{B}{\mu} \right)_+=0$ hence the conclusion \eqref{eq:estimationvoisinage} for $\rho$ smaller then $1/\mu$ and $\epsilon \mu / B$. \noindent {\bf Item \ref{item:convergence}}: For non-constant $\rho_n$ and arbitrary fixed $\epsilon$ we obtain from \eqref{eq:dnplusone} $$d_{n+1} - \epsilon \le (1-\rho_n \mu) (d_n - \epsilon) + \rho_n (\rho_n B - \mu \epsilon).$$ When $n$ is large enough $\rho_n (\rho_n B - \mu \epsilon)\le 0$ and thus $$d_{n+1} - \epsilon \le (1-\rho_n \mu) (d_n - \epsilon),$$ therefore $$\left(d_{n+k} - \epsilon\right)_+ \le (1-\rho_n \mu) \left(d_n - \epsilon\right)_+.$$ Iterating such inequalities we obtain $$\left(d_{n+k} - \epsilon \right)_+ \le \prod_{\ell=n}^{n+k-1} (1-\rho_\ell \mu) \left(d_n - \epsilon \right)_+.$$ From the Lemma \ref{lemma:product} we obtain $\lim_{k \to \infty} \left(d_{k} - \epsilon \right)_+=0$ and since this is true for any $\epsilon$ the conclusion follows. \end{proof} \end{theorem} \begin{lemma} Let $\mu >0$ and $\rho_n$ a sequence of positive real numbers such that $\rho_n \to 0$ and $\sum_{n\ge 1} \rho_n = \infty$. Then for any $n \ge 0$: \begin{equation} \lim_{k\to \infty} \prod_{\ell= n}^{n+k} (1-\rho_\ell \mu) = 0. \end{equation} \label{lemma:product} \end{lemma} \begin{proof} Recall that for any $x \in ]0,1[$ we have $\log(1-x) \le -x$; then: \begin{equation} 0 \le \prod_{\ell= n}^{n+k} (1-\rho_\ell \mu) = e^{ \sum_{\ell= n}^{n+k} \log(1-\rho_\ell \mu)} \le e^{ \sum_{\ell= n}^{n+k} (-\rho_\ell \mu)} \overset{k\to \infty}{\longrightarrow} e^{- \infty} =0, \end{equation} which concludes the proof. \end{proof} \section{Concluding remarks} We make here some remarks concerning the hypothesis and the use in Neural Networks. First, consider the hypothesis $\sum_n \rho_n= \infty$; at first it may seem strange but this is not really so. Note that in particular it is true when $\rho_n$ is a constant. But in general, if we forget the stochastic part\footnote{This can be made precise when the stochastic part is added, see \cite{ayadi2020stochastic}.}, one can interpret the SGD as following some continuous time dynamics of the type $X'(t) = - \nabla \mathcal{L}(X)$; for the simple quadratic function $\mathcal{L}(X)= \alpha \|X\|^2 / 2$ the dynamics is $X'(t) = -\alpha X(t)$ with solution $X(t) = e^{-\alpha t}X(0)$ needs an infinite 'time' $t$ to converge to the minimum $X_*=0_N$. Or here $\sum_n \rho_n$ is the discrete version of the time and thus it is not a surprise to need infinite time to obtain $X_*$ with infinite precision. On the other hand if a finite precision is needed one can just take a constant time step as indicated in the theorem\footnote{but in this case one may spend a too long time to wait for the convergence to this small neighborhood to arrive see \cite{ayadi2020stochastic} for some ways to accelerate the convergence.}. On the other hand, an important example that satisfies \eqref{eq:hyprhon} is $\rho_n = \frac{c_1}{c_2 + n}$, with $c_1, c_2 > 0$. In general giving a functional form for $\rho_n$ is termed 'choosing a decay rate', but it may not be clear what the best decay rate is in general.
{ "timestamp": "2021-03-29T02:15:47", "yymm": "2103", "arxiv_id": "2103.14350", "language": "en", "url": "https://arxiv.org/abs/2103.14350", "abstract": "We give here a proof of the convergence of the Stochastic Gradient Descent (SGD) in a self-contained manner.", "subjects": "Machine Learning (stat.ML); Machine Learning (cs.LG); Probability (math.PR)", "title": "The convergence of the Stochastic Gradient Descent (SGD) : a self-contained proof", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9857180643966651, "lm_q2_score": 0.8128673223709251, "lm_q1q2_score": 0.8012580036187682 }
https://arxiv.org/abs/1404.2710
On Collatz' Words, Sequences and Trees
Motivated by a recent work of Trümper we consider the general Collatz word (up-down pattern) and the sequences following this pattern. The recurrences for the first and last sequence entries are given, obtained from repeated application of the general solution of a binary linear inhomogeneous Diophantine equation. These recurrences are then solved. The Collatz tree is also discussed.
\section{Introduction} The Collatz map $C$ for natural numbers maps an odd number $m$ to $3\,m\,+\, 1$ and an even number to \dstyle{\frac{m}{2}}. The {\sl Collatz} conjecture \cite{Lagarias}, \cite{WikiC}, \cite{WeissteinC} is the claim that every natural number $n$ ends up, after sufficient iterations of the map $C$, in the trivial cycle $(4,\, 2,\, 1)$. Motivated by the work of {\sl Tr\"umper} ~\cite{Truemper} we consider a general finite {\sl Collatz} word on the alphabet $\{u,\,d\}$, where $u$ (for `up') indicates application of the map $C$ on an odd number, and $d$ (for `down') for applying the map $C$ on an even number. The task is to find all sequences which follow this word pattern (to be read from the left to the right). These sequences will be called $CS$ (for {\sl Collatz} sequence also for the plural) realizing the CW (for {\sl Collatz} word also for the plural) under consideration. This problem was solved by {\sl Tr\"umper} \cite{Truemper} under the restriction that the first and last sequence entries are odd. Here we shall not use this restriction.The solution will be given in terms of recurrence relations for the first and last entries of the $CS$ for a given $CW$. This involves a repeated application of the general solution on positive numbers of the linear inhomogeneous {\sl Diophant}ine equation $a\,x\,+\, b\,y\,=\, c$, with $a\,=\, 3^m$ and $b\,=\, 2^n$ and given integer $c$. Because $gcd(3,2)\,=\, 1$ one will always have a countable infinite number of solutions. This general solution depends on a non-negative integer parameter $k$. We believe that our solution is more straightforward than the one given in \cite{Truemper}. \section{Collatz words, sequences and the Collatz tree}\label{section2} The {\sl Collatz} map $C:\ \mathbb{N} \to \mathbb{N},\ m \mapsto 3\,m\,+\, 1$ if $m$ is odd, \dstyle{m \mapsto \frac{m}{2}} if $m$ is even, leads to an increase $u$ (for `up') or decrease $d$ (for `down'), respectively. Finite {\sl Collatz} words over the alphabet $\{u,\,d\}$ are considered with the restriction that, except for the one letter word $u$, every $u$ is followed by a $d$, because $2\,m\,+\, 1 \,\mapsto\, 2\,(3\,m+1)$. This is the reason for introducing (with \cite{Truemper}) also $s\, :=\, ud$. Thus $s$ stands for $2\,m\,+\, 1\,\mapsto\, 3\,m\,+\, 1$. The general finite word is encoded by an $(S+1)-$tuple $\vec n_{S}\,=\, [n_0,\,n_1,\,...,\,n_S ]$ with $S\,\in\, \mathbb N$. \begin{eqnarray} \label{CW} CW(\vec n_{S+1}) &\,=\,& d^{n_0}\,s\,d^{n_1-1}\,s\, \cdots\, s\,d ^{n_S-1} \\ &\,=\, & (d^{n_0}\,s)\,(d^{n_1-1}\,s)\, \cdots\, \,(d ^{n_{S-1}-1}\,s)\, d ^{n_S-1}\ , \nonumber \end{eqnarray} with $n_0\,\in\, {\mathbb N}_0\, :=\, {\mathbb N}\, \cup\, \{0\} $, $n_i\,\in\, \mathbb N$, for $i\,=\, 1,\,2,\,...,\, S$. The number of $u$ (that is of $s\,=\, ud$) letters in the word $CW(\vec n_{S})$, or $CW(S)$, for short, is $S$ (which is why we have used $\vec n_S$ not $\vec n_{S+1}$ for the $S+1$ tuple), and the number of $d$ is \dstyle{D(S)\, :=\, \sum_{j=0}^{S}\,n_j }. In \cite{Truemper} $n_0\,=\, 0$ (start with an odd number), $y\,=\, S$ and $x\,=\, D(S)$.\par\smallskip\noindent Some special words are not covered by this notation: first the one letter word $u$ with the {\sl Collatz} sequence ($CS$) of length two $CS(u;\,k)\,=\, [2\,k\,+\, 1,\, 2\,(3\,k\,+\, 2)]$, and $CW([n_0])\,=\, d^{n_0}$ with the family of sequences $CS([n_0];k) \,=\, [2^{n_0}\,k,2^{n_0 - 1}\,k,\,...,\,1\,k] $ with $k\,\in\, {\mathbb N}_0$. \par\smallskip\noindent A {\sl Collatz} sequence $CS$ realizing a word $CW(\vec n_{S})$ is of length $L\,=\, D\,+\, S\,+\, 1$ and follows the word pattern from the left to the right. $CS(\vec n_{S})\,=\, [c_1,\, c_2,\,...,\,c_L]$. For example, $CW([1,2,1])\,=\, dsds$ with $S\,=\, 2$, $D(2)\,=\, 2$, and length $L\,=\, 7$ with $SC_0(\vec n_S) \,=\, [2, 1, 4, 2, 1, 4, 2]$ is the first of these sequences (for non-negative integers), the one with smallest start number $c_1$. In order to conform with the notation used in \cite{Truemper} we shall use for the start number $c_1\,=\, M$ and for the last number $c_L \,=\, N$. However, in \cite{Truemper} $M$ and $N$ are restricted to be odd which will not be the case here. Later one can get the words with odd start number M by choosing $n_0\,=\, 0$. In order to have also $N$ odd one has to pick from $SC_k([0,n_1,...,n_S])$, for $k\,\in\, \mathbb N_0$ only the odd members.\par\smallskip\noindent In \cite{Truemper} the monoid of Collatz words, with the unit element $e\,=\,$ {\it empty word} is treated. This will not be considered in this work. Also the connection to the $3\,m\,-\, 1$ problem will not be pursued here.\par\smallskip\noindent The {\sl Collatz} tree CT is an infinite (incomplete) ternary tree, starting with the root, the number $8$ on top at level $l=0$. Three branches, labeled $L$, $V$ and $R$ can be present: If a node (vertex) has label $n\,\equiv\, 4\,(mod\,6))$ the out-degree is $2$ with the a left edge (branch) labeled $L$ ending in a node with label \dstyle{\frac{n-1}{3}} and a right edge (label $R$) ending in the node labeled $2\,n$. In the other cases, $n\,\equiv\, 0,\,1,\,2,\,3,\,5,\,(mod\,6)$, with out-degree $1$, a vertical edge (label $V$) ends in the node labeled $2\,n$. The root labeled $8$ stands for the trivial cycle $8$\, repeat$(4,\,2,\,1)$. See the {Figure 1} for $CT_7$ with only the first eight levels. It may seem that this tree is left-right symmetric (disregarding the node labels) but this is no longer the case starting at level $l\,=\, 12$. At level $l=10$ the $mod\, 6$ structure of the left and right part of $CT$, also taking into account the node labels, is broken for the first time, but the node labels $4 (mod\, 6)$ are still symmetric. At the next level $l\,=\, 11$ the left-right symmetry concerning the labels $4 (mod\, 6)$ is also broken, leading at level $l\,=\, 12$ to a symmetry breaking in the branch structure of the left and right part of $CT$. Thus at level $l\,=\, 12$ the number of nodes becomes odd for the first time: $15$ nodes on the left side versus $14$ nodes on the right one. See rows $ l \,+\, 3$ of \seqnum{A127824} for the node labels of the first levels, and \seqnum{A005186}$(l+3)$ for the number of nodes. The number of $4\, (mod\, 6)$ nodes at level $l$ is given in \seqnum{A176866}$(l+4)$. \par\smallskip\noindent A $CS$ is determined uniquely from its start number $M$. Therefore no number can appear twice in $CT$, except for the numbers $1,\,2,\,4$ of the (hidden) trivial cycle. The Collatz conjecture is that every natural number appears in CT at some level ($1, 2,$ and $4$ are hidden in the root $8$). A formula for $l\,=\, l(n)$ would prove the conjecture. \par\smallskip\noindent Reading $CT$ from bottom to top, beginning with some number $M$ at a certain level $l$, recording the edge labels up to level $l=0$, leads to a certain $L,V,R$-sequence. E.g., $M\,=\, 40$ at level $l\,=\, 5$ generates the length $5$ sequence $[V,R,V,L,V]$. This is related to the CS starting with $M\,=\, 40$, namely $[40,20,10,5,16,8]$, one of the realizations of the CW $d,d,d,u,d\,=\, d^3s$, with $S\,=\, 1$ and $\vec n_1\,=\, [3,1]$. (Later it will be seen that this is the realizations with the third smallest start number, the smaller once are $8$ and $24$). One has to map $V$ and $R$ to $d$ and $L$ to $u$. This shows that the map from a $L,V,R$-sequence to a CW is not one to one. The numbers $n\,\equiv\, 4\, (mod\ 6)$ except $4$ (see \seqnum{A016957}) appear exactly in two distinct CS. For example, $64 \,\equiv\, 4\,(mod\, 6)$ shows up in all $CS$ starting at any vertex which descends from the bifurcation at $64$, {\it e.g.},\, $21,\,128$; $42,\,256$; $84,\,85,\,512$; \, {\it etc.}\, \par\bigskip\noindent \par\smallskip\noindent \parbox{16cm}{\begin{center} {\includegraphics[height=10cm,width=.8\linewidth]{CollatzTree}} \end{center} } \par\smallskip\noindent \hskip 6cm {\bf Figure: Collatz Tree $\bf CT_7$} \section{Solution of a certain linear inhomogeneous Diophantine equation}\label{section3} The derivation of the recurrence relations for the start and end numbers $M$ and $N$ of {\sl Collatz} sequences ($CS$) with prescribed up-down pattern (realizing a given $CW$) we shall need the general solution of the following linear and inhomogeneous {\sl Diophant}ine equation. \begin{equation}\label{Diophant} D(m,n;c):\hskip 2cm 3^m\,x\,-\, 2^n\,y\,=\, c(m,n),\ \ m\,\in\, \mathbb N_0, \ n\,\in\, \mathbb N_0,\ c(m,n) \,\in\, \mathbb Z\ . \end{equation} It is well known \cite{NZM}, pp. 212-214, how to solve the equation $a\,x\,+\, b\,y\,=\, c$ for integers $a,\, b$ (not $0$) and $c$ provided $g\,=\, gcd(a,\,b)$ divides $c$ (otherwise there is no solution) for integers $x$ and $y$. One will find a sequence of solutions parameterized by $t\,\in\, \mathbb Z$. Then one has to restrict the $t$ range to obtain all positive solutions. The procedure is to find first a special solution $(x_0,\, y_0)$ of the equation with $c\,=\, g$. Then the general solution is $(x\,=\, \frac{c}{g}\, x_0\,+\, \frac{b}{g}\, t,\, y\,=\, \frac{c}{g}\, y_0\,-\, \frac{a}{g}\, t)$ with $t\,\in\, Z$. The proof is found in \cite{NZM}. For our problem $g\,=\, gcd(3^m,\, 2^n)\,=\, 1$ for non-negative $m,\, n$ which will divide any $c(m,\, n)$. \par\bigskip\noindent \begin{lemma}\label{SolDiophant} {\bf Solution of $\bf D(m,n;c)$}\par\smallskip\noindent {\bf a)} A special positive integer solution of D(m,n;1) is\par\smallskip\noindent \begin{eqnarray}\label{x0y0} y_0(m,n) &\,=\,& \left(\frac{3^m\,+\, 1}{2}\right)^{n\,+\, 3^{m-1}}\, (mod\, 3^m)\ , \nonumber\\ x_0(m,n) &\,=\,& \frac{1\,+\, 2^n\,y_0(m,n)}{3^m}\ . \end{eqnarray} {\bf b)} The general solution with positive $x$ and $y$ is\par\smallskip\noindent \begin{eqnarray}\label{Solxy} x(m,n) &\,=\,& c(m,n)\,x_0(m,n)\,+\, 2^n\,t_{min}(m,n;sign(c)) \,+\, 2^n\,k \ ,\nonumber \\ y_0(m,n) &\,=\,& c(m,n)\,y_0(m,n)\,+\, 3^m\,t_{min}(m,n;sign(c)) \,+\, 3^m\,k \ , \end{eqnarray} with $k\,\in\, \mathbb N_0$, and \begin{equation}\label{tmin} t_{min}(m,n;sign(c))\,=\, {\Caseszwei{\ceil{\abs{c(m,n)}\,\frac{x_0(m,n)}{2^n}}}{${\rm if} \ c\,<\, 0 \ ,$} {\ceil{-c(m,n)\,\frac{y_0(m,n)}{3^m}}} {${\rm if} \ c\,\geq\, 0\ .$}} \end{equation} \end{lemma} \noindent For the proof we shall use the following {\sl Lemma}: \begin{lemma}\label{Anm} $ A(n,m)\, :=\, {\binomial{n-1}{m-1}}\, \frac{gcd(m,n)}{m}$\, is a positive integer for $ m\,=\, 1,\,2\, ...\, n,\ n\,\in\, \mathbb N$\, . \end{lemma} \begin{proof}\par\smallskip\noindent [due to {\sl Peter Bala}, see \seqnum{A107711}, history, Feb 28 2014]:\par\noindent This is the triangle \seqnum{A107711} with A(0,0) =1. By a rearrangement of factors one also has \dstyle{A(n,m)\,=\, {\binomial{n}{m}}\,\frac{gcd(n,m)}{n}}. Use $gcd(n,m)\,lcm(m,n)\,=\, n\, m$ ({\it e.g.},\, \cite{FR}), theorem 2.2.2., pp. 15-16, where also the uniqueness of the $lcm$ is shown). \dstyle{A(n,m)\,=\, \frac{a(n,m)}{lcm(n,m)}} with \dstyle{a(n,m)\,=\, {\binomial{n}{m}}\,m}, a positive integer because the binomial is a combinatorial number. $m\,|\, a(n,m)$ and $n\,|\, a(n,m)$ because \dstyle{a(n,m) \,=\, n\,{\binomial{n-1}{m-1}}} by a rearrangement. Hence $a(n,m)\,=\, k_1\,m\,=\, k_2\,n$, {\it i.e.},\, $a(n,m)$ is a common multiple of $n$ and $m$ (call it $cm(n,m)$). $lcm(n,m)\,|\, a(n,m)$ because $lcn(n,m)$ is the (unique) lowest $cm(n,m)$. Therefore \dstyle{\frac{a(n,m)}{lcm(n,m)}\,\in\, \mathbb N}, since only natural numbers are in the game. \end{proof}\noindent Now to the proof of {\sl Lemma\ 1}.\par\noindent \begin{proof} \par\smallskip\noindent {\bf a)} \dstyle{x_0(m,n) \,=\, \frac{1\,+\, 2^n\,y_0(m,n)}{3^m}} is a solution of $D(m,n;1)$ for any $y_0(m,n)$. The given $y_0(m,n)$ is a positive integer $\,\in\, \{1,\,2,\,...,\, 3^{m}\,-\, 1\},\ m\,\in\, \mathbb N$ and $y_0(0,n)\,=\, 1$ for $n\,\in\, \mathbb N_0$. One has to prove that $x_0(m,n)$ is a positive integer. This can be done by showing that $1\,+\, 3^n\,y_0(n,m)\,\equiv\, 0\,(mod\,3^m)$ for $m\,\in\, \mathbb N$. One first observes that \dstyle{\frac{3^m\,+\, 1}{2}\,\equiv\, \frac{1}{2}\,(mod\, 3^m)}, because obviously \dstyle{2\,\frac{3^m\,+\, 1}{2} \,\equiv\, 1\, (mod\, 3^m)} ($2$ is a unit in the ring $\mathbb Z_{3^m}$). For $m\,=\, 0$ one has $x_0(0,n) \,=\, 1\,+\, 2^n$, $n\,\in\, \mathbb N_0$, which is positive. In the following $m\,\in\, \mathbb N$. \begin{equation} 1\,+\, 2^n\,\left(\frac{3^m\,+\, 1}{2}\right)^{n+3^{m-1}} \,\equiv\, 1 + 2^n\, \left(\frac{1}{2}\right)^{n+3^{m-1}} \,\equiv\, 1\,+\, \left(\frac{1}{2}\right)^{3^{m-1}}\, (mod\, 3^m)\ . \end{equation} Now we show that \dstyle{L(m)\, :=\, \left(\frac{3^m\,+\, 1}{2}\right)^{3^{m-1}}\,\equiv\, 0\, (mod\, 3^m)} by using \dstyle{\frac{3^m\,+\, 1}{2} \,=\, 3\, k(m)\,-\, 1} with \dstyle{k(m)\, :=\, \frac{3^{m-1}\sspp1}{2}}, a positive integer. The binomial theorem leads with $ a(m)\,=\, 3^{m\,-\, 1}$ to \begin{eqnarray} (3\,k(m)\,-\, 1)^{a(m)} &\,=\,& \sum_{j=0}^{a(m)-1}\,{\binomial{a(m)}{j}}\, (-1)^j\, (3\,k(m))^{a(m)-j} \nonumber \\ &\,=\,& 3^m\,\Sigma_1(m) \,+\, \Sigma_2(m), {\rm with} \\ \Sigma_1(m) &\,=\,& \sum_{j=0}^{a(m)-m}\, (-1)^j\,{\binomial{a(m)}{j}}\,k(m)^{a(m)-j}\,3^{a(m)-m-j}, \ {\rm and} \\ \Sigma_2(m) &\,=\,& \sum_{j=a(m)-m+1}^{a(m)-1}\, (-1)^j\,{\binomial{a(m)}{j}}\,(3\, k(m))^{a(m)-j}\ . \end{eqnarray} $\Sigma_1(m)$ is an integer because of $a(m)-j\,\geq\, a(m)-m-j\,\geq\, 0$ and the integer binomial, hence $L(m)\,\equiv\, \Sigma_2 \,(mod\, 3^m)$. Rewriting $\Sigma_2$ with $j'\,=\, j-a(m)+m-1$, using also the symmetry of the binomial, one has \begin{eqnarray} \Sigma_2(m)&\,=\,& \sum_{j=0}^{m-2}\,{\binomial{a(m)}{j}}\, (-1)^{j-m}\,(3\,k(m))^{m-1-j}\nonumber \\ &\,=\,& \sum_{j=1}^{m-1}\,(-1)^{j+1}\, {\binomial{a(m)}{j}}\, (3\,k(m))^j \,=\, 3^m\,\widehat\Sigma_2(m)\, \ {\rm with}\\ \widehat\Sigma_2(m)&\,=\,& \sum_{j=1}^{m-1}\,(-1)^{1+j}\,{\binomial{a(m)}{j}}\,k(m)^j\,3^{j-m}\nonumber \\ &\,=\,& \sum_{j=1}^{m-1}\,(-1)^{1+j}\,k(m)^j\,{\binomial{a(m)-1}{j-1}}\,\frac{1}{j}\,3^{j-1}\ . \end{eqnarray} In the last step a rearrangement of the binomial has been applied, remembering that $a(m)\,=\, 3^{m-1}$. It remains to be shown that \dstyle{A_{m,j}\, :=\, 3^{j-1}\,{\binomial{3^{m-1}-1}{j-1}}\,\frac{1}{j}} is a(positive) integer for $j\,=\, 1,\,2,\,...\,m-1$. Here {\sl Lemma 2} comes to help. Consider there $A(3^{m-1},j)$ for $j\,=\, 1,\,2,\,...,\,m-1$ ($m=0$ has been treated separately above), which is a positive integer. If $3\, \not|\, j$ then $3^{j-1}\,A(3^{m-1},j) \,=\, A_{m,j}$, hence a positive integer. If $j\,=\, 3^k\,J$, with $k\,\in\, \mathbb N$ the largest power of $3$ dividing $j$ then $gcd(3,J)\,=\, 1$, and $j\,=\, 3^k\,J \sspleq m-1\,<\, 3^{m-1}$ and $gcd(3^{m-1},3^k\,J)\,=\, 3^q$ with $q\,=\, min(k,m-1)$. \end{proof}\noindent \begin{proof} \par\smallskip\noindent {\bf b)} The general integer solution of ~\ref{Diophant} is then (see \cite{NZM}, pp. 212-214; note that there $b>0$, here $b<0$, and we have changed $t\mapsto -t$) \begin{eqnarray} x\,=\, \hat x(m,n;t)&\,=\,& c(m,n)\,x_0(m,n)\,+\, 2^n\,t\, ,\nonumber \\ y\,=\, \hat y(m,n;t)&\,=\,& c(m,n)\,y_0(m,n)\,+\, 3^m\,t\, , \ t\,\in\, \mathbb Z\,. \end{eqnarray} In order to find all positive solutions for $x$ and $y$ one has to restrict the $t$ range, depending on the sign of $c$. If $c(m,n)\,\geq\, 0$ then, because $x_0$ and $y_0$ are positive and \dstyle{\frac{x_0(m,n)}{2^n}\,=\, \frac{y_0(m,n)}{3^m}\,+\, \frac{1}{2^n\,3^m}},\ \dstyle{t\,>\, -\frac{c(m,n)\,x_0(m,n)}{2^n}} and \dstyle{t\,>\, -\frac{c(m,n)\,y_0(m,n)}{3^m}}, {\it i.e.},\, \par\noindent \dstyle{t\,\geq\, \ceil{max\left( - \frac{c(m,n)\,x_0(m,n)}{2^n},\, -\frac{c(m,n)\,y_0(m,n)}{3^m}\right)}} \dstyle{\,=\, \ceil{-c(m,n)\,min\left(\frac{x_0(m,n)}{2^n},\, \frac{y_0(m,n)}{3^m} \right)}} \par\noindent \dstyle{\,=\, \ceil{-c(m,n)\, \frac{y_0(m,n)}{3^m} }\,=\, t_{min}(m,n;+)}. \par\noindent If $c(m,n)\,<\, 0$ then \dstyle{t\,\geq\, \ceil{ \abs{c(m,n)}\, max\left(\frac{x_0(m,n)}{2^n},\, \frac{y_0(m,n)}{3^m}\right)}\,=\, \ceil{\abs{c(m,n)}\, \frac{x_0(m,n)}{2^n}}} $\,=\, t_{min}(m,n;-)$. Thus with $t\,=\, t_{min}(m,n;sign(c)) \,+\, k$, with $k\,\in\, \mathbb N_0$ one has the desired result. Note that $(x_0(m,n),\, y_0(m,n))$ is the smallest positive solution of the equation $D(m,n;1)$, eq.~\ref{Diophant}, because, for $c(m,n)\,=\, 1$, $t_{min}(m,n;+) \,=\, \ceil{-\frac{y_0(m,n)}{3^m}}$, but with $y_0(m,n)\,\in\, \{1,\,2,\,...\,3^m-1\}$ this is $0$. \end{proof} \noindent A proposition on the periodicity of the solution $y_0(m,n)$ follows.\par\smallskip\noindent \begin{proposition} \label{y0Periodicity} {\bf Periodicity of $\bf y_0(m,n)$ in $\bf n$}\par\smallskip\noindent {\bf a)} The sequence $y_0(m,n)$ is periodic in $n$ with primitive period length $L_0\,=\, \varphi(3^m)$, for $m\,\in\, \mathbb N_0\, $ with {\sl Euler}'s totient function $\varphi(n)\,=\,$\seqnum{A000010}$(n)$, where $\varphi(1)\, :=\, 1$. \par\smallskip\noindent {\bf b)} The sequence $x_0(m,n\,+\, L_0(m))\,=\, q(m)\,x_0(m,n)\,-\, r(m)$, $m\,\in\, \mathbb N_0$, with $q(m)\, :=\, 2^{\varphi(3^m)}$ and \dstyle{r(m) \, :=\, \frac{2^{\varphi{(3^m)}}\,-\, 1}{3^m}}. See \seqnum{A152007}.\par\smallskip\noindent {\bf c)} The set $Y_0(m)\, :=\, \{y_0(m,n) \,|\, n\,=\, 0,\,1,\,...,\, \varphi(3^m)\,-\, 1\} $, is, for $m\,\in\, \mathbb N_0$, a representation of the set $RRS(3^m)$, the smallest positive restricted residue system modulo $3^m$. See \cite{TApostol} for the definition. The multiplicative group modulo $3^m$, called \dstyle{\mathbb Z_{3^m}^{\times}\,=\, (\mathbb Z/3^m\,\mathbb Z)^\times} is congruent to the cyclic group $C_{\varphi(3^m)}$. See, e.g., \cite{WikiRRS},\par\smallskip\noindent \end{proposition} \begin{proof}\par\smallskip\noindent {\bf a)} By {\sl Euler}'s theorem ({\it e.g.},\, \cite{FR}, theorem 2.4.4.3 on p. 32) $a^{\varphi(n)} \,\equiv\, 1\, (mod\, n)$, provided $gcd(a,n)\,=\, 1$. Now \dstyle{gcd\left(\frac{3^m+1}{2},3^m\right) \,=\, gcd\left(\frac{3^m+1}{2},3\right)\,=\, 1} because \dstyle{\frac{3^m+1}{2} \,\equiv\, \frac{1}{2}\, (mod\, 3^m)} (see above) and hence \dstyle{\frac{3^m+1}{2} \,\not\equiv \, 0\, (mod\, 3^m)}. This shows that $L_0(m)$ is a period length, but we have to show that it is in fact the length of the primitive period, {\it i.e.},\, we have to prove that the order of \dstyle{\frac{3^m+1}{2}} modulo $3^m$ is $L_0(m)$. (See {\it e.g.},\, \cite{FR}, Definition 2.4.4.1. on p.31, for the order definition.) In other words we want to show that \dstyle{\frac{3^m+1}{2}} is a primitive root (of $1$) modulo $3^m$. Assume that $k(m)$ is this order (the existence is certain due to {\sl Euler}'s theorem), hence $(\frac{1}{2})^{k(m)} \,\equiv\, 1\, (mod\, 3^m)$ and $k(m)\,|\, L_0(m)$. It is known that the module $3^m$ possesses primitive roots, and the theorem on the primitive roots says that there are precisely $\varphi(\varphi(3^m))$ incongruent ones ({\it e.g.},\, \cite{NZM}, pp. 205, 207, or \cite{Nagell}, theorem 62, 3., p. 104 and theorem, 65, p. 107). In our case this number is $\varphi(2\cdot 3^{m-1}) \,=\, 2\cdot 3^{m-2}$ if $m \,\geq\, 1$. The important point, proven in \cite{Nagell}, theorem 65.3 on p. 107, is that if we have a primitive root $r$ modulo an odd prime, here $3$, then, if $r^{3-1}\,-\, 1$ is not divisible by $3^2$, it follows that $r$ is in fact a primitive root for any modulus $3^q$, with $q\,\in\, \mathbb N_0 $. One of the primitive roots modulo $3$ is $2$, because $2^2\,=\, 4\,\equiv\, 1\, (mod\, 3)$ and $2^1\,\not\equiv \, 1\, (mod\ 3)$. Also $2^{3-1}\,-\, 1 \,=\, 3$ is not divisible by $3^2$, hence $2$ is a primitive root of any modulus $3^q$ for $q\,\in\, \mathbb N_0$. From this we proof that \dstyle{\frac{3^m+1}{2}\,\equiv\, \frac{1}{2}\,(mod\, 3^m)} is a primitive root modulo $3^m$. Consider \dstyle{\left(\frac{3^m+1}{2}\right)^k\,\equiv\, \frac{1}{2^k}\,(mod\, 3^m)} for $k\,=\, 1,\,2,\,...,\, \varphi(3^m)$. In order to have \dstyle{\left(\frac{1}{2}\right )^k \,\equiv\, 1\, (mod\, 3^m)} one needs $2^k\,\equiv\, 1\, (mod\, 3^m)$. But due to \cite{Nagell} theorem 65.3. p. 107, for $p\,=\, 3$, a primitive root modulo $3^m$ is $2$, and the smallest positive $k$ is therefore $\varphi(3^m)$, hence \dstyle{\frac{3^m+1}{2}} is a primitive root (of $1$) of modulus $3^m$. \par\smallskip\noindent {\bf b)} \dstyle{x_0(m,n\,+\, \varphi(3^m))\,=\, \frac{1\,+\, 2^n\,2^{\varphi(3^m)}\,y_0(m,n)}{3^m}} from the periodicity of $y_0$. Rewritten as \dstyle{\frac{2^{\varphi(3^m)}\,\left( ( 2^{-\varphi(3^m)} \,-\, 1) \,+\, (1\,+\, 2^n\,y_0(m,n) \right)} {3^m}\,=\,} \dstyle{-\frac{1}{3^m}\,(2^{\varphi(3^m)}\,-\, 1) \,+\, 2^{\varphi(3^m)}\,x_0(m,n)\,=\, } \dstyle{ q(m)\, x_0(m,n) \,-\, r(m)} with the values given in the {\sl Proposition}.\par\smallskip\noindent {\bf c)} This follows from the reduced residue system modulo $3^m$ for $m\,\in\, \mathbb N_0$, \par\noindent \dstyle{\left\{ \left(\frac{1}{2}\right)^0,\,\left(\frac{1}{2}\right)^1,\, ...,\, \left(\frac{1}{2}\right)^{\varphi(3^m)\sspm1} \right\}}, because \dstyle{\frac{1}{2}} is a primitive root modulo $3^m$ (from part {\bf b)}). With $a(m)\, :=\, \frac{3^m\,+\, 1}{2}$ one has $1\,=\, gcd(a(m),3)\,=\, gcd(a(m),3^m)\,=\, gcd(a(m)^{b(m)},3^m)$ with $b(m)\, :=\, 3^{m\,-\, 1}$, also\par\noindent \dstyle{\left\{a(m)^{b(m)}\, \left(\frac{1}{2}\right)^0,\,a(m)^{b(m)}\, \left(\frac{1}{2}\right)^1,\,...,\,a(m)^{b(m)}\, \left(\frac{1}{2}\right)^{\varphi(3^m)\sspm1}\right\}} is a reduced residue system modulo $3^m$ (see \cite{TApostol}, theorem 5.16, p. 113). Thus \par\noindent \dstyle{Y_0(m)\,\equiv\, \{a(m)^{b(m)}\,1,\, a(m)^{b(m)+1},\,...,\, a(m)^{b(m) \,+\, \varphi(3^m)\,-\, 1}\}} is a reduced residue system modulo $3^m$. Therefore this gives a permutation of the reduced residue system modulo $3^m$ with the smallest positive integers sorted increasingly. \end{proof} \par\smallskip\noindent \begin{example} For $m\,=\, 3$, $\varphi(3^3)\,=\, 2\cdot 3^2\,=\, 18\,=\, L_0(3)$, \par\noindent $\{y_0(3,n)\}_{n=0}^{17}\,=\, \{26,\, 13,\, 20, \,10,\, 5,\, 16,\, 8,\, 4,\, 2, \,1,\, 14,\, 7,\, 17,\, 22,\, 11,\, 19,\, 23,\, 25\}$ a permutation of the standard reduced residue system modulo $27$, obtained by resorting the found system increasingly. See \seqnum{A239125}. For $m=1,\, 2$ and $ 4$ see \seqnum{A007583},\ \seqnum{A234038} and \seqnum{A239130} for the solutions $(x_0(m,n),\, y_0(m,n))$. \end{example} \par\smallskip\noindent \section{Recurrences and their solution}\label{section4} After these preparations it is straightforward to derive the recurrence for the start and end numbers $M$ and $N$ for any given $CW(\vec n_S)$, for $S\,\in\, \mathbb N$. \par\smallskip\noindent {\bf A)} We first consider the case of words with $n_S\,=\, 1$. {\it i.e.},\, $\vec n_S\,=\, [n_0,\,n_1\,,...\,n_{S-1},\,1]$. This is the word $CW(\vec n_S )\,=\, \rprod_{j=0}^{S-1}\, d^{n_j}\, s$ (with an ordered product, beginning with $j=0$ at the left-hand side). In order to simplify the notation we use $M(S)$, $N(S)$, $y_0(S)$, $x_0(S)$, and $c(S)$ for $ M(\vec n_S )$, $N(\vec n_S)$ , $y_0(S,n_S)$, $x_0(S,n_S)$ and $c(S,n_S)$, respectively. For $S=1$, the input for the recurrence, one has \begin{equation} M(1;k) \,=\, 2^{n_0}\,(2\,k+1)\ \text{and}\ N(1;k) \,=\, 3\,k\,+\, 2,\ {\text for}\ k \,\in\, \mathbb N\ , \end{equation} because there are $n_0$ factors of $2$ from $d^{n_0}$, and then an odd number $2\,k\,+\, 1$ leads after application of $s$ to $3\,k\,+\, 2$. Thus $M(1)\,=\, 2^{n_0}$ and $N(1)\,=\, 2$.\par\smallskip\noindent \begin{proposition} {\bf Recurrences for $\bf M(S)$ and $\bf N(S)$ with $\bf n_S\,=\, 1$} \label{Rec}\par\smallskip\noindent {\bf a)} The coupled recurrences for $M(S,t)$ and $N(S,t)$, the first and last entry of the {\sl Collatz} sequences $CS(\vec n_S;t)$ for the word $ CW(\vec n_S)$ with $\vec n_S\,=\, [n_0,\,n_1,\,...\, n_{S-1},1]$ ($n_S\,=\, 1$) are \par\smallskip\noindent \begin{eqnarray} M(S,t)\,=\, M(S)&\,+\,& 2^{\hat D(S)}\, t\ , \nonumber \\ N(S,t)\,=\, N(S)&\,+\,& 3^S\,t\, ,\ {\rm with}\ t\,\in\, \mathbb Z\ , \end{eqnarray} where \dstyle{\hat D(S)\, :=\, \sum_{j=0}^{S-1}\,n_j} (we prefer to use a new symbol for the $n_S\,=\, 1$ case), and the recurrences for $M(S)$ and $\widetilde N(S)\,=\, N(S)\,-\, 2$ are \begin{eqnarray} M(S) \,=\, M(S-1) &\,+\,& 2^{\hat D(S-1)}\,c(S-1)\, x_0(S-1) \ , \nonumber \\ \widetilde N(S) &\,=\,& 3\,y_0(S-1)\, c(S-1)\, \end{eqnarray} with \begin{equation} c(S-1) \,=\, 2\,(2^{n_{S-1}-2}\,-\, 1) \,-\, \widetilde N(S-1) \,=:\, A(S-1) \,-\, \widetilde N(S-1)\ . \end{equation} The recurrence for $c(S)$ is\par\smallskip\noindent \begin{equation} c(S) \,=\, -3\,y_0(S-1)\,c(S-1) \,+\, A(S)\ , S\,\geq\, 2, \end{equation} and the input is $M(1)\,=\, 2^{n_0}$, $\widetilde N(1)\,=\, 0$ and $c(1) = A(1)$.\par\smallskip\noindent {\bf b)} The general positive integer solution is \begin{eqnarray} M(S;k) &\,=\, & M(S) \,+\, 2^{\hat D(S)}\,t_{min}(S-1) \,+\, 2^{\hat D(S)}\,k\,,\nonumber \\ N(S;k) &\,=\, & 2\,+\, \widetilde N(S) \,+\, 3^S\,t_{min}(S-1) \,+\, 3^S\,k, \ k\,\in\, \mathbb N_0 \, , \end{eqnarray} where \begin{equation} t_{min}(S) \,=\, t_{min}(S, n_S,sign(c(S))) \,=\, {\Caseszwei{\ceil{\abs{c(S)}\,\frac{x_0(S)}{2^{n_S}}}}{${\rm if} \ c(S)\,<\, 0 \ ,$} {\ceil{-c(S)\,\frac{y_0(S)}{3^S}}} {${\rm if} \ c(S)\,\geq\, 0\ .$}} \end{equation} \end{proposition} \begin{corollary} \begin{eqnarray} M(S;k) &\,\equiv\,& M(S) \,+\, 2^{\hat D(S)}\,t_{min}(S-1)\, (mod\, 2^{\hat D(S)})\, , \nonumber \\ N(S;k) &\,\equiv\,& \widetilde N(S) \,+\, 3^S\,t_{min}(S-1)\, (mod\, 3^S)\, . \end{eqnarray} \end{corollary} \par\smallskip\noindent In Terras' article \cite{Terras} the first congruence corresponds to {\sl theorem 1.2}, where the encoding vector $E_k(n)$ refers to the modified {\sl Collatz} tree using only $d$ and $s$ operations. \begin{proof} \par\smallskip\noindent {\bf a)} By induction over $S$. For $S\,=\, 1$ the input $M(1)\,=\, 2^{n_0}$, $N(1)\,=\, 2$ or $\widetilde N(1)\,=\, 0$ provides the start of the induction. Assume that part {\bf a)} of the proposition is true for $S$ values $1,\,2,\,...,\,S-1$. To find $M(S)$ one has to make sure that $d^{n_{S-1}}\,s$ can be applied to $N(S-1;k)$, the end number of step $S-1$ sequence $CS(\vec n_{S-1};t)$ which is $N_{int}(S-1,t)\,=\, N(S-1) +3^{S-1}\, t$, with integer $t$, by the induction hypothesis. This number has to be of the form $2^{n_{S-1}-1}\,(2\, m\sspp1)$ (one has to have an odd number after $n_{S-1}$ $d-$steps such that $s$ can be applied). Thus \dstyle{3^{S-1}\,t \,-\, 2^{n_{S-1}}\,m \,=\, 2^{n_{S-1}-1}\,-\, N(S-1) \,=\, A(S-1)\,-\, \widetilde N(S-1)\,=:\, c(S-1)}, where $\widetilde N(S-1)\,=\, N(S-1) \,-\, 2$ and $ A(S-1)\,=\, 2\,(2^{n_{S-1}-2}\,-\, 1)$. Due to {\sl Lemma}~\ref{SolDiophant} the general solution, with $t\,\to\, x(S-1,n_{S-1};t)\, \hat =\, x(S-1;t)$, $m\,\to\, y(S-1,n_{S-1};t)\, \hat =\, y(S-1;t)$, to shorten the notation, is \begin{eqnarray} t\,\to\, x(S;t) &\,=\,& c(S-1)\,x_0(S-1) \,+\, 2^{n_{S-1}}\,t \, , \nonumber \\ m\,\to\, y(S;t) &\,=\,& c(S-1)\,y_0(S-1) \,+\, 3^{S-1}\,t\, , \ t\,\in\, \mathbb Z\ . \end{eqnarray} Therefore the first entry of the sequence $CS(\vec n_S;t)$ is \dstyle{M(S;t)\,=\, M(s-1,x(S-1,t))} which is \begin{equation} M_{int}(S;t)\,=\, M(S-1) \,+\, 2^{\hat D(S-1)}\,c(S-1)\,x_0(S-1) \,+\, 2^{\hat D(S)}\, t\, , \end{equation} hence \dstyle{M(S)\,=\, M(S-1)\,+\, 2^{\hat D(S-1)}\,c(S-1)\, x_0(S-1)}, the claimed recurrence for $M(S)$.\par\noindent The last member of $CS(\vec n_{S-1};t)$ is $3\,m+2$ (after applying $s$ on $2\,m\,+\, 1$ from above). Thus \dstyle{N_{int}(S;t) \,=\, 3\,y(S;t)\,+\, 2}, or \dstyle{N_{int}(S;t)\,-\, 2\,=\, 3\,c(S-1)\,y_0(S-1)\,+\, 3^S\, t}. Therefore, $\widetilde N(S) \,=\, N(S)\,-\, 2 \,=\, 3\,c(S-1)\,y_0(S-1)$ the claim for the $\widetilde N$ recurrence. Note that the remainder structure of eqs. $(20)$ and $(21)$, expressed also in the {\sl Corollary}, has also been verified by this inductive proof. The recurrence for $c(S)\,=\, A(S) \,-\, \widetilde N(S)$ follows from the one for $\widetilde N(S)$. \par\smallskip\noindent {\bf b)} Positive integer solutions from $M_{int}(S;t)$ and $N_{int}(S;t)$ of part {\bf a)} are found from the second part of {\sl Lemma} ~\ref{SolDiophant} applied to the equation $3^{S-1}\, x \,-\, 2^{n_{S-1}}\,y\,=\, c(S-1)$, determining $t_{min}(S-1)$ as claimed. This leads finally to the formulae for $M(S;k)$ and $N(S;k)$ with $k\,\in\, \mathbb N_0$. \end{proof} \begin{example} {\bf $\bf (sd)^{S-1}\,s$ Collatz sequences}\par\smallskip\noindent Here $n_0 \,=\, 0 \,=\, n_S $ and $n_{j}\,=\, 2$ for $j\,=\, 1,\,2\, ...,\,S-1$. The first entries $M(S;k)$ and the last entries $N(S;k)$ of the {\sl Collatz} sequence $CS([0,2,...,2];k)$ (with $S-1$ times a $2$), whose length is $3\,S$, are $M(S;k)\,=\, 1\,+\, 2^{2\,S-1}\,k$ and $N(S;k)\,=\, 2 \,+\, 3^S\,k$. For $S=3$ a complete {\sl Collatz} sequence $CS([0,2,2];3)$ of length $9$ is $[97, 292, 146, 73, 220, 110, 55, 166, 83]$ which is a special realization of the word $sdsds$ with start number $M(3;3)\,=\, 97$ ending in $N(3;3)\,=\, 83$. Note that for this $u-d$ pattern the start and end numbers have remainders $ M(S;0) = M(1;0) \,=\, 1$ and $N(S;0)\,=\, N(1;0) \,=\, 2$. See the tables \seqnum{A240222} and \seqnum{A240223}. \end{example}\par\smallskip\noindent The recurrences for $M(S)\, \hat =\, M(\vec n_{S-1})$, $\tilde N(S)\, \hat =\, \tilde N(\vec n_{S-1})$ or $N(S)\, \hat =\, N(\vec n_{S-1})$ and $c(S)\, \hat =\, c(\vec n_{S-1}$ are solved by iteration with the given inputs $M(1)\,=\, 2^{n_0}$, $\tilde N(1)\,=\, 0$ and $c(1)\,=\, A(1)\,=\, 2\,(2^{n_0-2}\,-\, 1)$. \par\smallskip\noindent \begin{proposition} {\bf Solution of the recurrences for $\bf n_s\,=\, 1$} \label{SolRec}\par\smallskip\noindent The solution of the recurrences of {\sl Proposition}~\ref{Rec} with the given inputs are, for $S\,\in\, \mathbb N$:\par\smallskip\noindent \begin{eqnarray} c(S)&\,=\,& A(S) \,+\, \sum_{j=1}^{S-1} \,(-3)^j\,A(S-j)\,\prod_{l=1}^j\,y_0(S-l) \, , \nonumber \\ \tilde n(S)&\,=\,& A(S) \,-\, c(S) \,=\, -\sum_{j=1}^{S-1}\,(-1)^j\,A(S-j)\, \prod_{l=1}^{j}\,y_0(S-l)\, , \nonumber \\ N(s)&\,=\,& \tilde N(S)\,+\, 2\, , \nonumber \\ M(S) &\,=\,& 2^{n_0}\,+\, \sum_{j=1}^{S-1}\,R(S-j)\, , \end{eqnarray} with \dstyle{\hat D(S)\, :=\, 1\,+\, \sum_{j=0}^{S-1}\, n_j},\ \dstyle{A(S)\, :=\, 2\,(2^{n_S-2}\,-\, 1)},\ \dstyle{R(S)\, :=\, 2^{\hat D(S)} \,x_0(S)\,c(S)}\ and $y_0(S)\, \hat =\, y_0(S,n_S)$, $x_0(S)\, \hat =\, x_0(S,n_S)$, given in {\sl Lemma} ~\ref{SolDiophant}. \end{proposition} \begin{proof}\par\smallskip\noindent This is obvious. \end {proof} \par\noindent {\bf B)} The general case $n_S \,\geq\, 1$ can now be found by appending the operation $d^{n_S-1}$ to the above result. This leads to the following {\sl theorem}.\par\smallskip\noindent \begin{theorem} \label{GenSol} {\bf The general case $\bf {\overrightarrow n}_S$} \par\smallskip\noindent For the Collatz word $CW(\vec n_S)\,=\, d^{n_0}\,\rprod_{j=1}^{S}\,(s\,d^{n_j-1}) \,=\, d^{n_0}\, s\,\rprod_{j=1}^{S}\,(d^{n_j-1}\, s)\,d^{n_S-1}$ (the ordered product begins with $j=1$ on the left-hand side) with $n_0\,\in\, \mathbb N_0$ , $n\,\in\, \mathbb N$, the first and last entries of the corresponding Collatz sequences $\{CS(\vec n_S;k)\}$, of length $L(S) \,=\, n_0 \,+\, 2\,\sum_{j=1}^S n_j$, for $k\,\in\, \mathbb N_0$, are\par\smallskip\noindent \begin{eqnarray} M(\vec n_S;k) &\,=\,& M(S) \,-\, 2^{\hat D(S)}\,N(S)\,x_0(S,n_S-1)\,+\, 2^{D(S)}\,t_{min}(S,n_S-1,sign(c_{new}(S)) \nonumber\\ && \,+\, 2^{D(S)}\,k\, , \nonumber \\ N(\vec n_S;k) &\,=\,& c_{new}(S)\,y_0(S,n_S-1)\,+\, 3^S\, t_{min}(S,n_S-1,sign(c_{new}(S))\sspp3^S\, k\ , \end{eqnarray} \end{theorem}\noindent with $c_{new}(S)\, :=\, -N(S)$, $\hat D(S) \,=\, 1\,+\, \sum_{j=0}^{S-1}\,n_j$, $D(S)\,=\, \sum_{j=0}^S\, n_j$. \begin {proof}\par\smallskip\noindent In order to be able to apply to the {\sl Collatz} sequences $CS([n_0,n_1,...,n_S-1,1])$ (with the results from part A) above) the final $d^{n_S-1}$ operation one needs for the last entries $N_{int}(S;t) \,=\, N(S) \,+\, 3^S\,t \,=\, 2^{n_S-1}\,m$ with some (even or odd) integer m. The new last entries of $CS([n_0,n_1,...,n_S];t)$ will then be $m$. The general solution of $3^S\,-\, 2^{n_S-1}\,m\,=\, -N(S)\,=:\, c_{new}(S)$ is according to {\sl Lemma}~\ref{SolDiophant} \par\smallskip\noindent \begin{eqnarray} t\,\to\, x(S;t)&\,=\,& c_{new}(S)\, x_0(S,n_S-1)\,+\, 2^{n_S-1}\,t,\, \nonumber \\ m\,\to\, y(S;t)&\,=\,& c_{new}(S)\, y_0(S,n_S-1)\,+\, 3^S\,t,\,\ t\,\in\, \mathbb Z\, . \end{eqnarray} This leads to positive integer solutions after the shift $t\,\to\, t_{min}\,+\, k$, with \par\noindent $t_{min} \,=\, t_{min}(S,n_S-1,sign(c_{new}(S)))$ to the claimed result $N(S;k)$ for the new last number of $CS(\vec n;k)$, with $k\,\in\, \mathbb N_0$. The new start value $M(S;k)$ is obtained by replacing $t\,\to\, x(S;t)$ in the old $M_{int}(S;t)$ (with $n_S\,=\, 1$). $M(S;k)\,=\, M_{int}(S,x(S;t))$ with $t\,\to\, t_{min}\,+\, k$, also leading to the claimed formula. \end{proof} \par\smallskip\noindent The remainder structure modulo $2^{D(S)}$ for $M(\vec n_S,k)$ and modulo $3^S$ for $N(\vec n_S,k)$ is manifest.\par\smallskip\noindent The explicit sum versions of the results for case $n_S\,=\, 1$, given in {\sl Proposition}~\ref {GenSol}, can be inserted here. \begin{example} {$\bf ud^{m}\,=\, sd^{m-1}$}\par\smallskip\noindent For $m \,=\, 1,\,2,\, 3$ and $ k\,=\, 0,\,1,\,...,\,10$ one finds for $N([0,m],k)$:\par\noindent $[2, 5, 8, 11, 14, 17, 20, 23, 26, 29, 32]$,\ $[1, 4, 7, 10, 13, 16, 19, 22, 25, 28, 31]$,\ \par\noindent $ [2, 5, 8, 11, 14, 17, 20, 23, 26, 29, 32]$, and for $M([0,m],k)$:\par\noindent $[1,3,5,7,9,11,13,15,17,19,21]$,\ $[1,5,9,13,17,21,25,29,33,37,41]$,\ \par\noindent $[5,13,21,29,37,45,53,61,69,77,85]$. Only the odd members of $N([0,m],k)$, that is the odd indexed entries, and the corresponding $M([0,m],k)$ appear in \cite{Truemper}, example 2.1. See \seqnum{A238475} for $M([0,2\,n],k)$ and \seqnum{A238476} for $M([0,2\,n\,-\, 1],k)$. The odd $N([0,2\,n],k)$ values are the same for all $n$, namely $5\,+\, 6\,k$, and $N([0,2\,n-1],k)\,=\, 1\,+\, 6\,k$ for all $n\,\in\, \mathbb N$. \end{example} \begin{example} {$\bf (ud)^{n}\,=\, s^{S}, S\,\in\, \mathbb N$}\par\smallskip\noindent $\vec n_S\,=\, [0,1,...,1]$ with $S$ times a $1$. For $S \,=\, 1,\,2,\, 3$ and $ k\,=\, 0,\,1,\,...,\,10$ one finds $N(\vec n_S,k)$\ \par\noindent $[5,\,8,\,11,\,14,\,17,\,20,\,23,\,26,\,29,\,32]$,\ $[17,\,26,\,35,\,44,\,53,\,62,\,71,\,80,\,89,\,98]$,\ \par\noindent $[53,\,80,\,107,\,134,\,161,\,188,\,215,\,242,\,269,\,296]$, and for $M(\vec n_S,k)$:\par\noindent $[3,\, 5,\, 7, \,9,\, 11,\, 13,\, 15,\, 17,\, 19, \,21]$,\ $ [7,\, 11,\, 15,\, 19, \,23,\, 27, \,31, \,35,\, 39,\, 43]$,\ \par\noindent $[15, \,23,\, 31, \,39, \,47, \,55, \,63, \,71, \,79, \,87]$,. For odd $N$ entries, and corresponding $M$ entries this is \cite{Truemper}, example 2.1. See \seqnum{A239126} for these $M$ values, and \seqnum{A239127} for these $N$ values, which are here $S$ dependent. \end{example} \par\smallskip\noindent In conclusion the author does not think that the knowledge of all {\sl Collatz} sequences with a given up-down pattern (a given {\sl Collatz} word) will help to prove the {\sl Collatz} conjecture. Nevertheless the problem considered in this paper is a nice application of a simple {\sl Diophan}tine equation.\par\bigskip\noindent {\bf Acknowlegement}\par\smallskip\noindent Thanks go to {\sl Peter Bala} who answered the author's question for a proof that all triangle \seqnum{A107711} entries are non-negative integers. See the history there, Feb 28 2014. \par\bigskip\noindent \par\bigskip\noindent
{ "timestamp": "2014-04-11T02:05:00", "yymm": "1404", "arxiv_id": "1404.2710", "language": "en", "url": "https://arxiv.org/abs/1404.2710", "abstract": "Motivated by a recent work of Trümper we consider the general Collatz word (up-down pattern) and the sequences following this pattern. The recurrences for the first and last sequence entries are given, obtained from repeated application of the general solution of a binary linear inhomogeneous Diophantine equation. These recurrences are then solved. The Collatz tree is also discussed.", "subjects": "Number Theory (math.NT)", "title": "On Collatz' Words, Sequences and Trees", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9857180631379973, "lm_q2_score": 0.8128673223709251, "lm_q1q2_score": 0.8012580025956383 }
https://arxiv.org/abs/2301.02645
The Generalized Kauffman-Harary Conjecture is True
For a reduced alternating diagram of a knot with a prime determinant $p,$ the Kauffman-Harary conjecture states that every non-trivial Fox $p$-coloring of the knot assigns different colors to its arcs. In this paper, we prove a generalization of the conjecture stated nineteen years ago by Asaeda, Przytycki, and Sikora: for every pair of distinct arcs in the reduced alternating diagram of a prime link with determinant $\delta,$ there exists a Fox $\delta$-coloring that distinguishes them.
\section{History of the alternation conjecture} In 1998, Louis H. Kauffman and Frank Harary formulated the following conjecture \cite{HK}: \begin{conjecture*} Let $D$ be a reduced, alternating diagram of a knot $K$ having determinant $p$, where $p$ is prime. Then every non-trivial $p$-coloring of $D$ assigns different colors to different arcs. \end{conjecture*} This conjecture is now known as the Kauffman-Harary conjecture. It was proved for rational knots \cite{KLa, PDDGS}, Montesinos knots \cite{APS}, some Turk's head knots \cite{DMMS}, and for algebraic knots \cite{DS}. In 2009, Thomas W. Mattman and Pablo Solis proved this conjecture using the notion of pseudo colorings. A generalization of this conjecture, known as the generalized Kauffman-Harary (GKH) conjecture, was formulated by Marta M. Asaeda, Adam S. Sikora, and the fifth author in 2004 \cite{APS}. They proved this conjecture for Montesinos links in the same paper. In this paper, we prove it in full generality. \ The paper is structured as follows. In the next section we introduce the GKH conjecture and we prove it in Section \ref{proofconjecture}. In Section \ref{nonprimealt}, we reformulate and prove the conjecture for non-prime alternating links. We illustrate the results with some examples in Section \ref{exfoxsection}. In the last section, we discuss pseudo colorings followed by some open questions. \section{Preliminaries} In this section, we state the original and alternate versions of the GKH conjecture. The difference between the original and generalized versions of the conjecture is that the former is about links with prime determinant, while the generalized version is about links with determinant not necessarily prime. It is important to note that the only link whose determinant is prime is the Hopf link. \begin{conj*} If $D$ is a reduced alternating diagram of a prime link $L$, then different arcs of $D$ represent different elements of $H_1(M^{(2)}_L,\mathbb Z)$, where $M^{(2)}_L$ denotes the double branched cover of $S^3$ branched along $L$. \end{conj*} The GKH conjecture was formulated in \cite{APS} using the homology of the double branched cover of $S^3$ branched along $L$. In this paper we use a diagrammatic version of this conjecture by using the universal\footnote{Analogous to the fundamental group and the fundamental quandle, this group is often called the fundamental group of Fox colorings.} group of Fox colorings $Col(D)$ for a prime link $L$ with diagram $D$. \begin{definition} The group $\boldsymbol{Col(D)}$ is the abelian group whose generators are indexed by the arcs of $D$, denoted by $arcs(D)$, and whose relations are $2b-a-c=0$ given by the crossings of $D$. More precisely, $$Col(D) = \displaystyle \bigg\{ \text {arcs}(D) \ | \ \ \vcenter {\hbox{ \begin{overpic}[scale = .08]{CrossingMatrixRelation.jpg} \put(26, 26.5){\tiny{$b$}} \put(0, -1){\tiny{$b$}} \put(14, -1){\tiny{$c=2b-a$}} \put(0, 26.5){\tiny{$a$}} \end{overpic} }} \ \ \ \ \ \ \bigg\}.$$ \end{definition} It is known that $Col(D)= \mathbb Z \oplus H_1(M^{(2)}_L,\mathbb Z)$ (see, for example, \cite{Prz1}). \begin{definition}\label{coltrivialdefi} Let $Col^{trivial}(D) \cong \mathbb Z$ be the group of trivial colorings of $D$. This group is embedded in $Col(D)$ and the quotient group $\displaystyle \frac{Col(D)}{Col^{trivial}(D)}$ is called the \textbf{reduced group of Fox colorings}. We denote it by $Col^{red}(D)$. \end{definition} Notice that, for a diagram $D$ of a link $L$, $Col^{red}(D)=H_1\big(M^{(2)}_L,\mathbb Z\big)$ and for non-split alternating links, this group is finite with non-zero determinant. \ The first two statements of the following conjecture are equivalent to the original GKH conjecture, while part $(c)$ offers an extension. \begin{conjecture}[Alternate forms of the generalized Kauffman-Harary conjecture]\label{ConAPS} \ Let $D$ be a reduced alternating diagram of an alternating prime link and let $\delta(D)$ denote the absolute value of its determinant. \begin{enumerate} \item [\namedlabel{a}{(a)}] Let $\mathbb{Z}^{|arcs|}$ denote the free abelian group $\mathbb{Z}^{|arcs|} = \{arcs(D) \mid \emptyset \}$. Consider the map $\mathbb{Z}^{|arcs|} \xrightarrow{\beta} Col(D).$ Then $\beta$ is injective on the arcs of $D$, that is, $\beta(a_i) \neq \beta(a_j)$ for $i \neq j$. \item [\namedlabel{b}{(b)}] The diagram $D$ has $t$ Fox $\delta(D)$-colorings $y_1, y_2, \hdots, y_t$, such that for every pair of distinct arcs $a_i, a_j$, there exists $y_k$ such that $y_k(a_i) \neq y_k(a_j)$. \item [\namedlabel{c}{(c)}] If $Col^{red}(D) = \mathbb Z_{n_1} \oplus \mathbb Z_{n_2} \oplus \cdots \oplus \mathbb Z_{n_s}$ with $n_{i+1} | n_i$, then there are $s$ Fox $n_1$-colorings that distinguish all the arcs of $D$. Note that, $s$ is strictly less than the number of crossings of $D$. \end{enumerate} \end{conjecture} \begin{remark} Parts \ref{a} and \ref{b} of Conjecture \ref{ConAPS} are equivalent to each other, since for a finite group $G$, we have $G \cong Hom(G, \mathbb{Z}_{n_1})\cong Hom(G, \mathbb{Z}_{\delta(D)})$, where $G=\mathbb{Z}_{n_1}\oplus \mathbb Z_{n_2} \oplus \cdots \oplus \mathbb{Z}_{n_s}$, with $n_{i+1} | n_i$ and $\delta(D) =n_{1}n_{2}\cdots n_{s}$. In particular, $Hom(Col^{red}(D), \mathbb{Z}_{\delta(D)}) \cong Hom(Col^{red}(D), \mathbb{Z}_{n_1}) \cong Col^{red}(D)$. Thus, we can work with a group or its dual. To distinguish elements in the group we often analyze its homomorphisms (dual elements) into the given ring. See \cite{lang}, for example. \end{remark} \section{Proof of the generalized Kauffman-Harary conjecture}\label{proofconjecture} The proof of the GKH conjecture is organized as follows. First, we define the crossing matrix $C'(D)$ and coloring matrix $L(D)$ of a link diagram $D$. Following \cite{MS} we prove that every column of the coloring matrix represents a non-trivial Fox $\delta(D)$-coloring. Then using the fact that the coloring matrix of the mirror image of $D$ is the transpose of $L$, we prove part \ref{b}, and equivalently, part \ref{a} of Conjecture \ref{ConAPS}. Additionally, we show that the columns of the coloring matrix generate the group $Col^{red}(D)$ and use this fact to prove part \ref{c} of Conjecture \ref{ConAPS}. \begin{definition} A \textbf{Fox} $\boldsymbol{k}$\textbf{-coloring} of a diagram $D$ is a function $f: \mathit{arcs}(D) \to \mathbb{Z}_{k}$, satisfying the property that every arc is colored by an element of $\mathbb{Z}_{k}=\left\lbrace 0, 1, 2, 3, \dots, k-1\right\rbrace $ in such a way that at each crossing the sum of the colors of the undercrossings is equal to twice the color of the overcrossing modulo $k$. That is, if at a crossing $v$ the overcrossing is colored by $b$, and the undercrossings are colored by $a$ and $c$, then $2b-a-c \equiv0$ modulo $k$. See Figure \ref{CrossingMatrixRelation} for an illustration. The group of Fox $k$-colorings of a diagram $D$ is denoted by $\mathit{Col}_{k}(D)$ and the number of Fox $k$-colorings is denoted by $\mathit{col}_{k}(D)$. Analogous to Definition \ref{coltrivialdefi}, we divide the group $Col_{k}(D)$ by the group of trivial colorings and denote the quotient group by $Col^{red}_{k}(D)$. \end{definition} The matrix describing the space of colorings $Col(D)$ is referred to, by Mattman and Solis, as the crossing matrix for a fixed arbitrary ordering of the crossings \cite{MS}. Here we do not assume that the diagram is alternating. \begin{definition} Fix an ordering of the crossings of a reduced link diagram $D$. Then the set of arcs inherits the order of the set of crossings. In this way, the over-arc has the same index as the crossing. The \textbf{crossing matrix}\footnote{The alternative, more descriptive, name could be {\it unreduced fundamental Fox colorings matrix.}} of $D$, denoted by $C'(D)$, is an $n \times n$ matrix such that each row corresponds to a crossing that gives the relation $2b-a-c=0$ (see Figure \ref{CrossingMatrixRelation}). The entries of the matrix are defined as follows\footnote{It is possible that two under-arcs at a crossing are not distinct. Then the relation $2b-a-c=0$ becomes $2b-2a=0$. For instance, this may occur for the Hopf link.}: \begin{minipage}{0.5\textwidth} $$ C_{ij}' = \left\{ \begin{array}{rr} 2 & \text{if } a_i \ \text{is the over-arc at } c_i, \\ -1 & \text{ if } a_j \text{ is an under-arc at } c_i \text{ }(i\neq j), \text{and} \\ 0 & \text{ otherwise}. \end{array} \right.$$ \end{minipage} \begin{minipage}{0.47\textwidth} \centering \includegraphics[scale=0.45]{FoxrelationOverpic.png} \captionof{figure}{Fox coloring relation at crossing $v$.} \label{CrossingMatrixRelation} \end{minipage} \end{definition} \begin{figure}[ht] \centering $$\vcenter{\hbox{ \begin{overpic}[scale = .44]{Crossingchange.png} \put(80, 125){$c_k$} \put(80, 83){$c_i$} \put(80, 40){$c_j$} \put(92, 70){$a_i$} \put(65, 135){$a_k$} \put(65, 50){$a_j$} \put(385, 125){$c_k$} \put(360, 125) {$\overline{a_k}$} \put(385, 83){$c_i$} \put(369, 95){$\overline{a_i}$} \put(385, 40){$c_j$} \put(360, 40){$\overline{a_j}$} \end{overpic} }}$$ \caption{Neighborhood of the crossing $c_i$ in $D$ (on the left) and $\overline{D}$.} \label{mirrorcrossing} \end{figure} The following lemma holds only for alternating links and plays an important role in the proof of the GKH conjecture. \begin{lemma}\label{Mirror} Let $D$ be a reduced alternating link diagram with crossing matrix $C'(D)$ and let $\overline{D}$ be its mirror image. Then the matrix $C'^{T}$ is a crossing matrix for $\overline{D}$. \end{lemma} \begin{proof} Denote the crossings of the diagram $D$ by $c_1, \dots,c_n$ and let the over-arc at the crossing $c_i$ be denoted by $a_i$. Notice that, in the matrix $C'(D)$ all entries on the diagonal are $2$. We obtain $\overline{D}$ by crossing-change operations and we keep the ordering and names of the crossings. Now, let $\overline{a_i}$ denote the over-arc at the crossing $c_i$ in the diagram $\overline{D}$. In the row corresponding to the crossing $c_i$, suppose the columns corresponding to the arcs $a_j$ and $a_k$ have $-1$ as entries. Then in the matrix $C'(\overline{D})$, the column corresponding to $\overline{a_i}$ must have entries $-1$ in the rows corresponding to the crossings $c_j$ and $c_k$; see Figure \ref{mirrorcrossing}. \end{proof} Recall that if $\delta(D)\neq0$, then $Col^{red}(D)$ is a finite group whose invariant factor decomposition is $Col^{red}(D) = \mathbb Z_{n_1}\oplus \mathbb Z_{n_2}\oplus \cdots \oplus \mathbb Z_{n_s}$, with $n_{i+1} | n_i$ for all $i$. Notice that, $s$ is the minimum number of generators of this group and $n_{1}$ is the annihilator of the group. Let $C(D)$ denote the reduced crossing matrix of $D$, which is the matrix obtained from $C'(D)$ by removing its last row and last column. We call the arc corresponding to the last column of $C'(D)$ the \textbf{base arc}. This matrix describes the group $Col^{red}(D)$. The matrix $C^{-1}(D)$ is a matrix with rational entries. However, $n_1C^{-1}(D)$ is an integral matrix, which we denote by $L_{n_1}(D).$ Observe that the columns of $L_{n_1}(D)$ modulo $n_1$ represent Fox $n_1$-colorings of the diagram $D$ after coloring the base arc by color $0$. \ The following result also holds for reduced non-alternating links. \begin{theorem}\label{generatorlemma} Let $D$ be a reduced diagram of a link with non-zero determinant. Then the columns of $L_{n_1}(D)$ modulo $n_1$ generate the space of Fox $n_1$-colorings of $D$. \end{theorem} \begin{proof} Let $C(D)$ be the reduced crossing matrix of $D$ and let $Col^{red}(D) = \mathbb Z_{n_1}\oplus \mathbb Z_{n_2}\oplus \cdots \oplus \mathbb Z_{n_s}$, with $n_{i+1} | n_i$ for all $i$. After row and column operations, $C(D)$ can be reduced to its Smith normal form, denoted by $C_{SNF}(D)$, given below. \begin{equation*} C_{SNF}(D) = \begin{pmatrix} n_{1} & & && \\ &n_{2} & &&&\text{\Huge0}& \\ & & \ddots&&& \\ & &&n_s && \\ &&&&1 && \\ &\text{\Huge0}&&&&\ddots &&\\ & & &&&& 1 \end{pmatrix} \end{equation*} Its inverse matrix, $C^{-1}_{SNF}(D)$, with entries in $\mathbb{Q}$ has the following form. \begin{equation*} C^{-1}_{SNF}(D) = \begin{pmatrix} 1/n_{1} & & && \\ &1/n_{2} & &&&\text{\Huge0}& \\ & & \ddots&&& \\ & &&1/n_s && \\ &&&&1 && \\ &\text{\Huge0}&&&&\ddots &&\\ & & &&&& 1 \end{pmatrix} \end{equation*} Thus, we obtain the following integral matrix $L^{SNF}_{n_1}(D)$. \begin{equation*} L^{SNF}_{n_1}(D)=n_1 C^{-1}_{SNF}(D)= \begin{pmatrix} n_1/n_{1} & & && \\ &n_1/n_{2} & &&&\text{\Huge0}& \\ & & \ddots&&& \\ & &&n_1/n_s && \\ &&&&n_1 && \\ &\text{\Huge0}&&&&\ddots &&\\ & & &&&& n_1 \end{pmatrix} \end{equation*} \ Now, the $i^{\mathit{th}}$ column $(0, 0, \dots, n_{1}/n_{i}, \dots, 0)^T$ of $L^{SNF}_{n_1}(D)$ modulo $n_1$ with $i \le s$ generates the subgroup $\mathbb{Z}_{n_i}$ of $\mathbb{Z}_{n_1}$. Since $Col^{red}_{n_1}(D) = Hom(\mathbb Z_{n_1}\oplus \mathbb Z_{n_2}\oplus \cdots \oplus \mathbb Z_{n_s}, \mathbb Z_{n_1})$, therefore, the columns of $L^{SNF}_{n_1}$ generate the group $Col^{red}_{n_1}(D)$, as desired.\\ \end{proof} For alternating diagrams we can prove the following stronger result, which proves part \ref{b}, and equivalently, part \ref{a} of Conjecture \ref{ConAPS}. \begin{theorem}\label{mainlemma} Let $D$ be a reduced alternating diagram of a prime link. For any two arcs $a_i$ and $a_j$, there exists a column of $L_{n_1}(D)$ which distinguishes them. \end{theorem} \begin{proof} Suppose the arcs (indexing rows) of the coloring matrix $n_{1}C^{-1}(D) = L_{n_1}(D)$ are given by $a_1$, $a_2$, \dots, $a_{n-1}$ as shown below. $$L_{n_1}(D)=n_{1}C^{-1}(D) = \begin{pNiceMatrix}[first-row,last-row,first-col,last-col] \ \ \ \ \ & & & & \\ \textcolor{blue}{a_1:} \ \ \ \ \ \ & c_{1,1} & c_{1,2} & \cdots & c_{1,n-1} & \ \ \\ \textcolor{blue}{a_2:} \ \ \ \ \ \ & c_{2,1} & c_{2,2} & \cdots & c_{2,n-1} & \ \ \\ \textcolor{blue}{\vdots \ \ \ } \ \ \ \ \ \ & \vdots & \vdots & \vdots & \vdots & \ \ \\ \textcolor{blue}{a_{n-1}:}\ \ \ \ \ \ & c_{n-1,1} & c_{n-1,2} & \cdots & c_{n-1,n-1} & \ \ \\ \textcolor{blue}{a_n:} \ \ \ \ \ \ & 0 & 0 & \cdots & 0 & \ \ \\ \end{pNiceMatrix}$$ Recall that, the reduced crossing matrix $C(D)$ is obtained from $C'(D)$ by removing its last row and last column. Now, each column of $L_{n_1}(D)$ colors the the remaining first $n-1$ arcs of the diagram. For a complete Fox $n_1$-coloring of $D$ we color the last (base) arc $a_n$ by color $0$. If any $c_{i,1}=0 \mod n_1$ for $i < n$, then column $C_1$ modulo $n_1$ cannot distinguish between the arcs $a_n$ and $a_i$. If all the entries of the rows corresponding to $a_i$ and $a_j$ are not identical modulo $n_1$, then they can be automatically distinguished by the column in which they are different. \ \textbf{Step 1:} If there is no column $C_j$ of $L_{n_1}(D)$ such that $c_{i,j} \neq 0$ mod $n_1$, then every entry in the $i^{\mathit{th}}$ row is $0$ mod $n_1$. It follows that in the transpose matrix $L_{n_1}^{T}(D)$, the column $C_{i}^{T}$ is the zero column modulo $n_1$. This would result in the existence of a pseudo coloring of $\overline{D}$ (see Definition \ref{pseudodef} and \cite{MS}), which is a contradiction. Thus, the base arc $a_n$ can be distinguished from any other arc by some column in $L_{n_1}(D)$ modulo $n_1$. \ \textbf{Step 2:} Furthermore, if there are two arcs $a_i$ and $a_j$ with the same color in every column of $L_{n_1}(D)$, then we choose arc $a_j$ as the base arc, which implies that the colors of $a_i$ are equal to zero. So we are back to Step 1. \end{proof} The next theorem proves part \ref{c} of Conjecture \ref{ConAPS}, which is a more general version of Theorem \ref{mainlemma}. \begin{theorem}\label{mainlemma2} If $Col^{red}(D) = \mathbb Z_{n_1} \oplus \mathbb Z_{n_2} \oplus \cdots \oplus \mathbb Z_{n_s}$, with $n_{i+1} | n_i$, then there are $s$ Fox $n_1$-colorings (not necessarily corresponding to the columns of the coloring matrix) which distinguish all arcs. That is, for every pair of arcs of $D$, one of these $n_{1}$-colorings distinguishes them. \end{theorem} \begin{proof} Denote the generators of the group $Col^{red}(D)$ by $a_1$, $a_2$, \dots, $a_s$. Every generator $a_i$ is a linear combination of some columns of the coloring matrix $L_{n_1}(D)$ modulo $n_1$ (see Theorem \ref{generatorlemma}). Therefore, they correspond to some coloring of the diagram $D$. Hence, for every pair of arcs there is a column of $L_{n_1}(D)$ modulo $n_1$ that distinguishes them. \end{proof} \begin{corollary}\label{Coro} If $Col^{red}(D)$ is the cyclic group $\mathbb{Z}_{n_1}$,\hfill \begin{itemize} \item [(a)] then there exists a non-trivial Fox $n_1$-coloring that distinguishes all arcs. \item [(b)] Additionally, if $n_1$ is a prime number, then the original Kauffman-Harary conjecture holds. That is, every non-trivial Fox $n_1$-coloring distinguishes all arcs. \end{itemize} \end{corollary} \begin{proof} Part (a) follows directly from Theorem \ref{mainlemma2}, for $s=1$. Part (b) follows because every non-zero element of $\mathbb{Z}_{n_1}$ is its generator. \end{proof} \section{Non-prime alternating links}\label{nonprimealt} Theorems \ref{mainlemma} and \ref{mainlemma2} do not hold as stated for the connected sum of alternating links\footnote{The connected sum of alternating links, is an alternating link. For example, see \cite{PBIMW}.} (see part \ref{(a)} of Lemma \ref{nonprimelemma}). In Theorem \ref{connectedsumnonprime}, we present a version of the GKH conjecture which holds for non-prime alternating links. \begin{lemma}\cite{Prz1} \label{nonprimelemma} Let $D= D_1 \ \# \ D_2$ be the connected sum of two link diagrams. Then, \begin{itemize} \item [\namedlabel{(a)}{(a)}] the arcs connecting the two components represent the same element in $Col(D)$, and \item[(b)] $Col^{red}(D_1 \ \# \ D_2) \cong Col^{red}(D_1)\oplus Col^{red}(D_2) $. \end{itemize} \end{lemma} \begin{theorem}\label{connectedsumnonprime} Let $D= D_1 \ \# \ \ D_2 \ \# \ \cdots \ \# \ D_n$, where $D_i$ is a reduced alternating diagram of a prime link $L_i$, for $i=1, 2, \dots, n$. Then, \begin{itemize} \item [(a)] for any pair of arcs different from arcs joining $D_i$ with $D_{i+1}$, there exists a Fox $n_1$-coloring which distinguishes them, and \item [(b)] there are $t$ ($t \leq s$) Fox $n_1$-colorings such that any pair of arcs different from the ones joining $D_i$ with $D_{i+1}$, is distinguished by one of them. \end{itemize} \end{theorem} \begin{proof} This result follows from Theorems \ref{generatorlemma} and \ref{mainlemma}, and Lemma \ref{nonprimelemma}. \end{proof} \begin{remark} Theorem \ref{connectedsumnonprime} was formulated for connected sums of diagrams. However, from William W. Menasco's result (see \cite{Men,Hos}), it follows that if an alternating diagram represents the connected sum of alternating links, then it is already a connected sum of diagrams. \end{remark} \begin{example} Let $D$ be an alternating diagram of the square knot, that is $D= \overline{3}_1 \ \# \ 3_1$, with reduced crossing matrix $C(D)$ (see Figure \ref{FoxSquareknot}). Then $Col^{red}(\overline{3}_1 \ \# \ 3_1)= \mathbb Z_3 \oplus \mathbb Z_3$. Observe that columns 3 and 5 of $L_3(D)$ modulo $3$ (Figure \ref{matricesSquareKnot}) distinguish all pairs of arcs except the ones connecting $\overline 3_1$ with $3_1.$ Also, the third row (corresponding to the third crossing in the chosen ordering and, therefore, to the third arc) has all zero entries. That is, the third arc cannot be distinguished from the base arc. \begin{figure}[H] \begin{subfigure}{.4\textwidth} $ C(D) = \begin{pmatrix*}[r] 2 & -1 & 0 & 0 & 0\\ -1 & 2 & -1 & 0 & 0\\ -1 & -1 & 2 & 0 & 0\\ 0 & 0 & -1 & 2 & -1 \\ 0 & 0 & 0 & -1 & 2 \end{pmatrix*}$ \end{subfigure} \begin{subfigure}{.3\textwidth} \centering $$\vcenter{\hbox{ \begin{overpic}[scale = 1.7]{squareknot} \put(55, 60){$(1,0)$} \put(29, 52){$(2,0)$} \put(37, 135){$(0,0)$} \put(133, 65){$(0,1)$} \put(105, 54){$(0,2)$} \put(30, 5){$(0,0)$} \end{overpic} }}$$ \end{subfigure} \caption{The reduced crossing matrix for the square knot (on the left). The square knot $\overline{3}_1 \ \# \ 3_1$ with two Fox $3$-colorings distinguishing every pair of arcs (on the right).} \label{FoxSquareknot} \end{figure} \begin{figure}[h] $ L_3(D) = 3 C^{-1}(D) = \begin{pmatrix} 3 & 2 & 1 & 0 & 0\\ 3 & 4 & 2 & 0 & 0\\ 3 & 3 & 3 & 0 & 0\\ 2 & 2 & 2 & 2 & 1\\ 1 & 1 & 1 & 1 & 2 \end{pmatrix} \ \ \ \ \ L_3(D) \ \text{mod 3}= \begin{pmatrix} 0 & 2 & 1 & 0 & 0\\ 0 & 1 & 2 & 0 & 0\\ 0 & 0 & 0 & 0 & 0\\ 2 & 2 & 2 & 2 & 1\\ 1 & 1 & 1 & 1 & 2 \end{pmatrix} $ \caption{Matrices $L_3(D)$ and $L_3(D)$ modulo $3$ for the square knot.}\label{matricesSquareKnot} \end{figure} \end{example} \section{Examples of Fox colorings}\label{exfoxsection} In this section we study examples of alternating link diagrams and their Fox colorings. For the structure of the group $Col^{red}(D)=H_{1}(M_{D}^{(2)}, \mathbb{Z})$ for knots up to 10 crossings, see Appendix C in \cite{BZ}. \begin{example} Kauffman and Harary showed that the knot $7_7$ is a counterexample to their conjecture for a knot with non-prime determinant \cite{HK}. We have, $det \ ( 7_7)= 21$ and $Col^{red}(7_7)=\mathbb{Z}_{21}$.\footnote{It was noticed in \cite{KLa} that the Kauffman-Harary conjecture holds for any rational (2-bridge) knot without restrictions on the determinant of the knot. However, as they note, the formulation of the conjecture needs to be changed from ``every non-trivial Fox $D$-coloring" to ``there exists a Fox $D$-coloring." See Corollary \ref{Coro}.} See Figure \ref{7_7 figure} for a Fox $21$-coloring distinguishing all arcs. \begin{figure}[h] $L(7_7)=\begin{pmatrix} 24 & 20 & 12 & 10 & 16 & 11 \\ 12 & 24 & 6 & 12 & 15 & 9 \\ 15 & 16 & 18 & 8 & 17 & 13 \\ 6 & 5 & 3 & 13 & 4 & 8 \\ 18 & 22 & 9 & 11 & 26 & 10 \\ 12 & 10 & 6 & 5 & 8 & 16 \\ \end{pmatrix} \ \ \ \ \ L(7_7) \ \text{mod} \ 21=\begin{pmatrix} 3 & 20 & 12 & 10 & 16 & 11 \\ 12 & 3 & 6 & 12 & 15 & 9 \\ 15 & 16 & 18 & 8 & 17 & 13 \\ 6 & 5 & 3 & 13 & 4 & 8 \\ 18 & 1 & 9 & 11 & 5 & 10 \\ 12 & 10 & 6 & 5 & 8 & 16 \\ \end{pmatrix} $ \caption{Matrices $L(7_7)$ and $L(7_7)$ modulo $21$. Some non-trivial Fox $21$-colorings of $7_7$ do not distinguish all arcs; for example, columns 1 or 3 of $L$ modulo $21$. However, columns $2$, $4$, $5$, and $6$ distinguish all arcs.} \end{figure} \begin{figure}[h] $$\vcenter{\hbox{ \begin{overpic}[scale = 0.8]{7_7.png} \put(170, 130){$1$} \put(142, 100){$2$} \put(53, 138){$4$} \put(75, 100){$7$} \put(2, 20){$12$} \put(70, 40){$20$} \put(195, 45){$0$} \end{overpic} }}$$ \caption{The knot $7_7$ with a Fox $21$-coloring which distinguishes all arcs.}\label{7_7 figure} \end{figure} \end{example} \begin{example} Consider the family of links obtained by closing the braids $(\sigma_1\sigma_2^{-1})^{n}$. These links are sometimes called Turk's head links and can also be obtained by drawing the Tait diagrams of the wheel graphs $W_n$. The closed formula for the determinant of the $D(W_n)$ is given in \cite{Prz-Goeritz}. Examples for $n=5$ and $n=6$ are drawn in Figure \ref{W5andW6} and their reduced groups of Fox colorings are as follows: \begin{itemize} \item [(a)] For $n=5$, $D(W_5)$ is $10_{123}$ in Rolfsen's table \cite{Rol}. $Col^{red}(D(W_5)) = \mathbb Z_{11} \oplus \mathbb Z_{11} $. \item [(b)] For $n=6$, $D(W_6)$ is the link $12^{3}_{474}$ in Thistlethwaite's tables \cite{Prz-Goeritz, Thi}. $Col^{red}(D(W_6)) = \mathbb Z_{40} \oplus \mathbb Z_{8}$. \end{itemize} \begin{figure}[ht] \centering \includegraphics[width=1\linewidth]{W_5.png} \caption{The knot $D(W_5)$ with two Fox colorings distinguishing all arcs (on the left), and the link $D(W_6)$ with two Fox colorings distinguishing all arcs (on the right).} \label{W5andW6} \end{figure} \end{example} \begin{example} The group $Col^{red}$ for pretzel links is given in Proposition $7$ in \cite{APS} and its generalization to Montesinos links is given in Proposition $8$ in \cite{APS}. Here we show two examples together with their coloring matrices modulo $n_1$. \begin{itemize} \item[(a)] Let $P(3,3,3,3,3)$ be a pretzel knot with $15$ crossings. Its group $Col^{red}$ is equal to $\mathbb{Z}_{15} \oplus \mathbb{Z}_3 \oplus \mathbb{Z}_3 \oplus \mathbb{Z}_3$. See its coloring matrix, $L(P(3,3,3,3,3))$ modulo $15$ in Figure \ref{coloringpretzel1}. \item[(b)] Let $P(3,3,3,6)$ be a pretzel knot with $15$ crossings. Its group $Col^{red}$ is equal to $\mathbb{Z}_{21} \oplus \mathbb{Z}_3 \oplus \mathbb{Z}_3$. See its coloring matrix, $L(P(3,3,3,6))$ modulo $21$ in Figure \ref{coloringpretzel2}. \end{itemize} \end{example} \begin{figure} $L(P(3,3,3,3,3))\ mod \ 15=\left( \begin{array}{cccccccccccccc} 5 & 1 & 1 & 1 & 12 & 12 & 7 & 8 & 3 & 3 & 14 & 14 & 14 & 10 \\ 10 & 5 & 10 & 10 & 10 & 10 & 5 & 10 & 0 & 10 & 10 & 10 & 10 & 10 \\ 10 & 1 & 11 & 1 & 12 & 12 & 7 & 8 & 13 & 3 & 4 & 14 & 14 & 10 \\ 10 & 12 & 12 & 7 & 14 & 14 & 9 & 6 & 11 & 11 & 13 & 3 & 3 & 10 \\ 10 & 1 & 1 & 1 & 7 & 12 & 7 & 8 & 13 & 13 & 4 & 4 & 14 & 10 \\ 10 & 12 & 12 & 12 & 14 & 9 & 9 & 6 & 11 & 11 & 13 & 13 & 3 & 0 \\ 10 & 8 & 8 & 8 & 6 & 6 & 11 & 4 & 9 & 9 & 7 & 7 & 7 & 5 \\ 5 & 7 & 7 & 7 & 9 & 9 & 4 & 11 & 6 & 6 & 8 & 8 & 8 & 10 \\ 0 & 3 & 13 & 13 & 11 & 11 & 6 & 9 & 9 & 14 & 12 & 12 & 12 & 10 \\ 10 & 14 & 4 & 4 & 13 & 13 & 8 & 7 & 12 & 7 & 1 & 1 & 1 & 10 \\ 10 & 3 & 3 & 13 & 11 & 11 & 6 & 9 & 14 & 14 & 7 & 12 & 12 & 10 \\ 10 & 14 & 14 & 4 & 3 & 13 & 8 & 7 & 12 & 12 & 1 & 11 & 1 & 10 \\ 10 & 10 & 10 & 10 & 10 & 0 & 10 & 5 & 10 & 10 & 10 & 10 & 5 & 10 \\ 10 & 14 & 14 & 14 & 3 & 3 & 8 & 7 & 12 & 12 & 1 & 1 & 1 & 5 \\ \end{array} \right)$ \caption{$L(P(3,3,3,3,3))$ modulo $15$. The colorings given by the columns 3, 4, and 10 distinguish all arcs.}\label{coloringpretzel1} \end{figure} \begin{figure} $L(P(3,3,3,6)) \ mod \ 21=\left( \begin{array}{cccccccccccccc} 9 & 5 & 1 & 10 & 11 & 18 & 18 & 5 & 7 & 3 & 20 & 1 & 1 & 14 \\ 1 & 6 & 11 & 12 & 9 & 16 & 16 & 20 & 14 & 19 & 3 & 18 & 18 & 14 \\ 14 & 7 & 0 & 14 & 7 & 14 & 14 & 14 & 0 & 14 & 7 & 14 & 14 & 14 \\ 6 & 8 & 10 & 16 & 5 & 12 & 12 & 8 & 7 & 9 & 11 & 10 & 10 & 14 \\ 15 & 13 & 11 & 5 & 16 & 9 & 9 & 13 & 14 & 12 & 10 & 11 & 11 & 7 \\ 13 & 15 & 17 & 9 & 12 & 12 & 19 & 15 & 14 & 16 & 18 & 17 & 3 & 0 \\ 11 & 17 & 2 & 13 & 8 & 15 & 8 & 17 & 14 & 20 & 5 & 2 & 16 & 14 \\ 13 & 15 & 17 & 9 & 12 & 19 & 19 & 8 & 14 & 16 & 18 & 3 & 3 & 14 \\ 5 & 16 & 6 & 11 & 10 & 17 & 17 & 2 & 0 & 11 & 1 & 20 & 20 & 14 \\ 18 & 17 & 16 & 13 & 8 & 15 & 15 & 17 & 7 & 6 & 5 & 16 & 16 & 14 \\ 10 & 18 & 5 & 15 & 6 & 13 & 13 & 11 & 14 & 1 & 9 & 12 & 12 & 14 \\ 12 & 16 & 20 & 11 & 10 & 17 & 3 & 2 & 14 & 18 & 1 & 13 & 20 & 14 \\ 14 & 14 & 14 & 7 & 14 & 0 & 14 & 14 & 14 & 14 & 14 & 14 & 7 & 14 \\ 12 & 16 & 20 & 11 & 10 & 3 & 3 & 16 & 14 & 18 & 1 & 20 & 20 & 7 \\ \end{array} \right)$ \caption{$L(P(3,3,3,6))$ modulo $21$. The colorings given by columns 1 and 6 distinguish all arcs.}\label{coloringpretzel2} \end{figure} \section{Odds and ends}\label{odds} \subsection{Pseudo colorings} An important tool in our proof of Theorem \ref{mainlemma} is the idea of pseudo colorings. In \cite{MS} and in this paper, it is shown that no pseudo colorings exist for reduced, prime, alternating link diagrams. However, the existence of pseudo colorings can be used to see how far a diagram is from being an alternating link diagram. In this section, we briefly explore this concept. In \cite{MS}, Proposition 3.2 depends on the fact that for reduced alternating diagrams the rows of the crossing matrix add to zero. This does not hold for non-alternating diagrams, as we illustrate in the following examples. \begin{definition}\label{pseudodef} Let $D$ be a link diagram and $\epsilon \in \{-1,+1\}$. Following Mattman and Solis \cite{MS}, we define an $\boldsymbol{\epsilon}$\textbf{-pseudo coloring} of $D$ as colorings of the arcs of $D$ such that, at all but two crossings the Fox coloring convention $2b - a - c = 0$ is satisfied. We denote the other two crossings by $c_{+1}$ and $c_{\epsilon},$ where the coloring conventions are $2b - a - c = +1$ and $2b - a - c = \epsilon,$ respectively. To obtain the pseudo colorings as defined in \cite{MS}, put $\epsilon=-1$. \end{definition} For an alternating link diagram $D$, our convention was to order crossings first and then, the set of arcs inherits the order of the set of crossings. Compare Definition \ref{CrossingMatrixRelation}. The reason for such a choice is that $C'(\overline{D})$ is the same as $C'(D)^{T}$. This does not work for non-alternating link diagrams. \ In general, we can arbitrarily order crossings and arcs. In Figure \ref{orderarcs} we give an example of ordering crossings and arcs for the knot $8_{19}$. We first choose a base point and an orientation (shown by an arrow on the left-hand side of Figure \ref{orderarcs}). Starting at this base point, we move along the knot and order crossings. Next, arcs can be ordered arbitrarily with the base arc always being the last one. In Figure \ref{orderarcs} the first coordinate gives the number of the crossing and the second one gives the number of the arc. \begin{figure}[ht] \centering \begin{subfigure}{.5\textwidth} $$\vcenter{\hbox{ \begin{overpic}[scale = 1.7]{8_19aspretzel} \put(103, 72){$1,5$} \put(103, 45){$2,1$} \put(25, 30){$3,7$} \put(25, 58){$4,4$} \put(25, 85){$5,6$} \put(62, 85){$6,3$} \put(65, 58){$7,8$} \put(65, 31){$8,2$} \end{overpic} }}$$ \end{subfigure}% \begin{subfigure}{.5\textwidth} $$\vcenter{\hbox{ \begin{overpic}[scale = 1.7]{8_19aspretzel} \put(70, 105){$0$} \put(70, 13){$-1$} \put(27, 27){$0$} \put(-3, 40){$0$} \put(-3, 75){$0$} \put(65, 124){$0$} \put(69, 70){$0$} \put(110, 31){$0$} \put(47, 20){$c_{+1}$} \put(103, 45){$c_{+1}$} \end{overpic} }}$$ \end{subfigure} \caption{The torus knot $T(3,4)$ ($8_{19}$ in Rolfsen's table \cite{Rol}) depicted as the pretzel knot $P(3,3,-2)$ showing ordering of crossings and arcs (on the left). On the right, there is pseudo coloring given by the second column; compare Remark \ref{remarkpseudo}.} \label{orderarcs} \end{figure} \ In the following example, we analyze non-split, non-prime alternating diagrams. \begin{example}\label{Ex52} Let $D = D_1 \ \# \ D_2$ be a non-split, non-prime alternating link diagram. $D$ always has a $-1$-pseudo coloring using color $1$ on $D_1$ and color $0$ on $D_2$. We illustrate this idea for the square knot $\overline{3}_1 \ \# \ 3_1$ in Figure \ref{PseudoSquare}. \begin{figure}[h] \centering $$\vcenter{\hbox{ \begin{overpic}[scale = 1.7]{squareknot} \put(57, 70){$1$} \put(32, 78){$1$} \put(40, 132){$0$} \put(5,103){$c_{+1}$} \put(108, 48){$0$} \put(134, 60){$0$} \put(32, 4){$1$} \put(85,23){$c_{-1}$} \end{overpic} }}$$ \caption{$-1$-pseudo coloring of the square knot with the $+1$-crossing denoted by $c_{ +1}$ and the $-1$-crossing denoted by $c_{ -1}$.}\label{PseudoSquare} \end{figure} \end{example} On the other hand, non-alternating link diagrams often have $-1$-pseudo colorings and $+1$-pseudo colorings. See Examples \ref{Ex53} and \ref{Ex819}. If the determinant of a knot with diagram $D$ is equal to $1$, we have $L(D) = C^{-1}(D)$ and every column of $C^{-1}(D)$ colors the first $n-1$ arcs of the diagram. Then for a complete $\epsilon$-pseudo coloring of $D$, we color the last (base) arc $a_n$ by color $0$. \begin{example}\label{Ex53} Consider the braid word $\sigma^{3}_{2}\sigma^{}_{1}\sigma^{-1}_{3}\sigma^{-2}_{2}\sigma^{}_{1}\sigma^{-1}_{2}\sigma^{}_{1}\sigma^{-1}_{3}$ whose closure is the Conway knot. The determinant of this knot is $1$ and its crossing matrix $C'(D)$ is given in Figure \ref{CMCK}. The $+1$-pseudo coloring given by column $4$ and the $-1$-pseudo coloring given by column $1$ in the matrix shown in Figure \ref{pseudoConwayKnot} are illustrated in Figure \ref{CKpic} on the left and on the right, respectively. \end{example} \begin{figure}[ht] \centering \begin{subfigure}{.5\textwidth} $$\vcenter{\hbox{ \begin{overpic}[scale = 0.9]{CK} \put(55, 150){$2$} \put(35.5, 118){$-8$} \put(40, 108){$-5$} \put(42, 86){$-2$} \put(5, 222){$5$} \put(1, 110){$0$} \put(20, 210){$9$} \put(18, 165){$6$} \put(18, 149){$3$} \put(60, 100){$-3$} \put(3, 72){$1$} \put(50, 60){$c_{+1}$} \put(14, 76){$c_{+1}$} \end{overpic} }}$$ \label{CKpicA} \end{subfigure}% \begin{subfigure}{.5\textwidth} $$\vcenter{\hbox{ \begin{overpic}[scale = 0.9]{CK} \put(55, 150){$6$} \put(35.5, 118){$-33$} \put(38.5, 108){$-21$} \put(42, 86){$-9$} \put(5, 225){$21$} \put(1, 110){$0$} \put(20, 212){$39$} \put(13, 165){$26$} \put(13, 149){$13$} \put(60, 100){$-13$} \put(3, 72){$3$} \put(50, 60){$c_{-1}$} \put(55, 130){$c_{+1}$} \end{overpic} }}$$ \label{CKpicB} \end{subfigure} \caption{The Conway knot with $+1$-pseudo coloring (on the left) and with $-1$-pseudo coloring (on the right). The last crossing $c_{+1}$ in the left figure changes to $c_{-1}$ in the right figure.}\label{CKpic} \end{figure} \begin{figure}[ht] $$C'(D)=\left( \begin{array}{ccccccccccc} -1 & -1 & 0 & 0 & 0 & 0 & 0 & 0 & 2 & 0 & 0 \\ 0 & -1 & -1 & 0 & 2 & 0 & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & 2 & 0 & -1 & -1 & 0 & 0 & 0 & 0 & 0\\ 2 & 0 & -1 & -1 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 2 & 0 & -1 & -1 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & -1 & -1 & 0 & 0 & 0 & 0 & 0 & 2 \\ -1 & 0 & 0 & 0 & 0 & 2 & 0 & 0 & 0 & 0 & -1 \\ 0 & 0 & 0 & 0 & 0 & 0 & -1 & -1 & 0 & 2& 0 \\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 2 & 0 & -1 & -1 \\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & -1 & -1 & 0 & 2 \\ 2& 0 & 0&0 & 0 & 0&0 &0 &-1 & -1 &0 \\ \end{array} \right)$$ \caption{The crossing matrix of the Conway knot. Notice that the rows of the crossing matrix satisfy the linear equation $R_{1}-R_2-R_3-R_4-R_5-R_6-R_7+R_8+R_9+R_{10}+R_{11}=0.$}\label{CMCK} \end{figure} \begin{figure}[h] $$L(D)= \begin{pNiceMatrix}[r,first-row,last-row,first-col] \ \ \ \ \ & \text{ \ } & \text{ \ } & \text{ \ } & \text{ \ } & \text{ \ } & \text{ \ } & \text{ \ } & \text{ \ } & \text{ \ } & \text{ \ } \\ \ \ \ \ \ \ \ & \textbf{6} & -6 & -2 & \textbf{2} & -4 & -10 & -3 & 4 & 8 & 12 \ \ \\ \ \ \ \ \ \ & \textbf{-33} & 32 & 12 & \textbf{-8} & 22 & 52 & 17 & -22 & -44 & -66 \ \ \\ \ \ \ \ \ \ & \textbf{-9} & 9 & 4 & \textbf{-2} & 6 & 14 & 5 & -6 & -12 & -18 \ \ \\ \ \ \ \ \ \ & \textbf{21} & -21 & -8 & \textbf{5} & -14 & -34 & -11 & 14 & 28 & 42 \ \ \\ \ \ \ \ \ \ &\textbf{-21} & 21 & 8 & \textbf{-5} & 14 & 33 & 11 & -14 & -28 & -42 \\ \ \ \ \ \ \ &\textbf{3} & -3 & -1 & \textbf{1} & -2 & -5 & -1 & 2 & 4 & 6 \\ \ \ \ \ \ \ & \textbf{39} & -39 & -15 & \textbf{9} & -27 & -63 & -21 & 26 & 52 & 78 \\ \ \ \ \ \ \ & \textbf{13} & -13 & -5 & \textbf{3} & -9 & -21 & -7 & 9 & 18 & 26 \\ \ \ \ \ \ \ & \textbf{-13} & 13 & 5 & \textbf{-3} & 9 & 21 & 7 & -9 & -18 & -27 \\ \ \ \ \ \ \ &\textbf{26} & -26 & -10 & \textbf{6} & -18 & -42 & -14 & 18 & 35 & 52 \\ & \textbf{0} & 0 & 0 & \textbf{0} & 0 & 0 & 0 & 0 & 0 & 0 \ \ \\ \end{pNiceMatrix}$$ \caption{Coloring matrix for the Conway knot. The last row of zeroes correspond to the coloring of the base arc.}\label{pseudoConwayKnot} \end{figure} \begin{example}\label{Ex819} Consider the torus knot $T(3,4)$ with diagram $D$ and crossings and arcs ordered as illustrated in Figure \ref{orderarcs} . Its crossing matrix is shown in Figure \ref{CprimeMatrix819}. Three columns of $C^{-1}(D)$ (shown in Figure \ref{CMatrix819}) are integral and they yield $\epsilon$-pseudo colorings. Column 5 gives a $-1$-pseudo coloring (shown on the right in Figure \ref{8_19aspretzel}) and columns 1 and 2 give -1-pseudo colorings. The $+1$-pseudo coloring corresponding to column $1$ is shown on the left of Figure \ref{8_19aspretzel}. \begin{figure}[ht] \centering \includegraphics[width=0.85\linewidth]{8_19aspretzel.png} \caption{The torus knot $T(3,4)$ ($8_{19}$ in the Rolfsen's table \cite{Rol}) depicted as the pretzel knot $P(3,3,-2)$. } \label{8_19aspretzel} \end{figure} \begin{figure}[h] $C'(D)=\left( \begin{array}{rrrrrrrr} 2& 0 &0 &0 & -1 & -1 & 0&0 \\ -1 & -1 & 0 &0 &2 &0 &0 &0 \\ 0& 0 & 0 &0 &2 &0 &-1&-1 \\ 0&0 &0 &-1 &-1 &0 &2 &0 \\ 0& 0 &0 & 2&0 &-1 &-1 &0 \\ 2& 0 & -1 &-1 &0 & 0&0 &0 \\ -1 & 0 & 2 &0 & 0& 0&0 & -1 \\ 0& -1 & -1 &0 &0 &0 &0 & 2 \\ \end{array} \right)$ \caption{The crossing matrix $C'(P(3,3,-2))$. The rows satisfy the linear relation $R_1 +R_2 - R_3 -R_4 -R_5 -R_6 -R_7 - R_8 = 0$.} \label{CprimeMatrix819} \end{figure} \setlength{\tabcolsep}{20pt} \renewcommand{\arraystretch}{1.5} \begin{figure}[ht] $\displaystyle C^{-1}(D)=\displaystyle \left( \begin{array}{rrrrrrr} -2 & 0 & -\frac{2}{3} & \frac{2}{3} & 2 & \frac{10}{3} & \frac{5}{3} \\ 0 & -1 & \frac{4}{3} & \frac{2}{3} & 0 & -\frac{2}{3} & -\frac{1}{3} \\ -1 & 0 & -\frac{1}{3} & \frac{1}{3} & 1 & \frac{5}{3} & \frac{4}{3} \\ -3 & 0 & -1 & 1 & 3 & 4 & 2 \\ -1 & 0 & \frac{1}{3} & \frac{2}{3} & 1 & \frac{4}{3} & \frac{2}{3} \\ -4 & 0 & -\frac{5}{3} & \frac{2}{3} & 3 & \frac{16}{3} & \frac{8}{3} \\ -2 & 0 & -\frac{1}{3} & \frac{4}{3} & 2 & \frac{8}{3} & \frac{4}{3} \\ \end{array} \right)$ \caption{$C^{-1}(D)$ corresponding to $T(3,4)$ with three integral columns.} \label{CMatrix819} \end{figure} \end{example} Non-alternating link diagrams always have $\epsilon$-pseudo colorings, as we describe in the following remark. \begin{remark}\label{remarkpseudo} Let $D$ be a non-alternating link diagram. \begin{enumerate} \item[(1)] Every integral column of $C^{-1}(D)$ leads to some $\epsilon$-pseudo coloring. \item[(2)] $D$ has an $\epsilon$-pseudo coloring. This follows from the fact that every non-alternating diagram has a tunnel of length at least two. Now, we can color $D$ by coloring one of the arcs of the tunnel by color $-1$ and all other arcs by color $0$ to get the $+1$-pseudo coloring. An example of such a coloring is shown on the right-hand side of Figure \ref{orderarcs}. \end{enumerate} \end{remark} \subsection{Future directions} The Kauffman-Harary conjecture was extended to the case of virtual knots by Mathew Williamson \cite{Wil} and proved by Zhiyun Cheng \cite{Che}. A natural question is to ask whether the conjecture in \cite{APS} holds for virtual links whose determinants are not prime. Another path of further research is to look for a natural generalization to non-alternating diagrams using a set theoretic Yang-Baxter operator or a general Yang-Baxter operator. \ An interesting prospect is to approach the generalized Kauffman-Harary conjecture from the perspective of incompressible surfaces in the double branched cover $M^{(2)}_L$ of $S^3$ branched along $L$. This was outlined in \cite{APS} with the hope of proving the GKH conjecture. Now that the GKH conjecture is proved, we can proceed in the opposite direction and analyze incompressible surfaces in $M^{(2)}_L$. \section*{Acknowledgements} The first author acknowledges the support of Dr. Max Rössler, the Walter Haefner Foundation, and the ETH Zürich Foundation. The third author acknowledges the support of the National Science Foundation through Grant DMS-2212736. The fourth author was supported by the American Mathematical Society and the Simons Foundation through the AMS-Simons Travel Grant. The fifth author was partially supported by the Simons Collaboration Grant 637794.
{ "timestamp": "2023-01-09T02:13:55", "yymm": "2301", "arxiv_id": "2301.02645", "language": "en", "url": "https://arxiv.org/abs/2301.02645", "abstract": "For a reduced alternating diagram of a knot with a prime determinant $p,$ the Kauffman-Harary conjecture states that every non-trivial Fox $p$-coloring of the knot assigns different colors to its arcs. In this paper, we prove a generalization of the conjecture stated nineteen years ago by Asaeda, Przytycki, and Sikora: for every pair of distinct arcs in the reduced alternating diagram of a prime link with determinant $\\delta,$ there exists a Fox $\\delta$-coloring that distinguishes them.", "subjects": "Geometric Topology (math.GT)", "title": "The Generalized Kauffman-Harary Conjecture is True", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9857180673335565, "lm_q2_score": 0.8128673155708975, "lm_q1q2_score": 0.8012579993031613 }
https://arxiv.org/abs/2106.12190
Closed-Form, Provable, and Robust PCA via Leverage Statistics and Innovation Search
The idea of Innovation Search, which was initially proposed for data clustering, was recently used for outlier detection. In the application of Innovation Search for outlier detection, the directions of innovation were utilized to measure the innovation of the data points. We study the Innovation Values computed by the Innovation Search algorithm under a quadratic cost function and it is proved that Innovation Values with the new cost function are equivalent to Leverage Scores. This interesting connection is utilized to establish several theoretical guarantees for a Leverage Score based robust PCA method and to design a new robust PCA method. The theoretical results include performance guarantees with different models for the distribution of outliers and the distribution of inliers. In addition, we demonstrate the robustness of the algorithms against the presence of noise. The numerical and theoretical studies indicate that while the presented approach is fast and closed-form, it can outperform most of the existing algorithms.
\section{Introduction} Principal Component Analysis (PCA) has been extensively used for linear dimensionality reduction. While PCA is useful when the data has low intrinsic dimension, its output is sensitive to outliers in the sense that the subspace found by PCA can arbitrarily deviate from the true underlying subspace even if a small portion of the data is corrupted. In addition, locating the outlying components is of great interest in many applications. There are two different robust PCA problems corresponding to two different models for the data corruption. The first problem, known as low rank plus sparse matrix decomposition, assumes that a random subset of the elements of data are corrupted and the corrupted elements are not concentrated in any column/row of the data~\cite{lamport22,lamport1}. In the second problem, a subset of columns of the data \textcolor{black}{are affected by the data corruption}~\cite{lamport10,zhang2014novel,lerman2014fast,fischler1981random,li1985projection,choulakian2006l1,feng2012robust,mccoy2011two,hardt2013algorithms,zhang2016robust,you2017provable,markopoulos2014optimal}. This paper focuses on the column-wise model, i.e., it is assumed that data matrix $\mathbf{D} \in \mathbb{R}^{M_1 \times M_2}$ can be expressed as $$ \mathbf{D} = ( [\mathbf{B} \hspace{.2cm} \mathbf{A}] ) \: \mathbf{T} \:, $$ where $\mathbf{A} \in \mathbb{R}^{M_1 \times n_i}$, $\mathbf{B} \in \mathbb{R}^{M_1 \times n_o}$, $\mathbf{T}$ is an unknown permutation matrix, and $[\mathbf{B} \hspace{.2cm} \mathbf{A}]$ represents the concatenation of matrices $\mathbf{A}$ and $\mathbf{B}$. The columns of $\mathbf{A}$ lie in an $r$-dimensional subspace $\mathcal{U}$. The columns of $\mathbf{B}$ do not lie entirely in $\mathcal{U}$, i.e., the $n_i$ columns of $\mathbf{A}$ are the inliers and the $n_o$ columns of $\mathbf{B}$ are the outliers. The output of a robust PCA method is an estimate for $\mathcal{U}$. If $\mathcal{U}$ is estimated accurately, the outliers can be located by projecting the data points on the complement of $\mathcal{U}$. \subsection{Summary of Contributions} Most of the existing robust PCA algorithms require a large number of iterations each with high computational complexity and most of them are not supported with thorough performance guarantees. We present closed-form and provable robust PCA methods which mostly outperform the existing methods. The main contributions can be summarized as follows. \\ $\bullet$ It is proved that Innovation Value introduced in~\cite{rahmani2019outlier2} is equivalent to Leverage Score if a quadratic cost function is used to find the optimal directions instead of the $\ell_1$-norm based cost function. This interesting connection is used to establish several theoretical guarantees \\ $\bullet$ Inspired by the explanation of Leverage Scores with Innovation Search, a new robust PCA method, which uses a symmetric measure of similarity, is presented. The presented closed-form methods mostly outperform the existing methods while they only include one singular value decomposition plus one matrix multiplication. \\ $\bullet$ Theoretical performance guarantees under several different models for the distribution of the outliers and inliers are presented. Furthermore, the robustness to the presence of noise is studied and it is shown that the algorithm can provably distinguish the outliers when the data is noisy. \subsection{Notation} Given a matrix $\mathbf{A}$, $\| \mathbf{A} \|$ denotes its spectral norm, $\| \mathbf{A} \|_F$ denotes its Frobenius norm, and $\mathbf{A}^T$ is the transpose of $\mathbf{A}$. For a vector $\mathbf{a}$, $\| \mathbf{a} \|_p$ denotes its $\ell_p$-norm and $\mathbf{a}(i)$ its $i^{\text{th}}$ element. For a matrix $\mathbf{A}$, $\mathbf{a}_i$ denotes its $i^{\text{th}}$ column and $\| \mathbf{A} \|_{1,2} = \sum_{i} \| \mathbf{a}_i \|_2$. $\mathbb{S}^{M_1 - 1}$ indicates the unit $\ell_2$-norm sphere in $\mathbb{R}^{M_1}$. Matrix $\mathbf{D} = \mathbf{U}^{'} \Sigma \mathbf{V}$ where $\mathbf{U}^{'} \in \mathbb{R}^{M_1 \times r_d}$ is the matrix of left singular vectors, $\Sigma \in \mathbb{R}^{r_d \times r_d}$ is a diagonal matrix whose diagonal values are equal to the non-zero singular values of $\mathbf{D}$, the rows of $\mathbf{V} \in \mathbb{R}^{r_d \times M_2}$ are equal to the right singular vectors, and $r_d$ is the rank of $\mathbf{D}$. The orthonormal matrix $\mathbf{U} \in \mathbb{R}^{M_1 \times r}$ is defined as a basis for $\mathcal{U}$. Note that $\mathbf{U}^{'}$ is a basis for the entire data, $\mathbf{U}$ is a basis for the inliers, $r_d$ is the rank of $\mathbf{D}$, and $r$ is the rank of $\mathbf{A}$. The subspace $\mathcal{U}^{\perp}$ is defined as the complement of $\mathcal{U}$. \section{Related Work} \label{sec:related_work} Robust PCA is a well-known problem and many approaches were developed for this problem. In this section, we briefly review some of the previous works on robust PCA. Some of the earliest approaches to robust PCA are based on robust estimation of the data covariance matrix, such as the minimum covariance determinant, the minimum volume ellipsoid, and the Stahel-Donoho estimator~\cite{lamport47,feng2012robust,xu2010principal}. However, these methods mostly compute a full SVD or eigenvalue decomposition in each iteration and their performance greatly degrades when $\frac{n_i}{n_o} < 0.5$. In addition, they lack performance guarantees with structured outliers. Another approach is to replace the Frobenius Norm in the cost function of PCA with $\ell_1$-norm~\cite{lamport18,lamport29}, as $\ell_1$-norm was shown to be robust to the presence of the outliers~\cite{decod,lamport29}. In order to leverage the column-wise structure of the outliers, the authors of~\cite{lamport21} replaced the $\ell_1$-norm minimization problem used in~\cite{lamport29} with an $\ell_{1,2}$-norm minimization problem. In~\cite{lerman2015robustnn} and~\cite{zhang2014novel}, the optimization problem used in~\cite{lamport21} was relaxed to two different convex optimization problems. The authors of~\cite{lerman2015robustnn,zhang2014novel} provided sufficient conditions \textcolor{black}{under which} the optimal points of the convex optimization problems proposed in~\cite{lerman2015robustnn,zhang2014novel} are guaranteed to yield an exact basis for $\mathcal{U}$ . The approach presented in~\cite{tsakiris2015dual} focused on the scenario in which the data is predominantly unstructured outliers and the number of outliers is larger than $M_1$. In~\cite{tsakiris2015dual}, it is essential to assume that the outliers are randomly distributed on $\mathbb{S}^{M_1 - 1}$ and the inliers are distributed randomly on the intersection of $\mathbb{S}^{M_1 - 1}$ and $\mathcal{U}$. The outlier detection method proposed in~\cite{soltanolkotabi2012geometric} assumes that the outliers are randomly distributed on $\mathbb{S}^{M_1 - 1}$ and a small number of them are not linearly dependent which means that~\cite{soltanolkotabi2012geometric} is not able to detect the linearly dependent outliers and the outliers which are close to each other. Another approach is based on decomposing the given $\mathbf{D}$ into a low rank matrix and a column sparse matrix where the column sparse matrix models the presence of the outliers \cite{cherapanamjeri2017thresholding,lamport10}. However, this approach requires the number of outliers to be significantly smaller than the number of the outliers and the solver algorithms that are used to decompose the data need to compute a SVD of the data in each iteration. The main shortcomings of the previous methods are sensitivity to structured outliers and the lack of comprehensive theoretical guarantees. The Coherence Pursuit method, proposed in~\cite{rahmani2017coherence}, was shown (theoretically and numerically) to be robust to different types of outliers. However, Coherence Pursuit can miss outliers which carry weak innovation with respect to the inliers. The iSearch algorithm, proposed in~\cite{rahmani2019outlier2}, was shown to notably outperform Coherence Pursuit in detecting outliers with weak innovation . Similar to Coherence Pursuit, the robustness of iSearch against different types of outliers were supported with several theoretical guarantees. However, in contrast to Coherence Pursuit which is a closed-form method, iSearch needs to run an iterative and computationally expensive solver to find the directions of innovation. \textit{This paper presents robust PCA methods which have the advantages of both (CoP and iSearch) algorithms: while they are closed-form algorithms, their ability in distinguishing outliers are on a par with iSearch. In the rest of this section, Coherence Pursuit and iSearch are reviewed in more details.} \begin{comment} In this section, some of the previous works on robust PCA are briefly reviewed. Since the presented approach is more related to Coherence Pursuit ~\cite{rahmani2017coherence} and iSearch ~\cite{rahmani2019outlier2}, these two methods are reviewed with more details. The earliest approaches to robust PCA were mostly based on a robust estimation of the data covariance matrix~\cite{lamport47}. These methods compute a full SVD or eigenvalue decomposition in each iteration and their performance greatly degrades when $\frac{n_i}{n_o} < 0.5$. Another approach to enhance the robustness of PCA against outliers is to replace the Frobenius Norm in the cost function of PCA with $\ell_1$-norm~\cite{lamport18,lamport29} because $\ell_1$-norm were shown to be robust to the presence of the outliers~\cite{decod,lamport29}. The authors of~\cite{lamport21} modified the $\ell_1$-norm minimization problem used in~\cite{lamport29} and replaced it with an $\ell_{1,2}$-norm minimization problem to leverage the column-wise structure of the outliers. In~\cite{lerman2015robustnn} and~\cite{zhang2014novel}, the optimization problem used in~\cite{lamport21} was relaxed to two different convex optimization problems. The approach presented in~\cite{tsakiris2015dual} assumes that the data is predominantly unstructured outliers and the outliers/inliers are randomly distributed in $\mathbb{S}^{M_1-1}$/$\mathbb{S}^{M_1-1}\cap \mathcal{U}$ In~\cite{lamport10}, a convex optimization problem was proposed which decomposes the data into a low rank component and a column sparse component. The approach presented in~\cite{lamport10} is provable but it requires $n_o$ to be significantly smaller than $n_i$. The outlier detection method proposed in~\cite{soltanolkotabi2012geometric} utilizes sparse self representation and it assumes that the outliers are randomly distributed on $\mathbb{S}^{M_1 - 1}$ and a small number of them are not linearly dependent. \end{comment} \vspace{0.05in} \noindent \textbf{Coherence Pursuit (CoP):} CoP ~\cite{rahmani2017coherence} assigns a value, termed Coherence Value, to each data point and $\mathcal{U}$ is recovered using the span of the data points with the highest Coherence Values. The Coherence Value corresponding to data point $\mathbf{d}_i$ represents the similarity between $\mathbf{d}_i$ and the rest of the data points. CoP uses inner-product to measure the similarity between data points and it distinguishes the outliers based on the fact that an inlier bears more resemblance to the rest of the data than an outlier. \vspace{0.05in} \noindent \textbf{Innovation Search (iSearch)~\cite{rahmani2019outlier2}:} Innovation Pursuit was initially proposed as a data clustering algorithm~\cite{rahmani2017subspacedi,rahmani2015innovation}. The authors of~\cite{rahmani2017subspacedi,rahmani2015innovation} showed that Innovation Pursuit can notably outperform the self-representation based clustering methods (e.g. Sparse Subspace Clustering~\cite{lamport7}) specifically when the clusters are close to each other. Innovation Pursuit computes an optimal direction corresponding to each data point $\mathbf{d}_i$ which can be written as the optimal point of \begin{eqnarray} \underset{ \mathbf{c}}{\min} \: \: \| \mathbf{c}^T \mathbf{D} \|_1 \quad \text{subject to} \quad \mathbf{c}^T \mathbf{d}_i = 1 \:. \label{eq:el1_innov} \end{eqnarray} If $\mathbf{C}^{*} \in \mathbb{R}^{M_1 \times M_2}$ contains all the optimal directions, Innovation Pursuit builds the adjacency matrix as $\mathbf{Q} + \mathbf{Q}^T $ where $\mathbf{Q} = |\mathbf{D}^T \mathbf{C}^{*}|$. In~\cite{rahmani2019outlier2}, it was shown that the optimal directions can be utilized for outlier detection too. The approach proposed in~\cite{rahmani2019outlier2}, termed iSearch, assigns an Innovation Value to each data point and it distinguishes the outliers as the data points with the higher Innovation Values. The Innovation Value assigned to $\mathbf{d}_i$ is computed as $$ \frac{1}{\| \mathbf{D}^T \mathbf{c}_i^{*} \|_1} $$ where $\mathbf{c}_i^{*}$ is the optimal point of (\ref{eq:el1_innov}). iSearch needs to run an iterative solver to find the optimal directions. In contrast, the methods presented in this paper are closed-form and they can be hundreds of time faster. \vspace{0.05in} \noindent \textbf{Leverage Statistics:} In regression, Leverage Scores are defined as the diagonal values of the hat matrix $\mathbf{X}^T (\mathbf{X} \mathbf{X}^T)^{-1} \mathbf{X} $ where $\mathbf{X} \in \mathbb{R}^{m_1 \times m_2}$ is the design matrix~\cite{everitt2002cambridge}. Assuming that the rank of $\mathbf{X}$ is equal to $m_1$, then the $i^{\text{th}}$ leverage score is equal to $\| {\mathbf{v}_x}_i \|_2^2$ where the rows of $\mathbf{V}_x \in \mathbb{R}^{m_1 \times m_2}$ are equal to the right singular vectors of $\mathbf{X}$ and ${\mathbf{v}_x}_i$ is the $i^{\text{th}}$ column of $\mathbf{V}_x$. Leverage has been typically used in the regression framework and there are few works which focused on using it for the robust PCA problem~\cite{mejia2017pca,naes1989leverage}. For instance, \cite{mejia2017pca} utilized leverage to reject the outlying time points in an functional magnetic resonance images (fMRI) run. However, there is not still an analysis and full understating of Leverage in the robust PCA setting. We show that Innovation Value introduced in~\cite{rahmani2019outlier2} is equivalent to Leverage Score if a quadratic cost function is used to find the optimal directions. This interesting connection is used to establish several theoretical guarantees and to design a new robust PCA method. \begin{algorithm} \caption{Asymmetric Normalized Coherence Pursuit (ANCP) for Robust PCA } { \textbf{Input.} The inputs are the data matrix $\mathbf{D} \in \mathbb{R}^{M_1 \times M_2}$ and $r$ which is the dimension of the recovered subspace. \smallbreak \textbf{1.} Normalize the $\ell_2$-norm of the columns of $\mathbf{D}$, i.e., set $\mathbf{d}_i$ equal to $\mathbf{d}_i / \| \mathbf{d}_i \|_2$ for all $1 \le i \le M_2$. \smallbreak \textbf{2.} Rows of $\mathbf{V} \in \mathbb{R}^{r_d \times M_2 }$ are equal to the first $r_d$ right singular vectors of $\mathbf{D}$ where $r_d$ is the rank of $\mathbf{D}$. \smallbreak \textbf{3.} Define $\mathbf{x} \in \mathbb{R}^{M_2}$, vector of Normalized Coherence Values, as \begin{eqnarray} \mathbf{x}(i) = \frac{1}{\| \mathbf{v}_i \|_2^2} \: . \label{eq:ANCP} \end{eqnarray} \smallbreak \textbf{4. } Construct matrix $\mathbf{Y}$ from the columns of $\mathbf{D}$ corresponding to the largest elements of $\mathbf{x}$ such that they span an r-dimensional subspace. \smallbreak \textbf{ Output:} The column-space of $\mathbf{Y}$ is the identified subspace. } \end{algorithm} \section{Robust PCA with Leverage and Innovation Search} \label{sec:proposed} The table of Algorithm 1 and the table of Algorithm 2 show the presented methods along with the used definitions. Algorithm 1 utilizes Leverage Scores to rank the data points and the data points with the minimum Leverage Scores are used to build a basis for $\mathcal{U}$. In this section, we show the underlying connection between Algorithm 1 and iSearch and the motivation for naming Algorithm 1 ``Asymmetric Normalized Coherence Pursuit'' is explained. In addition, we explain the motivations behind the design of Algorithm 2 based on the connection between Algorithm 1 and iSearch. Both algorithms are closed-form and they are faster than most of the existing methods. The computation complexity of ANCP is $\mathcal{O}( r_d M_1 M_2)$ and the computation complexity of SNCP is $\mathcal{O}( r_d M_1 M_2 + r_d M_2^2)$. The presented robust PCA algorithms use the data points corresponding to the largest Normalized Coherence Values to form the basis matrix $\mathbf{Y}$. If the inliers are distributed uniformly at random in $\mathcal{U}$, then $r$ data points corresponding to the $r$ largest Normalized Coherence Values span $\mathcal{U}$ with high probability. However, in real data, the inliers form some clustering structure and the algorithm should continue adding new columns to $\mathbf{Y}$ until the columns of $\mathbf{Y}$ span an $r$-dimensional subspace. It means that we need to check the singular values of $\mathbf{Y}$ multiple times. Two techniques can be utilized to \textcolor{black}{avoid these extra steps}~\cite{rahmani2019outlier2,rahmani2017coherence22f}. The first approach is based on leveraging side information that we mostly have about the population of the outliers. In most of the applications, we can have an upper-bound on $n_o/M_2$ because outliers are mostly associated with rare events. If we know that the number of outliers is less than $y$ $\%$ of the data, matrix $\mathbf{Y}$ can be constructed using $(1 - y)$ $\%$ of the data columns which are corresponding to the largest Normalized Coherence Values. The second technique is to use the adaptive column sampling method proposed in~\cite{rahmani2017coherence22f} which uses subspace projection to avoid sampling redundant columns. \begin{algorithm} \caption{Symmetric Normalized Coherence Pursuit (SNCP) for Robust PCA} \textbf{Input.} The inputs are the data matrix $\mathbf{D} \in \mathbb{R}^{M_1 \times M_2}$ and $r$ which is the dimension of the recovered subspace. \smallbreak \textbf{1.} Similar to Step 1 in Algorithm 1. \smallbreak \textbf{2.} Define $\mathbf{V} \in \mathbb{R}^{r_d \times M_2 }$ as in Algorithm 1. \smallbreak \textbf{3.} The vector of Normalized Coherence Values is defined as $$\mathbf{x}(i) = \sum_{j=1}^{M_2} \frac{({\mathbf{v}_i}^T \mathbf{v}_{j})^2}{ \|\mathbf{v}_i\|_2^2 \|\mathbf{v}_j\|_2^2 }\:.$$ \smallbreak \textbf{4. } Construct matrix $\mathbf{Y}$ as in Algorithm 1. \smallbreak \textbf{ Output:} The column-space of $\mathbf{Y}$ is the identified subspace. } \end{algorithm} \subsection{Explaining Leverage Score for Robust PCA Using Innovation Search}\label{sec:expl} Algorithm 1 ranks the data points based on the inverse of their leverage scores. The following lemma shows that Leverage Score is directly related to Innovation Value. \begin{lemma} Suppose rows of $\mathbf{V} \in \mathbb{R}^{r_d \times M_2 }$ are equal to the first $r_d$ right singular vectors of $\mathbf{D}$ where $r_d$ is the rank of $\mathbf{D}$. Define $\mathbf{c}_i^{*}$ as the optimal point of \begin{eqnarray} \underset{ \mathbf{c}}{\min} \: \: \| \mathbf{c}^T \mathbf{D} \|_2 \quad \text{subject to} \quad \mathbf{c}^T \mathbf{d}_i = 1 \:. \label{eq:el_2_inno} \end{eqnarray} Then, $ \| \mathbf{D}^T \mathbf{c}_i^{*} \|_2^2 = \frac{1}{\| \mathbf{v}_i \|_2^2} \:. $ \label{lem:equi} \end{lemma} \noindent Lemma~\ref{lem:equi} indicates that if a quadratic cost function is used to compute the optimal directions in iSearch, Innovation Values are equivalent to Leverage Scores. Accordingly, we can use the idea of Innovation Search to explain the Leverage Score based robust PCA method. First suppose that $\mathbf{d}_i$ is an outlier which means that $\mathbf{d}_i$ has a non-zero projection on $\mathcal{U}^{\perp}$. Since most of the data points are inliers, the optimization problem utilizes the projection of $\mathbf{d}$ in $\mathcal{U}^{\perp}$ and finds the optimal direction near $\mathcal{U}^{\perp}$ to minimize $\| \mathbf{A}^T \mathbf{c}_i\|_2^2$. In sharp contrast, when $\mathbf{d}_i$ is an inlier, the linear constraint strongly discourages the optimal direction to be close to $\mathcal{U}^{\perp}$. Thus, $\| \mathbf{A}^T \mathbf{c}_i^{*}\|_2^2$ is notably larger when $\mathbf{d}_i$ is an inlier comparing to $\| \mathbf{A}^T \mathbf{c}_i^{*}\|_2$ when $\mathbf{d}_i$ is an outlier. Accordingly, since $\frac{1}{\| \mathbf{v}_i \|_2^2} = \| \mathbf{D}^T \mathbf{c}_i^{*}\|_2^2 = \|\mathbf{A}^T \mathbf{c}_i^{*} \|_2^2 + \|\mathbf{B}^T \mathbf{c}_i^{*} \|_2^2$, $1/\| \mathbf{v}_i \|_2^2$ is much larger when $\mathbf{d}_i$ is an inlier comparing to the same value when $\mathbf{d}_i$ is an outlier because $\|\mathbf{A}^T \mathbf{c}_i^* \|_2^2$ is much larger when $\mathbf{d}_i$ is an inlier. The following Lemma indicates that $\| \mathbf{D}^T \mathbf{c}_i^{*} \|_2^2$ can be written as the sum of the similarities between the columns of $\mathbf{V} \in \mathbb{R}^{r_d \times M_2}$. \begin{lemma} Define $\mathbf{c}_i^{*}$ and $\mathbf{V}$ as in Lemma~\ref{lem:equi}. Then, \begin{eqnarray} \| \mathbf{D}^T \mathbf{c}_i^{*} \|_2^2 = \sum_{j=1}^{M_2} \left(\frac{ \mathbf{v}_i^T \mathbf{v}_j }{ \|\mathbf{v}_i\|_2 \|\mathbf{v}_i\|_2 }\right)^2 \:. \label{eq:sum_asymmm} \end{eqnarray} \label{lm:secondlm} \end{lemma} Thus, Algorithm 1 is inherently similar to CoP but Algorithm 1 utilizes the coherency between the columns of $\mathbf{V}$. In other word, the functionality of Algorithm 1 is similar to that of a CoP algorithm which is applied to a data matrix whose non-zero singular values are normalized to 1. This is the motivation to name the presented algorithms Normalized Coherence~Pursuit. \subsection{Symmetric Normalized Coherence Pursuit} The measure of similarity used in (\ref{eq:sum_asymmm}) is not symmetric. In other word, the similarity between $\mathbf{d}_i$ and $\mathbf{d}_j$ which is computed as $\left(\frac{ \mathbf{v}_i^T \mathbf{v}_j }{ \|\mathbf{v}_i\|_2 \|\mathbf{v}_i\|_2 }\right)^2 $ is not equal to the similarity between $\mathbf{d}_j$ and $\mathbf{d}_i$ which can be written as $\left(\frac{ \mathbf{v}_i^T \mathbf{v}_j }{ \|\mathbf{v}_j\|_2 \|\mathbf{v}_j\|_2 }\right)^2$. Accordingly, we modify the measure of similarity used in (\ref{eq:sum_asymmm}) into a symmetric measure of similarity. The Normalized Coherence Value corresponding to $\mathbf{d}_i$ using the symmetric measure of similarity is defined as $ \mathbf{x}(i) = \sum_{j=1}^{M_2} \left(\frac{ \mathbf{v}_i^T \mathbf{v}_j }{ \|\mathbf{v}_i\|_2 \|\mathbf{v}_j\|_2 }\right)^2 \:. $ Algorithm 2 uses the symmetric measure of similarity to compute the Normalized Coherence Values. The numerical experiments show that utilizing the symmetric function can notably improve the performance in most of the cases. \begin{comment} \begin{figure}[h!] \begin{center} \mbox{\hspace{-0.15in} \includegraphics[width=1.55in]{figs/noise_values.eps}\hspace{-0.15in} \includegraphics[width=1.55in]{figs/noise_values2.eps} } \end{center} \vspace{-0.15in} \caption{ In both plots, $n_i=100$, $n_o=500$, $r= 4$, $M_1 = 50$, the last 100 columns are the inliers, and the outliers are sampled uniformly at random from $\mathbb{S}^{M_1-1}$. In the left plot the data is not noisy and in the right plot $\mathbf{D} = [\mathbf{B} \:\:(\mathbf{A} + \mathbf{E})]$ where $\mathbf{E}$ represents the presence of noise and $\text{SNR} = \frac{\|\mathbf{A}\|_F^2}{\| \mathbf{E} \|_F^2} = 4$. } \label{fig:example_noise} \end{figure} \end{comment} \section{Theoretical Studies} In this section, we present analytical performance guarantees for Normalize Coherence Pursuit under different models for the distribution of the outliers. The connection between Innovation Values and Leverage Scores is utilized to analyze the ANCP method and we leave the analysis of SNCP to future works. In the following subsections, the performance guarantees with unstructured outliers, linearly dependant outliers, noisy inliers, and clustered outliers is provided. Moreover, in contrast to most of the previous methods whose guarantees are limited to randomly distributed inliers, Normalized Coherence Pursuit is supported with theoretical guarantees even when the inliers are clustered. The following sections provide the theoretical results and each theorem is followed by a short discussion which highlights the important aspects of that theorem. To simplify the exposition and notation, in the presented results, it is assumed without loss of generality that $\mathbf{T}$ is equal to the identity matrix, i.e, $\mathbf{D} = [\mathbf{B} \hspace{.2cm} \mathbf{A}]$. The subspace $\mathcal{U}$ is recovered using the span of the data points with the largest Normalized Coherence Values. A sufficient condition which guarantees exact recovery of $\mathcal{U}$ is that the minimum of the Normalized Coherence Values corresponding to the inliers is larger than the maximum of the Normalized Coherence Values corresponding to the outliers, i.e., \begin{eqnarray} \begin{aligned} &\min \left( \{ \mathbf{x}(i) \}_{i=n_o+1}^{M_2} \right) > \max \left( \{ \mathbf{x}(i) \}_{i=1}^{n_o} \right) \:. \end{aligned} \label{cond:main_cond} \end{eqnarray} \noindent This is not a necessary condition but it is easier to guarantee. In addition, we define $$\psi = \max \left( \left\{ \frac{1}{\| \mathbf{b}_i^T \mathbf{R}\|_2^2} \right\}_{i=1}^{n_o} \right)$$ where $\mathbf{R}$ is an orthonormal basis for $\mathcal{U}^{\perp}$ and $\mathbf{b}_i$ is the $i^{\text{th}}$ column of $\mathbf{B}$. The parameter $\psi$ indicates how close the outliers are to $\mathcal{U}$. \vspace{0.05in} \noindent \textbf{Proof Strategy:} Although the presented theorems consider different scenarios for the distribution of the inliers/outliers and different techniques are required to guarantee (\ref{cond:main_cond}) in each case, but a similar strategy is used in the proofs of all the results. In contrast to Cop which analyzed $\{ |\mathbf{d}_i^T \mathbf{d}_j|\}_{i,j}$ based on the distribution of the data, it is not straightforward to directly bound $\{ |\mathbf{v}_i^T \mathbf{v}_j|\}_{i,j}$. In addition, in contrast to iSearch which leveraged the fact that the optimal direction of (\ref{eq:el1_innov}) is mostly orthogonal to $\mathcal{U}$ when $\mathbf{d}_i$ is an outlier, the optimal direction obtained by (\ref{eq:el_2_inno}) is not necessarily orthogonal to $\mathcal{U}$ when $\mathbf{d}_i$ is an outlier. In the proofs of the presented results, we utilized the geometry of the problem in which the optimal direction of (\ref{eq:el_2_inno}) is not close to $\mathcal{U}$ when $\mathbf{d}_i$ is an outlier. Specifically, corresponding to outlier $\mathbf{d}_i$, we define $\mathbf{d}_i^{\perp} = \frac{\mathbf{R} \mathbf{R}^T \mathbf{d}_i}{\| \mathbf{d}_i^T \mathbf{R} \|_2^2}$ and by definition $\mathbf{d}_i^T \mathbf{d}_i^{\perp} = 1$. According to the definition of $\mathbf{c}_i^{*}$ as the optimal point of (\ref{eq:el_2_inno}) and according to Lemma~\ref{lem:equi}, $\frac{1}{\| \mathbf{v}_i \|_2^2} \le \| \mathbf{D}^T \mathbf{d}_i^{\perp} \|_2^2 $. We utilized this inequality to derive the sufficient conditions which prove that (\ref{cond:main_cond}) holds with high probability. The detailed proofs of all the results are provided in the appendix. \subsection{Outliers Distributed on $\mathbb{S}^{M_1 -1}$} \label{sec:uns} In most of the previous works on the robust PCA problem, the performance of the outlier detection method is analyzed under the assumption that the outliers are randomly distributed on $\mathbb{S}^{M_1 - 1}$. This is a simple scenario because the outliers are unstructured and the projection of each outlier on $\mathcal{U}$ is not strong with high probability given that $r$ is sufficiently small because when the outliers are randomly distributed as in Assumption \ref{assum_DistUni}, then $\mathbb{E} [\| \mathbf{U}^T \mathbf{b} \|_2^2] = \frac{r}{M}$ ( $\mathbf{b}$ is an outlying data point). The following assumption specifies the presumed model for the distribution of the inliers/outliers. \begin{assumption} The columns of $\mathbf{A}$ are drawn uniformly at random from $\mathcal{U} \cap \mathbb{S}^{M_1 -1}$ and the columns of $\mathbf{B}$ are drawn uniformly at random from $\mathbb{S}^{M_1 - 1}$. \label{assum_DistUni} \end{assumption} The following theorem provides the sufficient conditions to guarantee the exact recovery of $\mathcal{U}$. \begin{theorem} \label{theo:randomrandom} Suppose $\mathbf{D}$ follows Assumption 1. If $\mathbf{x}$ is defined as in (\ref{eq:ANCP}) and \begin{eqnarray} \begin{aligned} & \frac{n_i}{r} - \max \left( \frac{4}{3} \log \frac{2r}{\delta} , \sqrt{4 \frac{n_i}{r} \log \frac{2 r}{\delta}} \right) > \frac{(\psi - 1)n_o}{M_1} + \\ & \quad \max \left( \frac{8}{3} \log \frac{2M_1}{\delta} , \sqrt{16 \frac{n_o}{M_1} \log \frac{2 M_1}{\delta}} \right) \:, \end{aligned} \label{eq:suff_1} \end{eqnarray} then (\ref{cond:main_cond}) holds and $\mathcal{U}$ is recovered exactly with probability at least $1 - 3 \delta$. \label{theo:random} \end{theorem} Since the outliers are randomly distributed, the expected value of $\| \mathbf{b}_i^T \mathbf{R} \|_2^2$ is equal to $\frac{M_1- r}{M_1}$ which is nearly equal to~1 when $r/M_1$ is small~\cite{rahmani2017coherence,park2014greedy}. Thus, Theorem~\ref{theo:random} roughly indicates that if ${n_i}/{r}$ is sufficiently larger than $n_o/M_1$, Normalized Coherence Values can successfully distinguish the outliers. It is important to note that $n_i$ is scaled with $r$ while $n_o$ is scaled with $M_1$. It means that if $r$ is sufficiently small and if the outliers are unstructured, $\mathcal{U}$ can be recovered exactly even if $n_o$ is much larger than $n_i$. In addition, one can observe that when the outliers are unstructured, the requirements of Normalized Coherence Pursuit are similar with those of CoP~\cite{rahmani2017coherence}. In the next subsection, we observe a clear difference between their requirements when the outliers are close to $\mathcal{U}$. \subsection{Outliers in an Outlying Subspace} \label{sec:outlier_sub} Although Assumption~\ref{assum_DistUni} is a popular data model in the literature of robust PCA, it is not a realistic assumption in the practical scenarios. In practice, outliers can be structured and they are not completely independent from each other as it is assumed in Assumption~\ref{assum_DistUni}. For instance, in anomaly event detection, the outlying video frames are highly correlated or in misclassified data points identification, they can belong to the same cluster~\cite{gitlin2018improving}. In this section, we study the robustness against linearly dependant outliers. The following assumption specifies the presumed model for the outliers. \begin{assumption} Define subspace $\mathcal{U}_o$ with dimension $r_o$ such that $\mathcal{U}_o \notin \mathcal{U}$ and $\mathcal{U} \notin \mathcal{U}_o$. The columns of $\mathbf{A}$ are randomly distributed on $\mathcal{U} \cap \mathbb{S}^{M_1 -1}$ and the columns of $\mathbf{B}$ are randomly distributed on $\mathcal{U}_o \cap \mathbb{S}^{M_1 - 1} $ \label{asm:out} \end{assumption} The following theorem provides the sufficient condition to guarantee that (\ref{cond:main_cond}) holds. \begin{theorem} \label{theo:Linearly_dependant} Suppose $\mathbf{D}$ follows Assumption~\ref{asm:out}. If $\mathbf{x}$ is defined as in (\ref{eq:ANCP}) and \begin{eqnarray*} \begin{aligned} &\frac{n_i}{r} - \max \left( \frac{4}{3} \log \frac{2r}{\delta} , \sqrt{4 \frac{n_i}{r} \log \frac{2 r}{\delta}} \right) > \\ & \| \mathbf{U}_{o}^T \mathbf{U}^{\perp} \|\left( \frac{\psi n_o}{r_o} + \psi \max \left( \frac{4}{3} \log \frac{2r_o}{\delta} , \sqrt{\frac{n_o}{r_o} \log \frac{2 r_o}{\delta}} \right)\right) \end{aligned} \end{eqnarray*} then (\ref{cond:main_cond}) is satisfied and $\mathcal{U}$ is recovered exactly with probability at least $1 - 2 \delta$. \end{theorem} Theorem~\ref{theo:Linearly_dependant} roughly states that if $n_i/r$ is sufficiently larger than $n_o/r_o$, the exact recovery is guaranteed with high probability. If $r_o$ is comparable to $r$, then the number of inliers should be sufficiently larger than the number of outliers. This confirms our intuition about the outliers because if $r_o$ is comparable to $r$ and $n_o$ is also large, we cannot label the columns of $\mathbf{B}$ as outliers. It is informative to compare the requirements of Normalized Coherence Pursuit with that of CoP. The following theorem provides the sufficient conditions to guarantee that CoP successfully distinguishes the outliers. \begin{theorem} \label{theo:Linearly_dependant_CP} Suppose $\mathbf{D}$ follows Assumption~\ref{asm:out}. If \begin{eqnarray*} \begin{aligned} & \frac{n_i}{r} - \max \left( \frac{4}{3} \log \frac{2r}{\delta} , \sqrt{4 \frac{n_i}{r} \log \frac{2 r}{\delta}} \right) > \\ & \|\mathbf{U}^T \mathbf{U}_o \|^2 \left( \frac{n_i}{r} + \max \left( \frac{4}{3} \log \frac{2r}{\delta} , \sqrt{4 \frac{n_i}{r} \log \frac{2 r}{\delta}} \right) \right) + \\ & \frac{n_o}{r_o} + \max \left( \frac{4}{3} \log \frac{2r_o}{\delta} , \sqrt{4 \frac{n_o}{r_o} \log \frac{2 r_o}{\delta}} \right) \: , \end{aligned} \label{eq:cop_reqq} \end{eqnarray*} then the CoP method proposed in~\cite{rahmani2017coherence} recovers $\mathcal{U}$ exactly with probability at least $1- 3\delta$. \end{theorem} One can observe that the requirement of Theorem~\ref{theo:Linearly_dependant_CP} is much stronger than that of Theorem~\ref{theo:Linearly_dependant} because $n_i/r$ appears on the right hand side of the sufficient condition of Theorem~\ref{theo:Linearly_dependant_CP}. Theorem~\ref{theo:Linearly_dependant_CP} predicts that when $\mathcal{U}_o$ is close $\mathcal{U}$, the CoP Algorithm is more likely to fail. This is a correct prediction because when $\mathcal{U}$ and $\mathcal{U}_o$ are close, the inliers and the outliers are close to each other and their inner-product values are large. In addition, by comparing the sufficient conditions of Normalized Coherence Pursuit with that of iSearch~\cite{rahmani2019outlier2} with linearly dependant outliers, we can observe that the nature of the sufficient conditions are similar. In the presented experiments, it is shown that Normalized Coherence Pursuit is on a par with iSearch in identifying outliers with weak innovation while it is a closed-from algorithm and its running time is much faster. \begin{comment} \subsection{Clustered Outliers} In this section, we consider a different structure for the outliers. It is assumed that the outliers form a cluster outside the span of the inliers. Structured outliers are mostly associated with important rare events such as malignant tissues~\cite{karrila2011comparison} or web attacks~\cite{kruegel2003anomaly}. The following assumption specifies the presumed model for $\mathbf{B}$. \begin{assumption} Each column of $\mathbf{B}$ is formed as $\mathbf{b}_i = \frac{1}{\sqrt{1 + \eta^2}} ( \mathbf{q} + \eta \mathbf{f}_i)$. The unit $\ell_2$-norm vector $\mathbf{q}$ does not lie in $\mathcal{U}$, $\{\mathbf{f}_i \}_{i=1}^{n_o}$ are drawn uniformly at random from $\mathbb{S}^{M_1 - 1}$, and $\eta$ is a positive number. \label{asm:clus} \end{assumption} \noindent In Assumption~\ref{asm:clus}, the outliers form a cluster around vector $\mathbf{q}$ which does not lie in $\mathcal{U}$ and $\eta$ determines how close they are to each other. The following theorem provides the sufficient conditions to guarantee that Normalized Coherence Values distinguish the cluster of outliers. \begin{theorem} \label{theo:clustered_outliers} Suppose that the distribution of inliers/outliers follows Assumption-\ref{assum_DistUni}/Assumption-\ref{asm:clus}. If If $\mathbf{x}$ is defined as in (\ref{eq:ANCP}) and $\frac{n_i}{r} - \max \left( \frac{4}{3} \log \frac{2r}{\delta} , \sqrt{4 \frac{n_i}{r} \log \frac{2 r}{\delta}} \right) > n_o \frac{\psi \| \mathbf{q}^T \mathbf{U}^{\perp} \|_2^2}{1 + \eta^2} + \frac{\psi \eta^2 n_o}{(1 + \eta^2 )M_1} + \frac{\eta^2 \psi}{1 + \eta^2} \max \left( \frac{4}{3} \log \frac{2M_1}{\delta} , \sqrt{\frac{n_o}{M_1} \log \frac{2 M_1}{\delta}} \right) + \frac{ \eta \sqrt{\psi}}{1+ \eta^2} \| \mathbf{q}^T \mathbf{U}^{\perp} \|_2 \left( \frac{n_o}{\sqrt{M_1}} + 2\sqrt{n_o} + \sqrt{\frac{2 n_o \log \frac{1}{\delta}}{M_1 -1 }} \right) \:, $ then (\ref{cond:main_cond}) is satisfied and $\mathcal{U}$ is recovered exactly with probability at least $1 - 4 \delta$. \end{theorem} In sharp contrast to Theorem~\ref{theo:randomrandom}, $n_o$ is not scaled with $M_1$. This means that when $\eta$ is small (the outliers are close to each other), $n_i/r$ should be sufficiently larger than $n_o$. When $\eta$ goes to infinity, the distribution of outliers converges to the distribution of outliers in Assumption~\ref{assum_DistUni} and one can observe that the sufficient condition in Theorem~\ref{theo:clustered_outliers} converges to the sufficient condition in Theorem~\ref{theo:randomrandom}. \end{comment} \subsection{Noisy Inliers} \label{sec:nooooise} Although exact recovery of $\mathcal{U}$ is not feasible when the inliers are noisy but the Normalized Coherence Values can distinguish the outliers even in the strong presence of noise. In this section , we present a theorem which guarantees that (\ref{cond:main_cond}) holds with high probability if $n_i/r$ and Signal to Noise Ratio (SNR) are sufficiently large. The following assumption specifies the presumed model. \begin{assumption} The matrix $\mathbf{D}$ can be expressed as $$ \mathbf{D} = [\mathbf{B} \hspace{.2cm} \frac{1}{\sqrt{1+\sigma_n^2}}(\mathbf{A}+\mathbf{E})] \: \mathbf{T} \:. $$ The matrix $\mathbf{E} \in \mathbb{R}^{M_1 \times n_i}$ represents the presence of noise and it can be written as $\mathbf{E} = \sigma_n \mathbf{N}$ where the columns of $\mathbf{N} \in \mathbb{R}^{M_1 \times n_i}$ are drawn uniformly at random from $\mathbb{S}^{M_1 - 1}$ and $\sigma_n$ is a positive number which controls the power of the added noise. \label{asm:noiss} \end{assumption} Before we state the theorem, let us define vectors $\{ \mathbf{t}_i \}_{i=1}^{M_2}$ where $\mathbf{t}_i = \Sigma \mathbf{v}_i$ and the diagonal matrix $\Sigma$ contains the non-zero singular values of $\mathbf{D}$. Note that $\mathbf{d}_i = \mathbf{U}^{'} \mathbf{t}_i$ and $\| \mathbf{t}_i \|_2 = \| \mathbf{d}_i \|_2$. In addition, define $ t_{\min} = \min_i \left( \left\{ \frac{\| \Sigma^{-2} \mathbf{t}_i \|_2}{\mathbf{t}_i^T \Sigma^{-2} \mathbf{t}_i} \right\}_{i=n_o + 1}^{M_2} \right)$ and $ t_{\max} = \max_i \left( \left\{ \frac{\| \Sigma^{-2} \mathbf{t}_i \|_2}{\mathbf{t}_i^T \Sigma^{-2} \mathbf{t}_i} \right\}_{i=n_o + 1}^{M_2} \right) \:. $ \begin{theorem} \label{theo:noise} Suppose $\mathbf{A}$ and $\mathbf{B}$ follow Assumption~\ref{assum_DistUni} and $\mathbf{D}$ is formed according to Assumption~\ref{asm:noiss}. If $\mathbf{x}$ is defined as in (\ref{eq:ANCP}) and \begin{eqnarray*} \begin{aligned} & \frac{(\sqrt{1 + \sigma_n^2} - t_{\max} \sigma_n)^2}{1 + \sigma_n^2} \\ & \Bigg( \frac{n_i}{r} - \max \left( \frac{4}{3} \log \frac{2r}{\delta} , \sqrt{4 \frac{n_i}{r} \log \frac{2 r}{\delta}} \right) \Bigg) > 2 \sigma_n n_i t_{\max}^2 + \\ & \psi \left( \frac{n_o}{M_1} + \max \left( \frac{4}{3} \log \frac{2 M_1}{\delta} , \sqrt{4 \frac{n_o}{M_1} \log \frac{2 M_1}{\delta}} \right) \right) + \\ & \frac{\sigma_n^2 \psi}{1+\sigma_n^2 } \Bigg( \frac{n_i}{M_1} + \max \left( \frac{4}{3} \log \frac{2 M_1}{\delta} , \sqrt{4 \frac{n_i}{M_1} \log \frac{2 M_1}{\delta}} \right) \Bigg) \end{aligned} \end{eqnarray*} then (\ref{cond:main_cond}) holds with probability at least $1 - 3 \delta$. \end{theorem} In this section, we considered the unstructured outliers whose number can be much larger than $M_1$ and $n_i$. Consider the challenging scenario that the unstructured outliers dominate the data, thus the values of all the singular values are close to each other which indicates that the values of $t_{\min}$ and $t_{\max}$ are close to one. Thus, the sufficient condition of Theorem~\ref{theo:noise} roughly states that $\frac{(1- \sigma_n)^2}{1+ \sigma_n^2} \frac{n_i}{r}$ should be sufficiently larger than $n_i \sigma_n + \frac{n_o}{M_1}$. In practise, the algorithm works better than what the sufficient condition implies because the proof is based on considering the worst case scenarios. \begin{comment} \vspace{-0.1in} \begin{figure}[h!] \begin{center} \mbox{ \includegraphics[width=0.5\textwidth]{figs/phase.eps} } \end{center} \vspace{-0.15in} \caption{The phase transitions in presence of the unstructured outliers versus $n_i$ and $n_o$. White indicates correct subspace recovery and black designates incorrect recovery. In this experiment, the data follows Assumption~\ref{assum_DistUni} with $M_1=50$ and $r=4$. } \label{fig:phase} \end{figure} \end{comment} \subsection{Clustered Outliers} In this section, we consider a different structure for the outliers. It is assumed that the outliers form a cluster outside of the span of the inliers. Structured outliers are mostly associated with important rare events such as malignant tissues~\cite{karrila2011comparison} or web attacks~\cite{kruegel2003anomaly}. The following assumption specifies the presumed model for $\mathbf{B}$. \begin{assumption} Each column of $\mathbf{B}$ is formed as $\mathbf{b}_i = \frac{1}{\sqrt{1 + \eta^2}} ( \mathbf{q} + \eta \mathbf{f}_i)$. The unit $\ell_2$-norm vector $\mathbf{q}$ does not lie in $\mathcal{U}$, $\{\mathbf{f}_i \}_{i=1}^{n_o}$ are drawn uniformly at random from $\mathbb{S}^{M_1 - 1}$, and $\eta$ is a positive number. \label{asm:clus} \end{assumption} \noindent In Assumption~\ref{asm:clus}, the outliers form a cluster around vector $\mathbf{q}$ which does not lie in $\mathcal{U}$ and $\eta$ determines how close they are to each other. The following theorem provides the sufficient conditions to guarantee that Normalized Coherence Values distinguish the cluster of outliers. \begin{theorem} \label{theo:clustered_outliers} Suppose that the distribution of inliers follows Assumption ~\ref{assum_DistUni} and the distribution of outliers follows Assumption~\ref{asm:clus}. If $\mathbf{x}$ is defined as in (\ref{eq:ANCP}) and \begin{eqnarray*} \begin{aligned} & \frac{n_i}{r} - \max \left( \frac{4}{3} \log \frac{2r}{\delta} , \sqrt{4 \frac{n_i}{r} \log \frac{2 r}{\delta}} \right) > n_o \frac{\psi \| \mathbf{q}^T \mathbf{U}^{\perp} \|_2^2}{1 + \eta^2} + \\ & \frac{\psi \eta^2 n_o}{(1 + \eta^2 )M_1} + \frac{\eta^2 \psi}{1 + \eta^2} \max \left( \frac{4}{3} \log \frac{2M_1}{\delta} , \sqrt{\frac{n_o}{M_1} \log \frac{2 M_1}{\delta}} \right) + \\ &\frac{ \eta \sqrt{\psi}}{1+ \eta^2} \| \mathbf{q}^T \mathbf{U}^{\perp} \|_2 \left( \frac{n_o}{\sqrt{M_1}} + 2\sqrt{n_o} + \sqrt{\frac{2 n_o \log \frac{1}{\delta}}{M_1 -1 }} \right) \:, \end{aligned} \end{eqnarray*} then (\ref{cond:main_cond}) is satisfied and $\mathcal{U}$ is recovered exactly with probability at least $1 - 4 \delta$. \end{theorem} In sharp contrast to Theorem~\ref{theo:randomrandom}, $n_o$ is not scaled with $M_1$. This means that when $\eta$ is small (the outliers are close to each other), $n_i/r$ should be sufficiently larger than $n_o$. When $\eta$ goes to infinity, the distribution of outliers converges to the distribution of outliers in Assumption~\ref{assum_DistUni} and one can observe that the sufficient condition in Theorem~\ref{theo:clustered_outliers} converges to the sufficient condition in Theorem~\ref{theo:randomrandom}. \subsection{Outlier Detection in a Union of Subspaces} In practice, the inliers are not randomly distributed in a subspace and they mostly form some structures. In this section, we assume that the inliers are clustered. It is assumed that the columns of $\mathbf{A}$ form $m$ clusters and the data points in each cluster span a $d$-dimensional subspace. The following assumption provides the details. \begin{assumption} The matrix of inliers can be written as $\mathbf{A} = [\mathbf{A}_1 \: ... \: \mathbf{A}_m] \mathbf{T}_A$ where $\mathbf{A}_k \in \mathbb{R}^{M_1 \times {n_i}_k}$, $\sum_{k=1}^{m} {n_i}_k = n_i$, and $\mathbf{T}_A$ is an arbitrary permutation matrix. The columns of $\mathbf{A}_k$ are drawn uniformly at random from the intersection of subspace $\mathcal{U}_k$ and $\mathbb{S}^{M_1-1}$ where $\mathcal{U}_k$ is a $d$-dimensional subspace. In other word, the columns of $\mathbf{A}$ lie in a union of subspaces $\{ \mathcal{U}_k \}_{k=1}^m$ and $\left(\mathcal{U}_1 \oplus \: ... \oplus \mathcal{U}_m \right)= \mathcal{U}$ where $\oplus$ denotes the direct sum operator. \label{asm:union_of_sunb} \end{assumption} \noindent The following theorem provides the sufficient conditions to guarantee that the computed Normalized Coherence Values satisfy (\ref{cond:main_cond}) with high probability. \begin{theorem} \label{theo:Linearly_dependant_inliers} Suppose that the distribution of the inliers follows Assumption~\ref{asm:union_of_sunb} and the distribution of outliers follows Assumption~\ref{assum_DistUni}. If $\mathbf{x}$ is defined as in (\ref{eq:ANCP}) and \begin{eqnarray*} \begin{aligned} \vartheta \mathcal{A} > (\psi - 1)\frac{n_o}{M_1} + 2 \max \left( \frac{4}{3} \log \frac{2M_1}{\delta} , \sqrt{4 \frac{n_o}{M_1} \log \frac{2 M_1}{\delta}} \right) \end{aligned} \end{eqnarray*} where $ \vartheta = \underset{\mathbf{a} \in \mathcal{U} \atop \| \mathbf{a} \| = 1}{\inf} \sum_{k=1}^{m} \| \mathbf{a}^T \mathbf{U}_k \|_2^2$ and $ \mathcal{A} = \min_i \Bigg\{ \frac{ {n_i}_k }{d} - \max \left( \frac{4}{3} \log \frac{2 m d}{\delta} , \sqrt{4 \frac{{n_i}_k}{d} \log \frac{2 m d}{\delta}} \right) \Bigg\}_{i=1}^m, $ then (\ref{cond:main_cond}) is satisfied and $\mathcal{U}$ is recovered exactly with probability at least $1 - 3 \delta$. \end{theorem} Theorem~\ref{theo:Linearly_dependant_inliers} reveals an interesting property of the Normalized Coherence Values. According to the definition of $\mathcal{A}$, $\mathcal{A}$ is roughly equal to ${\min \{{n_i}_k \}_{k=1}^m}/{d}$. Thus, Theorem~\ref{theo:Linearly_dependant_inliers} states that when the inliers are clustered, the population of the cluster with the minimum population is the key factor. This property matches with our intuition about outlier detection because if there is a cluster with few number of data points, we could label them as outliers similar to the outliers modeled in Assumption~\ref{asm:out}. The parameter $\vartheta = \underset{\mathbf{a} \in \mathcal{U} \atop \| \mathbf{a} \| = 1}{\inf} \sum_{k=1}^{m} \| \mathbf{a}^T \mathbf{U}_k \|_2^2$ shows how well the inliers are diffused in $\mathcal{U}$. Clearly, if the inliers are present in all or most of the directions inside $\mathcal{U}$, a robust PCA algorithm is more likely to recover $\mathcal{U}$ correctly. However, the presented methods do not require the inliers to occupy all the directions in $\mathcal{U}$. The reason that $\vartheta$ appeared in the sufficient conditions is that the theorem guarantees the performance in the worst case scenarios. \section{Numerical Experiments} In this section, SNCP and ANCP are compared with the existing robust PCA approaches, including FMS~\cite{lerman2014fast}, GMS~\cite{zhang2014novel}, CoP~\cite{rahmani2017coherence}, iSearch~\cite{rahmani2019outlier2}, and R1-PCA~\cite{lamport21}, and their robustness against different types of outliers is examined with both real and synthetic data. \begin{remark} In the presented theoretical results, it was assumed that $r_d$ is known. When the data is noisy, one can utilize any rank estimation algorithm and the performance of the algorithms is not sensitive to the chosen rank as long as $r_d$ is sufficiently larger than $r$. In the presented experiments, we set $r_d$ equal to the number of singular values of $\mathbf{D}$ which are greater than $s_1/20$ where $s_1$ is the first singular value of $\mathbf{D}$. \end{remark} \subsection{Comparing Different Scores} In this experiment, we simulate a scenario in which the outliers are close to $\mathcal{U}$. Suppose $r=8$, $n_i = 180$, and $n_o = 40$. The outliers are generated as $[\mathbf{U} \:\: \mathbf{H}]\: \mathbf{G}$ where $\mathbf{H} \in \mathbb{R}^{M_1 \times 4}$ spans a random 4-dimensional subspace and the elements of $\mathbf{G} \in \mathbb{R}^{12 \times 20} $ are sampled independently from $\mathcal{N}(0,1)$. Fig.~\ref{fig:compare_scores} demonstrates Innovation Values, Coherence Values, and Normalized Coherence Values computed by ANCP and SNCP (the first 40 columns are outliers). In this figure, we show the inverse of Coherence Values and the inverse of Normalized Coherence Values to make them comparable to Innovation Values. One can observe that the scores computed by iSearch, ANCP, and SNCP can be reliably used to form a basis for $\mathcal{U}$ but the scores computed by CoP do not distinguish the outliers well enough. As it was predicted by Theorem~\ref{theo:Linearly_dependant_CP}, CoP can fail to distinguish the outliers when they are close to $\mathcal{U}$. The main reason is that CoP measures the similarity between the data points via a simple inner-product while iSearch and ANCP utilize the directions of innovation to measure the similarity between a data point and the rest of data. The functionality of SNCP is similar to that of ANCP while it uses a symmetric measure of similarity and the plots show that it distinguishes the outliers in a more clear way. \begin{comment} \vspace{-0.1in} \subsection{Abundance of Unstructured Outliers} \vspace{-0.1in} Theorem~\ref{theo:random} predicted that when the outliers are randomly distributed, the number of outliers can be much larger than the number of inliers provided that $n_i/r$ is sufficiently large. Suppose the data follows Assumption~\ref{assum_DistUni} with $M_1=50$ and $r=4$. Define $\hat{\mathbf{U}}$ as an orthonormal basis for the recovered subspace. A trial is considered successful if $ \frac{\| (\mathbf{I} - \mathbf{U} \mathbf{U}^T) \hat{\mathbf{U}} \|_F}{ \|\mathbf{U} \|_F } < 10^{-3} \: . $ Fig.~\ref{fig:phase} shows the phase transitions in which white means correct subspace recover and black designates incorrect recovery (the number of evaluation runs was 20). The phase transitions indicate that when $n_i/r$ is larger than 5, the algorithms can successfully recover $\mathcal{U}$ even if $n_o = 2000$. In addition, SNCP shows more robustness against the outliers when $n_i$ is small. \end{comment} \begin{figure}[t] \begin{center} \mbox{ \includegraphics[width=1.78in]{figs/vs1.eps}\hspace{-0.1in} \includegraphics[width=1.78in]{figs/vs2.eps}} \mbox{ \includegraphics[width=1.78in]{figs/vs3.eps}\hspace{-0.1in} \includegraphics[width=1.78in]{figs/vs4.eps} } \end{center} \vspace{-0.15in} \caption{The plots show the inverse of Normalized Coherence Values computed by SNCP and ANCP, the Innovation Values, and the inverse of Coherence Values. In this experiment, the first 40 columns are the outliers. } \label{fig:compare_scores} \end{figure} \subsection{Noisy Data} In this section, we examine the robustness of the robust PCA methods against noise. Suppose $\mathbf{B}$ follows Assumption~\ref{asm:out}, $r=5$, $r_o=10$, $M_1 = 200$, $n_i=100$, and $n_o=100$ where $\mathcal{U}_o$ is a random 10-dimensional subspace. We consider two models for the distribution of the inliers. The first model is random distribution on $\mathcal{U} \cap \mathbb{S}^{N-1}$ as described in Assumption~\ref{assum_DistUni}. In the second model, it is assumed that the inliers form a cluster in $\mathcal{U}$. The following assumption describes the second model. \begin{assumption} Each column of matrix $\mathbf{A}$ is formed as $\mathbf{a}_i = \frac{ \mathbf{U} \mathbf{s}_i}{\| \mathbf{U} \mathbf{s}_i \|_2} $ where $\mathbf{s}_i = \mathbf{w} + \gamma \mathbf{z}_i$, $\mathbf{w} \in \mathbb{R}^{r }$ is a unit $\ell_2$-norm vector, and $\{\mathbf{z}_i \}_{i=1}^{n_i}$ are sampled randomly from $\mathbb{S}^{r - 1}$. \label{asm:inliers_clus} \end{assumption} \noindent Since the data is noisy, exact subspace recovery is not feasible. Instead, we examine the probability that an algorithm distinguishes all the outliers correctly. Define vector $\mathbf{f} \in \mathbb{R}^{M_2 }$ such that $\mathbf{f}(k) = \| (\mathbf{I} - \hat{\mathbf{U}} \hat{\mathbf{U}}^T) \mathbf{d}_k \|_2$ where $\hat{\mathbf{U}}$ is the identified subspace. A trial is considered successful if \begin{eqnarray} \label{eq:out_conddi} \max \bigg( \{\mathbf{f}(k) \: \: : \: \: k > n_o \} \bigg) < \min \bigg( \{\mathbf{f}(k) \: \: : \: \: k \le n_o \} \bigg), \end{eqnarray} which means that the norm of projection of all the inliers on $\hat{\mathcal{U}}^{\perp}$ should be smaller than the corresponding values for the outliers. Define $\text{SNR} = \frac{\| \mathbf{A} \|_F^2}{\| \mathbf{E} \|_F^2 }$ where $\mathbf{E}$ is the noise component which is added to the inliers. Fig.~\ref{fig:lin_dep_out} shows the probability that (\ref{eq:out_conddi}) is valid versus SNR (the number of evaluation runs was 200). In the left plot, the distribution of inliers follows Assumption~\ref{assum_DistUni} and in the right plot it follows Assumption~\ref{asm:inliers_clus} with $\gamma = 0.2$. One can observe that SNCP outperforms most of the existing methods on both cases and the performance of iSearch and ANCP are close. In addition, by comparing the two plots, it can be observed that the performance of some of the robust PCA methods is sensitive to the distribution of the inliers. For instance, FMS outperforms most of the other methods when the inliers are randomly distributed but its performance degrades significantly when the inliers form a cluster in $\mathcal{U}$. \begin{figure}[h!] \begin{center} \mbox{ \includegraphics[width=1.8in]{figs/1000.eps}\hspace{-0.1in} \includegraphics[width=1.8in]{figs/210n.eps}} \end{center} \vspace{-0.15in} \caption{The outliers are linearly dependant and they lie in a 10-dimensional subspace. In the left plot, the inliers are randomly distributed in $\mathcal{U}$ (Assumption~\ref{assum_DistUni}) and in the right plot, the inliers form a cluster (Assumption~\ref{asm:inliers_clus} with $ \gamma = 0.2)$. } \label{fig:lin_dep_out} \end{figure} \subsection{ Identifying the Permuted Columns } The problem of regression with unknown permutation is similar to the conventional regression but the correspondence between input variables and labels is missing or erroneous. Suppose $\mathbf{X} \in \mathbb{R}^{d \times n}$ is the measurement matrix where $n$ is the number of measurements. Define $\mathbf{Y} \in \mathbb{R}^{m \times n}$ as the observation matrix which can be written as $\mathbf{Y} = \Theta \mathbf{X}$ where $\Theta \in \mathbb{R}^{m \times d}$ is the unknown matrix which is estimated by the regression algorithm. In the regression problem with unknown permutation, the observation matrix $\mathbf{Y}$ is affected by an unknown permutation matrix $\Pi$, i.e., matrix $\mathbf{Y}$ can be written as $\mathbf{Y} = \Theta \mathbf{X} \Pi$ where $\Pi \in \mathbb{R}^{n \times n}$. In this problem, it is assumed that $\Pi$ does not displace all the columns of $\Theta \mathbf{X}$ and only an unknown fraction of the columns are displaced. The authors of~\cite{slawski2019sparse} showed that this special regression problem can be translated into a robust PCA problem. Define matrix $\mathbf{Z} \in \mathbb{R}^{(d+m) \times n}$ as $\mathbf{Z} = ([\mathbf{X}^T \:\: \mathbf{Y}^T])^T$, i.e., each column of $\mathbf{Z}$ is equal to the concatenation of the corresponding columns of $\mathbf{X}$ and $\mathbf{Y}$. Suppose $n > d$ and assume that the rank of $\mathbf{X}$ is equal to $d$. If $\Pi$ is equal to the identity matrix, the rank of $\mathbf{Z}$ is equal to $d$. In contrast, when the columns of $\mathbf{Y}$ are displaced, the corresponding columns of $\mathbf{Z}$ do not lie in the $d$-dimensional subspace which the other columns of $\mathbf{Z}$ lie in. Therefore, the columns of $\mathbf{Z}$ which are corresponding to the displaced columns can be considered as outliers and a robust PCA method can be utilized to locate them. Once they are located and removed, the regression problem can be solved using the remaining measurements. In this experiment, the elements of $\mathbf{X}$ and $\Theta$ are sampled from $\mathcal{N} (0,1)$, $d = 10$, and $m = 10$. Define $n_i$ as the number of columns of $\mathbf{Y}$ which are not affected by the permutation matrix and define $n_o$ as the number of displaced columns. The robust PCA methods are applied to $\mathbf{Z}$ to find a basis for the $d$-dimensional subspace which is spanned by the columns of $\mathbf{Z}$ corresponding to the inliers. If this subspace is estimated accurately, all the displaced columns can be exactly located~\cite{slawski2019sparse}. Define Log-Recovery Error as $ \log_{10} \left( \frac{\| (\mathbf{I} - \mathbf{U} \mathbf{U}^T) \hat{\mathbf{U}} \|_F}{ \|\mathbf{U} \|_F } \right) $, where $\hat{\mathbf{U}}$ is an orthonormal basis for the recovered subspace. Fig.~\ref{fig:lin_dep_out2} shows Log-Recovery Error versus $n_o$ where $n_i$ is fixed equal to 200 (the number of evaluation runs was 400). This is a challenging subspace recovery task because the outliers can be close to the span of inliers and this is the main reason that CoP did not perform well. One can observe that SNCP and FMS yielded the best performance. \begin{figure}[h!] \begin{center} \mbox{ \includegraphics[width=2.1in]{figs/reg.eps} } \end{center} \vspace{-0.15in} \caption{This plot shows subspace recovery error versus the number of displaced measurements. The number of measurements which are not displaced are equal to $n_i=200$ and the total number of measurements are equal to $n_i + n_o$.} \label{fig:lin_dep_out2} \end{figure} \subsection{Unstructured Outliers} Theorem~\ref{theo:random} predicted that when the outliers are randomly distributed, the number of outliers can be much larger than the number of inliers provided that $n_i/r$ is sufficiently large. Suppose the data follows Assumption~\ref{assum_DistUni} with $M_1=50$ and $r=4$. Define $\hat{\mathbf{U}}$ as an orthonormal basis for the recovered subspace. A trial is considered successful if $ \frac{\| (\mathbf{I} - \mathbf{U} \mathbf{U}^T) \hat{\mathbf{U}} \|_F}{ \|\mathbf{U} \|_F } < 10^{-3} \: . $ Fig.~\ref{fig:phase} shows the phase transitions in which white means correct subspace recover and black designates incorrect recovery (the number of evaluation runs was 20). The phase transitions indicate that when $n_i/r$ is larger than 5, the algorithms can successfully recover $\mathcal{U}$ even if $n_o = 2000$. In addition, SNCP shows more robustness against the outliers when $n_i$ is small. \begin{figure}[h!] \begin{center} \mbox{ \includegraphics[width=0.5\textwidth]{figs/phase.eps} } \end{center} \vspace{-0.15in} \caption{ The phase transitions in presence of the unstructured outliers versus $n_i$ and $n_o$. White indicates correct subspace recovery and black designates incorrect recovery. In this experiment, the data follows Assumption~\ref{assum_DistUni} with $M_1=50$ and $r=4$. } \label{fig:phase} \end{figure} \subsection{Structured Outlier Detection in Real Data} The authors of~\cite{gitlin2018improving,rahmani2017coherence} proposed to use robust PCA to improve the accuracy of the clustering algorithms. The robust PCA method is applied to each identified cluster to find the miss-classified data points as outliers. We refer the reader to~\cite{gitlin2018improving,rahmani2017coherence} for further details. Similar to corresponding experiment in~\cite{rahmani2017coherence}, we use Hopkins155 dataset which makes the outlier detection problem challenging. In this dataset, the data points are linearly dependent and the clusters are close to each other. Therefore, the outliers are structured and they are close to the inliers. The clustering error of the clustering algorithm is $30 \%$ and we compute the final clustering error after applying the robust PCA methods and updating the clusters. Table~\ref{tab:accuracy} shows the clustering error after applying different robust PCA methods. One can observe that iSearch, ANCP, SNCP, and CoP yielded better performance and the main reason is that they leverage the clustering structure of the inliers and they are robust against structured outliers. \begin{comment} \vspace{-0.1in} \subsection{Running Time} \vspace{-0.1in} In this section, we study the running time of the robust PCA methods. For ANCP, SNCP, CoP, and iSearch, we used 50 data points to build the basis matrix (matrix $\mathbf{Y}$). Table \ref{tab:runnig_M2} shows the running times versus $M_2$ while $M_1 = 200$. Table~\ref{tab:runnig_M1} shows the running times versus $M_1$ while $M_2 = 1500$. In all the runs, $r = 5$ and $n_i = 200$. One can observe that CoP, SNCP, and ANCP are notably fast since they are single step algorithms. The running time of CoP and SNCP is longer than ANCP when $M_2$ is large because their computation complexity scale with $M_2^2$. GMS is also a fast algorithm when $M_1$ is small but its running time can be long when $M_1$ is large because its computation complexity scale with $M_1^3$. \end{comment} \begin{table}[h!] \centering \caption{Clustering error after using the robust PCA methods to detect the misclassified data points. } \begin{tabular}{| c | c | c | c |c|c|c|c|} \hline CoP & FMS & R1-PCA & GMS & iSearch & PCA & ANCP & SNCP \\ \hline 6.93 & 28.5 & 22.56 & 17.25 & 3.72 & 12.01 & 6.64 & 3.65 \\ \hline \end{tabular} \label{tab:accuracy} \end{table} \begin{figure*}[t] \begin{center} \mbox{ \includegraphics[width=1.22in]{figs/Ancp.eps}\hspace{-0.12in} \includegraphics[width=1.22in]{figs/sncp.eps}\hspace{-0.12in} \includegraphics[width=1.22in]{figs/svd.eps}\hspace{-0.12in} \includegraphics[width=1.22in]{figs/r1.eps}\hspace{-0.12in} \includegraphics[width=1.22in]{figs/isearch.eps}\hspace{-0.12in} \includegraphics[width=1.22in]{figs/FMS.eps} } \end{center} \vspace{-0.15in} \caption{\textcolor{black}{This figure shows the residual} values computed by different methods. Each element represents a frame of the video file and frames 65 to 80 are the outlying frames. } \label{fig:activity} \end{figure*} \subsection{Event Detection in Video} In this experiment, we utilize the robust PCA methods to identify an activity in a video files, i.e., the outlier detection methods identify the frames which contain the activity as the outlying frames/data-points. We use the Waving Tree video file~\cite{li2004statistical} where in this video a tree is smoothly waving and in the middle of the video, a person crosses the frame. The frames which only contain the background (the tree and the environment) are inliers and the few frames corresponding to the event (the presence of the person) are the outliers. The tree is smoothly waving and we use $r=3$ as the rank of inliers for all the methods. We use 100 frames where frames 65 to 80 are the outlying frames. In this interval (65 to 80), the person enters the frame from left, stay in the middle, and leaves from right. In this experiment, we vectorize each frame and form data matrix $\mathbf{D}$ by the vectorized frames. In addition, we reduce the dimensionality of $\mathbf{D}$ by projecting each column of $\mathbf{D}$ into the span of the first 50 left singular vectors. Thus, $\mathbf{D} \in \mathbb{R}^{50 \times 100}$. Define $\hat{\mathbf{U}}$ as the estimated subspace and define the residual value corresponding to data point $\mathbf{d}_i$ as ${\| \mathbf{d}_i - \hat{\mathbf{U}}\hat{\mathbf{U}}^T \mathbf{d}_i\|_2}$. The outliers are detected as the data points with the larger residual values. Fig.~\ref{fig:activity} shows the residual values computed by different methods. An important observation is that FMS, PCA (SVD), and R1-PCA clearly distinguished the first and the last outlying frames but they hardly distinguish the middle outliers (frames 69 to 74). The main reason is that in these frames the person does not move which means that these outlying frames are very similar to each other. Fig.~\ref{fig:activity} shows that the ANCP, SNCP, and iSearch successfully distinguish all the outlying frames since they are robust to structured and linearly dependant outliers. \subsection{Running Time} In this section, we study the running time of the robust PCA methods. For ANCP, SNCP, CoP, and iSearch, we used 50 data points to build the basis matrix (matrix $\mathbf{Y}$). Table \ref{tab:runnig_M2} shows the running times versus $M_2$ while $M_1 = 200$. Table~\ref{tab:runnig_M1} shows the running times versus $M_1$ while $M_2 = 1500$. In all the runs, $r = 5$ and $n_i = 200$. One can observe that CoP, SNCP, and ANCP are notably fast since they are single step algorithms. The running time of CoP and SNCP is longer than ANCP when $M_2$ is large because their computation complexity scale with $M_2^2$. GMS is also a fast algorithm when $M_1$ is small but its running time can be long when $M_1$ is large because its computation complexity scale with $M_1^3$. \begin{table}[h!] \centering \caption{Running time of the algorithms versus $M_2$ ($M_1 = 200$).} \begin{tabular}{| c | c | c | c |c|c|c|} \hline $M_2 $& SNCP & ANCP & iSearch & CoP & FMS & GMS \\ \hline 500 & 0.0228 & 0.0120 & 0.2660 & 0.016 & 0.1130 & 0.0872\\ \hline 1000 & 0.0427 & 0.0160 & 0.8983 & 0.0325 & 0.2440 & 0.1265\\ \hline 5000 & 0.2622 & 0.0428 & 19.4080 & 0.3930 & 0.6635 & 0.2926\\ \hline \end{tabular} \label{tab:runnig_M2} \end{table} \begin{table}[h!] \centering \caption{Running time of the algorithms versus $M_1$ ($M_2 = 1500$).} \begin{tabular}{| c | c | c | c |c|c|c|} \hline $M_1 $& SNCP & ANCP & iSearch & CoP & FMS & GMS \\ \hline 200 & 0.0614 & 0.0187 & 1.7279 & 0.0576 & 0.2978 & 0.1458 \\ \hline 500 & 0.1456 & 0.0727 & 2.1261 & 0.0710 & 0.9574 & 0.6399\\ \hline 1000 & 0.3145 & 0.2527 & 2.7695 & 0.0900 & 2.7731 & 2.5590 \\ \hline \end{tabular} \label{tab:runnig_M1} \end{table} \section{Conclusion} It was shown that Innovation Value under the quadratic cost function is equivalent to Leverage Score. Two closed-form robust PCA methods were presented where the first one was based on Leverage Score and the second one was inspired by the connection between Leverage Score and Innovation Value. Several theoretical performance guarantees for the robust PCA method under different models for the distribution of the outliers and the distribution of the inliers were presented. In addition, it was shown with both theoretical and numerical investigations that the algorithms are robust to the strong presence of noise. Although the presented methods are fast closed-form algorithms, it was shown that they often outperform most of the existing methods. \vspace{0.4in}
{ "timestamp": "2021-06-24T02:12:16", "yymm": "2106", "arxiv_id": "2106.12190", "language": "en", "url": "https://arxiv.org/abs/2106.12190", "abstract": "The idea of Innovation Search, which was initially proposed for data clustering, was recently used for outlier detection. In the application of Innovation Search for outlier detection, the directions of innovation were utilized to measure the innovation of the data points. We study the Innovation Values computed by the Innovation Search algorithm under a quadratic cost function and it is proved that Innovation Values with the new cost function are equivalent to Leverage Scores. This interesting connection is utilized to establish several theoretical guarantees for a Leverage Score based robust PCA method and to design a new robust PCA method. The theoretical results include performance guarantees with different models for the distribution of outliers and the distribution of inliers. In addition, we demonstrate the robustness of the algorithms against the presence of noise. The numerical and theoretical studies indicate that while the presented approach is fast and closed-form, it can outperform most of the existing algorithms.", "subjects": "Machine Learning (stat.ML); Machine Learning (cs.LG)", "title": "Closed-Form, Provable, and Robust PCA via Leverage Statistics and Innovation Search", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES\n\n", "lm_q1_score": 0.9857180685922242, "lm_q2_score": 0.8128673133042217, "lm_q1q2_score": 0.8012579980919878 }
https://arxiv.org/abs/2106.03070
Linear Rescaling to Accurately Interpret Logarithms
The standard approximation of a natural logarithm in statistical analysis interprets a linear change of \(p\) in \(\ln(X)\) as a \((1+p)\) proportional change in \(X\), which is only accurate for small values of \(p\). I suggest base-\((1+p)\) logarithms, where \(p\) is chosen ahead of time. A one-unit change in \(\log_{1+p}(X)\) is exactly equivalent to a \((1+p)\) proportional change in \(X\). This avoids an approximation applied too broadly, makes exact interpretation easier and less error-prone, improves approximation quality when approximations are used, makes the change of interest a one-log-unit change like other regression variables, and reduces error from the use of \(\log(1+X)\).
\section{The Traditional Interpretation of Logarithms}\label{the-traditional-interpretation-of-logarithms}} It is common practice in many statistical applications, especially in regression analysis, to transform variables using the natural logarithm \(\ln(X)\). This can be done for statistical reasons, for example to fit an apparent functional form in the data or to reduce skew and the impact of positive outliers in the variable \(X\). The logarithm transformation is also used for theoretical reasons, when theory dictates the model relates to proportional changes in \(X\) rather than linear changes. The standard interpretation of a log-transformed variable in a regression is that a linear increase of \(p\) in \(\ln(X)\) is equivalent to a \(p\times 100\%\) increase in \(X\). This is not literally true. A linear increase of \(p\) in \(\ln(X)\) is equivalent to an \((e^p - 1)\times 100\%\) increase in \(X\). The standard interpretation relies on the approximation \(e^p \approx 1+p\), or equivalently \(p \approx \ln(1+p)\), which is fairly accurate for small values of \(p\). \begin{equation} \ln(X) + p \approx \ln(X) + \ln(1+p) = \ln(X(1+p)) \end{equation} In this paper, I provide an very simple alternative approach to using and interpreting logarithms in the context of regression analysis, which solves three major problems with the standard approach. The first major problem with the standard approach is that the approximation loses quality relatively quickly as \(p\) grows. The error in approximation is equal to \[ 1+p - e^p \] which is always negative, and grows more negative with \(p\), such that this approximation always understates the proportional increase in \(X\) equivalent to a given linear increase in \(\ln(X)\). If \(\ln(X)\) is a treatment variable, the approximation will always overstate its effect. The quality of approximation is, subjectively, acceptable for small values of \(p\), but the error becomes large within ranges of interest. For linear increases in \(\ln(X)\) of .1, .2, or .3, respectively, interpretations of these changes as 10\%, 20\%, or 30\% increases in \(X\) would understate the actual change in \(X\) by about .5, 2.1, and 5 percentage points, respectively.\footnote{The approximation error expression also demonstrates that \(e\) does not hold any special property in regards to reducing approximation error, and is actually a poor choice of logarithmic base if the traditional approximation is to be used. Many other bases, like 2.6, produce nearly identical errors for small \(p\) and then dominate \(e\) afterwards. Base 2.35 is attractive in other ways: errors for base 2.35 are no larger in absolute value than .014 all the way up to \(p = .43\), and considerably improve on base \(e\) above that, although unfortunately performance relative to base-\(e\) is worst around \(p=.1\). If a researcher insists on the traditional approximation, I at least recommend the use of base-2.6 logarithm, which is a clear improvement on \(e\), or perhaps base-2.35 for something less sensitive to \(p\).} This leads naturally to the second problem with the standard interpretation, which is sociological in nature. The fact that the base-\(e\) approximation breaks down quickly as \(p\) increases does not appear to be universally known, nor is there a standard maximum \(p\) for which the approximation is considered acceptable. It is not uncommon to see papers describing 10\% or 20\% changes in a log-transformed variable, and it is doubtful that these authors would willingly inject biases of .5 or 2.1 percentage points into their analysis for no reason. Calculations that give exact interpretations using \(e^p\) or \(\ln(1+p)\) are available but are not universally applied even for larger percentage changes. Speculatively, this may be because the author assumes that approximation error is too small to bother, because the additional calculation could confuse a reader, or because in some fields it is not expected. It may even be the case that the researcher is not aware that the traditional interpretation \emph{is} an approximation. It is common in published studies and in econometric teaching materials to see the traditional \(\ln(X)+p \approx \ln(X(1+p))\) approximation discussed without reference to its approximate nature, implying by omission that it is an equality. At the undergraduate level, see for example Bailey (2017) Section 7.2. At the graduate level, see Greene (2008) Section 4.7, specifically Example 4.3. The third problem with the standard interpretation, when applied to variables on the right-hand side of a regression, is that it requires nonstandard interpretation of regression coefficients. Nearly all regression coefficients are understood in terms of one-unit changes in the associated predictor. Log-transformed variables are an exception to this. A one-unit change in \(\ln(X)\) describes what would in most cases be an unrealistically large increase in \(X\), and also would produce a large 71.8 percentage point error if the traditional approximation were applied. All three of these problems can be solved by simply changing the base of the logarithm.\footnote{After a literature search, I was unable to find previous studies making this same recommendation. However, given the long history of logarithms in regression, it seems unlikely that nobody has thought of the insight in this paper before. So I will not claim that this method is novel, but just that it is currently not widely known or applied.} Selecting a percentage increase \(p\times 100\%\) ahead of time and using \(\log_{1+p}(X)\) in place of \(\ln(X)\) means that a one-unit change in \(\log_{1+p}(X)\) is exactly equivalent to a \(p\times 100\%\) increase in \(X\). There is no need for approximation, and the exact interpretation can be written directly into a regression table, solving the first problem and much of the second. The use of base \(1+p\) can be restated as \(\ln(X)/\ln(1+p)\), framing the method in the easily-understood terms of linearly rescaling the variable by the constant \(1/\ln(1+p)\). The third problem is also solved because the relevant increase in \(\log_{1+p}(X)\) is 1, in line with other variables in the regression. Based on these results, I recommend the use of alternate logarithmic bases, or the ``linear rescaling'' approach, when using logarithms in statistical analysis, especially in regression. The benefits are most apparent when applied to variables on the right-hand side of a regression (the predictors/independent variables), but there are also benefits on the left-hand side (the outcome). Additionally, the use of linear rescaling helps ease some problems related to the use of \(\ln(1+X)\) with \(X\) variables that contain values of zero. \hypertarget{linearly-rescaling-logarithms}{% \section{Linearly Rescaling Logarithms}\label{linearly-rescaling-logarithms}} \hypertarget{exact-interpretation-of-logarithms}{% \subsection{Exact Interpretation of Logarithms}\label{exact-interpretation-of-logarithms}} \label{sec:exact} For any base \(b\), a linear increase of \(p\) in a logarithm \(\log_b(X)\) is equivalent to a proportional change in \(X\) of \(b^p\), or a percentage increase of \((b^p-1)\times 100\%\). Researchers could report exact interpretation of linear logarithmic increases using this formula. However, this practice is far from universal. Another approach to exact interpretation of linear logarithmic increases is to take advantage of the following feature of \(b^p-1\): \[ (1+p)^1 - 1 = p \] That is, a linear increase of \(1\) in \(\log_{1+p}(X)\) is exactly equivalent to a percentage increase of \(p\times100\%\) (or a proportional increase of \(1+p\)) in \(X\). This means that a researcher can select ahead of time the percentage increase in \(X\) that they are interested in, for example 10\% (although any other percentage would work as well), and then use \(\log_{1.1}(X)\) in their analysis instead of \(\ln(X)\). Then, a one-unit increase in \(\log_{1.1}(X)\) can be exactly interpreted as a 10\% increase in \(X\). Further, because of the change of base formula, \(\log_{1.1}(X)\) can be calculated as \(\ln(X)/\ln(1.1)\). \(\ln(1.1)\) is a constant, and so the researcher can achieve exact interpretation of the change by scaling the variable they're already using (\(\ln(X)\)) by a constant. Researchers using any estimation method should already be aware of the implications of scaling by a constant in that method, so the linear rescaling approach to changing the base should be understandable by both researchers and readers of research. The use of the change-of-base formula also allows another clear demonstration of what the choice of logarithm base does for interpretations of proportional change: \[ \frac{\ln(X(1+p))}{\ln(1+p)} = \frac{\ln(X) + \ln(1+p)}{\ln(1+p)} = \frac{\ln(X)}{\ln(1+p)}+1 \] There are several benefits to linear rescaling: \begin{itemize} \item It produces exact interpretations of linear increases in logarithms. \item It is, arguably, conceptually more simple than using \(b^p-1\) or \(log_b(1+p)\) to adjust the result after estimation, and avoids the introduction of another calculation where error may occur. \item The change of interest is one log unit, which is how most other variables are understood. \item The exact interpretation can be written directly onto a regression table rather than relying on supplemental calculations, as will be shown in the following sections. \end{itemize} The main downside of this approach is that it requires the choice of the percentage increase beforehand. In practice, this is unlikely to be a major hurdle, as researchers often report only a single percentage increase value anyway, typically a preselected value like 10\%, based on what a reasonable observable change in \(X\) would be. Additionally, if selecting a percentage increase beforehand is not realistic, or if multiple percentage increases are desired, exact interpretation is still available, as in the traditional method, using \((1+b)^{p}-1\). Linear rescaling also improves the process of adjusting the estimate, at least on the right-hand side of a regression (see Section \ref{sec:rhs}). Further, if an approximation is used instead of an exact interpretation, linear rescaling often produces better approximations than the traditional approximation. For example, if \(\log_{1.1}(X)\) has been used and the researcher wants to approximate a 20\% increase in \(X\) using a two-unit increase in \(\log_{1.1}(X)\), they will actually see the effects of a \(1.1^2=1.21\), or 21\%, increase, rather than 20\%. This is an error of one percentage point, compared to the traditional approach, which produces an error of 2.1 percentage points for a .2 increase. Figure \ref{fig:approxquality} examines the error in approximating different percentage increases with a traditional approximation, using a base-\(e\) logarithm where a \(p\)-unit increase in \(\ln(X)\) is taken to be a \(p\times 100\%\) increase in \(X\). I contrast the traditional approximations with approximations from the linear rescaling approach using two different bases: a base-\(1.1\) logarithm, where a \(p\)-unit increase in \(\log_{1.1}(X)\) is taken to be a \(p\times 10\%\) increase in \(X\), and a base-\(1.4\) logarithm, where a \(p\)-unit increase in \(\log_{1.4}(X)\) is taken to be a \(p\times 40\%\) increase in \(X\). \begin{figure} \centering \includegraphics[width=.75\textwidth]{approxquality-1.pdf} \caption{\label{fig:approxquality} Approximation Error with Traditional and Linear Rescaling Methods} \end{figure} The traditional base-\(e\) method slightly outperforms the base-1.1 linear rescaling method, by a miniscule degree, up to a linear change of .048 (although the linear rescaling method could outperform the traditional method for any linear change by selecting a different base than the ones shown in the graph). After .048, approximation with base-1.1 linear rescaling dominates the traditional approximation, especially near \(p=.1\). Both the traditional and base-1.1 approximations considerably outperform base-1.4 for small \(p\), but this is to be expected - base-1.4 is to be used when the change of interest is 40\%. In the region of \(p=.4\), the base-1.4 logarithm considerably outperforms the other two, and approximation errors with base-1.4 are relatively small for the entire graphed range up to \(p=.5\). While the linear rescaling method allows for easy access to exact interpretation, even in cases where approximation is used, the linear rescaling approximation error will be smaller than the approximation error for the traditional method as long as the linear rescaling base is near enough to the proportional change of interest. \hypertarget{linear-rescaling-on-the-right-hand-side}{% \subsection{Linear Rescaling on the Right Hand Side}\label{linear-rescaling-on-the-right-hand-side}}\label{sec:rhs} The benefits of linear rescaling are clearest when applied to a logarithmic transformation on the right-hand side of a regression. Considering the model: \[ Y = \beta_0 + \beta_1\ln(X) + \varepsilon \] the interpretation of \(\hat{\beta}_1\) is often given in a format similar to ``a 10\% increase in \(X\) is associated with a \(.1\times\hat{\beta}_1\) increase in \(Y\).'' Or for an exact interpretation, ``a 10\% increase in \(X\) is associated with a \(\ln(1.1)\times\hat{\beta}_1 = .0953\times\hat{\beta}_1\) increase in \(Y\).'' Under linear rescaling for a 10\% increase in \(X\), instead the model is: \[ Y = \beta_0 + \beta_1\frac{\ln(X)}{\ln(1.1)} + \varepsilon \] in which the interpretation of \(\hat{\beta}_1\) is the simpler ``a 10\% increase in \(X\) is associated with a \(\hat{\beta}_1\) increase in \(Y\).'' The interpretation is simple enough that it can be written directly into a regression table, as in Table \ref{tab:regtable} Column 1, rather than relying on additional in-text calculations. The row heading in the table itself ``\(X\) (10\% Increase)'' is able to convey that a 10\% change in \(X\) is associated with a \(\hat{\beta}_1\) change in \(Y\), and the table note provides more detail. The label ``\(\log_{1.1}(X)\) (10\% increase)'' may be preferred. If the researcher is interested in multiple percentage changes, adjustment to get exact interpretation for each percentage change is straightforward under linear rescaling, because of the change-of-base formula. If \(\log_{1.1}(X)\) has been used, but the researcher wants an exact interpretation for the linear change equivalent to a 20\% increase in \(X\), the researcher can multiply the coefficient \(\hat{\beta}_1\) by \(\ln(1.2)/\ln(1.1)\). This will work for any regression method where scaling a predictor by \(c\) has the result of scaling its coefficient by \(1/c\). \begin{table} \caption{\label{tab:regtable}Example Regression Table with Linear Rescaling} \centering \begin{tabular}[t]{lccc} \toprule & Y & Y (10\% Increase) & Y (10\% Increase) \\ \midrule X (10\% Increase) & 0.194*** & & 0.460***\\ & (0.003) & & (0.007)\\ X & & 2.001*** & \\ & & (0.038) & \\ \midrule Num.Obs. & 1000 & 1000 & 1000\\ \bottomrule \multicolumn{4}{l}{\rule{0pt}{1em}* p $<$ 0.1, ** p $<$ 0.05, *** p $<$ 0.01}\\ \multicolumn{4}{l}{\textsuperscript{} Variables marked with 10\% increase use a base-1.1 logarithm}\\ \multicolumn{4}{l}{transformation. Data is simulated.}\\ \end{tabular} \end{table} \hypertarget{linear-rescaling-on-the-left-hand-side}{% \subsection{Linear Rescaling on the Left Hand Side}\label{linear-rescaling-on-the-left-hand-side}} \label{sec:lhs} The benefits of linear rescaling are less clear on the left-hand side of the regression, since the proportional change of interest cannot be exactly chosen ahead of time. Still, there are benefits. Consider Table \ref{tab:regtable} Column 2, which uses the model \[ \frac{\ln(Y)}{\ln(1.1)} = \beta_0 + \beta_1X + \varepsilon \] The regression estimate is \(\hat{\beta}_1 = 2.001\). By itself, this does not easily lead to exact interpretation of the coefficient. At this point, the researcher can approximate the effect of \(2.001\) as a \(2.001\times 10\% \approx 20\%\) increase. As in Section \ref{sec:exact}, this will lead to less approximation error than in the traditional method provided that \(\hat{\beta}_1\) is not too far from \(1\). There is also the option to provide an exact interpretation, where a one-unit increase in \(X\) is associated with a \(b^{\hat{\beta}_1}\) proporional change in \(Y\). This is not much different from the process for getting exact interpretation using the traditional approach, although it may be somewhat easier to understand if \(p\) is a more natural object to think about than \(e\). A third option is to rerun the model with a different logarithmic base such that the \(\hat{\beta}_1\) will be near \(1\). Then, \(\hat{\beta}_1\) can be very accurately interpreted as a proportional increase of \(b\) in \(Y\). However, this is both laborious and would result in the coefficient of interest being oddly located in the logarithmic base. One note about left-hand side use is that the traditional approximation is known to perform particularly poorly, and exact interpretation is especially important, when the logarithm is on the left-hand side and a predictor of interest is binary (Halvorsen \& Palmquist, 1980). In theoretical terms this is because the derivative does not exist. In practical terms this is because, if the coefficient on the binary variable is large, the researcher cannot naturally select a linear change small enough for the approximation to perform well. Easy access to exact interpretation, and improved approximation when used, are especially important in this case. \hypertarget{linear-rescaling-on-both-sides}{% \subsection{Linear Rescaling on Both Sides}\label{linear-rescaling-on-both-sides}} \label{sec:bothsides} In the case of the log-log model \[ \ln(Y) = \beta_0 + \beta_1\ln(X) + \varepsilon \] linear rescaling on both the left and right-hand sides using the same bases will have no effect on \(\hat{\beta}_1\) or on its interpretation, and offers no major improvement over the traditional method, unless there is a reason to want different bases on the left and right-hand sides, or if there is interest in interpreting the coefficients on control variables with a linearly-rescaled left-hand side. There are some minor expositional benefits. Linear rescaling could be used here for consistency with other models that are not log-log. Rescaling can also make clear to an audience unfamiliar with log-log models how they can be interpreted. For example, it's not uncommon in a log-log model to still report a result like ``a 10\% increase in \(X\) is associated with a \(\hat{\beta}_1\times .1\)\% change in \(Y\).'' Herz \& Mejer (2016) is just one example of this. If the author wants the reader to think in terms of a percentage increase of a particular size in this way, linear rescaling can make that interpretation explicit on the regression table, as in Column 3 of Table \ref{tab:regtable}. \hypertarget{linear-rescaling-and-zeroes}{% \subsection{Linear Rescaling and Zeroes}\label{linear-rescaling-and-zeroes}} Researchers often want to apply a logarithmic transformation to a variable that can take values of zero. There are two common approaches to this: \(\ln(1+X)\), and the asymptotic hyperbolic sine transformation \(asinh(X) = \ln(X+\sqrt{X^2+1})\). Exact calculations for elasticity interpretations using the \(asinh(X)\) transformation are described in Bellemare \& Wichman (2020). Both \(\ln(1+X)\) and \(asinh(X)\) reduce skew and accept values of zero. However, if the researcher wishes to maintain a proportional-change interpretation (at values other than zero), there are problems with any sort of ad-hoc transformation like this, including a sensitivity to the scale of \(X\) and the fact that the zero-censored variable is treated as uncensored. For a left-hand side variable, poisson regression or a censoring model are likely to be superior to an ad-hoc transformation. However, the use of an ad-hoc transformation is still a concern on the right-hand side or for researchers who want to use standard linear regression for other reasons. In this section, I will assume that the researcher's goal is interpret a linear change in \(\ln(1+X)\) in terms of a proportional change in \(X\) (rather than a proportional change in \(1+X\)). In this case, exact interpretation is particularly important, whether performed using linear rescaling or \(e^p\), but there are several details that still separate the two approaches. Consider a one-unit increase in \(\ln(1+X)/\ln(1+p)\). This is equivalent to a proportional change of \(1+p\) in \(1+X\), or an absolute increase of \(p(1+X)\). If \(X\) increases by \(p(1+X)\), what proportional increase is that equivalent to? \begin{equation} \label{eq:linerror} X + p(1+X) = X(1+p + \frac{p}{X}) \end{equation} A one-unit linear increase in \(\ln(1+X)/\ln(1+p)\) interpreted as a \(1+p\) proportional change in \(X\) will get the proportional change wrong by \(\frac{p}{X}\). Similarly, a linear increase of \(p\) in \(\ln(1+X)\) is a proportional increase of \(e^p\) in \(1+p\). As above, \begin{equation} \label{eq:traderror} X+(e^p-1)(1+X) = X(e^p + \frac{e^p-1}{X}) \end{equation} A linear increase of \(p\) in \(\ln(1+X)\) interpreted as a proportional increase of \(e^p\) in \(X\) will get the proportion wrong by \(\frac{e^p-1}{X}\). For a given \(p\), since \(p \leq e^p-1 \ \forall \ p \geq 0\), the linear rescaling approach will always outperform the traditional approach, and by a greater margin as \(p\) increases. However, this is due to the fact that for a given \(p\), the traditional method describes a proportional change of \(e^p \geq 1+p\). For a given \emph{proportional change}, for example comparing a linear increase of \(1\) under linear rescaling to a linear increase of \(\ln(1+p)\) in the traditional method, both methods perform identically. There are still several points to recommend linear rescaling here. First, linear rescaling is an improvement if researchers using the traditional method first select a \(p\) of interest and then calculate \(e^p\), rather than selecting \(e^p\) directly (since, as above, \(p \leq e^p-1 \ \forall \ p \geq 0\)). Second, a bias of \(p/X\) may be easier to reason about than \((e^p-1)/X\). The comparison so far assumes that the researcher using the traditional approach uses the exact interpretation, where a linear increase of \(p\) is a proportional increase of \(e^p\). If they instead interpret a linear increase of \(p\) as a proportional change of \(1+p\) in \(X\) using the \(e^p \approx 1+p\) approximation, the error will be \[ e^p - (1+p) + \frac{e^p-1}{X} \] There are two problems here. First, the fact that linear rescaling and the traditional method perform identically for a given proportional increase doesn't matter, as by using the approximation the researcher has chosen to fix \(p\), under which linear rescaling outperforms the traditional method. Second, the use of the approximation adds the traditional approximation error \(e^p-(1+p)\) on top of \(\frac{e^p-1}{X}\), further increasing error relative to the linear rescaling method, and making the error grow even faster in \(p\). For either the linear rescaling or traditional approaches, the recommendation from Bellemare \& Wichman (2020) to scale \(X\) upwards in reference to \(asinh(X)\), elaborated upon in Aihounton \& Henningsen (2019) to determine optimal scaling values, is also implied by these results for \(\ln(1+X)\), as the interpretation error declines proportionally with larger absolute values of \(X\). \hypertarget{exact-interpretation-with-zeroes}{\subsection{Exact interpretation with zeroes}\label{exact-interpretation-with-zeroes}} As an aside (since it does not relate to linear rescaling in particular), a researcher using either the linear rescaling or traditional method could use one of the error formulas in Equations \ref{eq:linerror} or \ref{eq:traderror} to adjust their proportional-change interpretation, or decide whether the error is small enough to ignore, for a given \(p\) and \(X\). This may be especially useful in the calculation of elasticities, since it allows proportional changes in both \(Y\) and \(X\) to be recovered from proportional changes in \(1+Y\) and \(1+X\). For example, in the log-log model \[\frac{\ln(1+Y)}{\ln(1+p_Y)} = \beta_0 + \beta_1\frac{\ln(1+X)}{\ln(1+p_X)} + \varepsilon \] a \(1+p_X\) proportional change in \(1+X\) is associated with a \((1+p_Y)^{\hat{\beta}_1}\) proportional change in \(1+Y\). For the specific values \(X = X_0\) and \(Y = Y_0\), this means that a \(1 + p_X + \frac{p_X}{X_0}\) proportional change in \(X\) is associated with a \((1+p_Y)^{\hat{\beta}_1} + \frac{(1+p_Y)^{\hat{\beta}_1}-1}{Y_0}\) proportional change in \(Y\). If desired, \(p_X\) and \(X_0\) can be selected ahead of time so that \(1 + p_X + \frac{p_X}{X_0}\) is a round number. Similar calculations follow for the \(\ln(1+X)\)-linear and linear-\(\ln(1+X)\) cases. The only task remaining at that point is calculation of standard errors for this nonlinear function of \(\hat{\beta}_1\). The delta method is one acceptable approach, at least in large samples.\footnote{Keep in mind that this adjustment does not account for zero-censoring of the variables, which is still present and may still harm performance for logarithms on the left-hand side relative to using a model that properly accounts for censoring.} \hypertarget{example-applications}{\section{Example Applications} \label{sec:examples}} In this section I refer to several published studies that use natural logarithm transformations in regression analysis, and discuss how those papers might have been different using linear rescaling. The first study I look at, which is from economics, is Eren, Onda, \& Unel (2019). This study looks at the impact of foreign direct investment (FDI) on entrepreneurship in the United States, covering business-creation, business-destruction, and self-employment rates as outcome variables. The authors use a natural logarithmic transformation of FDI, and the headline result specified in the abstract is ``A 10\% increase in FDI decreases the average monthly rate of business creation and destruction by roughly 4 and 2.5\% (relative to the sample mean), respectively,'' where these figures refer to results in states without Right-to-Work laws. Despite the percentage interpretation given for the effects, business creation and destruction are not log-transformed, only FDI is. The linear effect of log FDI is reported as a percentage of the sample means of business creation and destruction. This study makes use of the traditional approximation. The result for business creation comes from a regression of business creation rates on \(\ln(FDI)\) (lagged two years), where the coefficient on \(\ln(FDI)\) is -1.083. They interpret this as a 10\% increase in FDI reducing the business creation rate by .1083 (or perhaps they round the effect to .11 before proceeding, this is not clear), which is roughly 4\% of the mean (\(.1083/2.889 = .0375\) or \(.11/2.889=.038\)). Under linear rescaling with a base of \(1.1\), the coefficient on \(\log_{1.1}(FDI)\) would be \(.1032\), which indicates an effect roughly 3.5\% of the mean (\(.1032/2.889=.0357\)), keeping in mind that the abstract reports its other effect to the half-percent level of precision. In this case, the use of the approximation made the effect seem larger than it was. The correction is not enormous, but still there is no particular reason the correct result could not have been in the original study. Traditional exact interpretation or linear rescaling would have avoided this problem. However, linear rescaling in this case offers the additional benefit over traditional exact approximation that, if the dependent variable had been divided by its mean, the coefficient on \(\log_{1.1}(FDI)\) would have been the value of interest \(-.0357\) directly, and no further calculation would have been necessary. This ability to put the value of interest on the table also applies to the 2.5\% result, although in this case the substantive reported conclusion does not change at the half-percent level of precision (the effect to two decimal places drops from to 2.55\% to 2.43\%). The second study is from public health. Kim \& Leigh (2020) look at the effect of wages on obesity rates, finding that low wages increase body mass and the prevalence of obesity. This study uses traditional exact interpretation, so linear rescaling cannot change its results, but it may change how they are presented. Their instrumental variables estimate with BMI as a dependent variable reads ``the coefficient on ln-wage was statistically significant (\(p < 0.01\)) and its value was \(-3.3\). Standard errors appear in parentheses. This coefficient suggests that a 10\% increase in wages is associated with 0.32 decline in BMI.'' The .32 appears to be a slight miscalculation and can be derived from \(\ln(1.1)\times(-3.3) = -.3145\), which should round to \(-.31\), not \(-.32\). Under linear rescaling with a base of \(1.1\), this additional calculation would not need to appear in the text, and the coefficient on \(\log_{1.1}(Wage)\) would be rounded to \(-.3145\) or \(-.315\), which would also avoid the slight miscalculation. The authors also report the effect of a 100\% increase in wages as a 2.29 decline in BMI (\(\ln(2)\times(-3.3) = -2.29\)). To achieve this with linear rescaling, the authors either could rerun analysis using a log base of 2, or, more realistically, retain the calculation in the text for this additional result, adjusting the 10\% estimate by \(-.3145\times\ln(2)/\ln(1.1) = -2.29\). This simplified presentation for the BMI results could be similarly applied in analysis of the obesity rate dependent variable. The final study I will mention is an example of how the complex calculations necessary to produce exact interpretations of logarithms using the traditional method can lead to error. Lin, Teymourian, \& Tursini (2018) look at the effect of the natural logarithm of sugar and processed food imports on the prevalence of obesity and overweight. In one model the coefficient on logged imports is .085, which is interpreted as ``10\% increase in import is associated with approximately 0.004 increase in average BMI.'' However, the effect should instead be of a .0081 increase in BMI using exact interpretation, or .0085 using the traditional approximation, an effect more than twice as large. Similar errors are made in another table, where a coefficient of .004 is again interpreted as having about half the appropriate effect size for a 10\% increase. They also provide an interpretation of a 50\% increase, which is about 40\% the appropriate size. If the authors had used linear rescaling, they would not have needed to produce these calculations, and there would have been no potential for this error to occur. Instead, the effect size of interest would have automatically been produced by the statistics software. \hypertarget{conclusion}{% \section{Conclusion}\label{conclusion}} Despite their wide use across many fields, both in the context of regression and in other statistical applications, logarithms are frequently misinterpreted to greater or lesser degrees. This is partly due to the standard interpretation of logarithms, which relies on an approximation that produces non-negligible errors outside of a fairly narrow range of percentage increases. It is possible to skip the approximation and instead accurately interpret a linear increase \(p\) in \(\ln(X)\) as a proportional increase of \(e^p\) in \(X\). However, this practice is not universal, and carries a chance of error. The interpretation of logarithms can be improved by changing the base of the logarithm to \(1+p\), where \(1+p\) is a proportional change of interest. The change of base can be achieved by linearly rescaling \(\ln(X)\) to \(\ln(X)/\ln(1+p)\). This rescaling offers a way of producing exact interpretations of logarithms, or in some cases approximations with less error than the traditional approximation. Linear rescaling can be easier for a reader to understand on a regression table. The one-unit log increase that accompanies a given percentage increase can be understood in the same way as changes in untransformed variables. In order to interpret the coefficient the reader does not need to search for additional calculations in the text, nor does the author need to perform and provide them. Crucially, rescaling is as easy to perform and explain as the traditional approximation, and does not in most cases require the additional post-analysis calculations that usually go along with exact interpretations under the traditional method. All three of the studies covered in Section \ref{sec:examples} contained either errors or misrepresentations of their results because of these post-analysis calculations. I did not select these studies because I knew they had problems or anticipated any; it just so happens that the first three studies I selected as good candidates for demonstrating the method all had problems that could have been avoided with linear rescaling. Ease of use for the researcher is important, because it may help sidestep some of the reasons why researchers do not already report exact results, and may help avoid calculation errors with traditional exact interpretation. Researchers should in general be producing exact interpretations of logarithms and be aware of the extent of error in the traditional approximation. Because it is as easy as the traditional approximation, linear rescaling is an attractive way of providing exact results. \hypertarget{references}{% \section*{References}\label{references}} \addcontentsline{toc}{section}{References} \hypertarget{refs}{} \begin{CSLReferences}{1}{0} \leavevmode\vadjust pre{\hypertarget{ref-aihounton2019units}{}}% Aihounton, Ghislain BD, and Arne Henningsen. 2019. {``Units of Measurement and the Inverse Hyperbolic Sine Transformation.''} IFRO Working Paper. \leavevmode\vadjust pre{\hypertarget{ref-bailey2017real}{}}% Bailey, Michael A. 2017. \emph{Real Econometrics: The Right Tools to Answer Important Questions}. Oxford University Press. \leavevmode\vadjust pre{\hypertarget{ref-bellemare2020elasticities}{}}% Bellemare, Marc F, and Casey J Wichman. 2020. {``Elasticities and the Inverse Hyperbolic Sine Transformation.''} \emph{Oxford Bulletin of Economics and Statistics} 82 (1): 50--61. \leavevmode\vadjust pre{\hypertarget{ref-eren2019}{}}% Eren, Ozkan, Onda, Masayuki, and Bulent Unel. 2019. {``Effects of FDI on Entrepreneurship: Evidence from Right-to-work and Non-right-to-work States.''} \emph{Labour Economics} 58: 98--109. \leavevmode\vadjust pre{\hypertarget{ref-greene2003econometric}{}}% Greene, William H. 2008. \emph{Econometric Analysis}. Pearson Education. \leavevmode\vadjust pre{\hypertarget{ref-halvorsen1980interpretation}{}}% Halvorsen, Robert, and Raymond Palmquist. 1980. {``The Interpretation of Dummy Variables in Semilogarithmic Equations.''} \emph{American Economic Review} 70 (3): 474--75. \leavevmode\vadjust pre{\hypertarget{ref-herz2016}{}}% Herz, Benedikt, and Malwina Mejer. 2016. {``On the Fee Elasticity of the Demand for Trademarks in Europe.''} \emph{Oxford Economic Papers} 68 (4): 1039--1061. \leavevmode\vadjust pre{\hypertarget{ref-kim2010estimating}{}}% Kim, DaeHwan, and John Paul Leigh. 2010. {``Estimating the Effects of Wages on Obesity.''} \emph{Journal of Occupational and Environmental Medicine} 52 (5): 495--500. \leavevmode\vadjust pre{\hypertarget{ref-lin2018effect}{}}% Lin, Tracy Kuo, Teymourian, Yasmin, and Maitri Shila Tursini. 2018. {``The Effect of Sugar and Processed Food Imports on the Prevalence of Overweight and Obesity in 172 Countries.''} \emph{Globalization and health} 14 (1): 1--14. \end{CSLReferences} \end{document}
{ "timestamp": "2021-10-07T02:12:48", "yymm": "2106", "arxiv_id": "2106.03070", "language": "en", "url": "https://arxiv.org/abs/2106.03070", "abstract": "The standard approximation of a natural logarithm in statistical analysis interprets a linear change of \\(p\\) in \\(\\ln(X)\\) as a \\((1+p)\\) proportional change in \\(X\\), which is only accurate for small values of \\(p\\). I suggest base-\\((1+p)\\) logarithms, where \\(p\\) is chosen ahead of time. A one-unit change in \\(\\log_{1+p}(X)\\) is exactly equivalent to a \\((1+p)\\) proportional change in \\(X\\). This avoids an approximation applied too broadly, makes exact interpretation easier and less error-prone, improves approximation quality when approximations are used, makes the change of interest a one-log-unit change like other regression variables, and reduces error from the use of \\(\\log(1+X)\\).", "subjects": "Econometrics (econ.EM)", "title": "Linear Rescaling to Accurately Interpret Logarithms", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9799765575409525, "lm_q2_score": 0.8175744850834648, "lm_q1q2_score": 0.8012038294254106 }
https://arxiv.org/abs/2208.02559
Equivalence between Time Series Predictability and Bayes Error Rate
Predictability is an emerging metric that quantifies the highest possible prediction accuracy for a given time series, being widely utilized in assessing known prediction algorithms and characterizing intrinsic regularities in human behaviors. Lately, increasing criticisms aim at the inaccuracy of the estimated predictability, caused by the original entropy-based method. In this brief report, we strictly prove that the time series predictability is equivalent to a seemingly unrelated metric called Bayes error rate that explores the lowest error rate unavoidable in classification. This proof bridges two independently developed fields, and thus each can immediately benefit from the other. For example, based on three theoretical models with known and controllable upper bounds of prediction accuracy, we show that the estimation based on Bayes error rate can largely solve the inaccuracy problem of predictability.
\section*{Theorem} \begin{theorem} \label{theorem:1} Given a M-state time series, its predictability $\Pi$ is equivalent to the Bayes error rate $R$ in a M-classification problem as \begin{equation} \Pi = 1 - R, \label{eqn:1} \end{equation} if we treat each state as a class, and the series before the state as the feature. \end{theorem} \begin{proof} Denote $x_{n-1}=\omega^{1}\omega^{2}\cdots\omega^{n-1}$ the historical series from time $1$ to $n-1$, where $\omega^{i}\in \Omega$ and $\Omega$ is the set of $M$ states. Denote $Pr[\omega^{n}=\hat{\omega}^{n}|x_{n-1}]$ the probability that our predicted state $\hat{\omega}^{n}$ is equal to the actual state $\omega^{n}$ given $x_{n-1}$, and $\pi(x_{n-1})=\sup_{\omega}\left\{Pr[\omega^{n}=\omega|x_{n-1}] \right\}$ the probability of occurrence of the most probable state at time $n$, then the predictability of the $n$th state given the historical states $x_{n-1}$ is $\Pi(n)=\sum_{x_{n-1}}P(x_{n-1})\pi(x_{n-1})$, where $P(x_{n-1})$ is the probability of observing a particular history $x_{n-1}$, and the sum is taken over all possible histories of length $n-1$. Notice that, $\pi(x_{n-1})$ contains the full predictive power including the potential long-range correlations in the time series, while in practice we usually use shorter historical series, such as $\omega^{n-r}\omega^{n-r+1}\cdots\omega^{n-1}$ with $r$ a cutoff parameter, instead of the full historical series $\omega^{1}\omega^{2}\cdots\omega^{n-1}$, so that in a more general case, the probability of observing a particular history can be smaller than 1. The overall predictability $\Pi$ is then defined as the time averaged predictability for a sufficiently long time series \cite{song2010limits}, as \begin{align} \Pi = \lim\limits_{n\to \infty}\frac{1}{n}\sum_{i=1}^{n}\Pi(i), \end{align} where $x_0=\varnothing$ and $\Pi(1)$ is the predictability of the first state without any available information. Considering an $M$-classification problem with $n$ samples whose class labels are the $n$ states in the time series and whose features are series before the corresponding states. Table \ref{tbl:1} illustrates the one-to-one relationship between time series prediction and classification for an example series with $M=3$ and $n=5$. Denote $p\left(x | \omega_{j}\right)$ the conditional probability density of the feature $x \in X'$, $X'\subset X$ is the set of observed features, and $p\left(\omega_{j}\right)$ the prior probability of the class $\omega_{j}\in \Omega$ $(j = 1, 2, \cdots, M)$, the BER is expressed as: \begin{equation} R=1-\sum_{j=1}^{M} \int_{\Gamma_{j}} p\left(\omega_{j}\right) p\left(x | \omega_{j}\right) dx, \end{equation} with the partition $\Gamma_{j}$ defined as: \begin{equation} \Gamma_{j} \!\triangleq\! \left\{\! x\! \in\! X'\! \mid\! p\left(\omega_{j}\right) p\left(x | \omega_{j}\right)\!>\!\max _{\substack{k \neq j}}\left\{p\left(\omega_{k}\right) p\left(x | \omega_{k}\right)\right\}\! \right\}. \end{equation} Applying the Bayes formula $p(x)p\left(\omega_{j}|x\right) = p\left(\omega_j\right)p\left(x | \omega_{j}\right)$, where $p(x)$ is the prior probability of the feature $x$, we have \begin{equation} \label{eqn:5} \sum_{j=1}^{M} \int_{\Gamma_{j}} p\left(\omega_{j}\right) p\left(x | \omega_{j}\right) dx = \sum_{j=1}^{M} \int_{\Gamma_{j}^{'}} p\left(x\right) p\left(\omega_{j} | x\right) dx, \end{equation} and the partition $\Gamma_{j}$ is equivalent to the partition $\Gamma_{j}^{'}$ with \begin{equation} \Gamma^{'}_{j} \triangleq \left\{ x \in X' \mid p\left(\omega_{j}|x \right)>\max _{\substack{k \neq j}}\left\{p\left(\omega_{k}|x\right)\right\} \right\}. \end{equation} According to our setting, there is a one-to-one relationship between features and historical series, namely for each $x_{i-1}$ $(1 \leq i \leq n)$, there exist a certain $x \in X'$ s.t. $x_{i-1}=x$, and vice versa. As $p(x)$ is the probability of observing the feature $x$ in the feature set $X'$ while $P(x_{i-1})$ is the probability of observing the series $x_{i-1}$ in all historical series of length $i-1$, in the large limit of $n$, $p(x)=\frac{1}{n}P(x_{i-1})$. In addition, $\pi(x_{i-1})=p\left(\omega_{j}|x\in \Gamma^{'}_{j} \right)$ according to the definition of $\pi(\cdot)$. As a consequence, \begin{equation} \sum\limits_{j=1}^{M} \int_{\Gamma^{'}_{j}} p(x) p\left(\omega_{j} |x\right) dx = \lim\limits_{n\to \infty}\frac{1}{n}\sum\limits_{i=1}^{n}\sum\limits_{x_{i-1}}P(x_{i-1})\pi(x_{i-1}). \end{equation} Therefore, the time series predictability is equivalent to the Bayes error rate, with the relationship $\Pi=1-R$. \end{proof} \section*{Results} According to Theorem \ref{theorem:1}, we can directly take advantage of methods developed to calculate $R$ in real datasets to improve the estimation of $\Pi$. Considering a simple example with three states $\Omega = \{A,B,C\}$, where the next state only depends on the current state, according to the following Markovian transfer matrix \begin{equation} \begin{blockarray}{cccc} &A&B&C\\ \begin{block}{c[ccc]} A& q& \frac{2}{3}(1-q)& \frac{1}{3}(1-q)\\ B& \frac{1}{3}(1-q)& q& \frac{2}{3}(1-q)\\ C& \frac{2}{3}(1-q)& \frac{1}{3}(1-q)& q\\ \end{block} \end{blockarray}. \end{equation} Obviously, when $0.4 \le q \le 1$, the true predictability $T=q$. Time series with arbitrary length $n$ can be generated by Eq. 8. We set $r=1$ to extract the features. Take $\{ABBCA\cdots\}$ as an example, the corresponding (feature, class) set is $\{(A,B),(B,B),(B,C),(C,A),\cdots\}$. According to the entropy-based method, the estimated predictability $\bar{\Pi}$ is determined by \begin{equation} \label{eqn:9} H = -\bar{\Pi}\log_2\bar{\Pi} - (1-\bar{\Pi})\log_2(1-\bar{\Pi}) + (1-\bar{\Pi})\log_2(M-1), \end{equation} where $M=3$ and $H$ is the entropy of the next-moment state that can be estimated by the data (see details in \cite{song2010limits}). In the corresponding classification problem, the lower and upper bounds of $R$ can be obtained by the inequality \begin{equation} \label{eqn:10} \begin{gathered} \frac{M-1}{(M-2) M} \sum_{i=1}^{M}\left[1-p(\omega_i)\right] R_{i}^{M-1} \leq R^{M} \leq \min_{\alpha \in\{0,1\}} \frac{1}{M-2 \alpha} \sum_{i=1}^{M}\left[1-p(\omega_i)\right] R_{i}^{M-1}+\frac{1-\alpha}{M-2 \alpha}, \end{gathered} \end{equation} where $R^k$ is the BER for the $k$-classification subproblem and $R_{i}^{M-1}$ is the BER for the $(M-1)$-classification subproblem created by removing the $i$th class (see details in \cite{wisler2016empirically,renggli2021evaluating}). The upper and lower bounds of predictability can then be obtained through Eq. \ref{eqn:1} (Theorem \ref{theorem:1}), and the estimated predictability $\tilde{\Pi}$ is the average of the two bounds. Figure 1A shows how $\bar{\Pi}$ changes with increasing $n$ for three specific cases $q=0.4$, $q=0.6$ and $q=0.8$. The result confirms two above-mentioned disadvantages of the entropy-based method, namely $\bar{\Pi}$ is sensitive to the length $n$ and much larger than the true predictability $T=q$. To ensure the stability, we set $n=2^{15}$ and compare the entropy-based method (Eq. \ref{eqn:9}) and the BER-inspired method (Eq. \ref{eqn:10}). As shown in figure 1B, the latter remarkably and consistently outperforms the former. Considering a more complicated series generator with $M$ states $\Omega = \{S_1,S_2,\cdots,S_M\}$, where the next state $\omega^{t+1}$ is randomly drawn from the $M$ states with probability $1-q$, or determined by the two anterior states with probability $q$. In the latter case, if $\omega^{t-1}=S_i$ and $\omega^t=S_j$, then $\omega^{t+1}=S_k$, $k=i+j$ (if $k>M$, we set $k \leftarrow k-M$). Obviously, the true predictability is $T=q+(1-q)/M$. Figure 1C reports how $\bar{\Pi}$ changes with increasing $n$ for four specific cases $q=0.2$, $q=0.4$, $q=0.6$ and $q=0.8$, with $M=100$ fixed. As $1/M$ is much smaller than $q$ in the above four cases, $T \approx q$. Analogous to what found in figure 1A, $\bar{\Pi}$ is sensitive to $n$ and much larger than $T$ after being nearly stable ($n>2^{15}$). As shown in figure 1D, in most cases the BER-inspired method performs better than the entropy-based method, and only when the time series is highly predictable ($q\approx 1$, see the top right corner), the results of the entropy-based method and BER-inspired method are close to each other. To reveal the effects of parameters $r$ and $M$, we further consider the third generator where the next state $\omega^{t+1}$ is equal to $\omega^{t}$ with probability $q_{1}=0.1$, equal to $\omega^{t-1}$ with probability $q_{2}=0.2$, equal to $\omega^{t-2}$ with probability $q_{3}=0.3$. With probability $1-q_1-q_2-q_3$, $\omega^{t+1}$ is randomly drawn from $M$ states. The true predictability is $T=\max\{q_1,\cdots,q_r\}+(1-q_1-q_2-q_3)/M$, sensitive to $r$ and $M$. As shown in figure 1E, the original entropy-based method does not consider the impacts of parameter $r$ while the BER-inspired method can well capture the effects of the memory length $r$. As shown in figure 1F, both the entropy-based and BER-inspired methods capture the decreasing tendency of predictability as the increase of $M$. One can clearly observed from figures 1E and 1F that the entropy-based method will largely overestimate the predictability even for sufficiently long time series, while the BER-inspired method performs much better. \begin{figure*}[t] \centering \subfigure \label{fig:side:a \includegraphics[width=2in]{fig_a} } \subfigure \label{fig:side:b \includegraphics[width=2in]{fig_b} } \subfigure \label{fig:side:c \includegraphics[width=2in]{fig_c} } \subfigure \label{fig:side:d \includegraphics[width=2.08in]{fig_d} } \subfigure \label{fig:side:e \includegraphics[width=2in]{fig_e} } \subfigure \label{fig:side:f \includegraphics[width=2.07in]{fig_f} } \vspace{0pt} \caption{(A) How the estimated predictability $\bar{\Pi}$ by the entropy-based method changes with the increasing $n$ under the first series generator. (B) The performance of the entropy-based method (Eq. 9) and the BER-inspired method (Eq. 10) under the first generator. (C) How the estimated predictability $\bar{\Pi}$ by the entropy-based method changes with the increasing $n$ under the second series generator. (D) The performance of the entropy-based and BER-inspired methods under the second generator. For the BER-inspired method, $\tilde{\Pi}^{\textup{low}}$ and $\tilde{\Pi}^{\textup{up}}$ are the lower and upper bounds by Eq. (10), and $\tilde{\Pi}=\frac{1}{2}\left( \tilde{\Pi}^{\textup{low}}+\tilde{\Pi}^{\textup{up}} \right)$ is the estimated predictability. In plots (B) and (D), the shadow areas indicate to what extent the BER-inspired method outperforms the entropy-based method. (E) The performance of the entropy-based and BER-inspired methods under the thrid generator with varying $r$, with $M=20$ fixed. (F) The performance of the entropy-based and BER-inspired methods under the thrid generator with varying $M$, with $r=3$ fixed. In plots (E) and (F), the shadow areas denote the standard errors. In all comparisons between the entropy-based and BER-inspired methods, the length of time series is fixed as $n=2^{15}$, and the corresponding results are averaged over 10 independent runs.} \label{fig:results \end{figure*} \section*{Discussion} The direct value of knowing predictability is to decide whether it is worthwhile to improve the current predictors \cite{song2010limits,lu2015toward}. The embodiment of such value requires an accurate estimate of predictability. Unfortunately, the entropy-based method \cite{song2010limits} usually fails as it largely overestimates the true predictability (see, for example, figure \ref{fig:results}). The dissatisfactory performance partially comes from the approximation that only accounts for the entropy of the state with the maximum next-moment occurrence probability. At the same time, such approximation is an indispensable part that guarantees the computational feasibility. Therefore, it is difficult to overcome the observed disadvantages within the entropic framework \cite{smith2014refined,zhang2022beyond}. This paper uncovers the equivalence between predictability and a seemingly unrelated metric BER, and immediately provides a novel way to improve the estimation of predictability -- applying the BER-inspired methods. \section*{Acknowledgement} This work was supported in part by the National Natural Science Foundation of China (No. 61960206008, No. 62002294, No. 11975071) and the National Science Fund for Distinguished Young Scholars (No. 61725205).
{ "timestamp": "2022-08-05T02:10:13", "yymm": "2208", "arxiv_id": "2208.02559", "language": "en", "url": "https://arxiv.org/abs/2208.02559", "abstract": "Predictability is an emerging metric that quantifies the highest possible prediction accuracy for a given time series, being widely utilized in assessing known prediction algorithms and characterizing intrinsic regularities in human behaviors. Lately, increasing criticisms aim at the inaccuracy of the estimated predictability, caused by the original entropy-based method. In this brief report, we strictly prove that the time series predictability is equivalent to a seemingly unrelated metric called Bayes error rate that explores the lowest error rate unavoidable in classification. This proof bridges two independently developed fields, and thus each can immediately benefit from the other. For example, based on three theoretical models with known and controllable upper bounds of prediction accuracy, we show that the estimation based on Bayes error rate can largely solve the inaccuracy problem of predictability.", "subjects": "Information Theory (cs.IT); Data Analysis, Statistics and Probability (physics.data-an)", "title": "Equivalence between Time Series Predictability and Bayes Error Rate", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9799765552017675, "lm_q2_score": 0.8175744761936437, "lm_q1q2_score": 0.8012038188011363 }
https://arxiv.org/abs/1212.1412
Antiderivatives Exist without Integration
We present a proof that any continuous function with domain including a closed interval yields an antiderivative of that function on that interval. This is done without the need of any integration comparable to that of Riemann, Cauchy, or Darboux. The proof is based on one given by Lebesgue in 1905.
\section*{Lebesgue's Construction} Throughout, let $f$ be a function that is continuous on an interval $[a, b]$. We use Lebesgue's notation and give a modern rendition of his construction. Suppose a set of $n$ points $\{(a_{0},d_{0}), (a_{1},d_{1}), \dots, (a_{n},d_{n})\}$, $a=a_{0} < a_{1} < \dots < a_{n}=b$ are points of the interval $[a,b]$. If desired, allow $f(a_{i}) = d_{i}, i = 0, 1, \dots, n$. Lebesgue defines a continuous function $\phi$ with domain including $[a,b]$ such that for $i=0,1,\dots,n-1$, there are numbers $m_{i},b_{i}$ such that $\phi(x) = m_{i}x+b_{i}$ for each $x$ in $[a_{i},a_{i+1}]$, and $m_{i}a_{i+1}+b_{i}=m_{i+1}a_{i+1}+b_{i+1}$ holds for $i=0,1,\dots,n-2$. Lebesgue defines an antiderivative $\Phi$ for $\phi$ on $[a,b]$ as follows: \begin{enumerate} \item Let $\Phi_{0}(x) = (m_{0}/2)x^{2} + b_{0}x - (m_{0}/2)a_{0}^{2}-b_{0}a_{0}$ for each $x$ in $[a_{0},a_{1}]$. \item Define $\Phi_{1}(x) = (m_{1}/2)x^{2} + b_{1}x + \Phi_{0}(a_{1}) - (m_{1}/2)a_{1}^{2}-b_{1}a_{1}$ for each $x$ in $[a_{1},a_{2}]$. Inductively, define $\Phi_{i}, i=2,3,\dots,n-1$. \item Since $\Phi_{0}(a_{0}), \Phi_{1}(a_{1}), \dots, \Phi_{n-1}(a_{n-1})$ are well defined, the function $\Phi$ is defined as follows: \begin{equation*} \Phi(x) = \begin{cases} \displaystyle \frac{m_{0}}{2}x^{2} + b_{0}x - \frac{m_{0}}{2}a_{0}^{2} - b_{0}a_{0}, & \text{if $x \in [a_{0},a_{1}]$}; \\ & \\ \displaystyle \frac{m_{i}}{2}x^{2} + b_{i}x +\Phi_{i-1}(a_{i})- \frac{m_{i}}{2}a_{i}^{2} - b_{i}a_{i}, & \text{if $x \in [a_{i},a_{i+1}]$},i=1,2,\dots,n-1. \end{cases} \end{equation*} \end{enumerate} The constructed function $\Phi$ consists of $n$ second degree polynomials whose left and right slopes at $a_{1},a_{2},\dots,a_{n-1}$, respectively, are equal. From the above construction, it can be shown that there is a function $F$ whose derivative is $f$. This is the unique contribution of Lebesgue. The reader may see how the proof might proceed from this point; however, we give an outline below. \section*{Outline of Proof} For the following, we use partition of $[a,b]$ to mean any finite collection of subintervals of $[a,b]$ that are non-overlapping and whose union is $[a,b]$. A refinement $P^{\prime}$ of a partition $P$ is merely another partition of $[a,b]$ such that each end point of each member of $P$ is also an end point of a member of $P^{\prime}$. For $n=1,2,\dots$, we let $P_{n}$ denote a regular partition of $[a,b]$ with $2^{n-1}$ members each having length $(b-a)/2^{n-1}$. This makes $P_{m}$ a refinement of $P_{n}$ for positive integers $m,n$ where $m \ge n$. The functions $\phi_{n}$ and $\Phi_{n}$ will be based on the partition $P_{n}$ for $n =1,2, \dots$. When $P_{n}$, $\phi_{n}$, and $\Phi_{n}$ are used, assume that the subscript is a positive integer unless otherwise stated. We remind the reader that any continuous function achieves extrema on any closed interval in its domain. This makes the following definition non-vacuous. \begin{definition} Suppose $P$ is a partition of $[a,b]$. By the {\bf oscillation} of $f$ on $\delta \in P$ (written $\omega_{\delta}$), we mean the real number \[\omega_{\delta} = \underset{\delta}{\max} f - \underset{\delta}{\min} f.\] Moreover, by the {\bf total oscillation} of $f$ on a partition $P$ of $[a,b]$ (written $\Omega(P)$), we mean the maximum value of the finite set of oscillations $\{\omega_{\delta} : \delta \in P\}$. \end{definition} Whenever each of $m$ and $n$ is a positive integer with $m \ge n$ and $\Omega_{n} < \epsilon$ for some $\epsilon > 0$, then $\Omega_{m} < \epsilon$. (Remember that $P_{m}$ is a refinement of $P_{n}$.) This result is an application of the previous definitions of oscillation and total oscillation. We state the following lemmas without proofs each ultimately being an application of the Heine-Borel Theorem or the basic definitions of oscillation and total oscillation given above. \begin{lemma} \label{ep2} For each $\epsilon >0$, there is a positive integer $n$ such that $\Omega_{m} < \epsilon$ for each positive integer $m \ge n$; that is, $\Omega_{n} \rightarrow 0$ as $n \rightarrow \infty$. \end{lemma} \begin{lemma} \label{Omega4} For each positive integer $n$, $ |f(x) - \phi_{n}(x)| \le \Omega_{n}$ for each $x$ in $[a,b]$; that is, $\phi_{n}(x)$ converges to $f(x)$ for each $x \in [a,b]$. \end{lemma} \begin{comment} \begin{proof} Suppose $x$ is any point in $[a,b]$, $P_{n}$ is any partition of $[a,b]$, $\delta$ is any member of $P_{n}$. Because $f$ is continuous on $[a,b]$, $\underset{\delta}{\min} f$ and $\underset{\delta}{\max} f$ both exist on $[a,b]$, positioning $f(x)$ between $\underset{\delta}{\min} f$ and $\underset{\delta}{\max} f$. Since $\phi_{n}$ is a line segment agreeing with $f$ at its end points, $\phi_{n}(x)$ lies between $\underset{\delta}{\min} f$ and $\underset{\delta}{\max} f$. Thus, we see that $\underset{\delta}{\min} f - \underset{\delta}{\max} f \le f(x) - \phi_{n}(x) \le \underset{\delta}{\max} f - \underset{\delta}{\min} f$. With the definition of $\omega_{\delta}$, we have \begin{equation*} \label{Exist:Eq} |f(x) - \phi_{n}(x)| \le \underset{\delta}{\max} f - \underset{\delta}{\min} f = \omega_{\delta}. \end{equation*} Therefore, from the definition of $\Omega_{n}$, \begin{equation*}\label{Omega} |f(x) - \phi_{n}(x)| \le \Omega_{n}.\end{equation*} In summary, for each positive integer $n$, $ |f(x) - \phi_{n}(x)| \le \Omega_{n}$ for each $x$ in $[a,b]$. \end{proof} \end{comment} \begin{theorem} The sequence $\{\Phi_{n}\}$ converges uniformly to a function $F$ that is an antiderivative of $f$. \end{theorem} \begin{proof} By Lemma~\ref{Omega4} and Theorem 7.9 of \cite{Rudin}, we know that $\phi_{n} \rightarrow f$ uniformly on $[a,b]$. By Theorem 7.17 of \cite{Rudin}, $\Phi_{n}$ converges uniformly to $F$ and \[\phi_{n}(x) \rightarrow F^{\prime}(x)\] for each $x \in [a,b]$. Thus, $F^{\prime} = f$ on $[a,b]$. \end{proof} Lebesgue finished his paper by proving that the integral of $f$ exists on $[a,b]$. He did this by applying a construction that he developed in his proof that $F^{\prime} = f$.
{ "timestamp": "2012-12-07T02:04:39", "yymm": "1212", "arxiv_id": "1212.1412", "language": "en", "url": "https://arxiv.org/abs/1212.1412", "abstract": "We present a proof that any continuous function with domain including a closed interval yields an antiderivative of that function on that interval. This is done without the need of any integration comparable to that of Riemann, Cauchy, or Darboux. The proof is based on one given by Lebesgue in 1905.", "subjects": "History and Overview (math.HO)", "title": "Antiderivatives Exist without Integration", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9799765575409524, "lm_q2_score": 0.817574471748733, "lm_q1q2_score": 0.801203816357686 }
https://arxiv.org/abs/1601.06332
An upper bound on the size of diamond-free families of sets
Let $La(n,P)$ be the maximum size of a family of subsets of $[n]=\{1,2,...,n\}$ not containing $P$ as a (weak) subposet. The diamond poset, denoted $B_{2}$, is defined on four elements $x,y,z,w$ with the relations $x<y,z$ and $y,z<w$. $La(n,P)$ has been studied for many posets; one of the major open problems is determining $La(n,B_{2})$.Studying the average number of sets from a family of subsets of $[n]$ on a maximal chain in the Boolean lattice $2^{[n]}$ has been a fruitful method. We use a partitioning of the maximal chains and introduce an induction method to show that $La(n,B_{2})\leq(2.20711+o(1))\binom{n}{\left\lfloor \frac{n}{2}\right\rfloor }$, improving on the earlier bound of $(2.25+o(1))\binom{n}{\left\lfloor \frac{n}{2}\right\rfloor }$ by Kramer, Martin and Young.
\section*{\refname}} \usepackage{chngcntr} \newenvironment{boldproof}[1][\proofname] {\par\pushQED{\qed}\normalfont\topsep6\p@\@plus6\p@\relax\trivlist\item[\hskip\labelsep\bfseries#1\@addpunct{.}]\ignorespaces}{\popQED\endtrivlist\@endpefalse} \@ifundefined{showcaptionsetup}{}{% \PassOptionsToPackage{caption=false}{subfig}} \usepackage{subfig} \makeatother \addto\captionsamerican{\renewcommand{\corollaryname}{Corollary}} \addto\captionsamerican{\renewcommand{\definitionname}{Definition}} \addto\captionsamerican{\renewcommand{\examplename}{Example}} \addto\captionsamerican{\renewcommand{\lemmaname}{Lemma}} \addto\captionsamerican{\renewcommand{\propositionname}{Proposition}} \addto\captionsamerican{\renewcommand{\remarkname}{Remark}} \addto\captionsamerican{\renewcommand{\theoremname}{Theorem}} \addto\captionsenglish{\renewcommand{\corollaryname}{Corollary}} \addto\captionsenglish{\renewcommand{\definitionname}{Definition}} \addto\captionsenglish{\renewcommand{\examplename}{Example}} \addto\captionsenglish{\renewcommand{\lemmaname}{Lemma}} \addto\captionsenglish{\renewcommand{\propositionname}{Proposition}} \addto\captionsenglish{\renewcommand{\remarkname}{Remark}} \addto\captionsenglish{\renewcommand{\theoremname}{Theorem}} \providecommand{\corollaryname}{Corollary} \providecommand{\definitionname}{Definition} \providecommand{\examplename}{Example} \providecommand{\lemmaname}{Lemma} \providecommand{\propositionname}{Proposition} \providecommand{\remarkname}{Remark} \providecommand{\theoremname}{Theorem} \addto\captionsamerican{\renewcommand{\observationname}{Observation}} \addto\captionsenglish{\renewcommand{\observationname}{Observation}} \providecommand{\observationname}{Observation} \providecommand{\theoremname}{Theorem} \begin{document} \begin{comment} If \pgfversion<3.0.0, UPDATE TikZ. \end{comment} \begin{comment} GENERATING EXTERNALIZED CHARTS (IF THEY ARE CHANGED): 1. Uncomment the line below 2. Export to tex (plain) 3. Run these commands: latex \textendash jobname diamond\_free\_families\_figure1 diamond\_free\_families.tex dvips diamond\_free\_families\_figure1.dvi ps2eps diamond\_free\_families\_figure1.ps latex \textendash jobname diamond\_free\_families\_figure2 diamond\_free\_families.tex dvips diamond\_free\_families\_figure2.dvi ps2eps diamond\_free\_families\_figure2.ps 4. Delete diamond\_free\_families\_figure1.ps and diamond\_free\_families\_figure2.ps 5. Comment out the line below again, export to tex, and run the regular latex CHANGE THE \textbackslash{}pgfrealjobname COMMAND IN THE PREAMBLE IF YOU CHANGE THE FILE NAME \end{comment} \compiletikztrue \global\long\def\mathcal{F}{\mathcal{F}} \global\long\def\mathcal{X}{\mathcal{X}} \global\long\def\mathcal{\mathcal{\mathcal{B}}}{\mathcal{\mathcal{\mathcal{B}}}} \global\long\def\mathcal{C}{\mathcal{C}} \global\long\def\mathcal{Y}{\mathcal{Y}} \global\long\def\mathcal{Z}{\mathcal{Z}} \global\long\def\mathcal{A}{\mathcal{A}} \global\long\def\mathcal{D}{\mathcal{D}} \global\long\def\mathcal{Q}{\mathcal{Q}} \global\long\def\mathcal{H}{\mathcal{H}} \global\long\def\textnormal{if }{\textnormal{if }} \global\long\def\textnormal{ and }{\textnormal{ and }} \global\long\def\textnormal{ or }{\textnormal{ or }} \global\long\def\operatorname{La}{\operatorname{La}} \global\long\def\mathbf{N}{\mathbf{N}} \global\long\def\mathbf{R}{\mathbf{R}} \global\long\def\mathnote#1{} \author{D\'aniel Gr\'osz\thanks{Department of Mathematics, University of Pisa. e-mail: \protect\href{mailto:groszdanielpub@gmail.com}{groszdanielpub@gmail.com}}\and Abhishek Methuku\thanks{Department of Mathematics, Central European University, Budapest, Hungary. e-mail: \protect\href{mailto:abhishekmethuku@gmail.com}{abhishekmethuku@gmail.com}}\and Casey Tompkins\thanks{Alfr\'ed R\'enyi Institute of Mathematics, Hungarian Academy of Sciences. e-mail: \protect\href{mailto:ctompkins496@gmail.com}{ctompkins496@gmail.com}}} \title{An upper bound on the size of diamond-free families of sets} \maketitle \begin{abstract} Let $\operatorname{La}(n,P)$ be the maximum size of a family of subsets of $[n]=\{1,2,\ldots,n\}$ not containing $P$ as a (weak) subposet. The diamond poset, denoted $\mathcal{Q}_{2}$, is defined on four elements $x,y,z,w$ with the relations $x<y,z$ and $y,z<w$. $\operatorname{La}(n,P)$ has been studied for many posets; one of the major open problems is determining $\operatorname{La}(n,\mathcal{Q}_{2})$. It is conjectured that $\operatorname{La}(n,\mathcal{Q}_{2})=(2+o(1))\binom{n}{\left\lfloor n/2\right\rfloor }$, and infinitely many significantly different, asymptotically tight constructions are known. Studying the average number of sets from a family of subsets of $[n]$ on a maximal chain in the Boolean lattice $2^{[n]}$ has been a fruitful method. We use a partitioning of the maximal chains and introduce an induction method to show that $\operatorname{La}(n,\mathcal{Q}_{2})\leq(2.20711+o(1))\binom{n}{\left\lfloor n/2\right\rfloor }$, improving on the earlier bound of $(2.25+o(1))\binom{n}{\left\lfloor n/2\right\rfloor }$ by Kramer, Martin and Young. \end{abstract} \section{Introduction} Let $[n]=\{1,2,\dots,n\}$. The Boolean lattice $2^{[n]}$ is defined as the family of all subsets of $[n]=\{1,2,\ldots,n\}$, and the\emph{ $i$th level} of $2^{[n]}$ refers to the collection of all sets of size $i$. In 1928, Sperner proved the following well-known theorem. \begin{thm}[Sperner \cite{sperner1928satz}] \emph{}If $\mathcal{F}$ is a family of subsets of $[n]$ such that no set contains another ($A,B\in\mathcal{F}$ implies $A\not\subset B$), then $\left|\mathcal{F}\right|\le\binom{n}{\left\lfloor n/2\right\rfloor }$. Moreover, equality occurs if and only if $\mathcal{F}$ is a level of maximum size in $2^{[n]}$. \end{thm} \begin{defn} Let $P$ be a finite poset, and $\mathcal{F}$ be a family of subsets of $[n]$. We say that $P$ is contained in $\mathcal{F}$ as a (weak) subposet if there is an injection $\varphi:P\rightarrow\mathcal{F}$ satisfying $x_{1}<_{p}x_{2}\Rightarrow\varphi(x_{1})\subset\varphi(x_{2})$ for every $x_{1},x_{2}\in P$. $\mathcal{F}$ is called $P$-free if $P$ is not contained in $\mathcal{F}$ as a weak subposet. We define the corresponding extremal function as $\operatorname{La}(n,P):=\max\{\left|\mathcal{F}\right|:\mathcal{F}\textnormal{ is \ensuremath{P}-free}\}$. \end{defn} A $k$-chain, denoted by $P_{k}$, is defined to be the poset on the set $\{x_{1},x_{2},\dots,x_{k}\}$ with the relations $x_{1}\le x_{2}\le\dots\le x_{k}$. Using the above notation, Sperner's theorem can be stated as $\operatorname{La}(n,P_{2})=\binom{n}{\left\lfloor n/2\right\rfloor }$. Let $\Sigma(n,k)$ denote the sum of the $k$ largest binomial coefficients of order $n$. An important generalization of Sperner's theorem due to Erd\H os\emph{ }\cite{erdos1945lemma} states that $\operatorname{La}(n,P_{k+1})=\Sigma(n,k)$. Moreover, equality occurs if and only if $\mathcal{F}$ is the union of $k$ of the largest levels in $2^{[n]}$. \begin{defn}[Posets $\mathcal{Q}_{2},V$ and $\Lambda$] The diamond poset, denoted $\mathcal{Q}_{2}$ (or $\mathcal{D}_{2}$ or $\mathcal{\mathcal{\mathcal{B}}}_{2}$), is a poset on four elements $\{x,y,z,w\}$, with the relations $x<y,z$ and $y,z<w$. That is, $\mathcal{Q}_{2}$ is a subposet of a family of sets $\mathcal{A}$ if there are different sets $A,B,C,D\in\mathcal{A}$ with $A\subset B,C$ and $B,C\subset D$. (Note that $B$ and $C$ are not necessarily unrelated.) The $V$ poset is a poset on $\{x,y,z\}$ with the relations $x\leq y,z$; the $\Lambda$ poset is defined on $\{x,y,z\}$ with the relations $x,y\leq z$. That is, the $\Lambda$ is a subposet of a family of sets $\mathcal{A}$ if there are different sets $B,C,D\in\mathcal{A}$ with $B,C\subset D$. \end{defn} The general study of forbidden poset problems was initiated in the paper of Katona and Tarj\'an \cite{katona1983extremal} in 1983. They determined the size of the largest family of sets containing neither a $V$ nor a $\Lambda$. They also gave an estimate on the maximum size of $V$-free families: $\left(1+\frac{1}{n}+o\!\left(\frac{1}{n}\right)\right)\binom{n}{\left\lfloor n/2\right\rfloor }\le\operatorname{La}(n,V)\le\left(1+\frac{2}{n}\right)\binom{n}{\left\lfloor n/2\right\rfloor }$. This result was later generalized by De Bonis and Katona \cite{krfork} who obtained bounds for the $r$-fork poset, $V_{r}$ defined by the relations $x\le y_{1},y_{2},\dots,y_{r}$. Other posets for which $\operatorname{La}(n,P)$ has been studied include complete two level posets, batons \cite{thanh}, crowns $O_{2k}$ (cycle of length $2k$ on two levels, asymptotically solved except for $k\in\{3,5\}$ \cite{evencrown,lu2014crown}), butterfly \cite{de2005largest}, skew-butterfly \cite{methuku2014exact}, the $\mbox{N}$ poset \cite{griggs2008}, harp posets $\mathcal{H}(l_{1},l_{2},\dots,l_{k})$, defined by $k$ chains of length $l_{i}$ between two fixed elements \cite{griggs2012diamond}, and recently the complete 3 level poset $K_{r,s,t}$ \cite{patkos} among others. (See \cite{griggs_survey} for a nice survey by Griggs and Li.) One of the first general results is due to Bukh \cite{bukh2009set} who determined the asymptotic value of $\operatorname{La}(n,P)$ for all posets whose Hasse diagram is a tree: If $T$ is a finite poset whose Hasse diagram is a tree of height $h(T)\ge2$, then $\operatorname{La}(n,T)=(h(T)-1)\binom{n}{\lfloor n/2\rfloor}\left(1+O\!\left(\frac{1}{n}\right)\right).$ Using more general structures instead of chains for double counting, Burcsi and Nagy \cite{burcsi2013method} obtained a weaker version of this theorem for general posets showing that $\operatorname{La}(n,P)\le\left(\frac{\left|P\right|+h(P)}{2}-1\right)\binom{n}{\lfloor n/2\rfloor}$. Later this was generalized by Chen and Li \cite{chen2014note} and recently this general bound was improved by the authors of the present article \cite{GrMeToGeneral}. The most investigated poset for which even the asymptotic value of $\operatorname{La}(n,P)$ has yet to be determined is the diamond $\mathcal{Q}_{2}$ which is the topic of our paper. The two middle levels of the Boolean lattice do not contain a diamond, so $\operatorname{La}(n,\mathcal{Q}_{2})\geq(2-o(1))\binom{n}{\left\lfloor n/2\right\rfloor }$. Czabarka, Dutle, Johnston and Sz\'ekely \cite{diamond_constructions} gave infinitely many asymptotically tight constructions by using random set families defined from posets based on Abelian groups. Such constructions suggest that the diamond problem is hard. Using a simple and elegant argument, Griggs, Li and Lu \cite{griggs2012diamond} showed that $\operatorname{La}(n,\mathcal{Q}_{2})<2.296\mbox{\ensuremath{\binom{n}{\left\lfloor n/2\right\rfloor }}}$. Some time after they had announced this bound, Axenovich, Manske and Martin \cite{diamond_228} improved the upper bound to $2.283\mbox{\ensuremath{\binom{n}{\left\lfloor n/2\right\rfloor }}}$. This bound was further improved to $2.273\binom{n}{\left\lfloor n/2\right\rfloor }$ by Griggs, Li and Lu \cite{griggs2012diamond}. The best known upper bound on $\operatorname{La}(n,\mathcal{Q}_{2})$ is $(2.25+o(1))\binom{n}{\left\lfloor n/2\right\rfloor }$ due to Kramer, Martin and Young \cite{diamond225}. \begin{defn} A \emph{maximal chain} or, for the rest of this article, simply a \emph{chain} of the Boolean lattice is a sequence of sets $\emptyset,\{x_{1}\},\{x_{1},x_{2}\},\{x_{1},x_{2},x_{3}\},\dots,[n]$ with $x_{1},x_{2},x_{3}\ldots\in[n]$. We refer to $\{x_{1},\ldots,x_{i}\}$ as the $i$th set on the chain. In particular, we refer to $\{x_{1}\}$ as the first set on the chain, or just say that the chain starts with the element $x_{1}$ (as a singleton). We refer to $x_{i}$ as the $i$th element added to form the chain. \end{defn} \begin{defn} The \emph{Lubell function} of a family of sets $\mathcal{F}\subseteq2^{[n]}$ is defined as \[ l(n,\mathcal{F})=\sum_{F\in\mathcal{F}}\frac{1}{\binom{n}{\left|F\right|}}. \] The notation is shortened to just $l(\mathcal{F})$ when there is no ambiguity as to the dimension of the Boolean lattice. \end{defn} \begin{observation} The Lubell function of a family $\mathcal{F}$ is the average number of sets from $\mathcal{F}$ on a chain, taken over all $n!$ chains. In particular, the Lubell function of a level is 1, and the Lubell function of an antichain $\mathcal{F}$ is the number of chains containing a set from $\mathcal{F}$ divided by $n!$. The Lubell function is additive across a union of disjoint families of sets. Furthermore, $\left|\mathcal{F}\right|\leq l(\mathcal{F})\binom{n}{\left\lfloor n/2\right\rfloor }$ (\cite{LYMlubell}). \end{observation} The Lubell function was derived from the celebrated YMBL inequality which was independently discovered by Yamamoto, Meshalkin, Bollob\'as and Lubell. Using the Lubell function terminology, it states that \begin{namedthm}[YMBL inequality \textup{\textmd{(Yamamoto, Meshalkin, Bollob\'as, Lubell \cite{yamamoto1954logarithmic,meshalkin1963generalization,bollobas1965generalized,LYMlubell})}}] If $\mathcal{F}\subseteq2^{[n]}$ is an antichain, then $l(\mathcal{F})\leq1$. \end{namedthm} For a poset $P$, let $\overline{l}(n,P)$ be the maximum of $l(n,\mathcal{F})$ over all families $\mathcal{F}\subseteq2^{[n]}$ which are both $P$-free and contain the empty set. Let $\overline{l}(P)=\limsup_{n\to\infty}\overline{l}(n,P)$. Griggs, Li and Lu proved that \begin{lem}[Griggs, Li and Lu \cite{griggs2012diamond}] \label{lem:minpartition} \[ \operatorname{La}(n,\mathcal{Q}_{2})\leq\left(\overline{l}(\mathcal{Q}_{2})+o(1)\right)\binom{n}{\left\lfloor n/2\right\rfloor }. \] \end{lem} Kramer, Martin and Young used flag algebras to prove that \begin{lem}[Kramer, Martin and Young \cite{diamond225}] \label{lem:lubell2.25}$\overline{l}(\mathcal{Q}_{2})=2.25$ \end{lem} thereby proving \begin{thm}[Kramer, Martin and Young \cite{diamond225}] \label{thm:La2.25} \[ \operatorname{La}(n,\mathcal{Q}_{2})\leq(2.25+o(1))\binom{n}{\left\lfloor n/2\right\rfloor }. \] \end{thm} The following construction shows that $\bar{l}(\mathcal{Q}_{2})\ge2.25$ in Lemma \ref{lem:lubell2.25}. There are other constructions known as well. \begin{example} Let $\mathcal{F}\subseteq2^{[n]}$ consist of all the sets of the following forms: $\emptyset,\{e\},\{e,o\},\{o_{1},o_{2}\}$ where $e$ denotes any even number in $[n]$, and $o$, $o_{1}$ and $o_{2}$ denote any odd numbers in $[n]$. This family is diamond-free, and $l(\mathcal{F})=2.25\pm o(1)$. \end{example} \begin{example} \label{exa:canonical}This construction is a generalization of the previous one. Let $A\subseteq[n]$ with $\left|A\right|=an$. Let $\mathcal{F}\subseteq2^{[n]}$ consist of all the sets of the following forms: $\emptyset,\{e\},\{e,o\},\{o_{1},o_{2}\}$ where now $e$ denotes any element of $A$, while $o$, $o_{1}$ and $o_{2}$ denote any elements of $[n]\setminus A$. This family is diamond-free, and $l(\mathcal{F})=2+a-a^{2}\pm o(1)$. This family contains all size 2 sets that do not form a diamond with $\emptyset$ and the singletons, so all maximal diamond-free families on levels 0, 1 and 2 that contain $\emptyset$ are of this form. \end{example} The following restriction of the problem of diamond-free families has been investigated: How big can a diamond-free family be if it can only contain sets from the middle three levels of $2^{[n]}$ (denoted $\mathcal{\mathcal{\mathcal{B}}}(n,3)$)? Better bounds are known with this restriction. Axenovich, Manske and Martin showed that \begin{thm}[Axenovich, Manske and Martin \cite{diamond_228}] \label{thm:3levels2.207}If $\mathcal{F}\subseteq\mathcal{\mathcal{\mathcal{B}}}(n,3)$ is diamond-free, then $\left|\mathcal{F}\right|\leq\bigl(2.20711+o(1)\bigr)\binom{n}{\left\lfloor n/2\right\rfloor }$. \end{thm} Later, Manske and Shen improved it to $2.1547\mbox{\ensuremath{\binom{n}{\left\lfloor n/2\right\rfloor }}}$ in \cite{manske2013three} and recently, Balogh, Hu, Lidick\'y and Liu gave the best known bound of $2.15121\binom{n}{\left\lfloor n/2\right\rfloor }$ in \cite{diamond21512} using flag algebras. \begin{defn} We call a chain \emph{maximal\textendash non-maximal (MNM)} with respect to (w.r.t.) $\mathcal{F}$ if it contains a set from $\mathcal{F}$, and the biggest set contained in $\mathcal{F}$ on the chain is not maximal in $\mathcal{F}$ (i.e., there are other sets from $\mathcal{F}$ containing it on some other chains). \end{defn} It is easy to see that an $\emptyset$-free family is $\Lambda$-free if and only if the family we get by adding $\emptyset$ is diamond-free; adding $\emptyset$ increases the Lubell function by 1. In Section \ref{sec:invertedV} of this paper, we prove the following lemma: \begin{lem} \label{lem:lubell}Let $\mathcal{F}\subseteq2^{[n]}$ be a $\Lambda$-free family that does \emph{not} contain the empty set, nor any set of size bigger than $n-n'$ for some $n'\in\mathbf{N}$ (that can be chosen independently of $n$). Assume that there are $cn!$ MNM chains w.r.t.\ $\mathcal{F}$. Then $l(\mathcal{F})\leq1-\min\!\left(c+\frac{1}{n'},\frac{1}{4}\right)+\sqrt{\min\!\left(c+\frac{1}{n'},\frac{1}{4}\right)}+\frac{3}{n'}$. \end{lem} It is easy to see that in Example \ref{exa:canonical} the number of MNM-chains is approximately $a^{2}n!$ (so $a\approx\sqrt{c}$): these are the chains whose second set is $\{e_{1},e_{2}\}$ with $e_{1},e_{2}\in A$. Thus, this lemma is (asymptotically) sharp, and states that for a given number of MNM chains, Example \ref{exa:canonical} cannot be beaten (with some restriction on the sizes of the sets). Barring the requirement that the topmost $n'$ levels be empty, Lemma \ref{lem:lubell} is a generalization of Lemma \ref{lem:lubell2.25}. The proof of Lemma \ref{lem:minpartition} in \cite{diamond225} actually works with the restriction of Lemma \ref{lem:lubell} concerning the topmost sets (that there is no set of size bigger than $n-n'$) with $n'=n/2-n^{2/3}$, immediately giving a new proof of Theorem \ref{thm:La2.25}. Our proof of Lemma \ref{lem:lubell} includes an intricate induction step and a (non-combinatorial) lemma about functions involving a lot of elementary algebra and calculus; but it does not require flag algebras, and it does not use details of the structure of $\mathcal{F}$ above the second level (except inside the induction). Section \ref{sec:diamond} of this paper uses Lemma \ref{lem:lubell} to prove our main theorem: \begin{thm} \label{thm:main}$\operatorname{La}(n,\mathcal{Q}_{2})\leq\left(\frac{\sqrt{2}+3}{2}+o(1)\right)\binom{n}{\left\lfloor n/2\right\rfloor }<\bigl(2.20711+o(1)\bigr)\binom{n}{\left\lfloor n/2\right\rfloor }$. \end{thm} The proof is inspired by the proof of Theorem \ref{thm:3levels2.207} (the same bound when restricted to 3 levels) as in \cite{diamond_228}, using the idea of grouping chains by the smallest set contained in $\mathcal{F}$ on a chain (as developed in \cite{griggs2012diamond} and \cite{diamond225}).\global\let\addcontentslinesave\addcontentsline\renewcommand{\addcontentsline}[3]{} \section{$\Lambda$-free families \textendash{} Proof of Lemma \ref{lem:lubell}\label{sec:invertedV}\addcontentslinesave{toc}{section}{\ref{sec:invertedV} Inverted V-free families \textendash{} Proof of Lemma \ref{lem:lubell}}} \subsection{Definitions and main lemma} \begin{defn} \global\let\addcontentsline\addcontentslinesave\label{def:functions}We define the following functions:% \begin{comment} For $x,c\in[0,\infty)$, \[ f(x,c)=\begin{cases} 1-x & \textnormal{if }\frac{1}{2}\leq x\textnormal{ or } c\leq x^{2}\\ x^{2}-2x+1-c+\sqrt{c} & \textnormal{if } x\leq\frac{1}{2}\textnormal{ and } x^{2}\leq c\leq\frac{1}{4}\\ x^{2}-2x+1.25 & \textnormal{if } x\leq\frac{1}{2}\textnormal{ and }\frac{1}{4}\leq c \end{cases} \] \begin{tikzwhenenabled} \begin{figure} \subfloat[Values of $f(x,c)$ plotted in $x$, for $c=0,0.0125,0.025,\ldots,0.25$. Note that the $1-x$ part of the plots coincide.]{\tikz \datavisualization [scientific axes, x axis={length=6cm,ticks and grid={step=0.25}}, y axis={length=8.5cm,ticks and grid={step=0.25}}, visualize as smooth line/.list={100,201.0,202.0,....0,220.0}] data [set=100,format=function] { var c : {0}; var x_ : interval [0:0.6] samples 2; func x = \value{x_}; func y = 1-\value{x_}; } data [format=function] { var i : {1,2,...,20}; func c = \value i/80; var x_ : interval [0:sqrt(\value c)] samples 11; func x = \value{x_}; func y = \value{x_}^2-2*\value{x_}+1-\value c+sqrt(\value c); func set = 200+\value i; };}\hspace*{\fill}\subfloat[Values of $f(x,c)$ plotted in $c$, for $x=0,0.05,0.01,\ldots,1$.]{\tikz \datavisualization [scientific axes, x axis={length=8cm,ticks and grid={step=0.25}}, y axis={length=10cm,ticks and grid={step=0.25}}, visualize as smooth line/.list={100.0,101.0,....0,120.0,200.0,201.0,....0,209.0,300.0,301.0,....0,309.0}] data [format=function] { var i : {0,1,...,20}; func x_ = \value i/20; var c : {0,((\value{x_}<0.5)?\value{x_}^2:1)}; func x = \value{c}; func y = 1-\value{x_}; func set = 100+\value i; } data [format=function] { var i : {0,1,...,9}; func x_ = \value i/20; var c : interval[\value{x_}^2:0.25] samples 20-2*\value i; func x = \value{c}; func y = \value{x_}^2-2*\value{x_}+1-\value c+sqrt(\value c); func set = 200+\value i; } data [format=function] { var i : {0,1,...,9}; func x_ = \value i/20; var c : interval[0.25:1] samples 2; func x = \value{c}; func y = \value{x_}^2-2*\value{x_}+1.25; func set = 300+\value i; };} \end{figure} \end{tikzwhenenabled} \end{comment} \begin{itemize} \item For $x\in[0,1],c\in[0,\infty)$, \[ f(x,c)=\begin{cases} 1-x+\left(\frac{1}{4\left(x-x^{2}\right)}-1\right)c & \textnormal{if } x\leq\frac{1}{2}\textnormal{ and } c<4\left(x-x^{2}\right)^{2}\\ x^{2}-2x+1-c+\sqrt{c} & \textnormal{if } x\leq\frac{1}{2}\tand4\left(x-x^{2}\right)^{2}\leq c\leq\frac{1}{4}\\ x^{2}-2x+1.25 & \textnormal{if } x\leq\frac{1}{2}\textnormal{ and }\frac{1}{4}\leq c\\ 1-x & \textnormal{if }\frac{1}{2}\leq x. \end{cases} \] \begin{tikzwhenenabled} \begin{figure} \subfloat[Values of $f(x,c)$ plotted in $x$, for $c=0,0.0125,0.025,\ldots,0.25$ (bottom to top). Note that the $x\geq0.5$ part of the plots coincide.]{\beginpgfgraphicnamed{diamond_free_families_figure1} \tikz \datavisualization [scientific axes, x axis={length=6cm,ticks and grid={step=0.25}}, y axis={length=8.5cm,ticks and grid={step=0.25}}, visualize as smooth line/.list={100,201.0,202.0,....0,220.0,301.0,302.0,....0,319.0}] data [set=100,format=function] { var c : {0}; var x_ : interval [0:0.6] samples 2; func x = \value{x_}; func y = 1-\value{x_}; } data [format=function] { var i : {1,2,...,20}; func c = \value i/80; var x_ : interval [0:(1-sqrt(1-2*sqrt(\value c)))/2] samples 11; func x = \value{x_}; func y = \value{x_}^2-2*\value{x_}+1-\value c+sqrt(\value c); func set = 200+\value i; } data [format=function] { var i : {1,2,...,19}; func c = \value i/80; var x_ : interval [(1-sqrt(1-2*sqrt(\value c)))/2:0.5] samples 11; func x = \value{x_}; func y = 1-\value{x_}+(1/(4*(\value{x_}-\value{x_}^2))-1)*\value c; func set = 300+\value i; }; \endpgfgraphicnamed}\hspace*{\fill}\subfloat[Values of $f(x,c)$ plotted in $c$, for $x=0,0.05,0.01,\ldots,1$ (top to bottom).]{\beginpgfgraphicnamed{diamond_free_families_figure2} \tikz \datavisualization [scientific axes, x axis={length=8cm,ticks and grid={step=0.25}}, y axis={length=10cm,ticks and grid={step=0.25}}, visualize as smooth line/.list={110.0,111.0,....0,120.0,200.0,201.0,....0,209.0,300.0,301.0,....0,309.0,401.0,402.0,....0,409.0}] data [format=function] { var i : {10,11,...,20}; func x_ = \value i/20; var c : {0,1)}; func x = \value{c}; func y = 1-\value{x_}; func set = 100+\value i; } data [format=function] { var i : {1,2,...,9}; func x_ = \value i/20; var c : {0,4*(\value{x_}-\value{x_}^2)^2}; func x = \value{c}; func y = 1-\value{x_}+(1/(4*(\value{x_}-\value{x_}^2))-1)*\value c; func set = 400+\value i; } data [format=function] { var i : {0,1,...,9}; func x_ = \value i/20; var c : interval[4*(\value{x_}-\value{x_}^2)^2:0.25] samples 20-2*\value i; func x = \value{c}; func y = \value{x_}^2-2*\value{x_}+1-\value c+sqrt(\value c); func set = 200+\value i; } data [format=function] { var i : {0,1,...,9}; func x_ = \value i/20; var c : interval[0.25:1] samples 2; func x = \value{c}; func y = \value{x_}^2-2*\value{x_}+1.25; func set = 300+\value i; }; \endpgfgraphicnamed} \end{figure} \end{tikzwhenenabled} \item For $x\in[0,1),c\in[0,\infty),a\in[0,1-x),\tilde{a}\in\left[0,\min\!\left(a,\frac{c}{x+a}\right)\right]$, \[ g(x,c,a,\tilde{a})=a+(1-x-a)f\!\left(x+a,\frac{c-\tilde{a}(x+a)}{1-x-a}\right)+2\tilde{a}(1-x-a). \] \item For $x\in[0,1),c\in[0,\infty),a\in[0,1-x),\tilde{a}\in\left[0,\min\!\left(a,\frac{c}{x+a}-x\right)\right]$, \[ h(x,c,a,\tilde{a})=a+(1-x-a)f\!\left(x+a,\frac{c-(x+\tilde{a})(x+a)}{1-x-a}\right)+2\tilde{a}(1-x-a)+x-3x(x+a). \] \end{itemize} \end{defn} \begin{lem} \label{lem:functions}The functions above satisfy the following conditions: \begin{enumerate} \item \label{enu:x=00003D0}For all $c\in[0,\infty)$, if $\tilde{c}=\min\!\left(c,\frac{1}{4}\right)$, then $f(0,c)=1-\tilde{c}+\sqrt{\tilde{c}}$. \item \label{enu:concave}$f(x,c)$ is concave and monotonously increasing in $c$, and monotonously decreasing in $x$. \item \label{enu:g}For all $x\in[0,1],c\in[0,\infty),a\in[0,1-x),\tilde{a}\in\left[0,\min\!\left(a,\frac{c}{x+a}\right)\right]:g(x,c,a,\tilde{a})\leq f(x,c)$. \item \label{enu:h}For all $x\in[0,1],c\in[0,\infty),a\in[0,1-x),\tilde{a}\in\left[0,\min\!\left(a,\frac{c}{x+a}-x\right)\right]:h(x,c,a,\tilde{a})\leq f(x,c)$. \item \label{enu:1-x}For all $c\in[0,\infty):1-x\leq f(x,c)$. \end{enumerate} \end{lem} We prove Lemma \ref{lem:functions} in Appendix \ref{sec:functions_proof}. Rather that proving Lemma \ref{lem:lubell} directly, we prove a strengthening of it \textendash{} Lemma \ref{lem:ind}. This strengthened version involves additional parameters, $X$ and $\mathcal{X}$, and their functions $x$, $\alpha$ and $\mu$, which we introduce in order to make the inductive proof possible. Lemma \ref{lem:lubell} is a special case of Lemma \ref{lem:ind} with $X=\mathcal{X}=\emptyset$. In the rest of Section \ref{sec:invertedV}, we prove Lemma \ref{lem:ind}. \begin{lem} \label{lem:ind}Let $\mathcal{F}\subseteq2^{[n]}$ be a $\Lambda$-free family which does not contain $\emptyset$, nor any set larger than $n-n'$ for some $n'\in\mathbf{N}$. Let us assume that we are given a ``forbidden'' set $X\subseteq[n]$, with $x=\frac{\left|X\right|}{n}$. Also, let $\mathcal{X}\subset2^{[n]}$ be a ``forbidden'' antichain in which each set contains exactly one element of $X$ (and may or may not be a singleton). Let us assume that the sets in $\mathcal{F}$ are disjoint from $X$, and unrelated to every set in $\mathcal{X}$. Let $\alpha=l(\mathcal{X})$, and let $\mu n!$ be the number of chains which start with an element of $X$ as a singleton, but do not contain any set in $\mathcal{X}$. Assume, furthermore, that there are $cn!$ MNM chains w.r.t.\ $\mathcal{F}$. Then $l(\mathcal{F})\leq f\!\left(x,c+\mu+\frac{1}{n'}\right)-(\alpha-\mu-x)+\frac{3}{n'}$. \end{lem} First we verify the base case of the induction. \begin{prop} \label{prop:indbase}Lemma \ref{lem:ind} holds for $n\leq n'$. \end{prop} \begin{proof} $\mathcal{F}=\emptyset$. $\mathcal{X}$ is an antichain, so, by the YMBL inequality, $\alpha\leq1$. By Lemma \ref{lem:functions} Point \ref{enu:concave}\ and Point \ref{enu:1-x}, $f\!\left(x,c+\mu+\frac{1}{n'}\right)-(\alpha-\mu-x)+\frac{3}{n'}\geq f(x,c+\mu)-(\alpha-\mu-x)\geq1-x-(\alpha-x)\geq0=l(\mathcal{F})$. \end{proof} From now on we assume $n'\leq n-1$. \begin{nameddef}[Notation] Let $A\subseteq[n]\setminus X$ be the set of elements of $[n]$ that appear as singletons in $\mathcal{F}$, and let $a=\frac{\left|A\right|}{n}$. Let $\mathcal{\mathcal{\mathcal{B}}}$ be the family of those sets in $\mathcal{F}$ which contain at least one element of $A$, but which are not singletons. Let $\beta$ be the Lubell function of $\mathcal{\mathcal{\mathcal{B}}}$. Let $\mathcal{C}$ be the family of those sets in $\mathcal{F}$ which only contain elements of $[n]\setminus X\setminus A$. Let $\tilde{A}=\bigl\{ e\in A:(\exists B\in\mathcal{\mathcal{\mathcal{B}}}:e\in B)\bigr\}$, and let $\tilde{a}=\frac{\left|\tilde{A}\right|}{n}$. Let $c_{0}n!$ be the number of chains that start with $\{e\}$ as a singleton for some $e\in\tilde{A}$, but do not contain any set from $\mathcal{\mathcal{\mathcal{B}}}$. Let $\nu n!$ be the number of chains that start with $\{e\}$ as a singleton for some $e\in\tilde{A}$, continue with an element of $[n]\setminus X\setminus A$ as the second element added to form the chain, yet do not contain any set from $\mathcal{\mathcal{\mathcal{B}}}$. \global\long\def\overline{1}{\overline{1}} \global\long\def\underline{1}{\underline{1}} Let $\overline{1}=\frac{n}{n-1}>1$ and $\underline{1}=\frac{(x+a)n-1}{(x+a)(n-1)}\leq1$. These correction factors will account for the difference from the asymptotic behavior. (They are both typically close to 1. If $x+a=0$, let $\underline{1}=1$; it is irrelevant as it will always be multiplied by $x+a$.)% \begin{comment} Unfortunately we have to be precise because we use induction. Asymptotically $n\approx n-1,n-1\approx n-2,\ldots,\left\lfloor n/2\right\rfloor +1\approx\left\lfloor n/2\right\rfloor $. \end{comment} \end{nameddef} \begin{table} \caption{Summary of notation} \centering{}% \begin{tabular}{cll|l} $X$ & \multicolumn{2}{>{\raggedright}p{0.7\columnwidth}|}{``Forbidden'' set (sets in $\mathcal{F}$ are disjoint from it \textendash{} parameter of Lemma \ref{lem:ind})} & $x=\left|X\right|/n$\tabularnewline $\mathcal{X}$ & \multicolumn{2}{>{\raggedright}p{0.7\columnwidth}|}{``Forbidden'' antichain (sets in $\mathcal{F}$ are unrelated to sets in it \textendash{} parameter of Lemma \ref{lem:ind})} & $\alpha=l(\mathcal{X})$\tabularnewline $\mu$ & \multicolumn{2}{>{\raggedright}p{0.7\columnwidth}}{$\#\left\{ \text{chains containing \ensuremath{\{d\}} for \ensuremath{d\in X} but no set from \ensuremath{\mathcal{X}}}\right\} /n!$} & \tabularnewline $c$ & \multicolumn{2}{>{\raggedright}p{0.7\columnwidth}}{$\#\{\text{MNM chains}\}/n!$ (parameter of Lemma \ref{lem:lubell} / Lemma \ref{lem:ind})} & \tabularnewline \hline $A$ & $\text{\ensuremath{\left\{ e\in[n]:\{e\}\in\mathcal{F}\right\} }}$ & \multirow{3}{*}{$\left.\vphantom{\mbox{\begin{tabular}{c} \ensuremath{\left\{ \right\} } \\ \ensuremath{\left\{ \right\} } \\ \ensuremath{\left\{ \right\} } \end{tabular}}}\right\} \left\{ \{e\}:e\in A\right\} \cup\mathcal{\mathcal{\mathcal{B}}}\cup\mathcal{C}=\mathcal{F}$} & $a=\left|A\right|/n$\tabularnewline $\mathcal{\mathcal{\mathcal{B}}}$ & $\left\{ B\in\mathcal{F}:(\left|B\right|\geq2,A\cap B\neq\emptyset)\right\} $ & & $\beta=l(\mathcal{\mathcal{\mathcal{B}}})$\tabularnewline $\mathcal{C}$ & $\left\{ C\in\mathcal{F}:C\subseteq[n]\setminus X\setminus A\right\} $ & & \tabularnewline $\tilde{A}$ & \multicolumn{2}{>{\raggedright}p{0.7\columnwidth}|}{$\bigl\{ e\in A:(\exists B\in\mathcal{\mathcal{\mathcal{B}}}:e\in B)\bigr\}$} & $\tilde{a}=\bigl|\tilde{A}\bigr|/n$\tabularnewline $\nu$ & \multicolumn{2}{>{\raggedright}p{0.7\columnwidth}}{$\#\bigl\{\text{chains containing }\ensuremath{\{e\}}\text{ and }\ensuremath{\{e,o\}}\text{ for }\ensuremath{e\in\tilde{A},o\in[n]\setminus X\setminus A}\text{ but no set from }\mathcal{\mathcal{\mathcal{B}}}\bigr\}/n!$} & \tabularnewline $c_{0}$ & \multicolumn{2}{>{\raggedright}p{0.7\columnwidth}}{$\#\bigl\{\text{chains containing \ensuremath{\{e\}} for \ensuremath{e\in\tilde{A}} but no set from \ensuremath{\mathcal{\mathcal{\mathcal{B}}}}}\bigr\}/n!$} & \tabularnewline \end{tabular} \end{table} \emph{Outline of the proof:} In Subsection \ref{subsec:On-the-structure}, we make some observations on the structure of $\mathcal{X}$ and $\mathcal{\mathcal{\mathcal{B}}}$. In Subsection \ref{subsec:Induction}, we will finish the proof by applying induction to the Boolean lattices $\bigl[\{o_{i}\},[n]\bigr]$ where $o_{i}\in[n]\setminus X\setminus A$. When applying Lemma \ref{lem:ind} by induction, we will use $X\cup A$ in the place of $X$, while sets from $\mathcal{X}$ and $\mathcal{\mathcal{\mathcal{B}}}$ will contribute to the family we use in the place of $\mathcal{X}$ (which we will denote by $\mathcal{X}'_{i}$). We know little about the parameters of each $\mathcal{X}'_{i}$, but we will be able to bound their sums. The relevant calculations are done in Subsection \ref{subsec:Chain-calculations}. \subsection{\label{subsec:On-the-structure}On the structure of $\protect\mathcal{X}$ and $\protect\mathcal{\mathcal{\mathcal{B}}}$} \begin{prop} \label{prop:Xform}Every $D\in\mathcal{X}$ is of the form $\{d,o_{1},\ldots,o_{k}\}$ with $d\in X,o_{1},\ldots,o_{k}\in[n]\setminus X\setminus A$ (where $k$ may be 0). \end{prop} \begin{proof} $D$ contains exactly one element of $X$ by definition. Let $e\in A$; then $e\notin D$ for otherwise $D$ and $\{e\}\in\mathcal{F}$ would be related. \end{proof} \begin{prop} \label{prop:Bform}Sets in $\mathcal{\mathcal{\mathcal{B}}}$ only contain one element of $A$. $\mathcal{\mathcal{\mathcal{B}}}$ is an antichain, and the sets in $\mathcal{\mathcal{\mathcal{B}}}$ are also unrelated to every set in $\mathcal{C}$. \end{prop} \begin{proof} If $e_{1}\in B\in\mathcal{\mathcal{\mathcal{B}}}$ with $e_{1}\in A$, and B was related to another set $S\in\mathcal{F}$, then $\{e_{1}\}$, $B$ and $S$ would form a $\Lambda$. This applies to any $S\in\mathcal{\mathcal{\mathcal{B}}}\cup\mathcal{C}$, as well as $S=\{e_{2}\}$ for any $e_{1}\neq e_{2}\in A$. \end{proof} \begin{prop} \label{prop:cbound}$\tilde{a}(x+a)\underline{1}\leq\tilde{a}(x+a)\underline{1}+\nu=c_{0}\leq c$, and thus $\tilde{a}\leq\frac{c}{(x+a)\underline{1}}$. \end{prop} \begin{proof} Any chain on which the singleton is $\{e\}$ and the second set is $\{e,d\}$ with $e\in\tilde{A}$ and $d\in X\cup A$ is always an MNM chain: $\{e,d\}$ and any set that contains it is forbidden from being in $\mathcal{\mathcal{\mathcal{B}}}$ either because it is not disjoint from $X$ (when $d\in X$), or because it would contain two elements of $A$ (when $d\in A$). The number of such chains is $\tilde{a}n\cdot(an+xn-1)\cdot(n-2)!=\tilde{a}(x+a)n!\underline{1}$. And out of the chains which start with $\{e\}$, and whose second set is $\{e,o\}$ with some $o\in[n]\setminus X\setminus A$, $\nu n!$ do not contain any set from $\mathcal{\mathcal{\mathcal{B}}}$. We have $c_{0}\leq c$ because a chain whose first set is $\{e\}$ for some $e\in\tilde{A}$, but does not contain any set from $\mathcal{\mathcal{\mathcal{B}}}$, is an MNM chain. \end{proof} For a family of sets $\mathcal{A}\subseteq2^{[n]}$, let $m(\mathcal{A})n!$ be the number of chains which start with an element of $X$ as a singleton and do not contain any set from $\mathcal{A}$. (For example, $m(\mathcal{X})=\mu$, and therefore $l(\mathcal{X})-m(\mathcal{X})=\alpha-\mu.)$ For a fixed $d\in X$, let $m_{d}(\mathcal{A})n!$ be the number of chains on which the singleton is $\{d\}$, and do not contain any element of $\mathcal{A}$. \begin{prop} \label{prop:uniform}For any $d\in X$, let $\mathcal{X}_{d}=\{D\in\mathcal{X}:d\in D\}$. We can assume without loss of generality that for any $d_{1},d_{2}\in X$, $\left\{ D\setminus\left\{ d_{1}\right\} :D\in\mathcal{X}_{d_{1}}\right\} =\left\{ D\setminus\left\{ d_{2}\right\} :D\in\mathcal{X}_{d_{2}}\right\} $. That is, if $\mathcal{X}$ does not satisfy this condition, we show a family $\hat{\mathcal{X}}$ which does, and also satisfies the conditions of Lemma \ref{lem:ind}'s statement (each set contains exactly one element of $X$, the sets are unrelated to each other and to every set in $\mathcal{F}$), and for which $f\!\left(x,c+m(\hat{\mathcal{X}})+\frac{1}{n'}\right)-\bigl(l(\hat{\mathcal{X}})-m(\hat{\mathcal{X}})-x\bigr)\leq f\!\left(x,c+\mu+\frac{1}{n'}\right)-(\alpha-\mu-x)$. \end{prop} \begin{proof} Let $d_{0}\in X$ be such that \begin{gather*} f\!\left(x,c+\left|X\right|m_{d_{0}}\!\left(\mathcal{X}_{d_{0}}\right)+\frac{1}{n'}\right)-\bigl(\left|X\right|l\!\left(\mathcal{X}_{d_{0}}\right)-\left|X\right|m_{d_{0}}\!\left(\mathcal{X}_{d_{0}}\right)-x\bigr)\\ =\min_{d\in X}\!\left[f\!\left(x,c+\left|X\right|m_{d}\!\left(\mathcal{X}_{d}\right)+\frac{1}{n'}\right)-\bigl(\left|X\right|l\!\left(\mathcal{X}_{d}\right)-\left|X\right|m_{d}\!\left(\mathcal{X}_{d}\right)-x\bigr)\right]. \end{gather*} Let $\hat{\mathcal{X}}=\bigl\{ D\setminus\{d_{0}\}\cup\{d\}:d\in X,D\in\mathcal{X}_{d_{0}}\bigr\}$. $\mathcal{X}=\bigsqcup_{d\in X}\mathcal{X}_{d}$, so $\alpha=\sum_{d\in X}l(\mathcal{X}_{d})$. It immediately follows from the definition of $\mathcal{X}_{d}$ that if a chain has $\{d\}$ as a singleton, and does not contain any set from $\mathcal{X}_{d}$, then it does not contain any set from $\mathcal{X}$. So $\mu=\sum_{d\in X}m_{d}(\mathcal{X}_{d})$. Similarly, $l(\hat{\mathcal{X}})=\left|X\right|l\!\left(\mathcal{X}_{d_{0}}\right)$ and $m(\hat{\mathcal{X}})=\left|X\right|m_{d_{0}}\!\left(\mathcal{X}_{d_{0}}\right)$. Since $f(x,c)$ is monotonously increasing and concave in $c$, using Jensen's inequality \begin{gather*} f\!\left(x,c+\left|X\right|m_{d_{0}}\!\left(\mathcal{X}_{d_{0}}\right)+\frac{1}{n'}\right)-\bigl(\left|X\right|l\!\left(\mathcal{X}_{d_{0}}\right)-\left|X\right|m_{d_{0}}\!\left(\mathcal{X}_{d_{0}}\right)-x\bigr)\\ \leq\frac{1}{\left|X\right|}\left[\sum_{d\in X}f\!\left(x,c+\left|X\right|m_{d}\!\left(\mathcal{X}_{d}\right)+\frac{1}{n'}\right)-\bigl(\left|X\right|\sum_{d\in X}l\!\left(\mathcal{X}_{d}\right)-\left|X\right|\sum_{d\in X}m_{d}\!\left(\mathcal{X}_{d}\right)-\left|X\right|x\bigr)\right]\\ \leq f\!\left(x,c+\mu+\frac{1}{n'}\right)-(\alpha-\mu-x). \end{gather*} Sets in $\hat{\mathcal{X}}$ contain exactly one element of $X$, and form an antichain. They are also unrelated to every set $S\in\mathcal{F}$: $S$ cannot contain any element of $X$, so it could only be related to a set in $\hat{\mathcal{X}}$ by being its subset. But $S$ must also be unrelated to every $D\in\mathcal{X}_{d_{0}}\subseteq\mathcal{X}$, so it cannot be a subset of $D\setminus\{d_{0}\}\cup\{d\}$ either. \end{proof} In fact we will only use the following simple corollary of Proposition \ref{prop:uniform}. In many parts of the rest of this section we will treat the two cases of the corollary below separately. \begin{cor} With the assumption of Proposition \ref{prop:uniform}, \begin{itemize} \item either $\mathcal{X}=\bigl\{\{d\}:d\in X\bigr\}$ (we refer to it as the \textbf{singletons case}), \item or $\mathcal{X}$ does not contain any singleton (referred to as the \textbf{no singleton case}). \end{itemize} \end{cor} \begin{proof} Let $d_{1}\in X$. (If $X=\emptyset$, both statements trivially hold.) If $\{d_{1}\}\in\mathcal{X}_{d_{1}}=\{D\in\mathcal{X}:d_{1}\in D\}$, then $\mathcal{X}_{d_{1}}=\left\{ \{d_{1}\}\right\} $, because sets in $\mathcal{X}_{d_{1}}$ are unrelated. So either $\mathcal{X}_{d_{1}}=\left\{ \{d_{1}\}\right\} $ or $\mathcal{X}_{d_{1}}$ does not contain any singleton, yielding the two cases above by Proposition \ref{prop:uniform}. \end{proof} \begin{rem*} The fact that sets in $\mathcal{X}$ contain an element of $X$ implies that sets in $\mathcal{F}$ do not contain sets in $\mathcal{X}$. Now, let us consider what restrictions are imposed on $\mathcal{F}$ by the fact that sets in $\mathcal{F}$ are not contained in the sets in $\mathcal{X}$, beyond the other conditions of Lemma \ref{lem:ind} (namely that all the sets in $\mathcal{F}$ are disjoint from $X$). In the singletons case, clearly there are no such additional restrictions. However, in the no singleton case, there are two additional restrictions that are not already implied by the set $X$: \begin{itemize} \item The union of singletons in $\mathcal{F}$, $A\subseteq[n]\setminus\bigcup\mathcal{X}$. \item Sets in $\mathcal{C}$ must not be contained in sets in $\mathcal{X}$. Clearly this imposes a restriction only if $\mathcal{X}$ contains sets bigger than 2. \end{itemize} \end{rem*} \begin{example} \label{exa:nosingleton}Let $C\subseteq[n]\setminus X$, and let $\mathcal{X}=\bigl\{\{d,o\}:d\in X,o\in C\bigr\}$. Then $\alpha=l(\mathcal{X})=\frac{2\cdot xn\cdot\left|C\right|\cdot(n-2)!}{n!}=2x\frac{\left|C\right|}{n}\overline{1}$, and $\mu=\frac{xn\cdot(xn+an-1)\cdot(n-2)!}{n!}=x(x+a)\underline{1}$. The only restriction on $\mathcal{F}$ that this $\mathcal{X}$ creates is that the union of singletons $A\subseteq[n]\setminus X\setminus C$. In other words, let us assume that $\alpha=l(\mathcal{X})=2x\gamma\overline{1}$ for $\gamma\in\mathbf{R}$ (without assuming that $\mathcal{X}$ is of the above form). Then it is possible that $a=\frac{\left|A\right|}{n}$ can be as big as $1-x-\gamma$ with $\mathcal{X}$ not creating any restrictions on $\mathcal{C}$ (depending on the actual structure of $\mathcal{X}$, namely, if it is made up of sets of size $2$ as above; then $\alpha=2x(1-x-a)\overline{1}$ and $\mu=x(x+a)\underline{1}$). But if $a>1-x-\gamma$, then $\alpha=2x\gamma\overline{1}$ implies that $\mathcal{X}$ contains sets bigger than 2, and thus it creates restrictions on $\mathcal{C}$. So, in the no singleton case, one way to understand the calculations that follow is to check them for $\mathcal{X}=\bigl\{\{d,o\}:d\in X,o\in C\bigr\}$; then check what happens if $x$, $c$ and $a$ are fixed, but $\mathcal{X}$ is changed. \end{example} \subsection{\label{subsec:Chain-calculations}Chain calculations} Now we estimate the numbers of certain types of chains, in preparation for applying induction. \begin{prop} \label{prop:Xchainbound}In the no singleton case, $\left(\alpha-x+\mu\right)n!$ chains start with $\{o\}$ for some $o\in[n]\setminus X\setminus A$, and contain a set from $\mathcal{X}$. \end{prop} \begin{proof} A total of $\alpha n!$ chains contain a set $D\in\mathcal{X}$. By Proposition \ref{prop:Xform}, the singleton on such a chain is either from $X$ or $[n]\setminus X\setminus A$. The number of chains which start with an element of $X$ as their singleton and do not contain a set from $\mathcal{X}$ is $\mu n!$, so the number of chains which contain a set from $\mathcal{X}$, and which start with an element of $X$, is $(x-\mu)n!$. On the rest, the singleton is from $[n]\setminus X\setminus A$. \end{proof} \begin{prop} \label{prop:Bchainbound}$\left(\beta-\tilde{a}(1-x-a)\overline{1}+\nu\right)n!$ chains start with $\{o\}$ for some $o\in[n]\setminus X\setminus A$, and contain a set from $\mathcal{\mathcal{\mathcal{B}}}$. \end{prop} \begin{proof} A total of $\beta n!$ chains contain a set from $\mathcal{\mathcal{\mathcal{B}}}$. A set in $\mathcal{\mathcal{\mathcal{B}}}$ is of the form $\{e,o_{1},\ldots,o_{k}\}$ with $e\in\tilde{A},o_{1},\ldots,o_{k}\in[n]\setminus X\setminus A,k\geq1$. A chain that contains a $B\in\mathcal{\mathcal{\mathcal{B}}}$, and does not start with $\{o\}$ for some $o\in[n]\setminus X\setminus A$, must start with an element of $\tilde{A}$, and continue with an element of $[n]\setminus X\setminus A$ as the second element added to form the chain. There are $\tilde{a}n\cdot(1-x-a)n\cdot(n-2)!=\tilde{a}(1-x-a)\overline{1} n!$ such chains, out of which $\nu n!$ do not contain any set from $\mathcal{\mathcal{\mathcal{B}}}$. So $\left(\tilde{a}(1-x-a)\overline{1}-\nu\right)n!$ chains contain a set from $\mathcal{\mathcal{\mathcal{B}}}$ and start with an element of $\tilde{A}$. The rest start with $\{o\}$ for some $o\in[n]\setminus X\setminus A$. \end{proof} \begin{prop} \label{prop:mubound}In the no singleton case, $\mu\geq x(x+a)\underline{1}$; and the number of chains of the form $\emptyset,\{d\},\{d,o\},\ldots$ with $d\in X,o\in[n]\setminus X\setminus A$, which do not contain any set from $\mathcal{X}$, is $\bigl(\mu-x(x+a)\underline{1}\bigr)n!$. \end{prop} \begin{proof} A total of $\mu n!$ chains start with an element of $X$ and do not contain any set from $\mathcal{X}$. The chains of the form $\emptyset,\{d_{1}\},\{d_{1},d_{2}\},\ldots$ with $d_{1}\in X,d_{2}\in X\cup A$ never contain a set from $\mathcal{X}$ when $\mathcal{X}$ contains no singleton. The number of these chains is $xn\cdot(xn+an-1)\cdot(n-2)!=\bigl(x(x+a)\underline{1}\bigr)n!$. For the rest, the second element added to form the chain is from $[n]\setminus X\setminus A$. \end{proof} \begin{nameddef}[Notation] Let $X'=X\cup A$. Let $\mathcal{Y}=\bigl\{\{d,o\}:d\in X,o\in[n]\setminus X\setminus A\bigr\}$, and let $\mathcal{Z}=\bigl\{\{e,o\}:e\in A\setminus\tilde{A},o\in[n]\setminus X\setminus A\bigr\}\bigr\}$. In the \emph{singletons case}, let $\mathcal{X}'=\mathcal{Y}\sqcup\mathcal{\mathcal{\mathcal{B}}}\sqcup\mathcal{Z}$. (Note that here and in the rest of the paper, $\sqcup$ stands for a union of sets which are pairwise disjoint.) In the \emph{no singleton case}, let $\mathcal{X}'=\mathcal{X}\sqcup\mathcal{\mathcal{\mathcal{B}}}\sqcup\mathcal{Z}$. \end{nameddef} \begin{prop} The three families which make up $\mathcal{X}'$ are indeed disjoint in each case, and their union forms an antichain. \end{prop} \begin{proof} $\mathcal{\mathcal{\mathcal{B}}}$ is an antichain by Proposition \ref{prop:Bform}; $\mathcal{X}$ is an antichain by definition; and $\mathcal{Y}$ and $\mathcal{Z}$ are antichains because both consist of size 2 sets only. Let $D=\{d,o_{1},\ldots,o_{k}\}\in\mathcal{X}$, $Y=\{d,o\}\in\mathcal{Y}$, $B=\{e_{1},p_{1},\ldots,p_{l}\}\in\mathcal{\mathcal{\mathcal{B}}}$ and $Z=\{e_{2},q\}\in\mathcal{Z}$ with $d\in X$, $e_{1}\in\tilde{A}$, $e_{2}\in A\setminus\tilde{A}$, $o_{i},o,p_{i},q\in[n]\setminus X\setminus A$, and $l\geq1$. $B$ is unrelated to $D$ by definition, and to $Y$ because $d\notin B$ and $\left|B\right|\geq2$. $Z$ is unrelated to $D$ and $Y$ because $d\notin Z$ and $e_{2}\notin D,Y$; $Z$ is unrelated to $B$ because $e_{1}\notin Z$ and $e_{2}\notin B$. \end{proof} \begin{prop} \label{prop:indcondition}Sets in $\mathcal{C}$ are disjoint from $X'$, and they are unrelated to every set in $\mathcal{X}'$ (in both cases). \end{prop} \begin{proof} For every $C\in\mathcal{C}$, $C\subseteq[n]\setminus X'$ and it is unrelated to every set in $\mathcal{X}$ by definition. $C$ is unrelated to every set in $\mathcal{\mathcal{\mathcal{B}}}$ by Proposition \ref{prop:Bform}. It also cannot be a superset of a $Y\in\mathcal{Y}$ or a $Z\in\mathcal{Z}$, since those contain an element of $X$ or $A$; neither a proper subset of $Y$ or $Z$ because $\left|Y\right|=\left|Z\right|=2\leq\left|C\right|$. \end{proof} \begin{prop} \label{prop:indchainbound}The number of chains that start with an element of $[n]\setminus X'$ and contain a set from $\mathcal{X}'$ is \begin{itemize} \item at least $\left[x(1-x-a)\overline{1}+\left(\beta-\tilde{a}(1-x-a)\overline{1}+\nu\right)+(1-x-a)\left(a-\tilde{a}\right)\overline{1}\right]n!$ in the \emph{singletons case}, and \item at least $\left[\left(\alpha-x+\mu\right)+\left(\beta-\tilde{a}(1-x-a)\overline{1}+\nu\right)+(1-x-a)\left(a-\tilde{a}\right)\overline{1}\right]n!$ in the \emph{no singleton case}. \end{itemize} \end{prop} \begin{proof} The number of chains on which the singleton is $\{o\}$ with $o\in[n]\setminus X'=[n]\setminus X\setminus A$, and the second set is $\{o,d\}\in\mathcal{Y}$ with $d\in X$, is $xn\cdot(1-x-a)n\cdot(n-2)!=x(1-x-a)\overline{1} n!$. The number of chains on which the singleton is $\{o\}$, and the second set is $\{o,e\}\in\mathcal{Z}$ with $e\in A\setminus\tilde{A}$, is $(1-x-a)n\cdot(a-\tilde{a})n\cdot(n-2)!=(1-x-a)\left(a-\tilde{a}\right)\overline{1} n!$. The rest follows from Proposition \ref{prop:Xchainbound} and Proposition \ref{prop:Bchainbound}. \end{proof} \begin{prop} \label{prop:indmubound}The number of chains on which the singleton is $\{o\}$ with $o\in[n]\setminus X'$, the second set is $\{o,d\}$ with $d\in X'=X\cup A$, and which do not contain any set from $\mathcal{X}'$, is \begin{itemize} \item $\nu n!$ in the \emph{singletons case}, and \item $\bigl(\mu-x(x+a)\underline{1}+\nu\bigr)n!$ in the \emph{no singleton case}. \end{itemize} \end{prop} \begin{proof} Let $\mathcal{A}=\emptyset,A_{1},A_{2},\ldots,A_{n-1},[n]$ be a chain with $\emptyset\subset A_{1}\subset A_{2}\subset\ldots\subset A_{n-1}\subset[n]$. Let $\varphi(\mathcal{A})$ be the chain $\emptyset,A_{2}\setminus A_{1},A_{2},A_{3},\ldots,A_{n-1},[n]$. (In other words, in the order in which elements of $[n]$ are added to form the chain, the first two are swapped.) $\varphi$ is a bijection. It is easy to check that $\mathcal{X}'$ does not contain singletons. $\varphi$ is a bijection between chains of the form $\emptyset,\{o\},\{o,d\},\ldots$ containing no set from $\mathcal{X}'$, and chains of the form $\emptyset,\{d\},\{o,d\},\ldots$ containing no set from $\mathcal{X}'$, with $o\in[n]\setminus X'$ and $d\in X\cup A$. Below we classify the chains $\emptyset,\{d\},\{o,d\},\ldots$ based on what set $d$ belongs to and count them separately. \begin{itemize} \item For $d\in X$, $\{o,d\}\in\mathcal{Y}$ in the singletons case. In the no singleton case, $\bigl(\mu-x(x+a)\underline{1}\bigr)n!$ chains of the form $\emptyset,\{d\},\{o,d\},\ldots$ contain no set from $\mathcal{X}$ by Proposition \ref{prop:mubound}; these chains also contain no set from $\mathcal{\mathcal{\mathcal{B}}}$ or $\mathcal{Z}$, since sets from those do not contain any element of $X$. \item For $d\in\tilde{A}$, the number of chains of this form which contain no set from $\mathcal{\mathcal{\mathcal{B}}}$ is $\nu n!$; these chains also contain no set from $\mathcal{X},\mathcal{Y}$ or $\mathcal{Z}$, since sets from those contain no element of $\tilde{A}$. \item For $d\in A\setminus\tilde{A}$, $\{o,d\}\in\mathcal{Z}$. \end{itemize} Summing these cases, we get the statement of the proposition. \end{proof} \subsection{\label{subsec:Induction}Inductive step} \begin{nameddef}[Notation] Using standard notation for intervals, let $\left[A,[n]\right]$ denote the Boolean lattice $\left\{ S\subseteq[n]:A\subseteq S\right\} $. Let $[n]\setminus X'=[n]\setminus X\setminus A=\{o_{1},o_{2},\ldots,o_{(1-x-a)n}\}$; and for a family of sets $\mathcal{A}$, let $\mathcal{A}-o_{i}=\left\{ S\setminus\{o_{i}\}:S\in\mathcal{A}\right\} $. Let $\mathcal{C}'_{i}=\left(\mathcal{C}\cap\bigl[\{o_{i}\},[n]\bigr]\right)-o_{i}$, and $\mathcal{X}'_{i}=\left(\mathcal{X}'\cap\bigl[\{o_{i}\},[n]\bigr]\right)-o_{i}$. Let $\alpha'_{i}=l(n-1,\mathcal{X}'_{i})$. (Here the Lubell function on the Boolean lattice $2^{[n]\setminus\{o_{i}\}}$ of order $n-1$ is used.) \end{nameddef} $\mathcal{C}'_{i}\subseteq2^{[n]\setminus\{o_{i}\}}$ is a $\Lambda$-free family which does not contain $\emptyset$ (since $o_{i}\notin A$, so $\{o_{i}\}\notin\mathcal{F}$), nor any set larger than $n-1-n'$. Sets in $\mathcal{C}'_{i}$ are disjoint from $X'$, and are unrelated to sets in $\mathcal{X}'_{i}$ by Proposition \ref{prop:indcondition}. Moreover, every set in $\mathcal{X}'_{i}$ contains exactly one element of $X'$. Therefore, the conditions of Lemma \ref{lem:ind} are satisfied for the family $\mathcal{C}'_{i}\subseteq2^{[n]\setminus\{o_{i}\}}$ where the corresponding ``forbidden'' set is $X'\subseteq[n]\setminus\{o_{i}\}$, with $\frac{\left|X'\right|}{n-1}=(x+a)\overline{1}$ and the corresponding ``forbidden'' antichain is $\mathcal{X}'_{i}$. Since $\mathcal{X}'_{i}$ is an antichain, $\alpha'_{i}(n-1)!$ is the number of chains in $2^{[n]\setminus\{o_{i}\}}$ that contain a set from $\mathcal{X}'_{i}$. Chains of $2^{[n]\setminus\{o_{i}\}}$ correspond to chains of $2^{[n]}$ that start with $\{o_{i}\}$. So by Proposition \ref{prop:indchainbound}, in the \emph{singletons case} \[ \sum_{i=1}^{(1-x-a)n}\alpha'_{i}\geq\left[x(1-x-a)\overline{1}+\left(\beta-\tilde{a}(1-x-a)\overline{1}+\nu\right)+(1-x-a)\left(a-\tilde{a}\right)\overline{1}\right]n, \] and in the \emph{no singleton case} \[ \sum_{i=1}^{(1-x-a)n}\alpha'_{i}\geq\left[\left(\alpha-x+\mu\right)+\left(\beta-\tilde{a}(1-x-a)\overline{1}+\nu\right)+(1-x-a)\left(a-\tilde{a}\right)\overline{1}\right]n. \] Let $\mu'_{i}(n-1)!$ be the number of chains in the Boolean lattice $2^{[n]\setminus\{o_{i}\}}$ which start with an element of $X'$ as a singleton, but do not contain any set from $\mathcal{X}'_{i}$. By Proposition \ref{prop:indmubound}, in the \emph{singletons case } \[ \sum_{i=1}^{(1-x-a)n}\mu'_{i}=\nu n, \] and in the \emph{no singleton case} \[ \sum_{i=1}^{(1-x-a)n}\mu'_{i}=\bigl(\mu-x(x+a)\underline{1}+\nu\bigr)n. \] Let $c'_{i}(n-1)!$ be the number of MNM chains w.r.t.\ $\mathcal{C}'_{i}$ in $2^{[n]\setminus\{o_{i}\}}$. The corresponding $2^{[n]}$-chains, starting with $\{o_{i}\}$, are MNM chains w.r.t.\ $\mathcal{F}$. The total number of MNM chains w.r.t.\ $\mathcal{F}$ is $cn!$, out of which $c_{0}n!$ start with an element of $A$ as a singleton. By Proposition \ref{prop:cbound}, \[ \sum_{i=1}^{(1-x-a)n}c'_{i}=(c-c_{0})n=\left(c-\tilde{a}(x+a)\underline{1}-\nu\right)n. \] The following two examples are typical cases where, in the induction step for the $\mathcal{C}'_{i}$'s, we will get the singletons case and the no singleton case respectively. \begin{example} Let $X=\mathcal{X}=\emptyset$ and $\mathcal{\mathcal{\mathcal{B}}}=\bigl\{\{e,o\}:e\in A,o\in[n]\setminus A\bigr\}$. Then $\tilde{A}=A$, $\beta=2a(1-a)\overline{1}$, and $\nu=0$. $X'=A$, and $\mathcal{X}'_{i}=\bigl\{\{e\}:e\in A\bigr\}$. $\sum_{i=1}^{(1-a)n}\alpha'_{i}=a(1-a)\overline{1} n$, $\alpha'_{i}=a\overline{1}=\frac{\left|X'\right|}{n-1}$ and $\mu'_{i}=0$. $\sum_{i=1}^{(1-x-a)n}c'_{i}=\left(c-a^{2}\right)n$ and the average of the $c'_{i}$'s is $\frac{c-a^{2}}{1-a}$. \end{example} \begin{example} Let $X=\mathcal{X}=\emptyset$ and $\mathcal{\mathcal{\mathcal{B}}}\subseteq\hat{\mathcal{\mathcal{\mathcal{B}}}}:=\bigl\{\{e,o_{1},o_{2}\}:e\in A,o_{1},o_{2}\in[n]\setminus A\bigr\}$. Then $X'=A$, and $\mathcal{X}'_{i}\subseteq\bigl\{\{e,o\}:e\in A,o\in[n]\setminus A\setminus\{o_{i}\}\bigr\}$. Chains on $2^{[n]}$ of the form $\emptyset,\{e_{1}\},\{e_{1},o\},\{e_{1},o,e_{2}\},\ldots$ do not intersect $\mathcal{\mathcal{\mathcal{B}}}$. So $\sum_{i=1}^{(1-a)n}\mu'_{i}=\nu\geq a^{2}(1-a)\overline{1}^{2}\frac{\overline{1} a(n-1)-1}{\overline{1} a(n-2)}n$ (greater if $\mathcal{\mathcal{\mathcal{B}}}\subsetneqq\hat{\mathcal{\mathcal{\mathcal{B}}}}$), and the average of the $\mu'_{i}$'s is $\geq a^{2}\overline{1}^{2}\frac{\overline{1} a(n-1)-1}{\overline{1} a(n-2)}={x'}^{2}\frac{x'(n-1)-1}{x'((n-1)-1)}$ where $x'=\frac{\left|X'\right|}{n-1}$. In the case of $\mathcal{\mathcal{\mathcal{B}}}=\hat{\mathcal{\mathcal{\mathcal{B}}}}$, the size of the sets in $\mathcal{C}$ is at least 3, and the size of those in $\mathcal{C}'_{i}$ is at least 2. \begin{comment} TODO: EXPAND ON EXAMPLES? Yea, maybe expand or remove them althogether... \end{comment} \end{example} \begin{prop} \label{prop:lubell_induction} \[ l(\mathcal{C})=\frac{1}{n}\sum_{i=1}^{(1-x-a)n}l(n-1,\mathcal{C}'_{i})\quad\textnormal{and}\quad l(\mathcal{F})=a+\beta+l(\mathcal{C})=a+\beta+\frac{1}{n}\sum_{i=1}^{(1-x-a)n}l(n-1,\mathcal{C}'_{i}). \] (Still understanding the one parameter version $l(\mathcal{F})$ as $l(n,\mathcal{F})$ for a family $\mathcal{F}\subseteq2^{[n]}$.) \end{prop} \begin{proof} Every chain in the Boolean lattice $2^{[n]}$ that intersects $\mathcal{C}$ has an $\{o_{i}\}$ as a singleton, and thus corresponds to a chain in the Boolean lattice $\left[\{o_{i}\},[n]\right]-o_{i}$ that intersects $\mathcal{C}'_{i}$. \[ l(\mathcal{C})=\frac{1}{n!}\sum_{\mathcal{H}\text{ is a chain in }2^{[n]}}\left|\mathcal{H}\cap\mathcal{C}\right|=\frac{1}{n!}\sum_{i=1}^{(1-x-a)n}\sum_{\mathcal{H}\text{ is a chain in }\left[\{o_{i}\},[n]\right]-o_{i}}\left|\mathcal{H}\cap\mathcal{C}'_{i}\right|=\frac{1}{n}\sum_{i=1}^{(1-x-a)n}l(n-1,\mathcal{C}'_{i}). \] Let $\mathcal{A}=\bigl\{\{e\}:e\in A\bigr\}$. Then $\mathcal{F}=\mathcal{A}\sqcup\mathcal{\mathcal{\mathcal{B}}}\sqcup\mathcal{C}$. So $l(\mathcal{F})=l(\mathcal{A})+l(\mathcal{\mathcal{\mathcal{B}}})+l(\mathcal{C})$ with $l(\mathcal{A})=\frac{\left|A\right|}{n}=a$ and $l(\mathcal{\mathcal{\mathcal{B}}})=\beta$. \end{proof} We now prove Lemma \ref{lem:ind} (and thus Lemma \ref{lem:lubell}) using induction on $n$. According to Proposition \ref{prop:indbase}, Lemma \ref{lem:ind} holds for $n\leq n'$. By induction and Lemma \ref{lem:functions} Point \ref{enu:concave}, \begin{align*} l(n-1,\mathcal{C}'_{i}) & \leq f\!\left((x+a)\overline{1},c'_{i}+\mu'_{i}+\frac{1}{n'}\right)-\left(\alpha'_{i}-\mu'_{i}-(x+a)\overline{1}\right)+\frac{3}{n'}\\ & \leq f\!\left(x+a,c'_{i}+\mu'_{i}+\frac{1}{n'}\right)-\left(\alpha'_{i}-\mu'_{i}-(x+a)\overline{1}\right)+\frac{3}{n'}. \end{align*} So, by Proposition \ref{prop:lubell_induction}, we have \begin{align*} l(\mathcal{C}) & =\frac{1}{n}\sum_{i=1}^{(1-x-a)n}l(n-1,\mathcal{C}'_{i})\leq\frac{1}{n}\sum_{i=1}^{(1-x-a)n}f\!\left(x+a,c'_{i}+\mu'_{i}+\frac{1}{n'}\right)\\ & {}-\frac{1}{n}\left(\sum_{i=1}^{(1-x-a)n}\alpha'_{i}-\sum_{i=1}^{(1-x-a)n}\mu'_{i}-\sum_{i=1}^{(1-x-a)n}(x+a)\overline{1}\right)+\frac{1}{n}\cdot\frac{3(1-x-a)n}{n'}. \end{align*} We handle the case of $1-x-a=0$ separately. If $1-x-a=0$, $A=[n]\setminus X$ and, since any non-singleton $\{e_{1},e_{2},\ldots\}\in\mathcal{F}$ would form a $\Lambda$ with the singletons $\{e_{1}\},\{e_{2}\}\in\mathcal{F}$, we have $\mathcal{F}=\mathcal{A}$ and $l(\mathcal{F})=a=1-x$. This is only possible in the singletons case, since a non-singleton in $\mathcal{X}$ would have to contain elements of $[n]\setminus X\setminus A$. In the singletons case $\alpha=x$ and $\mu=0$, so $l(\mathcal{F})=1-x\leq f(x,c)\leq f\!\left(x,c+\mu+\frac{1}{n'}\right)-(\alpha-\mu-x)+\frac{3}{n'}$ by Lemma \ref{lem:functions} Point \ref{enu:1-x}. From now on, we assume that $1-x-a>0$. Since $f$ is concave in $c$, by Jensen's inequality, and since $f$ is monotonously decreasing in $x$, \begin{align*} l(\mathcal{C}) & \leq(1-x-a)f\!\left(x+a,\frac{\sum_{i=1}^{(1-x-a)n}c'_{i}+\sum_{i=1}^{(1-x-a)n}\mu'_{i}}{(1-x-a)n}+\frac{1}{n'}\right)\\ & {}-\left(\frac{1}{n}\sum_{i=1}^{(1-x-a)n}\alpha'_{i}-\frac{1}{n}\sum_{i=1}^{(1-x-a)n}\mu'_{i}-(1-x-a)(x+a)\overline{1}\right)+\frac{3(1-x-a)}{n'}. \end{align*} \emph{Correction term calculations} that we will use later (assuming $n'\leq n-1$):\emph{ } \begin{equation} \begin{gathered}(1-\underline{1})(x+\tilde{a})(x+a)+\frac{1-x-a}{n'}=\frac{(x+\tilde{a})(1-x-a)}{n-1}+\frac{1-x-a}{n'}\\ \leq\frac{(1+x+\tilde{a})(1-x-a)}{n'}\leq\frac{1-(x+a)^{2}}{n'}\leq\frac{1}{n'}. \end{gathered} \label{eq:correction_nosing_c} \end{equation} \emph{ \begin{equation} (1-\underline{1})\tilde{a}(x+a)+\frac{1-x-a}{n'}\leq(1-\underline{1})(x+\tilde{a})(x+a)+\frac{1-x-a}{n'}\leq\frac{1}{n'}.\label{eq:correction_singletons_c} \end{equation} \begin{equation} \begin{gathered}2\tilde{a}(1-x-a)(\overline{1}-1)+2(\overline{1}-1)x-\left(2(\overline{1}-1)+(\underline{1}-1)\right)x(x+a)+\frac{3(1-x-a)}{n'}\\ \leq\frac{2a+3x}{n-1}+\frac{3(1-x-a)}{n'}\leq\frac{3}{n'}. \end{gathered} \label{eq:correction_nosing} \end{equation} \begin{equation} 2\tilde{a}(1-x-a)(\overline{1}-1)+\frac{3(1-x-a)}{n'}\leq\frac{2a}{n-1}+\frac{3(1-x-a)}{n'}\leq\frac{3}{n'}.\label{eq:correction_singletons} \end{equation} } \emph{In the singletons case:} \begin{align*} l(\mathcal{F}) & \leq a+\beta+(1-x-a)f\!\left(x+a,\frac{\left(c-\tilde{a}(x+a)\underline{1}-\nu\right)+\nu}{1-x-a}+\frac{1}{n'}\right)\\ & {}-\Bigl(\left[x(1-x-a)\overline{1}+\left(\beta-\tilde{a}(1-x-a)\overline{1}+\nu\right)+(1-x-a)\left(a-\tilde{a}\right)\overline{1}\right]\\ & {}-\nu-(1-x-a)(x+a)\overline{1}\Bigr)+\frac{3(1-x-a)}{n'}\displaybreak[0]\\ & =a+(1-x-a)f\!\left(x+a,\frac{c-\tilde{a}(x+a)\underline{1}}{1-x-a}+\frac{1}{n'}\right)+2\tilde{a}(1-x-a)\overline{1}+\frac{3(1-x-a)}{n'}\displaybreak[0]\\ & =a+(1-x-a)f\!\left(x+a,\frac{c-\tilde{a}(x+a)+(1-\underline{1})\tilde{a}(x+a)+\frac{1-x-a}{n'}}{1-x-a}\right)+2\tilde{a}(1-x-a)\\ & {}+2\tilde{a}(1-x-a)(\overline{1}-1)+\frac{3(1-x-a)}{n'}. \end{align*} By Lemma \ref{lem:functions} Point \ref{enu:concave}~and Point \ref{enu:g}, and \eqref{eq:correction_singletons_c} and \eqref{eq:correction_singletons} in the Correction term calculations (note that in this case $\alpha=x$ and $\mu=0$), \[ l(\mathcal{F})\leq g\!\left(x,c+\frac{1}{n'},a,\tilde{a}\right)+\frac{3}{n'}\leq f\!\left(x,c+\frac{1}{n'}\right)+\frac{3}{n'}=f\!\left(x,c+\mu+\frac{1}{n'}\right)-(\alpha-\mu-x)+\frac{3}{n'}. \] (Note that $\tilde{a}\leq\frac{c}{(x+a)\underline{1}}$, so $0\leq\frac{c-\tilde{a}(x+a)\underline{1}}{1-x-a}\leq\frac{c+\frac{1}{n'}-\tilde{a}(x+a)}{1-x-a}$, and $\tilde{a}\leq\frac{c+\frac{1}{n'}}{x+a}$.) \emph{In the no singleton case:} \begin{align*} l(\mathcal{F}) & \leq a+\beta+(1-x-a)f\!\left(x+a,\frac{\left(c-\tilde{a}(x+a)\underline{1}-\nu\right)+\mu-x(x+a)\underline{1}+\nu}{1-x-a}+\frac{1}{n'}\right)\\ & {}-\Bigl(\left[\left(\alpha-x+\mu\mathnote{here}\right)+\left(\beta-\tilde{a}(1-x-a)\overline{1}+\nu\right)+(1-x-a)\left(a-\tilde{a}\right)\overline{1}\right]\\ & {}-\left[\mu-x(x+a)\underline{1}+\nu\right]-(1-x-a)(x+a)\overline{1}\Bigr)+\frac{3(1-x-a)}{n'}\displaybreak[0]\\ & =a+(1-x-a)f\!\left(x+a,\frac{c+\mu-(x+\tilde{a})(x+a)\underline{1}}{1-x-a}+\frac{1}{n'}\right)-(\alpha-x(x+a)\underline{1}\mathnote{here}-x)\\ & {}+2\tilde{a}(1-x-a)\overline{1}+(2\cdot\overline{1}-1)x-(2\cdot\overline{1}+\underline{1})x(x+a)+\frac{3(1-x-a)}{n'}\displaybreak[0]\\ & =a+(1-x-a)f\!\left(x+a,\frac{c+\mu-(x+\tilde{a})(x+a)+(1-\underline{1})(x+\tilde{a})(x+a)+\frac{1-x-a}{n'}}{1-x-a}\right)\\ & {}+2\tilde{a}(1-x-a)+x-3x(x+a)-(\alpha-x(x+a)\underline{1}\mathnote{here}-x)\\ & {}+2\tilde{a}(1-x-a)(\overline{1}-1)+2(\overline{1}-1)x-\left(2(\overline{1}-1)+(\underline{1}-1)\right)x(x+a)+\frac{3(1-x-a)}{n'}. \end{align*} By Lemma \ref{lem:functions} Point \ref{enu:concave}~and Point \ref{enu:h}, Proposition \ref{prop:mubound}. and \eqref{eq:correction_nosing_c} and \eqref{eq:correction_nosing} in the Correction term calculations, \[ l(\mathcal{F})\leq h\!\left(x,c+\mu+\frac{1}{n'},a,\tilde{a}\right)-(\alpha-\mu-x)+\frac{3}{n'}\leq f\!\left(x,c+\mu+\frac{1}{n'}\right)-(\alpha-\mu-x)+\frac{3}{n'}. \] (Note that $\tilde{a}\leq\frac{c}{(x+a)\underline{1}}$, so $0\leq\frac{c-\tilde{a}(x+a)\underline{1}}{1-x-a}\leq\frac{c+\mu+\frac{1}{n'}-(x+\tilde{a})(x+a)}{1-x-a}$, and $\tilde{a}\leq\frac{c+\mu+\frac{1}{n'}}{x+a}-x$.) \section{Diamond-free families \textendash{} Proof of Theorem \ref{thm:main}\label{sec:diamond}} Let $\mathcal{F}$ be a diamond-free family on $2^{[n]}$. We cite Lemma 1 from \cite{diamond_228}: \begin{lem}[Axenovich, Manske, Martin \cite{diamond_228}] \[ \sum_{\substack{k\in\{0,1,\ldots,n\}\\ \left|k-n/2\right|\geq n^{2/3} } }\binom{n}{k}\leq2^{n-\Omega(n^{1/3})}=2^{-\Omega(n^{1/3})}\binom{n}{\left\lfloor n/2\right\rfloor }. \] \end{lem} By this lemma, the number of sets in $\mathcal{F}$ in the top and bottom $n':=n/2-n^{\frac{2}{3}}$ levels is $o(1)\binom{n}{\left\lfloor n/2\right\rfloor }$, so, since we are bounding the cardinality of $\mathcal{F}$, we may assume that those levels do not contain any set from $\mathcal{F}$. \begin{nameddef}[Notation] For $c\in[0,1]$, let $\tilde{c}=\min\!\left(c,\frac{1}{4}\right)$, and let $f(c)=1-\tilde{c}+\sqrt{\tilde{c}}$. (This is equal to $f(0,c)$ as defined in Definition \ref{def:functions}.) For $A\in\mathcal{F}$, recall that $\left[A,[n]\right]$ denotes the Boolean lattice $\left\{ S\subseteq[n]:A\subseteq S\right\} $. A chain of this lattice is of the form $A\subset A_{\left|A\right|+1}\subset A_{\left|A\right|+2}\subset\ldots\subset A_{n-1}\subset[n]$. (When saying just ``chain'', we continue to mean a maximal chain in the Boolean lattice $2^{[n]}$.) Let \[ c(A)=\frac{1}{\left(n-\left|A\right|\right)!}\#\left\{ \begin{gathered}\mathcal{C}\textnormal{ is a chain in \ensuremath{\left[A,[n]\right]}}:\\ \mathcal{C}\textnormal{ is MNM w.r.t.\ }\mathcal{F}\cap\left[A,[n]\right] \end{gathered} \right\} . \] Further, we can assume without loss of generality that \[ C:=\frac{1}{n!}\#\left\{ \begin{gathered}\mathcal{C}\textnormal{ is a chain: \ensuremath{\mathcal{C}\cap\mathcal{F}=\emptyset} or }\\ \min(\mathcal{C}\cap\mathcal{F})\textnormal{ is not minimal in }\mathcal{F} \end{gathered} \right\} \geq\frac{1}{n!}\#\left\{ \begin{gathered}\mathcal{C}\textnormal{ is a chain: \ensuremath{\mathcal{C}\cap\mathcal{F}=\emptyset} or }\\ \mathcal{C}\textnormal{ is MNM w.r.t.\ }\mathcal{F} \end{gathered} \right\} . \] \end{nameddef} (If this does not hold, we can replace $\mathcal{F}$ with $\left\{ [n]\setminus A:A\in\mathcal{F}\right\} $: this family is diamond-free, has the same cardinality, and the opposite inequality holds.) Clearly $C\geq\dfrac{1}{n!}\#\left\{ \begin{gathered}\mathcal{C}\textnormal{ is a chain:}\\ \mathcal{C}\textnormal{ is MNM w.r.t.\ }\mathcal{F} \end{gathered} \right\} $. \[ l(\mathcal{F})=\frac{1}{n!}\sum_{\mathcal{C}\textnormal{ is a chain}}\#(\mathcal{C}\cap\mathcal{F})=\frac{1}{n!}\sum_{A\in\mathcal{F}}\sum_{\substack{\mathcal{C}\textnormal{ is a chain}\\ A=\min(\mathcal{C}\cap\mathcal{F}) } }\#(\mathcal{C}\cap\mathcal{F}), \] since each chain $\mathcal{C}$ will be counted when $A=\min(\mathcal{C}\cap\mathcal{F})$ \textendash{} except if $\mathcal{C}\cap\mathcal{F}=\emptyset$, but then $\#(\mathcal{C}\cap\mathcal{F})=0$. Continuing, \[ l(\mathcal{F})=\frac{1}{n!}\sum_{\substack{A\in\mathcal{F}\\ \exists\textnormal{ a chain \ensuremath{\mathcal{C}}: }A=\min(\mathcal{C}\cap\mathcal{F}) } }\#\left\{ \begin{gathered}\mathcal{C}\textnormal{ is a chain containing }A:\\ A=\min(\mathcal{C}\cap\mathcal{F}) \end{gathered} \right\} \frac{{\displaystyle \sum_{\substack{\mathcal{C}\textnormal{ is a chain}\\ A=\min(\mathcal{C}\cap\mathcal{F}) } }\#(\mathcal{C}\cap\mathcal{F}})}{\#\left\{ \begin{gathered}\mathcal{C}\textnormal{ is a chain containing }A:\\ A=\min(\mathcal{C}\cap\mathcal{F}) \end{gathered} \right\} }. \] Each chain on $\left[A,[n]\right]$ can be extended to a full $2^{[n]}$-chain in $\left|A\right|!$ ways. Furthermore, the Boolean lattice $\left[A,[n]\right]$ can be made equivalent to the Boolean lattice $2^{[n]\setminus A}$ by subtracting $A$ from each set; for $\mathcal{A}\subseteq\left[A,[n]\right]$, we denote $\mathcal{A}-A=\left\{ S\setminus A:S\in\mathcal{A}\right\} $. If $A=\min(\mathcal{C}\cap\mathcal{F})$, $\#(\mathcal{C}\cap\mathcal{F})=\#(\mathcal{C}\cap\left[A,[n]\right]\cap\mathcal{F})$. If $A$ is minimal in $\mathcal{F}$ (that is, on every chain), \begin{gather*} \frac{{\displaystyle \sum_{\substack{\mathcal{C}\textnormal{ is a chain}\\ A=\min(\mathcal{C}\cap\mathcal{F}) } }\#(\mathcal{C}\cap\mathcal{F})}}{\#\left\{ \begin{gathered}\mathcal{C}\textnormal{ is a chain containing }A:\\ A=\min(\mathcal{C}\cap\mathcal{F}) \end{gathered} \right\} }=\frac{\left|A\right|!{\displaystyle \sum_{\mathcal{C}\textnormal{ is a chain in }\left[A,[n]\right]}\#(\mathcal{C}\cap\mathcal{F}\cap\left[A,[n]\right])}}{\left|A\right|!\left(n-\left|A\right|\right)!}\\ =\frac{{\displaystyle \sum_{\mathcal{C}\textnormal{ is a chain in }2^{[n]\setminus A}}\#(\mathcal{C}\cap((\mathcal{F}\cap\left[A,[n]\right])-A))}}{\left(n-\left|A\right|\right)!}=l(n-\left|A\right|,(\mathcal{F}\cap\left[A,[n]\right])-A). \end{gather*} $(\mathcal{F}\cap\left[A,[n]\right])-A$ is diamond-free, so $((\mathcal{F}\cap\left[A,[n]\right])-A)\setminus\emptyset$ is $\Lambda$-free; and the top $n'$ levels are assumed to be empty. Using Lemma \ref{lem:lubell} as well as that $\frac{1}{n'}=\frac{1}{\Omega(n)}=o(1)$ and the subadditivity of the square root function, \begin{gather*} l\bigl(n-\left|A\right|,\bigl((\mathcal{F}\cap\left[A,[n]\right])-A\bigr)\setminus\emptyset\bigr)\leq1-\min\!\left(c(A)+\frac{1}{n'},\frac{1}{4}\right)+\sqrt{\min\!\left(c(A)+\frac{1}{n'},\frac{1}{4}\right)}+\frac{3}{n'}\\ \leq f(c(A))+\sqrt{\frac{1}{n'}}+\frac{3}{n'}=f(c(A))+o(1), \end{gather*} so $l\bigl(n-\left|A\right|,(\mathcal{F}\cap\left[A,[n]\right])-A\bigr)\leq1+f(c(A))+o(1)$. Whereas if $A$ is not minimal in $\mathcal{F}$, i.e. $\exists S\in\mathcal{F}$ such that $A\supsetneqq S$, then for any chain $\mathcal{C}$ for which $\min(\mathcal{C}\cap\mathcal{F})=A$, we have $\#(\mathcal{C}\cap\mathcal{F})\leq2$ (otherwise $S$ and three sets in $\mathcal{C}\cap\mathcal{F}$ would form a diamond), so $\frac{\sum_{\substack{\mathcal{C}\textnormal{ is a chain}\\ A=\min(\mathcal{C}\cap\mathcal{F}) } }\#(\mathcal{C}\cap\mathcal{F})}{\#\left\{ \substack{\mathcal{C}\textnormal{ is a chain through }A:\\ A=\min(\mathcal{C}\cap\mathcal{F}) } \right\} }\leq2$. \begin{align*} l(\mathcal{F}) & \leq\frac{1}{n!}\sum_{\substack{A\in\mathcal{F}\\ A\textnormal{ is minimal in }\mathcal{F} } }\#\left\{ \mathcal{C}\textnormal{ is a chain containing }A\right\} \left(1+f(c(A))+o(1)\right)\\ & {}+\frac{1}{n!}\sum_{\substack{A\in\mathcal{F}\\ A\textnormal{ is not minimal in }\mathcal{F} } }\#\left\{ \begin{gathered}\mathcal{C}\textnormal{ is a chain containing }A:\\ A=\min(\mathcal{C}\cap\mathcal{F}) \end{gathered} \right\} \cdot2\\ & \le2+\frac{1}{n!}\sum_{\substack{A\in\mathcal{F}\\ A\textnormal{ is minimal in }\mathcal{F} } }\#\left\{ \mathcal{C}\textnormal{ is a chain containing }A\right\} \left(f\left(c\left(A\right)\right)-1+o(1)\right). \end{align*} Since $f$ is concave, we can use Jensen's inequality with the weights $\frac{\#\left\{ \mathcal{C}\textnormal{ is a chain containing }A\right\} }{(1-C)n!}$ (where $A$ is minimal in $\mathcal{F}$). Notice that the sum of all the weights is $1$ because the sum of numerators is the total number of chains $\mathcal{C}$ where $\min(\mathcal{C}\cap\mathcal{F})$ is minimal in $\mathcal{F}$, that is, $(1-C)n!$. \begin{gather*} l(\mathcal{F})\leq2+(1-C)\left(f\!\left(\frac{{\displaystyle \sum_{\substack{A\in\mathcal{F}\\ A\textnormal{ is minimal in }\mathcal{F} } }\#\left\{ \mathcal{C}\textnormal{ is a chain containing }A\right\} c(A)}}{(1-C)n!}\right)-1+o(1)\right). \end{gather*} $c(A)$ is the fraction of the chains containing $A$ which are MNM, so $\#\left\{ \mathcal{C}\textnormal{ is a chain containing }A\right\} c(A)$ is the number of MNM chains through $A$. In the numerator, each MNM chain in the whole Boolean lattice is counted once, except if the minimal element on it is not a global minimal, then it is not counted. So the numerator is less than or equal to the total number of MNM chains in the Boolean lattice, which is at most $Cn!$. Substituting, we get \[ l(\mathcal{F})\leq2+(1-C)\left(f\!\left(\frac{n!C}{n!(1-C)}\right)-1+o(1)\right)=1+C+(1-C)f\!\left(\frac{C}{1-C}\right)+o(1). \] $C$ varies between 0 and 1. $\frac{C}{1-C}$ is increasing in $C$. Above $\frac{C}{1-C}=\frac{1}{4}$ (corresponding to $C=\frac{1}{5}$), $f(\frac{C}{1-C})$ is constant $\frac{5}{4}$, so $1+C+(1-C)f\!\left(\frac{C}{1-C}\right)=\frac{9}{4}-\frac{C}{4}$ is decreasing in C. So it is enough to take the maximum in the interval $\left[0,\frac{1}{5}\right]$: \begin{align*} \frac{\left|\mathcal{F}\right|}{\binom{n}{\left\lfloor n/2\right\rfloor }} & \leq l(\mathcal{F})\leq\max_{C\in\left[0,\frac{1}{5}\right]}\left(1+C+(1-C)\left(1-\frac{C}{1-C}+\sqrt{\frac{C}{1-C}}\right)\right)+o(1)\\ & =\frac{\sqrt{2}+3}{2}+o(1)<2.20711+o(1). \end{align*} \section*{Acknowledgments\phantomsection\addcontentsline{toc}{section}{Acknowledgements}} We thank the anonymous referees for their detailed comments which helped improve the presentation of our paper. We would also like to thank D\"om\"ot\"or P\'alv\"olgyi for helpful discussions concerning the $3$-level version of this problem. The research of the second and third authors was supported by the National Research, Development and Innovation Office \textendash{} NKFIH, grant K 116769. \bibliographystyle{plain} \phantomsection\addcontentsline{toc}{section}{\refname}
{ "timestamp": "2017-11-27T02:06:43", "yymm": "1601", "arxiv_id": "1601.06332", "language": "en", "url": "https://arxiv.org/abs/1601.06332", "abstract": "Let $La(n,P)$ be the maximum size of a family of subsets of $[n]=\\{1,2,...,n\\}$ not containing $P$ as a (weak) subposet. The diamond poset, denoted $B_{2}$, is defined on four elements $x,y,z,w$ with the relations $x<y,z$ and $y,z<w$. $La(n,P)$ has been studied for many posets; one of the major open problems is determining $La(n,B_{2})$.Studying the average number of sets from a family of subsets of $[n]$ on a maximal chain in the Boolean lattice $2^{[n]}$ has been a fruitful method. We use a partitioning of the maximal chains and introduce an induction method to show that $La(n,B_{2})\\leq(2.20711+o(1))\\binom{n}{\\left\\lfloor \\frac{n}{2}\\right\\rfloor }$, improving on the earlier bound of $(2.25+o(1))\\binom{n}{\\left\\lfloor \\frac{n}{2}\\right\\rfloor }$ by Kramer, Martin and Young.", "subjects": "Combinatorics (math.CO)", "title": "An upper bound on the size of diamond-free families of sets", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.97997655637136, "lm_q2_score": 0.8175744695262777, "lm_q1q2_score": 0.801203813223503 }
https://arxiv.org/abs/2002.00080
Convergence rate analysis and improved iterations for numerical radius computation
The main two algorithms for computing the numerical radius are the level-set method of Mengi and Overton and the cutting-plane method of Uhlig. Via new analyses, we explain why the cutting-plane approach is sometimes much faster or much slower than the level-set one and then propose a new hybrid algorithm that remains efficient in all cases. For matrices whose fields of values are a circular disk centered at the origin, we show that the cost of Uhlig's method blows up with respect to the desired relative accuracy. More generally, we also analyze the local behavior of Uhlig's cutting procedure at outermost points in the field of values, showing that it often has a fast Q-linear rate of convergence and is Q-superlinear at corners. Finally, we identify and address inefficiencies in both the level-set and cutting-plane approaches and propose refined versions of these techniques.
\section{Introduction} Consider the discrete-time dynamical system \begin{equation} \label{eq:ode_disc} x_{k+1} = Ax_k, \end{equation} where $A \in \mathbb{C}^{n \times n}$ and $x_k \in \mathbb{C}^n$. The asymptotic behavior of \eqref{eq:ode_disc} is of course characterized by the moduli of $\Lambda(A)$. Given the \emph{spectral radius} of~$A$, \begin{equation} \rho(A) \coloneqq \max \{ |\lambda| : \lambda \in \Lambda(A)\}, \end{equation} $\lim_{k \to \infty} \| x_k\| = 0$ for all $x_0$ if and only if $\rho(A) < 1$, with the asymptotic decay rate being faster the closer $\rho(A)$ is to zero. However, knowing the transient behavior of~\eqref{eq:ode_disc} is often of interest. Clearly, the trajectory of~\eqref{eq:ode_disc} is tied to powers of~$A$, since $x_k = A^k x_0$ and so \mbox{$\|x_k\| \leq \|A^k\| \|x_0\|$}. Indeed, a central theme of Trefethen and Embree's treatise on pseudospectra \cite{TreE05} is how large $\sup_{k \geq 0} \|A^k\|$ can be. One perspective is given by the \emph{field of values} (\emph{numerical range}) of $A$, \begin{equation} \label{eq:fov} \fovX{A} \coloneqq \{ x^* A x : x \in \mathbb{C}^n, \|x\| = 1\}. \end{equation} Consider the maximum of the moduli of points in $\fovX{A}$, i.e., the \emph{numerical radius} \begin{equation} \label{eq:numr} r(A) \coloneqq \max \{ | z | : z \in \fovX{A}\}. \end{equation} It is known that $\tfrac{1}{2}\|A\| \leq r(A) \leq \|A\|$; see \cite[p.~44]{HorJ91}. Combining the lower bound with the power inequality $r(A^k) \leq (r(A))^k$ \cite{Ber65,Pea66} yields \begin{equation} \label{eq:numr_ineq} \| A^k \| \leq 2 (r(A))^k. \end{equation} As $2(r(A))^k \leq \|A\|^k$ if and only if $r(A) \leq \sqrt[k]{0.5} \|A\|$, and $r(A) \leq \|A\|$ always holds, it follows that $2(r(A))^k$ is often a tighter upper bound for $\|A^k\|$ than $\|A\|^k$ is, and so the numerical radius can be useful in estimating the transient behavior of \eqref{eq:ode_disc}.\footnote{Per \cite{TreE05}, the \emph{pseudospectral radius} and the \emph{Kreiss constant} \cite{Kre62} also give information on the trajectory of~\eqref{eq:ode_disc}. For computing these quantities, see \cite{MenO05,BenM19} and \cite{Mit20a,Mit21}.} The concept of the numerical radius dates to at least 1961 (see \cite[p.~1005]{LewO20}), but methods to compute it seem to have only first appeared in the 1990s. Mathias showed that~$r(A)$ can be obtained by solving a semidefinite program~\cite{Mat93}, but doing so is expensive. Much faster algorithms were then proposed by He and Watson \cite{Wat96,HeW97}, but these methods may not converge to $r(A)$. In 2005, Mengi and Overton gave a fast globally convergent method to compute $r(A)$~\cite{MenO05} by combining an idea of He and Watson \cite{HeW97} with the level-set approach of Boyd, Balakrishnan, Bruinsma, and Steinbuch (BBBS) \cite{BoyB90,BruS90} for computing the $\Hcal_\infty$ norm. While Mengi and Overton observed that their method converged quadratically, this was only later proved in 2012 by G\"urb\"uzbalaban in his PhD thesis~\cite[section~3.4]{Gur12}. In 2009, Uhlig proposed a geometric approach to computing $r(A)$~\cite{Uhl09}.\footnote{In this same paper~ \cite[section~3]{Uhl09}, Uhlig also discussed how Chebfun~\cite{DriHT14} can be used to reliably compute~$r(A)$ with just a few lines of MATLAB, but that it is generally orders of magnitude slower than either his method or the one of Mengi and Overton; see also~\cite{GreO18}.} Uhlig's method is based on Johnson's cutting-plane technique for approximating $\fovX{A}$ \cite{Joh78}, which itself stems from the much earlier Bendixson-Hirsch theorem \cite{Ben02a} and fundamental results of Kippenhahn \cite{Kip51}. In \cite[Remark~3]{Joh78}, Johnson observed that while his $\fovX{A}$ boundary method could be adapted to compute $r(A)$, a modified version might be more efficient. These geometric approaches all work by computing a number of supporting hyperplanes to sufficiently approximate the boundary of $\fovX{A}$ or a region of it where $r(A)$ is attained. A major benefit of cutting-plane methods is that they only require computing $\lambda_\mathrm{max}$ of $n \times n$ Hermitian matrices. If $A$ is sparse, this can be done efficiently and reliably using, say, \texttt{eigs} in MATLAB\@. Hence, Uhlig's method can be used on large-scale problems while still being globally convergent. In contrast, at every iteration, the level-set approach requires solving a generalized eigenvalue problem of order $2n$, which by standard convention on work complexity, is an atomic operation with $\mathcal{O}(n^3)$~work. While Uhlig noted that convergence of his method can sometimes be quite slow~\cite[p.~344]{Uhl09}, his experiments in the same paper showed several problems where his cutting-plane method was decisively faster Mengi and Overton's level-set method. The paper is organized as follows. In~\cref{sec:background}, we give necessary preliminaries on the field of values, the numerical radius, and earlier $r(A)$ algorithms. We then identify and address some inefficiencies in the level-set method of Mengi and Overton and propose a faster variant in~\cref{sec:alg1}. We analyze Uhlig's method in \cref{sec:rate}, deriving (a) its overall cost when the field of values is a disk centered at the origin, and (b) a Q-linear local rate of convergence result for its cutting procedure. These analyses precisely show how, depending on the problem, Uhlig's method can be either extremely fast or extremely slow. In \cref{sec:alg2}, we identify an inefficiency in Uhlig's cutting procedure and address it via a more efficient cutting scheme whose exact convergence rate we also derive. Putting all of this together, we present our new hybrid algorithm in \cref{sec:hybrid}. We validate our results experimentally in \cref{sec:experiments} and give concluding remarks in~\cref{sec:conclusion}. \section{Preliminaries} \label{sec:background} We will need the following well-known facts~\cite{Kip51,HorJ91}: \begin{remark} \label{rem:fov} Given $A \in \mathbb{C}^{n \times n}$, \begin{enumerate}[leftmargin=30pt,label=(A\arabic*),font=\normalfont] \item $\fovX{A} \subset \mathbb{C}$ is a compact, convex set, \item if $A$ is real, then $\fovX{A}$ has real axis symmetry, \item if $A$ is normal, then $\fovX{A}$ is the convex hull of $\Lambda(A)$, \item $\fovX{A} = [\lambda_\mathrm{min}(A), \lambda_\mathrm{max}(A)]$ if and only if $A$ is Hermitian, \item the boundary of $\fovX{A}$, $\bfovX{A}$, is a piecewise smooth algebraic curve, \item if $v \in \bfovX{A}$ is a point where $\bfovX{A}$ is not differentiable, i.e., a corner, then $v \in \Lambda(A)$. Corners always correspond to two line segments in $\bfovX{A}$ meeting at some angle less than $\pi$ radians. \end{enumerate} \end{remark} \begin{definition} Given a nonempty closed set $\mathcal{D} \subset \mathbb{C}$, a point $\tilde z \in \mathcal{D}$ is (globally) \emph{outermost} if $|\tilde z| = \max\{|z| : z \in \mathcal{D}\}$ and \emph{locally outermost} if $\tilde z$ is an outermost point of $\mathcal{D} \cap \mathcal{N}$, for some neighborhood $\mathcal{N}$ of $\tilde z$. \end{definition} For continuous-time systems $\dot x = Ax$, we have the \emph{numerical abscissa} \begin{equation} \label{eq:numa} \alpha_\fovsym(A) \coloneqq \max \{ \Re z : z \in \fovX{A}\}, \end{equation} i.e., the maximal real part of all points in $\fovX{A}$. Unlike the numerical radius, computing the numerical abscissa is straightforward, as~\cite[p.~34]{HorJ91} \begin{equation} \label{eq:tan_line} \alpha_\fovsym(A) = \lambda_\mathrm{max} \left(\tfrac{1}{2} \left(A + A^*\right)\right). \end{equation} For $\theta \geq 0$, $\fovX{\eix{\theta} A}$ is $\fovX{A}$ rotated counter-clockwise about the origin. Consider \begin{equation} \label{eq:hmat} H(\theta) \coloneqq \tfrac{1}{2} \left(\eix{\theta} A + \emix{\theta} A^*\right), \end{equation} so $\numaX{\eix{\theta} A} = \lambda_\mathrm{max}(H(\theta))$ and $\alpha_\fovsym(A) = \lambda_\mathrm{max}(H(0))$. Let $\lambda_\theta$ and~$x_\theta$ denote, respectively, $\lambda_\mathrm{max}(H(\theta))$ and an associated normalized eigenvector. Furthermore, let $L_\theta$ denote the line $\{ \emix{\theta} (\lambda_\theta + \mathbf{i} t): t \in \mathbb{R}\}$ and $P_\theta$ the half plane $\emix{\theta} \{ z : \Re z \leq \lambda_\theta\}$. Then $L_\theta$ is a \emph{supporting hyperplane} for~$\fovX{A}$ and \cite[p.~597]{Joh78} \begin{enumerate}[leftmargin=30pt,label=(B\arabic*),font=\normalfont] \item $\fovX{A} \subseteq P_\theta$ for all $\theta \in [0,2\pi)$, \item $\fovX{A} = \cap_{\theta \in [0,2\pi)} P_\theta$, \item $z_\theta = x_\theta^*Ax_\theta \in L_\theta$ is a \emph{boundary point} of $\fovX{A}$. \end{enumerate} As $H(\theta + \pi) = -H(\theta)$, $P_{\theta+\pi}$ can also be obtained via $\lambda_\mathrm{min}(H(\theta))$ and an associated eigenvector. The Bendixson-Hirsch theorem is a special case of these properties, defining the bounding box of $\fovX{A}$ for $\theta = 0$ and~\mbox{$\theta = \tfrac{\pi}{2}$}. When $\lambda_\mathrm{max}(H(\theta))$ is simple, the following result of Fiedler~\cite[Theorem~3.3]{Fie81} gives a formula for computing the \emph{radius of curvature $\tilde r$ of $\bfovX{A}$ at~$z_\theta$}, defined as the radius of the \emph{osculating circle of $\bfovX{A}$ at~$z_\theta$}, i.e., the circle with the same tangent and curvature as $\bfovX{A}$ at $z_\theta$. At corners of $\bfovX{A}$, we say that~$\tilde r = 0$, while at other boundary points where the radius of curvature is well defined,\footnote{An example where $\tilde r$ is not well defined is given by $A = \begin{bsmallmatrix} J & 0 \\ 0 & J+I \end{bsmallmatrix}$ with $J = \begin{bsmallmatrix} 0 & 1 \\ 0 & 0 \end{bsmallmatrix}$. At~\mbox{$b=0.5 \in \bfovX{A}$}, two of the algebraic curves, a line segment and a semi-circle, comprising $\bfovX{A}$ meet, and $\bfovX{A}$ is only once differentiable at this non-corner boundary point. Here, the radius of curvature of $\bfovX{A}$ jumps from $0.5$ (for the semi-circular piece) to $\infty$ (for the line segment).} $\tilde r > 0$ and becomes infinite at points inside line segments in $\bfovX{A}$. Although the formula is given for $\theta=0$ and~$z_\theta = 0$, by simple rotation and shifting, it can be applied generally. See \cref{fig:demo_fov} for a depiction of the osculating circle of $\bfovX{A}$ at an outermost point in~$\fovX{A}$. \begin{theorem}[Fiedler] \label{thm:curvature} Let $H_1 = \tfrac{1}{2}(A + A^*)$, \mbox{$H_2 = \tfrac{1}{2\mathbf{i}}(A - A^*)$}, let $H_1^+$ be the Moore-Penrose pseudoinverse of $H_1$, and let $x_\theta$ be a normalized eigenvector corresponding to $\lambda_\mathrm{max}(H(\theta))$. Noting that $A = H_1 + \mathbf{i} H_2$ and $H(0) = H_1$, suppose that \mbox{$\lambda_\mathrm{max}(H_1) = 0$} and is simple, and that the associated boundary point \mbox{$z_\theta = x_\theta^* A x_\theta = 0$}, where $\theta=0$. Then the radius of curvature of $\bfovX{A}$ at $z_\theta$ is \begin{equation} \label{eq:curvature} \tilde r = -2 (H_2 x_\theta)^* H_1^+ (H_2 x_\theta). \end{equation} \end{theorem} \begin{figure} \centering \subfloat[The field of values and local curvature.]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{demo_fov}} \label{fig:demo_fov} } \subfloat[Level-set iterates for $h(\theta)$.]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{demo_levelset}} \label{fig:demo_levelset} } \caption{For a random matrix $A \in \mathbb{C}^{10 \times 10}$, the left and right panes respectively show $\bfovX{A}$ (blue curve) and $h(\theta)$ (blue plot). On the left, the following are also shown: $\Lambda(A)$ (black dots), a polygonal approximation $\mathcal{G}_j$ to $\fovX{A}$ (blue polygon), the outermost point in $\fovX{A}$ (small blue circle) with the corresponding supporting hyperplane (blue line) and osculating circle (dashed red circle), and the circle of radius~$r(A)$ centered at the origin (black dotted circle). On the right, three iterations of the level-set method are also shown (small circles and dash-dotted lines in red). } \label{fig:demo} \end{figure} Via \eqref{eq:tan_line} and \eqref{eq:hmat}, the numerical radius can be written as \begin{equation} \label{eq:nr_opt} r(A) = \max_{\theta \in [0,2\pi)} h(\theta) \qquad \text{where} \qquad h(\theta) \coloneqq \lambda_\mathrm{max} \left(H(\theta)\right), \end{equation} i.e., a one-variable maximization problem. Via $H(\theta + \pi) = -H(\theta)$, it also follows that \begin{equation} \label{eq:nr_opt_abs} r(A) = \max_{\theta \in [0,\pi)} \rho( H(\theta)). \end{equation} However, as \eqref{eq:nr_opt} and \eqref{eq:nr_opt_abs} may have multiple maxima, it is not straightforward to find a global maximizer of either, and crucially, assert that it is indeed a global maximizer in order to verify that $r(A)$ has been computed. Per (A3), we generally assume that $A$ is non-normal, as otherwise $r(A) = \rho(A)$. We now discuss earlier numerical radius algorithms in more detail. In 1996, Watson proposed two $r(A)$ methods~\cite{Wat96}: one which converges to local maximizers of~\eqref{eq:nr_opt_abs} and a second which lacks convergence guarantees but is cheaper (though they each do $\mathcal{O}(n^2)$ work per iteration). However, as both iterations are related to the power method, they may exhibit very slow convergence, and the cheaper iteration may not converge at all. Shortly thereafter~\cite{HeW97}, He and Watson used the second iteration (because it was cheaper) in combination with a new certificate test inspired by Byers’ distance to instability algorithm~\cite{Bye88}. This certificate either asserts that $r(A)$ has been computed to a desired accuracy or provides a way to restart Watson’s cheaper iteration with the hope of more accurately estimating~$r(A)$. However, He and Watson’s method is still not guaranteed to converge, since Watson’s cheaper iteration may not converge. Inspired by the BBBS algorithm for computing the $\Hcal_\infty$~norm, Mengi and Overton then proposed a globally convergent iteration for~$r(A)$ in 2005 by using He and Watson's certificate test in a much more powerful way. Given $\gamma \leq r(A)$, the test actually allows one to obtain the $\gamma$-level set of $h(\theta)$, i.e., $\{ \theta : h(\theta) = \gamma\}$. Assuming the level set is not empty, Mengi and Overton's method then evaluates~$h(\theta)$ at the midpoints of the intervals under~$h(\theta)$ determined by the $\gamma$-level~set points. Estimate $\gamma$ is then updated (increased) to the highest of these corresponding function values. This process is done in a loop, and as mentioned in the introduction, has local quadratic convergence. See \cref{fig:demo_levelset} for a depiction of this level-set iteration. The certificate (or level-set) test is based on \cite[Theorem~3.1]{MenO05}, which is a slight restatement of \cite[Theorem~2]{HeW97} from He and Watson. We omit the proof. \begin{theorem} \label{thm:level} Given $\gamma \in \mathbb{R}$, the pencil $R_\gamma - \lambda S$ has $\eix{\theta}$ as an eigenvalue or is singular if and only if $\gamma$ is an eigenvalue of $H(\theta)$ defined in \eqref{eq:hmat}, where \begin{equation} \label{eq:RS} R_\gamma \coloneqq \begin{bmatrix} 2\gamma I & -A^* \\ I & 0 \end{bmatrix} \quad \text{and} \quad S \coloneqq \begin{bmatrix} A & 0 \\ 0 & I \end{bmatrix}. \end{equation} \end{theorem} Per \cref{thm:level}, the $\gamma$-level set of $h(\theta) = \lambda_\mathrm{max}(H(\theta))$ is associated with the unimodular eigenvalues of $R_\gamma - \lambda S$, which can be obtained in $\mathcal{O}(n^3)$ work (with a significant constant factor). Note that the converse may not hold, i.e., for a unimodular eigenvalue of $R_\gamma - \lambda S$, $\gamma$ may correspond to an eigenvalue of $H(\theta)$ other than $\lambda_\mathrm{max}(H(\theta))$. Given any $\theta \in [0,2\pi)$, also note that $R_\gamma - \lambda S$ is nonsingular for all $\gamma > h(\theta)$. This is because if $\fovX{A}$ and a disk centered at the origin enclosing $\fovX{A}$ have more than $n$ shared boundary points, then $\fovX{A}$ is that disk; see \cite[Lemma~6]{TamY99}. Uhlig's method computes $r(A)$ via updating a bounded convex polygonal approximation to $\fovX{A}$ and set of known points in $\bfovX{A}$ respectively given by: \begin{equation*} \mathcal{G}_j \coloneqq \bigcap_{\theta \in \left\{\theta_1,\ldots,\theta_j\right\}} P_\theta \qquad \text{and} \qquad \mathcal{Z}_j \coloneqq \{ z_{\theta_1}, \ldots, z_{\theta_j} \}, \end{equation*} where $\fovX{A} \subseteq \mathcal{G}_j$ (see~\cref{fig:demo_fov} for a depiction), $0 \leq \theta_1 < \cdots < \theta_j < 2\pi$, and $z_{\theta_\ell} = x_{\theta_\ell}^* A x_{\theta_\ell}$ is a boundary point of $\fovX{A}$ on $L_{\theta_\ell}$ for $\ell = 1,\ldots,j$. Note that the corners of $\mathcal{G}_j$ are given by $L_{\theta_{\ell}} \cap L_{\theta_{\ell+1}}$ for $\ell = 1,\ldots,j-1$ and $L_{\theta_1} \cap L_{\theta_j}$. Given $\mathcal{G}_j$ and $\mathcal{Z}_j$, lower and upper bounds $l_j \leq r(A) \leq u_j$ are immediate, where \[ l_j \coloneqq \max \{ |b| : b \in \mathcal{Z}_j \} \quad \text{and} \quad u_j \coloneqq \max \{ |c| : c \text{ a corner of } \mathcal{G}_j \}, \] so we define the relative error estimate: \begin{equation} \label{eq:nr_err} \varepsilon_j \coloneqq \frac{ u_j - l_j }{ l_j}. \end{equation} By repeatedly cutting outermost corners of $\mathcal{G}_j$, and in turn, adding computed boundary points of $\fovX{A}$ to $\mathcal{Z}_j$, it follows that $\varepsilon_j$ must fall below a desired relative tolerance for some~$k \geq j$; hence, $r(A)$ can be computed to any desired accuracy. Uhlig's method achieves this via a greedy strategy. On each iteration, his algorithm chops off an outermost corner $c_j$ from $\mathcal{G}_j$, which is done via computing the supporting hyperplane~$L_{\theta_{j+1}}$ for~\mbox{$\theta_{j+1} = -\Arg(c_{j})$} and the boundary point $z_{\theta_{j+1}} = x_{\theta_{j+1}}^* A x_{\theta_{j+1}}$. Assuming that $c_j \not\in \fovX{A}$, the cutting operation results in \mbox{$\mathcal{G}_{j+1} \coloneqq \mathcal{G}_j \cap P_{\theta_{j+1}}$}, a smaller polygonal region excluding the corner~$c_j$, and \mbox{$\mathcal{Z}_{j+1} \coloneqq \mathcal{Z}_j \cup \{z_{\theta_{j+1}}\}$}; therefore, $\varepsilon_{j+1} \leq \varepsilon_j$. However, if~$c_j$ happens to be a corner of $\bfovX{A}$, then it cannot be cut from $\mathcal{G}_j$, and instead this operation asserts that $|c_j| = r(A)$, and so $r(A)$ has been computed. In \cref{sec:rate}, \Cref{fig:uhlig} depicts Uhlig's method when a corner is cut. \begin{remark} \label{rem:two_hp} Recall that the parallel supporting hyperplane $L_{\theta_{j+1} + \pi}$ and the corresponding boundary point $z_{\theta_{j+1} + \pi}$ can be obtained via an eigenvector $\tilde x_{\theta_{j+1}}$ of $\lambda_\mathrm{min}(H(\theta_{j+1}))$. If~$\tilde x_{\theta_{j+1}}$ is already available or relatively cheap to compute, there is little reason \emph{not} to also update~$\mathcal{G}_j$ and~$\mathcal{Z}_j$ using this additional information. \end{remark} \section{Improvements to the level-set approach} \label{sec:alg1} We now propose two straightforward but important modifications to make the level-set approach faster and more reliable. We need the following immediate corollary of \cref{thm:level}, which clarifies that \cref{thm:level} also allows all points in any $\gamma$-level set of $\rho(H(\theta))$ to be computed. \begin{corollary} \label{cor:remap} Given $\gamma \geq 0$, if $\rho(H(\theta)) = \gamma$, then there exists $\lambda \in \mathbb{C}$ such that $|\lambda| = 1$, $\det(R_\gamma - \lambda S) = 0$, and $\theta = f(\Arg(\lambda))$, where $f : (-\pi,\pi] \mapsto [0, \pi)$ is \begin{equation} \label{eq:remap} f(\theta) \coloneqq \begin{cases} \theta + \pi & \text{if } \theta < 0 \\ 0 & \text{if } \theta = 0 \text{ or } \theta = \pi \\ \theta & \text{otherwise}. \end{cases} \end{equation} \end{corollary} Thus, first we propose doing a BBBS-like iteration using $\rho(H(\theta))$ instead of $h(\theta)$, which also has local quadratic convergence. By an extension of the argument of Boyd and Balakrishnan~\cite{BoyB90}, near maximizers, $\rho(H(\theta))$ is unconditionally twice continuously differentiable with Lipschitz second derivative; see~\cite{MitO21}. Using $\rho(H(\theta))$ is also typically faster in terms of constant factors. This is because $\rho(H(\theta)) \geq h(\theta)$ always holds, $\rho(H(\theta)) \geq 0$ (unlike $h(\theta)$, which can be negative), and the optimization domain is reduced from $[0,2\pi)$ to $[0,\pi)$. Thus, every update to the current estimate~$\gamma$ computed via $\rho(H(\theta))$ must be at least as good as the one from using~$h(\theta)$ (and possibly much better), and there may also be fewer level-set intervals per iteration, which reduces the number of eigenproblems incurred involving $H(\theta)$. Second, we also propose using local optimization on top of the BBBS-like step at every iteration, i.e., the BBBS-like step is used to initialize optimization in order to find a maximizer of $\rho(H(\theta))$. The first benefit is speed, as optimization often results in much larger updates to estimate $\gamma$ and these updates are now locally optimal. This greatly reduces the total number of expensive eigenvalue computations done with $R_\gamma - \lambda S$, often down to just one; hence, the overall runtime can be substantially reduced since in comparison, optimization is cheap (as we explain momentarily). The second benefit is that using optimization also avoids some numerical difficulties when solely working with $R_\gamma - \lambda S$ to update $\gamma$. In their 1997 paper, He and Watson showed that the condition number of a unimodular eigenvalue of $R_\gamma - \lambda S$ actually blows up as $\theta$ approaches critical values of $h(\theta)$ or $\rho(H(\theta))$ \cite[Theorem~4]{HeW97},\footnote{The exact statement appears in the last lines of the corresponding proof on p.~335.} as this corresponds to a pair of unimodular eigenvalues of $R_\gamma - \lambda S$ coalescing into a double eigenvalue. Since this must always occur as a level-set method converges, rounding errors may prevent all of the unimodular eigenvalues from being detected, causing level-set points to go undetected, thus resulting in stagnation of the algorithm before it finds~$r(A)$ to the desired accuracy. He and Watson wrote that their analytical result was ``hardly encouraging" \cite[p.~336]{HeW97}, though they did not observe this issue in their experiments. However, an example of such a deleterious effect is shown in~\cite[Figure~2]{BenM19}, where analogous eigenvalue computations are shown to greatly reduce numerical accuracy when computing the \emph{pseudospectral abscissa}~\cite{BurLO03}. In contrast, optimizing $\rho(H(\theta))$ does not lead to numerical difficulties. This objective function is both Lipschitz (as $H(\theta)$ is Hermitian \cite[Theorem~II.6.8]{Kat82}) and smooth at its maximizers (as discussed above). Thus, local maximizers of $\rho(H(\theta))$ can be found using, say, Newton's method, with only a handful of iterations. Interestingly, in their concluding remarks \cite[p.~341--2]{HeW97}, He and Watson seem to have been somewhat pessimistic about using Newton's method, writing that while it would have faster local convergence than Watson's iteration, ``the price to be paid is at least a considerable increase in computation, and possibly the need of the calculation of higher derivatives, and for the incorporation of a line search." As we now explain, using, say, secant or Newton's method, is actually an overall big win. Also, note that with either secant or Newton, steps of length one are always eventually accepted; hence, the cost of line searches should not be a concern. \begin{algfloat}[t] \begin{algorithm}[H] \floatname{algorithm}{Algorithm} \caption{An Improved Level-Set Algorithm} \label{alg1} \begin{algorithmic}[1] \REQUIRE{ $A \in \mathbb{C}^{n \times n}$ with $n \geq 2$, initial guesses $\mathcal{M} = \{\theta_1,\ldots,\theta_q\}$, and \mbox{$\tau_\mathrm{tol} > 0$}. } \ENSURE{ $\gamma$ such that $|\gamma -r(A)| \leq \tau_\mathrm{tol} \cdot r(A)$. \\ \quad } \STATE $\psi \gets f(\Arg(\lambda))$ where $\lambda \in \Lambda(A)$ such that $|\lambda| = \rho(A)$ \COMMENT{0 if $\lambda = 0$} \STATE $\mathcal{M} \gets \mathcal{M} \cup \{ 0, \psi \}$ \WHILE { $\mathcal{M}$ is not empty } \STATE $\theta_\mathrm{BBBS} \gets \argmax_{\theta \in \mathcal{M}} \rho(H(\theta))$ \COMMENT{In case of ties, just take any one} \STATE $\gamma \gets$ maximization of $\rho(H(\theta))$ via local optimization initialized at $\theta_\mathrm{BBBS}$ \STATE $\gamma \gets \gamma (1 + \tau_\mathrm{tol})$ \STATE $\Theta \gets \{ f(\Arg(\lambda)) : \det (R_\gamma - \lambda S) = 0, |\lambda | = 1 \}$ \STATE $[\theta_1,\ldots,\theta_q] \gets \Theta~\text{sorted in increasing order with any duplicates removed}$ \STATE $\mathcal{M} \gets \{ \theta : \rho(H(\theta)) > \gamma ~\text{where}~ \theta = 0.5(\theta_\ell + \theta_{\ell+1}), \ell=1,\ldots,q-1\}$ \ENDWHILE \end{algorithmic} \end{algorithm} \algnote{ For simplicity, we forgo giving pseudocode to exploit possible normality of~$A$ or symmetry of~$\fovX{A}$, and assume that eigenvalues and local maximizers are obtained exactly and that the optimization solver is monotonic, i.e., it guarantees $\rho(H(\theta_\mathrm{BBBS})) \leq \rho(H(\theta_\star))$, where $\theta_\star$ is the maximizer computed in line~5. Recall that~$f(\cdot)$ is defined in~\eqref{eq:remap}, and note that the method reduces to a BBBS-like iteration using \eqref{eq:nr_opt_abs} if line~5 is replaced by \mbox{$\gamma \gets \rho(H(\theta_\mathrm{BBBS}))$}. Running optimization from other angles in $\mathcal{M}$ (in addition to $\theta_\mathrm{BBBS}$) every iteration may also be advantageous, particularly if this can be done via parallel processing. Adding zero to the initial set $\mathcal{M}$ avoids having to deal with any ``wrap-around" level-set intervals due to the periodicity of $\rho(H(\theta))$, while $\psi$ is just a reasonable initial guess for a global maximizer of \eqref{eq:nr_opt_abs}. } \end{algfloat} Suppose $\rho(H(\theta))$ is attained by a unique eigenvalue~$\lambda_j$ with normalized eigenvector $x_j$. Then by standard perturbation theory for simple eigenvalues, \begin{equation} \rho^\prime(H(\theta)) = \sgn(\lambda_j) \cdot x_j^*H^\prime(\theta)x_j = \sgn(\lambda_j) \cdot x_j^*\left(\tfrac{\mathbf{i}}{2} \left( \eix{\theta} A - \emix{\theta} A^* \right) \right) x_j. \end{equation} Thus, given $\lambda_j$ and $x_j$, the additional cost of obtaining $\rho^\prime(H(\theta))$ mostly amounts to the single matrix-vector product $H^\prime(\theta)x_j$. To compute $\rho^{\prime\prime}(H(\theta))$, we will need the following result for second derivatives of eigenvalues; see \cite{Lan64}. \begin{theorem} \label{thm:eigdx2} For $t \in \mathbb{R}$, let $A(t)$ be a twice-differentiable $n \times n$ Hermitian matrix family with, for $t=0$, eigenvalues $\lambda_1 \geq \ldots \geq \lambda_n$ and associated eigenvectors $x_1,\ldots,x_n$, with $\|x_k\| = 1$ for all $k$. Then assuming $\lambda_j$ is unique, \[ \lambda_j^{\prime\prime}(t) \bigg|_{t=0}= x_j^* A''(0) x_j + 2 \sum_{k \ne j} \frac{| x_k^* A'(0) x_j |^2}{\lambda_k - \lambda_j}. \] \end{theorem} Although obtaining the eigendecomposition of $H(\theta)$ is cubic work, this is generally negligible compared to the cost of obtaining all the unimodular eigenvalues of $R_\gamma - \lambda S$ when using \cref{thm:level} computationally; recall that $H(\theta)$ is an $n \times n$ Hermitian matrix, while $R_\gamma - \lambda S$ is a generalized eigenvalue problem of order $2n$. Moreover, $H^\prime(\theta)x_j$ would already be computed for~$\rho^\prime(H(\theta))$, while $ H^{\prime\prime}(\theta)x_j = -H(\theta)x_j = -\lambda_j x_j $, so there is no other work of consequence to obtain $\rho^{\prime\prime}(H(\theta))$ via \cref{thm:eigdx2}. \begin{table} \centering \caption{The running time of a given operation \emph{divided by} the running time of \texttt{eig($H(\theta)$)} for random $A \in \mathbb{C}^{n \times n}$. Eigenvectors were requested for~$H(\theta)$ (for computing derivatives and boundary points) but not for $R_\gamma - \lambda S$. For \texttt{eigs}, \texttt{k} is the number of eigenvalues requested, while \texttt{'LM'} (largest modulus), \texttt{'LR'} (largest real), and \texttt{'BE'} (both ends) specifies which eigenvalues are desired. } \setlength{\tabcolsep}{3pt} \begin{tabular}{c | r SS | SS | SSS | S} \toprule \multicolumn{2}{c}{} & \multicolumn{2}{c}{\texttt{eigs($H(\theta)$,k,'LM')}} & \multicolumn{2}{c}{\texttt{eigs($H(\theta)$,k,'LR')}} & \multicolumn{3}{c}{\texttt{eigs($H(\theta)$,k,'BE')}} & \multicolumn{1}{c}{\texttt{eig($R_\gamma$,$S$)}} \\ \cmidrule(lr){3-4} \cmidrule(lr){5-6} \cmidrule(lr){7-9} \cmidrule(lr){10-10} \multicolumn{1}{c}{} & \multicolumn{1}{c}{$n$} & \multicolumn{1}{c}{$\texttt{k}=1$} & \multicolumn{1}{c}{$\texttt{k}=6$} & \multicolumn{1}{c}{$\texttt{k}=1$} & \multicolumn{1}{c}{$\texttt{k}=6$} & \multicolumn{1}{c}{$\texttt{k}=2$} & \multicolumn{1}{c}{$\texttt{k}=4$} & \multicolumn{1}{c}{$\texttt{k}=6$} & \multicolumn{1}{c}{} \\ \midrule \multirow{4}{*}{\rotatebox[origin=c]{90}{Dense $A$}} & \multicolumn{1}{r|}{200} & 1.4 & 1.4 & 0.7 & 1.1 & 1.0 & 1.1 & 1.1 & 36.7 \\ & \multicolumn{1}{r|}{400} & 0.5 & 0.8 & 0.5 & 0.8 & 0.6 & 0.7 & 0.8 & 79.3 \\ & \multicolumn{1}{r|}{800} & 0.6 & 1.0 & 0.6 & 1.1 & 1.2 & 1.0 & 1.1 & 180.2 \\ & \multicolumn{1}{r|}{1600} & 0.3 & 0.6 & 0.3 & 0.5 & 0.6 & 0.7 & 0.6 & 196.3 \\ \midrule \multirow{4}{*}{\rotatebox[origin=c]{90}{Sparse $A$}} & \multicolumn{1}{r|}{200} & 4.0 & 3.7 & 2.1 & 3.5 & 3.4 & 3.8 & 3.6 & 45.7 \\ & \multicolumn{1}{r|}{400} & 2.0 & 3.1 & 1.8 & 3.1 & 3.1 & 3.2 & 3.1 & 80.6 \\ & \multicolumn{1}{r|}{800} & 0.7 & 1.2 & 0.6 & 1.1 & 1.4 & 1.1 & 1.2 & 167.9 \\ & \multicolumn{1}{r|}{1600} & 0.4 & 0.6 & 0.4 & 0.7 & 0.8 & 0.8 & 0.6 & 177.2 \\ \bottomrule \end{tabular} \label{tbl:eig} \end{table} Pseudocode for our improved level-set algorithm is given in~\cref{alg1}. We now address some implementation concerns. What method is used to find maximizers of~$\rho(H(\theta))$ depends on the relative costs of solving eigenvalue problems involving $H(\theta)$ and $R_\gamma - \lambda S$. \cref{tbl:eig} shows examples where 37--196 calls of \texttt{eig($H(\theta)$)} can be done before the total cost exceeds that of a single call of \texttt{eig($R_\gamma$,$S$)}. This highlights just how beneficial it can be to incur a few more computations with $H(\theta)$ to find local maximizers as \cref{alg1} does. Comparisons for computing extremal eigenvalues of~$H(\theta)$ via \texttt{eig} and \texttt{eigs} are also shown in the table. Such data inform whether or not the increased cost of needing to use \texttt{eig} in order to compute $\rho^{\prime\prime}(H(\theta))$ is offset by the advantages that second derivatives can bring, e.g., faster local convergence. Of course, fine-grained implementation decisions like these should ideally be made via tuning, as such timings are generally also software and hardware dependent. Nevertheless, \cref{tbl:eig} suggests that implementing \cref{alg1} using Newton's method via \texttt{eig} might be a bit more efficient than using the secant method for $n \leq 800$ or so.\footnote{ Subspace methods such as \cite{KreLV18} might also be used to find local maximizers of $h(\theta)$ or~$\rho(H(\theta))$ and would likely provide similar benefits in terms of accelerating the globally convergent algorithms in this paper. } There is one more subtle but important detail for implementing \cref{alg1}. Suppose that $\theta_\mathrm{BBBS}$ in line~4 is close to the argument of a (nearly) double unimodular eigenvalue of $R_\gamma - \lambda S$, where $\gamma = \rho(H(\theta_\mathrm{BBBS}))$. If rounding errors prevent this one or two eigenvalues from being detected as unimodular, the computed $\gamma$-level set of $\rho(H(\theta))$ may be incomplete, which again, can cause stagnation. As pointed out in~\cite[p.~372--373]{BurLO03} in the context of computing the pseudospectral abscissa, a robust fix is simple: explicitly add $a(\theta_\mathrm{BBBS})$ to $\Theta$ in line~7 if it appears to be missing. \section{Analysis of Uhlig's method} \label{sec:rate} In the next two subsections, we respectively (a) analyze the overall cost of Uhlig's method for so-called \emph{disk matrices} and (b) for general problems, establish how the exact Q-linear local rate of convergence of Uhlig's cutting strategy varies with respect to the local curvature of $\bfovX{A}$ at outermost points. A disk matrix is one whose field of values is a circular disk centered at the origin, and it is a worst-case scenario for Uhlig's method; as we show in this case, the required number of supporting hyperplanes to compute $r(A)$ blows up with respect to increasing the desired relative accuracy. Although relatively rare, disk matrices can arise from minimizing the numerical radius of parametrized matrices; see \cite{LewO20} for a thorough discussion. For concreteness here, we make use of the $n \times n$ Crabb matrix: \begin{equation} \label{eq:crabb} K_2 = \begin{bmatrix} 0 & 2 \\ 0 & 0 \end{bmatrix}, \ K_3 = \begin{bmatrix} 0 & \sqrt{2} & 0 \\ 0 & 0 & \sqrt{2} \\ 0 & 0 & 0 \end{bmatrix}, \ K_n = \begin{bmatrix} 0 & \sqrt{2} & & & & \\ & . & 1 & & & \\ & & . & . & & \\ & & & . & 1 & \\ & & & & . & \sqrt{2} \\ & & & & & 0 \\ \end{bmatrix}, \end{equation} where for all $n$, $r(K_n) = 1$ and $W(K_n)$ is the unit disk. However, note that not all disk matrices are variations of Jordan blocks corresponding to the eigenvalue zero. For other types of disk matrices and the history and relevance of $K_n$, see~\cite{LewO20}. \subsection{Uhlig's method for disk matrices} The following theorem completely characterizes the total cost of Uhlig's method for disk matrices with respect to a desired relative tolerance. Note that Uhlig's method begins with a rectangular approximation~$\mathcal{G}_4$ to $\fovX{A}$, which for a disk matrix, is a square centered at the origin. \begin{theorem} \label{thm:uhlig_disk} Suppose that $A \in \mathbb{C}^{n \times n}$ is a disk matrix with $r(A) > 0$ and that~$\fovX{A}$ is approximated by $\mathcal{G}_j$ with $j \geq 3$ and $\mathcal{G}_j$ a regular polygon, i.e., it is the intersection of $j$ half planes $P_{\theta_\ell}$, where $\theta_\ell = \tfrac{2\pi }{j} \ell$ for $\ell = 1,\ldots,j$. Then, \begin{enumerate}[label=(\roman*),font=\normalfont] \item $\varepsilon_j = \sec(\pi/j) - 1$, \item if $\varepsilon_j \leq \tau_\mathrm{tol}$, then $j \geq \left\lceil \tfrac{\pi}{\arcsec(1 + \tau_\mathrm{tol})} \right\rceil$, where $\tau_\mathrm{tol} > 0$ is the desired relative error. \end{enumerate} Moreover, if $\mathcal{G}_k$ is a further refined version of $\mathcal{G}_j$, so $\fovX{A} \subseteq \mathcal{G}_k \subseteq \mathcal{G}_j$, then \begin{enumerate}[label=(\roman*),font=\normalfont,resume] \item if $\varepsilon_k < \varepsilon_j$, then $k \geq 2j$. \item if $\varepsilon_k \leq \tau_\mathrm{tol} < \varepsilon_j$, then $k \geq j \cdot 2^d$, where $d = \left\lceil \log_2 \left(\tfrac{\pi}{j \arcsec(1 + \tau_\mathrm{tol})}\right) \right\rceil$. \end{enumerate} \end{theorem} \begin{proof} As $\fovX{A}$ is a disk centered at zero with radius $r(A)$ and $\bfovX{A}$ is a circle inscribed in the regular polygon $\mathcal{G}_j$, every boundary point in $\mathcal{Z}_j$ has modulus~$r(A)$, and so $l_j = r(A)$, and the moduli of the corners of~$\mathcal{G}_j$ are all identical. Consider the corner $c$ with $\Arg(c) = \pi/j$ and the right triangle defined by zero, $r(A)$ on the real axis, and $c$. Then~\mbox{$|c| = u_j = r(A) \sec(\pi/j)$}, and so~(i) holds. Statement~(ii) simply holds by substituting~(i) into $\varepsilon_j \leq \tau_\mathrm{tol}$ and then solving for $j$. For~(iii), as~$\varepsilon_j = |c|$ for any corner~$c$ of~$\mathcal{G}_j$, all $j$~corners of $\mathcal{G}_j$ must be refined to lower the error; thus,~$\mathcal{G}_k$ must have at least~$2j$~corners. Finally, as $\lim_{j \to \infty} \varepsilon_j = 0$, but the error only decreases when~$j$ is doubled, it follows that in order for $\varepsilon_k \leq \tau_\mathrm{tol}$ to hold, $k \geq j \cdot 2^d$ for some~$d \geq 1$. The smallest possible integer is obtained by replacing~$j$ in~(ii) with $j \cdot 2^d$ and solving for~$d$, thus proving~(iv). \end{proof} \begin{table} \centering \caption{ For any disk matrix $A$, the minimum number of supporting hyperplanes required to compute $r(A)$ to different accuracies is shown, where $\texttt{eps} \approx 2.22~\times~10^{-16}$. } \setlength{\tabcolsep}{6pt} \begin{tabular}{l | *{10}{S[table-format=9.0]}} \toprule \multicolumn{1}{c}{} & \multicolumn{3}{c}{\# of supporting hyperplanes needed}\\ \cmidrule(lr){2-4} \multicolumn{1}{c}{}& \multicolumn{1}{c}{Minimum} & \multicolumn{2}{c}{Uhlig's method}\\ \cmidrule(lr){2-2} \cmidrule(lr){3-4} \multicolumn{1}{c}{Relative Tolerance} & \multicolumn{1}{c}{}& \multicolumn{1}{c}{Starting with $\mathcal{G}_3$} & \multicolumn{1}{c}{Starting with $\mathcal{G}_4$} \\ \midrule $\tau_\mathrm{tol} = \texttt{1e-4}$ & 223 & 384 & 256 \\ $\tau_\mathrm{tol} = \texttt{1e-8}$ & 22215 & 24576 & 32768 \\ $\tau_\mathrm{tol} = \texttt{1e-12}$ & 2221343 & 3145728 & 4194304 \\ $\tau_\mathrm{tol} = \texttt{eps}$ & 149078414 & 201326592 & 268435456 \\ \bottomrule \end{tabular} \label{tbl:disk_thm} \end{table} Via \cref{thm:uhlig_disk}, we report the number of supporting hyperplanes needed to compute the numerical radius of disk matrices for increasing levels of accuracy in~\cref{tbl:disk_thm}, illustrating just how quickly the cost of Uhlig's method skyrockets. Consequently, the level-set approach is typically much faster for disk matrices or those whose fields of values are close to a disk centered at zero; indeed, since $h(\theta)=\rho(H(\theta))$ is constant for disk matrices, it converges in a single iteration. \subsection{Local rate of convergence} \label{sec:uhlig_conv} As we now explain, the local behavior of Uhlig's cutting procedure at an outermost point in $\fovX{A}$ can actually be understood analyzing one key example. For this analysis, we use Q-linear and Q-superlinear convergence, where ``Q" stands for ``quotient"; see \cite[p.~619]{NocW99}. \begin{definition} Let $b_\star$ be an outermost point of $\fovX{A}$ such that the radius of curvature $\tilde r$ of~$\bfovX{A}$ is well defined at $b_\star$. Then the \emph{normalized radius of curvature of~$\bfovX{A}$ at~$b_\star$} is~\mbox{$\mu \coloneqq \tilde r/ r(A) \in [0,1]$}. \end{definition} Note that if $\mu = 0$, $b_\star$ is a corner of $\fovX{A}$. If $\mu = 1$, then near~$b_\star$, $\bfovX{A}$ is well approximated by an arc of the circle with radius $r(A)$ centered at the origin. We show that the local convergence is precisely determined by the value of $\mu$ at~$b_\star$. In the upcoming analysis we use the following assumptions. \begin{assumption} \label{asm:wlog} We assume that $r(A) > 0$ and that it is attained at a non-corner $b_\star \in \bfovX{A}$ with $\Arg(b_\star) = 0$. \end{assumption} \cref{asm:wlog} is essentially without any loss of generality. Assuming $r(A) > 0$ is trivial, as it only excludes $A=0$. Since we are concerned with finding the local rate of convergence at $b_\star$, its location does not matter, and so we can assume a convenient one, that $b_\star$ is on the positive real axis. As will be seen, our analysis does not lose any generality by assuming that $b_\star$ is not a corner. \begin{assumption} \label{asm:c2} We assume that the current approximation $\mathcal{G}_j$ has been constructed using the supporting hyperplane $L_0$ passing through~$b_\star$, and so $b_\star \in \mathcal{Z}_j$, and that~$\bfovX{A}$ is twice continuously differentiable at $b_\star$. \end{assumption} \cref{asm:c2} is also quite mild. Although Uhlig's method may sometimes only encounter supporting hyperplanes for outermost points in the limit as it converges, as we explain in \cref{sec:alg2}, they can actually be easily and cheaply obtained and used to update $\mathcal{G}_j$ and $\mathcal{Z}_j$. Moreover, knowing outermost points and associated hyperplanes does not guarantee convergence. Uhlig's method only terminates once~$\varepsilon_k$ is sufficiently small for some~$k$, which means that its cost is generally determined by how quickly it can sufficiently approximate~$\bfovX{A}$ in \emph{neighborhoods about the outermost points}; per \cref{sec:background}, when $A$ is not a disk matrix, there can be up to $n$ such points. The smoothness assumption ensures that there exists a unique osculating circle of $\bfovX{A}$ at $b_\star$, and consequently, the disagreement of the osculating circle and~$\bfovX{A}$ decays at least cubicly as~$b_\star$ is approached; for more on osculation, see, e.g., \cite[chapter~2]{Kuh15}. \begin{keyremark} \label{key:why} By our assumptions, $b_\star \in \mathcal{Z}_j$, and $\bfovX{A}$ is twice continuously differentiable and has normalized radius of curvature~$\mu > 0$ at $b_\star$. Since $\partial \mathcal{G}_j$ is a piecewise linear approximation of $\bfovX{A}$, the local behavior of a cutting-plane method is determined by the resulting second-order approximation errors, with the higher-order errors being negligible. As $\bfovX{A}$ is curved at $b_\star$, these second-order errors must be non-zero on both sides of~$b_\star$. Now recall that the osculating circle of~$\bfovX{A}$ at $b_\star$ locally agrees with~$\bfovX{A}$ to at least second order. Hence, near~$b_\star$, the second-order errors of a cutting-plane method applied to $A$ \emph{are identical} to the second-order errors of applying the method to a matrix $M$ with $\bfovX{M}$ being the same as that osculating circle. Thus, to understand the local convergence rate of a cutting-plane method for general matrices, it actually suffices to study how the method behaves on~$M$. \end{keyremark} We now define our key example $M$ such that, via two real parameters, $\bfovX{M}$ is the osculating circle at the outermost point $b_\star$ \begin{figure} \centering \subfloat[Iteration $j$.]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{uhlig_j}} \label{fig:uhlig_j} } \subfloat[Iteration $j+1$.]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{uhlig_jp1}} \label{fig:uhlig_jp1} } \caption{Depiction of Uhlig's method for \cref{ex:circ} where $z = 0.3$ and $\tilde r = 0.7$. The dotted circle is the circle of radius $r(M) = 1$ centered at the origin. See \cref{ex:circ} for a complete description of the plots. } \label{fig:uhlig} \end{figure} \defs{s} \begin{example}[See \cref{fig:uhlig} for a visual description] \label{ex:circ} For $n \geq 2$, let \begin{equation} \label{eq:model} M = s I + \tilde r K_n, \end{equation} where $K_n \in \mathbb{C}^{n \times n}$ is any disk matrix with $r(K_n)=1$, e.g., \eqref{eq:crabb}, $s \geq 0$, and~$\tilde r > 0$. Clearly, $\fovX{M}$ is a disk with radius $\tilde r$ centered at $s$ on the real axis with outermost point~$b_\star = s + \tilde r = r$ and $\mu = \tfrac{\tilde r}{r} > 0$ at $b_\star$, where~$r \coloneqq r(M)$, a shorthand we will often use in \cref{sec:rate}~and~\cref{sec:alg2}. Thus, given any matrix~$A$ satisfying \cref{asm:wlog,asm:c2} with any $\mu \in (0,1]$, by choosing $s \geq 0$ and~$\tilde r > 0$ appropriately we have that~$\bfovX{M}$ agrees exactly with the osculating circle of~$\bfovX{A}$ at~$b_\star$. Assume that~$\theta_\star = 0$ and~$\theta_j \in (-\pi,0)$ have been used to construct~$\mathcal{G}_j$ and~$\mathcal{Z}_j$, i.e., supporting hyperplanes~$L_{\theta_\star}$ and~$L_{\theta_j}$ respectively pass through boundary points~$b_\star$ and~$b_j = s + \tilde r \eix{\phi_j}$ of $\fovX{M}$, where~\mbox{$\phi_j \coloneqq \Arg (b_j - s) \in (0,\pi)$}. Therefore, $b_\star = z_0 \in \mathcal{Z}_j$ and since $\fovX{M}$ is a disk, it also follows that $b_j=z_{\theta_j} \in \mathcal{Z}_j$, where~$\theta_j = -\phi_j$. Further assume that for all~\mbox{$\theta \in (\theta_j,\theta_\star)$}, $z_\theta \not\in \mathcal{Z}_j$, and so~$c_j \coloneqq L_{\theta_\star} \cap L_{\theta_j}$ is a corner of~$\mathcal{G}_j$. Now suppose~$c_j$ is cut, by any cutting procedure, Uhlig's or otherwise. This results in a boundary point~$b_{j+1}$ of $\fovX{M}$ being added to $\mathcal{Z}_j$ and $c_j$ is replaced by two new corners, $\hat c_{j+1}$ and~$c_{j+1}$, that respectively lie on~$L_{\theta_j}$ and $L_{\theta_\star}$; due to their orientation with respect to~$b_{j+1}$, we refer to $\hat c_{j+1}$ as a counter-clockwise (CCW) corner and~$c_{j+1}$ as a clockwise~(CW) corner. If~$c_{j+1}$ is then cut next, this produces two more corners,~$\hat c_{j+2}$ and~$c_{j+2}$, with $c_{j+2}$ also on~$L_{\theta_\star}$ and between $c_{j+1}$ and~$b_\star$. Note that if~$\{\phi_k\} \to 0$, then the sequence of CW corners $\{c_k\} = c_j,c_{j+1},c_{j+2},\ldots$ converges to $b_\star$. To understand the local behavior of a cutting-plane technique, we will analyze~$\{\phi_k\}$ and~$\{c_k\}$, i.e., the case when the cuts are applied to the CW corners that are sequentially generated~on~$L_{\theta_\star}$. \end{example} Some remarks on \cref{ex:circ} are in order, as it ignores the CCW corners $\hat c_j$, and cutting these CCW corners may also introduce new corners that need to be cut. However, since $\{c_k\}$ is a subsequence of all the corners generated by Uhlig's method to sufficiently approximate $\bfovX{M}$ between boundary points~$b_\star$ and~$b_j$, analyzing $\{c_k\}$ gives a lower bound on its local efficiency. Furthermore, this often describes the true local efficiency because, as will become clear, for many problems, there are either no or few CCW corners that requiring cutting. Finally, in \cref{sec:alg2}, we introduce an improved cutting scheme that guarantees only CW corners must be cut. \begin{lemma} \label{lem:uhlig_cuts} Recalling that $\phi_j \coloneqq \Arg (b_j - s)$, if Uhlig's cuts are sequentially applied to the corners $\{c_k\}$ described in \cref{ex:circ}, then for all $k \geq j$, \begin{equation} \label{eq:uhlig_angle} \phi_{k+1} = \arctan \big( \mu \tan \tfrac{1}{2}\phi_k \big). \end{equation} \end{lemma} \begin{proof} Since $\fovX{M}$ is a disk with tangents at $b_k$ and $b_\star$ determining $c_k$, first note that the equality \mbox{$\Arg(c_k - s) = \tfrac{1}{2}\phi_k$} holds. Then \mbox{$\tan \tfrac{1}{2} \phi_k = \tilde r^{-1}|c_k - b_\star|$}, and since the tangent at $b_\star$ is vertical, we also have that $\Arg(c_k) = \arctan(r^{-1} |c_k - b_\star|)$. Thus via substitution, \[ \Arg(c_k) = \arctan ( r^{-1} \tilde r \tan \tfrac{1}{2}\phi_k ) = \arctan ( \mu \tan \tfrac{1}{2}\phi_k ). \] The proof is completed since $\phi_{k+1} = \Arg (b_{k+1} - s)$ is also equal to $\Arg(c_k)$. \end{proof} \begin{theorem} \label{thm:uhlig_angle} The sequence $\{\phi_k\}$ produced by Uhlig's cutting procedure and described by recursion \eqref{eq:uhlig_angle} converges to zero Q-linearly with rate $\tfrac{1}{2} \mu$. \end{theorem} \begin{proof} First note that $\lim_{k\to\infty} \phi_k = 0 = \phi_\star$ and $\phi_k > 0$ for all $k \geq j$. Then \[ \lim_{k \to \infty} \frac{| \phi_{k+1} - \phi_\star |}{| \phi_{k} - \phi_\star |} = \lim_{k \to \infty} \frac{\phi_{k+1}}{\phi_{k}} = \lim_{k \to \infty} \frac{\arctan \left( \mu \tan \tfrac{1}{2} \phi_k \right)}{\phi_k}. \] Since the numerator and denominator both go to zero as $k\to\infty$, the result follows by considering the continuous version of the limit: \[ \lim_{\phi \to 0} \frac{\arctan \big( \mu \tan \tfrac{\phi}{2} \big)}{\phi} = \lim_{\phi \to 0} \frac{\mu \tan \tfrac{1}{2}\phi}{\phi} = \lim_{\phi \to 0} \frac{\tfrac{1}{2}\mu \phi }{\phi} = \tfrac{1}{2} \mu, \] where the first and second equalities are obtained, respectively, using small-angle approximations $\arctan x \approx x$ and $\tan x \approx x$ for $x\approx 0$. \end{proof} While \cref{thm:uhlig_angle} tells us how quickly $\{\phi_k\}$ will converge, we really want to estimate how quickly the error $\varepsilon_j$ becomes sufficiently small. For that, we must consider how fast the moduli of the corresponding outermost corners $c_k$ converge. \begin{theorem} \label{thm:uhlig_ub} Given the sequence $\{\phi_k\}$ from \cref{thm:uhlig_angle}, the corresponding sequence~$\{|c_k|\}$ converges to $r$ Q-linearly with rate $\tfrac{1}{4} \mu^2$. \end{theorem} \begin{proof} First note that \[ \cos \phi_{k+1} = \frac{r}{|c_k|} \qquad \text{and so} \qquad |c_k| = r \sec \phi_{k+1} > r \] for all $k\geq j$. Thus, we consider the limit \[ \lim_{k \to \infty} \frac{| |c_{k+1}| - r |}{| |c_k| - r |} = \lim_{k \to \infty} \frac{r \sec \phi_{k+2} - r}{r \sec \phi_{k+1} - r} = \lim_{k \to \infty} \frac{\sec \phi_{k+1} - 1}{\sec \phi_k - 1}, \] which when substituting in $\phi_{k+1} = \arctan \big( \mu \tan \tfrac{1}{2}\phi_k \big)$ becomes \[ \lim_{k \to \infty} \frac{\sec \left( \arctan \left( \mu \tan \tfrac{1}{2}\phi_k \right)\right) - 1}{\sec \phi_k - 1}. \] Since the numerator and denominator both go to zero as $k\to\infty$, we consider the continuous version of the limit, i.e., \[ \lim_{\phi \to 0} \frac{\sec \left( \arctan \left( \mu \tan \tfrac{1}{2}\phi \right)\right) - 1}{\sec \phi - 1} = \lim_{\phi \to 0} \frac{\sec \left( \mu \tan \tfrac{1}{2}\phi \right) - 1}{\sec \phi - 1} = \lim_{\phi \to 0} \frac{\sec \left( \tfrac{1}{2}\mu \phi \right) - 1}{\sec \phi - 1}, \] again using the small-angle approximations for $\arctan$ and $\tan$. As this is an indeterminant form, we will apply L'H{\^ o}pital's Rule. However, it will be convenient to first multiply by the following identity to eliminate the term $\sec \phi$ in the denominator: \[ \lim_{\phi \to 0} \; \frac{\cos \phi}{\cos \phi} \cdot \frac{\sec \left( \tfrac{1}{2}\mu \phi \right) - 1}{\sec \phi - 1} = \lim_{\phi \to 0} \frac{\cos \phi \left( \sec \left( \tfrac{1}{2} \mu \phi \right) - 1 \right)}{1 - \cos \phi} \eqqcolon \lim_{\phi \to 0} \frac{f_1(\phi)}{f_2(\phi)}. \] For $f_2(\phi)$, $f_2^\prime(\phi) = \sin \phi$, while for $f_1(\phi)$, we have \begin{align*} f_1^\prime(\phi) &= -\sin \phi \left( \sec\left( \tfrac{1}{2}\mu \phi \right) - 1\right) + \tfrac{1}{2} \mu \cos \phi \cdot \sec\left( \tfrac{1}{2} \mu \phi \right) \cdot \tan \left( \tfrac{1}{2}\mu \phi \right)\\ &= -\sin \phi \left( \sec\left( \tfrac{1}{2}\mu \phi \right) - 1\right) + \tfrac{1}{2} \mu \cos \phi \cdot \frac{\sin \left(\tfrac{1}{2} \mu \phi \right) }{\cos^2 \left(\tfrac{1}{2}\mu \phi \right)}. \end{align*} As \mbox{$f_1^\prime(0) = 0$} and $f_2^\prime(0) = 0$, we still have an indeterminate form and so will again apply L'H{\^ o}pital's Rule. For the denominator, we have $f_2^{\prime\prime}(\phi) = \cos \phi$ and so $f_2^{\prime\prime}(0) = 1$. For the numerator, it will be convenient to write \mbox{$f_1^\prime(\phi) = -g_1(\phi) + \tfrac{1}{2}\mu g_2(\phi)$}, where \[ g_1(\phi) = \sin \phi \left( \sec\left( \tfrac{1}{2}\mu \phi \right) - 1\right) \quad \text{and} \quad g_2(\phi) = \cos \phi \cdot \frac{\sin \left(\tfrac{1}{2} \mu \phi \right)}{\cos^2 \left(\tfrac{1}{2}\mu \phi \right)}, \] and differentiate the parts separately. For $g_1(\phi)$, we have that \[ g_1^\prime(\phi) = \cos \phi \left(\sec \left(\tfrac{1}{2} \mu \phi \right) - 1\right) + \tfrac{1}{2}\mu \sin \phi \cdot \sec\left( \tfrac{1}{2}\mu \phi \right) \cdot \tan \left( \tfrac{1}{2}\mu \phi \right), \] and so $g_1^\prime(0) = 0 $. For $g_2(\phi)$, we have that \[ g_2^\prime(\phi) = -\sin \phi \cdot \frac{\sin \left( \tfrac{1}{2}\mu \phi \right) }{\cos^2 \left( \tfrac{1}{2}\mu \phi \right)} + \cos \phi \cdot \frac{\tfrac{1}{2}\mu \left( \cos^3 \left(\tfrac{1}{2}\mu \phi \right) + 2\sin^2 \left( \tfrac{1}{2}\mu \phi \right) \cdot \cos \left( \tfrac{1}{2}\mu \phi \right) \right) }{\cos^4 \left(\tfrac{1}{2}\mu \phi \right)}, \] and so $g_2^\prime(0) = \tfrac{1}{2}\mu$. Thus, $f_1^{\prime\prime}(0) = -g_1^\prime(0) + \tfrac{1}{2}\mu g_2^\prime(0) = \tfrac{1}{4}\mu^2$, proving the claim. \end{proof} As \cref{ex:circ} can model any $\mu \in (0,1]$, per \cref{key:why}, \cref{thm:uhlig_angle,thm:uhlig_ub} also accurately describe the local behavior of Uhlig's cutting procedure at outermost points in $\fovX{A}$ for any matrix $A$. Moreover, due to the squaring and one-quarter factor in \cref{thm:uhlig_ub}, the linear rate of convergence becomes very fast rather rapidly as~$\mu$ decreases from one, ultimately becoming superlinear if the outermost point is a corner ($\mu = 0$). We can also estimate the cost of approximating $\bfovX{A}$ about~$b_\star$, determining how many iterations will be needed until it is no longer necessary to refine corner $c_k$, i.e., the value of $k$ such that $|c_k| \leq r(A) \cdot ( 1 + \tau_\mathrm{tol} )$. For simplicity, it will now be more convenient to assume that~$j=0$ with $|c_0| = \betar(A)$ for some scalar~$\beta > (1 + \tau_\mathrm{tol})$. Via the Q-linear rate given by \cref{thm:uhlig_ub}, we have that \[ |c_k| - r(A) \leq (|c_0| - r(A)) \cdot \left(\tfrac{1}{4}\mu^2\right)^k, \] and so if \[ r(A) + (|c_0| - r(A)) \cdot \left(\tfrac{1}{4}\mu^2\right)^k \leq r(A) \cdot ( 1 + \tau_\mathrm{tol} ), \] then it follows that $|c_k| \leq r(A) \cdot ( 1 + \tau_\mathrm{tol} )$, i.e., it does not need to be refined further. By first dividing the above equation by $r(A)$ and doing some simple manipulations, we have that $|c_k|$ is indeed sufficiently close to $r(A)$~if \begin{equation} k \geq \frac{\log (\tau_\mathrm{tol}) - \log (\beta - 1) }{\log \left(\tfrac{1}{4}\mu^2\right)}. \end{equation} Using \cref{ex:circ} with $\beta = 100$, and $\tau_\mathrm{tol} = \texttt{1e-14}$, only $k \approx 27$, $14$, $7$, and $4$ iterations are needed, respectively, for $\mu = 1$, $0.5$, $0.1$, and $0.01$. This is indeed rather fast for linear convergence. Of course, if $\fovX{A}$ has more than one outermost point, the total cost of a cutting-plane method increases commensurately, since $\bfovX{A}$ must be well approximated about all of these outermost points. For disk matrices, all boundary points are outermost, and so the cost blows up, per \cref{thm:uhlig_disk}. \section{An improved cutting-plane algorithm} \label{sec:alg2} We now address some inefficiencies in Uhlig's method by giving an improved cutting-plane method. The two main components of this refined algorithm are as follows. First, any of the local optimization techniques from \cref{sec:alg1} also allows us to more efficiently locate outermost points in~$\fovX{A}$. This is possible because each outermost point is bracketed on~$\bfovX{A}$ by two boundary points of $\fovX{A}$ in $\mathcal{Z}_j$, and these brackets improve as $\mathcal{G}_j$ more accurately approximates $\fovX{A}$. Therefore, once $\mathcal{G}_j$ is no longer a crude approximation, these brackets can be used to initialize optimization to find global maximizers of~$h(\theta)$, and thus, globally outermost points of~$\fovX{A}$. Second, given a boundary point of $\fovX{A}$ that is also known to be locally outermost, we use a new cutting procedure that reduces the total number of cuts needed to sufficiently approximate $\bfovX{A}$ in this region. When this new cut cannot be invoked, we will fall back on Uhlig's cutting procedure. In the next three subsections, we describe our new cutting strategy, establish a Q-linear rate of convergence for it, and finally, show how these cuts can be sufficiently well estimated so that our theoretical convergence rate result is indeed realized in practice. Finally, pseudocode of our completed algorithm is given in~\cref{alg2}. \subsection{An optimal-cut strategy} \label{sec:opt_cut} Again consider \cref{ex:circ}. In \cref{fig:uhlig_jp1}, Uhlig's cut of corner $c_j$ between $b_j$ (with $|b_j| < r$) and $b_\star$ produces two new corners~$\hat c_{j+1}$ and $c_{j+1}$, but since $|\hat c_{j+1}| < r$ and $|c_{j+1}| > r$, it is only necessary to subsequently refine $c_{j+1}$. However, in \cref{fig:uhlig_two} we show another scenario where both of the two new corners produced by Uhlig's cut will require subsequent cutting as well. While \cref{thm:uhlig_angle,thm:uhlig_ub} indicate the number of iterations Uhlig's method needs to sufficiently refine the sequence $\{c_k\}$, they do not take into account that the CCW corners that are generated may also need to be cut. Thus, the total number of eigenvalue computations with $H(\theta)$ can be higher than what is suggested by these two theorems. However, comparing \cref{fig:uhlig_jp1,fig:uhlig_two} immediately suggests a better strategy, namely, to make the largest reduction in the angle $\phi_j$ such that the CCW corner~$\hat c_{j+1}$ (between $b_j$ and $c_j$ on the tangent line for $b_j$) does not subsequently need to be refined, i.e., such that $|\hat c_{j+1}| = r$. In \cref{fig:uhlig_two}, this ideal corner is labeled $d_j$, while the corresponding optimal cut for this same example is shown in \cref{fig:opt_cut}, where $d_j$ coincides with~$\hat c_{j+1}$, and so the latter is not labeled. \subsection{Convergence analysis of the optimal cut} \label{sec:opt_conv} Before describing how to compute optimal cuts, we derive the convergence rate of the sequence of angles~$\{\phi_k\}$ this strategy produces. Per \cref{key:why}, it again suffices to study \cref{ex:circ}. \begin{figure}[t] \centering \subfloat[Uhlig's cut.]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{cut_uhlig}} \label{fig:uhlig_two} } \subfloat[The optimal cut.]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{cut_optimal}} \label{fig:opt_cut} } \caption{Depictions of corner $c_j$ between boundary points $b_j$ and $b_\star$ being cut by Uhlig's procedure (left) and the optimal cut (right), where the latter always makes the largest possible reduction in $\phi_j$ such that only corner $c_{j+1}$ must be refined. } \label{fig:cuts_comp} \end{figure} \begin{lemma} \label{lem:dj} Given \cref{ex:circ}, additionally assume that $|b_j| < r$ and $\tilde r < r$, and so at $b_\star$, $\mu \in (0,1)$. Then the point on~$L_{\theta_j}$, the supporting hyperplane passing through $b_j$ with \mbox{$\theta_j = -\phi_j \in (-\pi,0)$}, that is closest to $c_j$ and has modulus $r$ is \begin{subequations} \begin{align} \label{eq:dlem} d_j &= b_j - \mathbf{i} t_j \eix{\phi_j}, \quad \text{where} \\ \label{eq:tlem} t_j &= -s \sin \phi_j + \sqrt{s^2 \sin^2 \phi_j + 2s \tilde r(1 - \cos \phi_j)} > 0. \end{align} \end{subequations} \end{lemma} \begin{proof} As $\phi_j \in (0,\pi)$, clearly \eqref{eq:dlem} must hold for some~$t_j > 0$. To obtain~\eqref{eq:tlem}, we use the fact that $|d_j|^2 = r^2$ and solve for $t_j$ using \eqref{eq:dlem}. Setting~$u = \eix{\phi_j}$, we have \begin{align*} 0 &= |d_j|^2 - r^2 = (b_j - \mathbf{i} t_j u) \big(\overline b_j + \mathbf{i} t_j \overline{u}) - r^2 = t_j^2 + \mathbf{i} (b_j \overline{u} - \overline b_j u)t_j + |b_j|^2 - r^2. \end{align*} which by substituting in the following two equivalences \begin{align*} b_j \overline{u} - \overline b_j u &= (s + \tilde r u) \overline{u} - (s + \tilde r \overline{u}) u = s (\overline{u} - u) = -\mathbf{i} 2s \sin \phi_j, \\ |b_j|^2 &= s^2 + \tilde r^2 + s \tilde r(u + \overline{u}) = s^2 + \tilde r^2 + 2s \tilde r \cos \phi_j, \end{align*} yields \[ 0 = t_j^2 + (2 s \sin \phi_j) t_j + (s^2 + \tilde r^2 - r^2 + 2s\tilde r \cos \phi_j). \] By substituting in $r^2 = (s + \tilde r)^2 = s^2 + \tilde r^2 + 2s\tilde r$, this simplifies further to \[ 0 = t_j^2 + (2 s \sin \phi_j) t_j + 2s \tilde r (\cos \phi_j - 1). \] Thus, by the quadratic formula, we obtain \eqref{eq:tlem}. \end{proof} \begin{lemma} \label{lem:opt_angle} Given the assumptions in \cref{lem:dj} and $t_j > 0$ from \eqref{eq:tlem}, also suppose that \mbox{$\phi_j \in (0,\tfrac{\pi}{2})$}. Then if optimal cuts are sequentially applied to the corners~$\{c_k\}$ described in \cref{ex:circ}, for all $k \geq j$, \begin{equation} \label{eq:opt_thetajp1} \phi_{j+1} = -\phi_j + 2\arctan \left( \frac{\tilde r \sin \phi_j - t_j \cos \phi_j}{\tilde r \cos \phi_j + t_j\sin\phi_j} \right). \end{equation} \end{lemma} \begin{proof} Let $\hat \phi = \phi_j - \Arg(d_j - s)$. Then it follows that \begin{equation} \label{eq:theta_jp1_proof} \phi_{j+1} = \phi_j - 2\hat \phi = -\phi_j + 2\Arg(d_j - s) = - \phi_j + 2\arctan \left( \frac{\Im (d_j - s)}{\Re (d_j - s)} \right), \end{equation} where the last equality follows because $\Re (d_j - s) > 0$, as $\Re b_j > s$. Using \eqref{eq:dlem} and substituting in $b_j = s + \tilde r\mathrm{e}^{\mathbf{i} \phi_j}$, we have that \[ d_j - s = \tilde r \mathrm{e}^{\mathbf{i} \phi_j} - \mathbf{i} t_j \mathrm{e}^{\mathbf{i} \phi_j} \qquad \Longleftrightarrow \qquad \begin{aligned} \Re (d_j - s) &= \tilde r \cos \phi_j + t_j \sin \phi_j,\\ \Im (d_j - s) &= \tilde r \sin \phi_j - t_j \cos \phi_j. \end{aligned} \] Substituting these into~\eqref{eq:theta_jp1_proof} completes the proof. \end{proof} Before deriving how fast $\{\phi_k\}$ converges, we show that it indeed converges to zero. \begin{lemma} \label{lem:opt_zero} Given the assumptions of \cref{lem:opt_angle}, the recursion \eqref{eq:opt_thetajp1} for optimal cuts produces a sequence of angles $\{\phi_k\}$ converging to zero. \end{lemma} \begin{proof} By construction, $\{\phi_k\}$ is monotone, i.e., $\phi_{k+1} < \phi_k$ for all $k$, and bounded below by zero, and so $\{\phi_k\}$ converges to some limit $l$. Per \eqref{eq:tlem}, the $t_j$ values appearing in \eqref{eq:opt_thetajp1} depend on $\phi_j$, so we define the analogous continuous function \begin{equation} \label{eq:t_cont} t(\phi) = -s \sin \phi + \sqrt{s^2 \sin^2 \phi + 2s \tilde r (1 - \cos \phi )}. \end{equation} Now by way of contradiction, assume that $l > 0$ and so $0 < l < \phi_j < \tfrac{\pi}{2}$. Thus, \[ \lim_{k\to\infty} \phi_{k+1} = l = -l + 2\arctan \left( \frac{\tilde r \sin l - t(l) \cos l}{\tilde r \cos l + t(l) \sin l} \right) \ \Leftrightarrow \ \tan l = \frac{\tilde r \sin l - t(l) \cos l}{\tilde r \cos l + t(l) \sin l}. \] By multiplying both sides by $\tilde r \cos l + t(l) \sin l$ and rearranging terms, we obtain the equality \mbox{$t(l)(\sin^2 l + \cos^2 l) = 0$}, and so $t(l) = 0$. However, \cref{lem:dj} states that $t(l) > 0$ should hold since $l \in (0,\tfrac{\pi}{2})$, a contradiction, and so $l=0$. \end{proof} We now have the necessary pieces to derive the exact rate of convergence of the angles produced by optimal cuts. \begin{theorem} \label{thm:opt_conv} The sequence $\{\phi_k\}$ produced by optimal cuts and described by recursion \eqref{eq:opt_thetajp1} converges to zero Q-linearly with rate $\tfrac{2(1 - \sqrt{1 - \mu})}{\mu} - 1$. \end{theorem} \begin{proof} By \cref{lem:opt_angle,lem:opt_zero}, \eqref{eq:opt_thetajp1} holds, $\phi_k \to \phi_\star = 0$, and $\phi_k \geq 0$ for all $k$, so \[ \lim_{k \to \infty} \frac{|\phi_{k+1} - \phi_\star|}{| \phi_{k} - \phi_\star |} = \lim_{k \to \infty} \frac{\phi_{k+1}}{\phi_{k}} = \lim_{k \to \infty} \frac{ -\phi_k + 2\arctan \left( \frac{\tilde r \sin \phi_k - t_k \cos \phi_k}{\tilde r \cos \phi_k + t_k \sin\phi_k} \right) }{\phi_k}. \] Using the continuous version of $t_k$ given in \eqref{eq:t_cont}, we instead consider the entire limit in continuous form: \begin{equation} \label{eq:opt_lim_cont} \lim_{\phi \to 0} \frac{ -\phi + 2\arctan \left( \frac{\tilde r \sin \phi - t(\phi) \cos \phi}{\tilde r \cos \phi + t(\phi) \sin\phi} \right) }{\phi} = -1 + \lim_{\phi \to 0} \frac{2}{\phi} \cdot \frac{ \tilde r \phi - t(\phi)\cos \phi}{\tilde r \cos\phi + t(\phi) \phi}, \end{equation} where the equality holds by using the small-angle approximations $\arctan x \approx x$ (as the ratio inside the $\arctan$ above goes to zero as $\phi \to 0$) and $\sin x \approx x$. Again using $\sin x \approx x$ as well as the small-angle approximation $1 - \cos x \approx \tfrac{1}{2}x^2$, we also have the small-angle approximation \begin{equation} \label{eq:t_approx} t(\phi) \approx -s \phi + \sqrt{s^2 \phi^2 + 2s\tilde r \big(\tfrac{1}{2}\phi^2\big)} = -\phi \left(s - \phi \sqrt{s^2 + s\tilde r}\right) = -\phi \left(s - \sqrt{s r} \right), \end{equation} where the last equality holds since $\tilde r = r - s$. Via substituting in \eqref{eq:t_approx}, the limit on the right-hand side of \eqref{eq:opt_lim_cont} is \[ \lim_{\phi \to 0} \frac{2}{\phi}\cdot \frac{\tilde r \phi + \phi \left(s - \sqrt{s r} \right) \cos \phi }{\tilde r \cos \phi - \phi \left(s - \sqrt{s r} \right) \phi } = \lim_{\phi \to 0} \frac{2 \left( \tilde r + \left(s - \sqrt{s r} \right) \cos \phi \right) }{\tilde r \cos \phi - \phi^2 \left(s - \sqrt{s r} \right) } \\ = \frac{ 2 \left( \tilde r + s - \sqrt{s r} \right) }{\tilde r}. \] Recalling that $\tilde r = \mu r$ and that $s = r - \tilde r = r - \mu r$, by substitutions we can rewrite the ratio above as \[ \frac{2\left(\mu r + (r - \mu r) - \sqrt{(r - \mu r) r} \right) }{\mu r} = \frac{2\left(r - r \sqrt{1 - \mu} \right) }{\mu r} = \frac{2\left(1 - \sqrt{1 - \mu} \right) }{\mu}. \] Subtracting one from the value above completes the proof. \end{proof} As we show momentarily, optimal cuts have a total lower cost than Uhlig's cutting procedure. Thus, there is no need to derive an analogue of \cref{thm:uhlig_ub} for describing the convergence rate of the moduli of corners~$c_k$ produced by the optimal-cut strategy. \subsection{Computing the optimal cut} \label{sec:compute_opt} Suppose that $b_\star \in \mathcal{Z}_j$ attains the value of~$l_j$, and that $b_\star$ is also locally outermost in~$\fovX{A}$, and let \mbox{$\gamma = |b_\star| \leq r(A)$}. Without loss of generality, we assume that $\Arg(b_\star) = 0$, and let $b_j \in \mathcal{Z}_j$ be the next known boundary point of $\fovX{A}$ with $\Arg(b_j) \in (0,\pi)$. We can model $\bfovX{A}$ between~$b_\star$ and~$b_j$ by fitting a quadratic that interpolates $\bfovX{A}$ at~$b_\star$ and~$b_j$. If this model is a good fit, then it can be used to estimate~$d_j$, and thus, also the optimal cut. Since $b_\star$ is also a locally outermost point of $\fovX{A}$ and $\Arg(b_\star) = 0$, we can interpolate these boundary points using the sideways quadratic (opening up to the left in the complex plane) \[ q(y) = q_2 y^2 + q_1 y + q_0, \] with the remaining degree of freedom used to specify that $q(y)$ should be tangent to~$\fovX{A}$ at~$b_\star$. Clearly, $q(y)$ cannot be a good fit if $L_{\theta_j}$, the supporting hyperplane passing through $b_j$, is increasing from left to right in the complex plane; hence, we also assume that \mbox{$\theta_j \in (-\tfrac{\pi}{2},0)$}. Let~\mbox{$\theta_\dagger \in (\theta_j,0)$} denote the angle of the supporting hyperplane for the optimal cut, e.g., for \cref{ex:circ}, the one that passes through $d_j$ and the boundary point $b_{j+1}$ between $b_j$ and $b_\star$. By our criteria, the equations \begin{equation} q(0) = \gamma, \quad q(\Im b_j) = \Re b_j, \quad \text{and} \quad q^\prime(0) = 0 \end{equation} determine the coefficients $q_0$, $q_1$, and $q_2$, and solving these yields \begin{equation} q_2 = \frac{ \Re b_j - \gamma}{(\Im b_j)^2}, \quad q_1 = 0, \quad \text{and} \quad q_0 = \gamma. \end{equation} \begin{algfloat}[t] \begin{algorithm}[H] \floatname{algorithm}{Algorithm} \caption{An Improved Cutting-Plane Algorithm} \label{alg2} \begin{algorithmic}[1] \REQUIRE{ $A \in \mathbb{C}^{n \times n}$ with $n \geq 2$ and $\tau_\mathrm{tol} > 0$. } \ENSURE{ $l$ such that $|l -r(A)| \leq \tau_\mathrm{tol} \cdot r(A)$. \\ \quad } \STATE $\psi \gets \Arg(\lambda)$ where $\lambda \in \Lambda(A)$ attains $\rho(A)$ \COMMENT{0 if $\lambda = 0$} \STATE $\mathcal{G} \gets P_{\theta_1} \cap \cdots \cap P_{\theta_4}$ where $\theta_\ell = \tfrac{\ell -1 }{2}\pi - \psi$ for $\ell = 1,2,3,4$ \STATE $\mathcal{Z} \gets \{ z_{\theta_\ell} : \ell = 1,2,3,4\}$ \STATE $l \gets \max \{ |b| : b \in \mathcal{Z} \}$, $u \gets \max \{ |c| : c \text{ a corner of } \mathcal{G} \}$ \WHILE {$\tfrac{ u - l }{l} > \tau_\mathrm{tol}$ } \STATE $L_{\theta_1} \gets$ supporting hyperplane for the boundary point in $\mathcal{Z}$ attaining~$l$ \STATE $\gamma \gets$ local max of $\rho(H(\theta))$ via optimization initialized at $\theta_1$ \STATE $\mathcal{G} \gets \mathcal{G} \cap P_{\theta_1} \cap \cdots \cap P_{\theta_q}$ for the $q$ angles $\theta_\ell$ encountered during optimization \STATE $\mathcal{Z} \gets \mathcal{Z} \cup \{ z_{\theta_\ell} : \ell = 1,\ldots,q\}$ \STATE $c \gets \text{ outermost corner of } \mathcal{G}$ \IF { the optimal cut should be applied to $c$ per \cref{sec:compute_opt} } \STATE $\theta \gets$ angle $\theta_\dagger$ is given by \eqref{eq:opt_cut_angle} (rotated and flipped as necessary) \ELSE \STATE $\theta \gets -\Arg(c)$ \COMMENT{Uhlig's cut} \ENDIF \STATE $\mathcal{G} \gets \mathcal{G} \cap P_\theta$ \STATE $\mathcal{Z} \gets \mathcal{Z} \cup \{ z_{\theta} \}$ \STATE $l \gets \max \{ |b| : b \in \mathcal{Z} \}$, $u \gets \max \{ |c| : c \text{ a corner of } \mathcal{G} \}$ \ENDWHILE \end{algorithmic} \end{algorithm} \algnote{ For simplicity, we forgo describing pseudocode to exploit possible normality of~$A$ or symmetry of $\fovX{A}$, and assume that $A\neq0$, eigenvalues and local maximizers are obtained exactly, optimization is monotonic, i.e., $l \leq \gamma$ is guaranteed, and there are no ties for the boundary point in line~6. Note that $\psi$ just specifies the rotation of the initial rectangular bounding box~$\mathcal{G}$ for~$\fovX{A}$. } \end{algfloat} We can assess whether $q(y)$ is a good fit for $\bfovX{A}$ about $b_\star$ by checking how close $q(y)$ is to being tangent to $\bfovX{A}$ at $b_j$, i.e., $q(y)$ is a good fit if \begin{equation} q^\prime(\Im b_j) \approx \tan \theta_j. \end{equation} If these two values are not sufficiently close, then we consider $q(y)$ a poor local approximation of $\bfovX{A}$ at $b_j$ (and $b_\star$) and use Uhlig's cutting procedure to update $\mathcal{G}_j$ and $\mathcal{Z}_j$. Otherwise, we assume that $q(y)$ does accurately model $\bfovX{A}$ in this region and do an optimal cut. To estimate $\theta_\dagger$, we need to determine the line \[ a(y) = a_1 y + a_0 \] such that $a(y)$ passes through $d_j$ for $y = \Im d_j$ and is tangent to $q(y)$ for some \mbox{$\tilde y \in (0,\Im d_j)$}. Thus, we solve the following set of equations: \begin{subequations} \begin{alignat}{4} \label{eq:lin1} \Re d_j &= a(\Im d_j) && \qquad \Longleftrightarrow \qquad & \Re d_j &= a_1 \Im d_j + a_0,\\ \label{eq:lin2} q(\tilde y) &= a(\tilde y) && \qquad \Longleftrightarrow \qquad & q_2 \tilde y^2 + q_0 &= a_1 \tilde y + a_0, \\ \label{eq:lin3} q^\prime(\tilde y) &= a^\prime(\tilde y) && \qquad \Longleftrightarrow \qquad & 2q_2 \tilde y &= a_1, \end{alignat} \end{subequations} to determine $a_0$, $a_1$ and $\tilde y$. This yields \begin{equation} \label{eq:lincoeffs} \tilde y = \Im d_j - \sqrt{(\Im d_j)^2 + \tfrac{q_0 - \Re d_j }{q_2}}, \ \ \ a_0 = -q_2 \tilde y^2 + q_0, \ \ \ \text{and} \ \ \ a_1 = \frac{\Re d_j - a_0}{\Im d_j}, \end{equation} where $a_1$ follows directly from \eqref{eq:lin1}, $a_0$ is obtained by substituting the value of $a_1$ given in~\eqref{eq:lin3} into \eqref{eq:lin2}, and $\tilde y$ follows from substituting the value of $a_0$ given in \eqref{eq:lincoeffs} into $a_1$ in \eqref{eq:lincoeffs} (so that $a_1$ now only has $\tilde y$ as an unknown), and then substituting this version of $a_1$ into \eqref{eq:lin3}, which results in a quadratic equation in~$\tilde y$. Since $q(y)$ is a sufficiently accurate local model of $\bfovX{A}$, it follows that \begin{equation} \label{eq:opt_cut_angle} \theta_\dagger \approx \arctan a_1. \end{equation} If $\gamma = r(A)$, we can also estimate the value of $\mu$ at~$b_\star$ via \begin{equation} \label{eq:mu_est} \mu_\mathrm{est} \coloneqq \frac{1}{2|q_2|\gamma}, \end{equation} as the osculating circle of $q(y)$ at $y=0$ has radius $\tfrac{1}{2}|q_2|$. While the value of $\mu$ at $b_\star$ might be computed using \cref{thm:curvature}, this would be much more expensive and it requires that $\lambda_\mathrm{max}(H(0))$ be simple, which may not hold. Detecting the normalized radius of curvature at outermost points via~\eqref{eq:mu_est} will be a key component of our hybrid algorithm in \cref{sec:hybrid}. Our formulas for computing $\theta_\dagger$ can be used for any outermost point simply by rotating and flipping the problem as necessary to satisfy the assumptions on~$b_\star$ and~$b_j$. To be robust against even small rounding errors, instead of $d_j$, we use \mbox{$(1 - \delta) d_j + \delta b_j$} for some small $\delta \in (0,1)$, i.e., a point slightly closer to $b_j$. For different values of $\mu \in [0,1)$, \cref{fig:conv_rates} plots the convergence rates for $\{\phi_k\}$ given by \cref{thm:uhlig_angle,thm:opt_conv}, while \cref{fig:total_cuts} shows the total number of cuts needed by each cutting strategy in order to sufficiently approximate~$\bfovX{M}$ near~$b_\star$. Uhlig's method is usually slightly more expensive but becomes significantly worse than optimal cutting for normalized curvatures $\mu \approx 0.84$ and higher, requiring about double the number of cuts at this transition point. A variant of \cref{fig:total_cuts} (not shown) also reveals that optimal cuts become slightly more expensive than Uhlig's cuts for~$\mu \approx 0.999961$ and~above, and so we only use the optimal cut when $\mu_\mathrm{est}$ is less than this value. \begin{figure} \centering \subfloat[Convergence rates of $\{\phi_k\}$.]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{convergence_rates}} \label{fig:conv_rates} } \subfloat[Cost to approximate $\bfovX{M}$ region.]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{total_cuts_disk}} \label{fig:total_cuts} } \caption{ Left: the respective convergence rates of $\{\phi_j\}$ given by \cref{thm:uhlig_angle,thm:opt_conv}. Right: for tolerance \mbox{$\tau_\mathrm{tol} = \texttt{1e-14}$}, the total number of cuts required to sufficiently approximate the region of~$\bfovX{M}$ specified by the supporting hyperplanes $L_{\theta_\star}$ and~$L_{\theta_j}$ respectively passing through $b_\star$ and $b_j$ and the corner $c_j$, where $\theta_\star = 0$ and $\theta_j = -\tfrac{\pi}{50}$. } \label{fig:rates_comp} \end{figure} \section{A hybrid algorithm} \label{sec:hybrid} \cref{tbl:eig} and our new analyses suggest that it would be far more efficient to combine level-set and cutting-plane techniques in a single hybrid algorithm rather than rely on either technique alone. For the smallest values of~$n$, the level-set approach would generally be most efficient, while for larger problem sizes, which approach would be fastest depends on the specific shape of $\fovX{A}$ and the normalized radius of curvature at outermost points. While we cannot know these things \emph{a priori}, per \eqref{eq:mu_est}, \cref{alg2} can estimate $\mu$ as it iterates, and so we can predict how many more cuts may be needed about a particular outermost point. The current approximation~$\mathcal{G}$ and $\mathcal{Z}$ can also be used to obtain cheaply computed estimates of how many more cuts will be needed to approximate regions of $\bfovX{A}$ to sufficient accuracy. Consequently, as \cref{alg2} iterates, we can maintain an evolving estimate of how many more cuts would be needed in order to compute $r(A)$ to the desired tolerance. Thus, our hybrid algorithm can automatically determine if the cutting-plane approach is likely to be fast or slow, and if the latter, automatically switch to the level-set approach. For example, in practice, \cref{alg1} often only requires one to two eigenvalue computations with $R_\gamma - \lambda S$ and several more with~$H(\theta)$. Hence, in conjunction with tuning/benchmark data, such as that shown in \cref{tbl:eig}, our hybrid algorithm can reliably estimate whether it will be faster to continue \cref{alg2} or immediately switch to \cref{alg1}, which will be warm-started using the angle of the supporting hyperplane that passes through the point in $\mathcal{Z}$ that attains $l$ in line~6, as well as the arguments of the corners to the left and right of this point. \section{Numerical validation} \label{sec:experiments} Experiments were done in MATLAB\ R2021a on a 2020 13" MacBook Pro with an Intel i5 1038NG7 quad-core CPU laptop, 16GB of RAM, and macOS v10.15. For each problem, all computed estimates for $r(A)$ agreed to at least 14 digits, the tolerance we set, so we only show cost comparisons here. We used \texttt{eig} for all eigenvalue computations\footnote{In practice, note the following recommendations. As $n$ increases, \texttt{eigs} should be preferred over \texttt{eig} for computing eigenvalues of $H(\theta)$, but this can be determined automatically via tuning. Relatedly, we suggest using \texttt{eigs} with $\texttt{k} > 1$ for robustness, as the desired eigenvalue may not always be the first to converge. For robustly identifying all the unimodular eigenvalues of $R_\gamma - \lambda S$, it is generally recommended to use structure-preserving eigensolvers, e.g., \cite{BenBMetal02,BenSV16}. } as (a) it sufficed to verify the benefits of our new methods and theoretical results and (b) this consistency simplifies the comparisons; e.g., \texttt{numr}, Mengi's implementation of his level-set method with Overton, also only uses \texttt{eig}. All code and data are included as supplementary material for reproducibility. Implementations of our new methods will be added to ROSTAPACK~\cite{rostapack}. We begin by comparing \cref{alg1} to \texttt{numr}. Per \cref{tbl:level}, \cref{alg1} generally only needed a single eigenvalue computation with $R_\gamma - \lambda S$ and at most two, and for $n \geq 300$, ranged from 4.1--6.5 times faster than \texttt{numr}. Even with optimization disabled, our iteration using $\rho(H(\theta))$ was still faster than~\texttt{numr}. \begin{table} \centering \caption{ For dense random $A$ matrices, the costs of \cref{alg1} and Mengi and Overton's method (\texttt{numr}) are shown. We tested \cref{alg1} with optimization enabled (Opt., done via Newton's method) and disabled (MP, for midpoints only). } \setlength{\tabcolsep}{6pt} \begin{tabular}{r SSc | ccc | SSS} \toprule \multicolumn{1}{c}{} & \multicolumn{3}{c}{\# of \texttt{eig($H(\theta)$)}} & \multicolumn{3}{c}{\# of \texttt{eig($R_\gamma$,$S$)}} & \multicolumn{3}{c}{Time (sec.)} \\ \cmidrule(lr){2-4} \cmidrule(lr){5-7} \cmidrule(lr){8-10} \multicolumn{1}{c}{} & \multicolumn{2}{c}{Alg.~\ref{alg1}} & \multicolumn{1}{c}{\texttt{numr}} & \multicolumn{2}{c}{Alg.~\ref{alg1}} & \multicolumn{1}{c}{\texttt{numr}} & \multicolumn{2}{c}{Alg.~\ref{alg1}} & \multicolumn{1}{c}{\texttt{numr}} \\ \cmidrule(lr){2-3} \cmidrule(lr){4-4} \cmidrule(lr){5-6} \cmidrule(lr){7-7} \cmidrule(lr){8-9} \cmidrule(lr){10-10} \multicolumn{1}{c}{$n$} & \multicolumn{1}{c}{Opt.} & \multicolumn{1}{c}{MP} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{Opt.} & \multicolumn{1}{c}{MP} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{Opt.} & \multicolumn{1}{c}{MP} & \multicolumn{1}{c}{} \\ \midrule \multicolumn{1}{r|}{ 100} & 9 & 10 & 39 & 1 & 4 & 6 & 0.1 & 0.1 & 0.2 \\ \multicolumn{1}{r|}{ 200} & 14 & 11 & 34 & 2 & 4 & 5 & 0.6 & 0.9 & 1.2 \\ \multicolumn{1}{r|}{ 300} & 17 & 9 & 26 & 1 & 4 & 5 & 1.1 & 3.6 & 4.4 \\ \multicolumn{1}{r|}{ 400} & 17 & 15 & 42 & 1 & 5 & 7 & 2.5 & 10.7 & 14.9 \\ \multicolumn{1}{r|}{ 500} & 7 & 8 & 54 & 1 & 3 & 7 & 4.3 & 12.0 & 28.0 \\ \multicolumn{1}{r|}{ 600} & 13 & 10 & 45 & 1 & 4 & 7 & 8.0 & 28.6 & 49.7 \\ \bottomrule \end{tabular} \label{tbl:level} \end{table} We now verify that our local convergence rate analyses from \cref{sec:uhlig_conv} and \cref{sec:opt_conv} do indeed hold for general matrices and that our procedure for computing optimal cuts is sufficiently accurate to realize the convergence rate given by \cref{thm:opt_conv}. First, we obtained 200 general examples with roughly equally spaced values of~\mbox{$\mu \in [0,1]$}. This was done by running optimization on $\min_{X} r(A+BXC)$, where $A \in \mathbb{C}^{10 \times 10}$ is diagonal, while $B \in \mathbb{C}^{10 \times 5}$, $C \in \mathbb{C}^{5 \times 10}$, and $X \in \mathbb{R}^{5 \times 5}$ are dense, and collecting the iterates. By starting at $X=0$, we obtain an example with~$\mu=0$; since~$A$ is diagonal, $\fovX{A}$ is a polygon. Since minimizing $r(A+BXC)$ often causes $\mu \to 1$ as optimization progresses~\cite{LewO20}, we also obtain a sequence of examples $A+BX_kC$ for iterates $\{X_k\}$ with various~$\mu$~values (computed via \cref{thm:curvature}). Generating new $A$, $B$, and $C$ matrices and running optimization from $X_0=0$ was repeated in a loop until the desired set of 200 general examples had been obtained. For each problem, we recorded the total number of cuts that Uhlig's cutting procedure and the optimal-cut strategy needed to sufficiently approximate the field of values boundary in a small neighborhood to one side of the outermost point in its field of values. More specifically, we performed an analogous experiment to the one we showed earlier for approximating a region of the boundary of \cref{ex:circ}. As can be seen by comparing \cref{fig:total_cuts,fig:alg2_exp}, for any given $\mu$, the total number of cuts needed on arbitrarily shaped fields of values is essentially the same as that needed for \cref{ex:circ}, thus validating the generality of our convergence rate analysis and the reliability of our method for computing optimal cuts. \begin{figure}[t] \centering \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{total_cuts_sof}} \caption{ For 200 arbitrarily shaped fields of values with roughly equally spaced values of $\mu \in [0,1]$ at outermost points, Uhlig's cutting procedure and optimal cuts were used to approximate the boundary region specified by supporting hyperplanes $L_{\theta_\star}$ and $L_{\theta_j}$ and the corner $c_j$ of $\mathcal{G}_j$ that they define, where~$L_{\theta_\star}$ passes through the outermost point~$b_\star$ and $\theta_j = \theta_\star - \tfrac{\pi}{50}$. See \cref{fig:total_cuts} for the analogous experiments for approximating the same-sized region of~$\bfovX{M}$ in the sense that~$\theta_\star - \theta_j = \tfrac{\pi}{50}$. } \label{fig:alg2_exp} \end{figure} For comparing our improved level-set and cutting-plane methods, we also set \cref{alg2} to do optimization via Newton's method, and per \cref{rem:two_hp}, had it add supporting hyperplanes for both $\lambda_\mathrm{max}$ and $\lambda_\mathrm{min}$ on every cut. For test problems, we used the Gear, Grcar, and FM examples used by Uhlig in \cite{Uhl09}, \texttt{randn}-based complex matrices, and $\mathrm{e}^{\mathbf{i} 0.25\pi}((1 - \mu) I + \mu K_n)$ with \mbox{$\mu=0.9999$} and $K_n$ from~\eqref{eq:crabb}, which is a rotated version of \cref{ex:circ} that we call Nearly Disk. In~\cref{tbl:both}, we again see that \cref{alg1} is well optimized in terms of its overall possible efficiency, as it often only required a single computation with $R_\gamma - \lambda S$ and at most two. As predicted by our analysis, we also see that the cost of \cref{alg2} is highly correlated with the value of $\mu$. On Gear ($\mu \approx 0$), \cref{alg2} was extremely fast, essentially showing Q-superlinear convergence. In fact, \cref{alg2} was much faster (2 to 87 times) than \cref{alg1} on Gear, Grcar, FM, and \texttt{randn}, as $\mu < 0.9$ for all of these problems. In contrast, for Nearly Disk ($\mu=0.9999$), \cref{alg2} was noticeably slower, with our level-set approach now being 5.9 to~13.2 times faster. \begin{table} \centering \caption{ The respective costs of \cref{alg1,alg2} are shown. The values of $\mu$ at outermost points are also shown, computed via \cref{thm:curvature}. } \setlength{\tabcolsep}{6pt} \begin{tabular}{lrc | rc | r | SS } \toprule \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & \multicolumn{3}{c}{\# of calls to \texttt{eig($\cdot$)}} & \multicolumn{2}{c}{Time (sec.)} \\ \cmidrule(lr){4-6} \cmidrule(lr){7-8} \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & \multicolumn{2}{c}{Alg.~\ref{alg1}} & \multicolumn{1}{c}{Alg.~\ref{alg2}} & \multicolumn{1}{c}{Alg.~\ref{alg1}} & \multicolumn{1}{c}{Alg.~\ref{alg2}} \\ \cmidrule(lr){4-5} \cmidrule(lr){6-6} \cmidrule(lr){7-7} \cmidrule(lr){8-8} \multicolumn{1}{l}{Problem} & \multicolumn{1}{c}{$n$} & \multicolumn{1}{c}{$\mu$} & \multicolumn{1}{c}{$H(\theta)$} & \multicolumn{1}{c}{$R_\gamma - \lambda S$} & \multicolumn{1}{c}{$H(\theta)$} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} \\ \midrule Gear & \multicolumn{1}{|r|}{ 320} & $1.194 \times 10^{-6}$ & 2 & 1 & 5 & 1.1 & 0.2 \\ Gear & \multicolumn{1}{|r|}{ 640} & $1.499 \times 10^{-7}$ & 5 & 1 & 4 & 14.2 & 0.4 \\ Gear & \multicolumn{1}{|r|}{1280} & $1.878 \times 10^{-8}$ & 5 & 1 & 4 & 216.0 & 2.5 \\ \midrule Grcar & \multicolumn{1}{|r|}{ 320} & $ 0.6543$ & 34 & 1 & 30 & 1.4 & 0.4 \\ Grcar & \multicolumn{1}{|r|}{ 640} & $ 0.6544$ & 33 & 1 & 31 & 14.8 & 2.3 \\ Grcar & \multicolumn{1}{|r|}{1280} & $ 0.6544$ & 26 & 1 & 28 & 196.5 & 15.0 \\ \midrule FM & \multicolumn{1}{|r|}{ 320} & $ 0.1851$ & 11 & 1 & 20 & 1.2 & 0.2 \\ FM & \multicolumn{1}{|r|}{ 640} & $ 0.1836$ & 8 & 1 & 20 & 10.0 & 0.8 \\ FM & \multicolumn{1}{|r|}{1280} & $ 0.1829$ & 9 & 1 & 20 & 80.1 & 5.3 \\ \midrule \texttt{randn} & \multicolumn{1}{|r|}{ 320} & $ 0.7576$ & 7 & 1 & 49 & 1.4 & 0.7 \\ \texttt{randn} & \multicolumn{1}{|r|}{ 640} & $ 0.8663$ & 21 & 2 & 76 & 22.9 & 4.4 \\ \texttt{randn} & \multicolumn{1}{|r|}{1280} & $ 0.7971$ & 23 & 2 & 56 & 187.7 & 26.5 \\ \midrule Nearly Disk & \multicolumn{1}{|r|}{ 320} & $ 0.9999$ & 6 & 1 & 1571 & 1.4 & 18.7 \\ Nearly Disk & \multicolumn{1}{|r|}{ 640} & $ 0.9999$ & 6 & 1 & 1567 & 12.1 & 81.6 \\ Nearly Disk & \multicolumn{1}{|r|}{1280} & $ 0.9999$ & 7 & 1 & 1566 & 104.3 & 618.1 \\ \bottomrule \end{tabular} \label{tbl:both} \end{table} Finally, we benchmark our hybrid algorithm. We tested our three algorithms on $n=400$ and $n=800$ examples with $\mu \in \{0.1,0.2,\ldots,0.9,0.99,\ldots,0.9999999\}$. Since minimizing $r(A+BXC)$ to generate such matrices would be prohibitively expensive for these values of $n$ and $\mu$, we instead generated examples of the form~\mbox{$T_{n,\mu} = \eix{\theta} \begin{bsmallmatrix} M & 0 \\ 0 & D\end{bsmallmatrix}$}, where $\theta \in [0,2\pi)$ was chosen randomly, $M$ is an instance of \cref{ex:circ} with the desired value of $\mu$ and dimension $n-100$, and $D \in \mathbb{C}^{100 \times 100}$ is a complex diagonal matrix. In order to make $h(\theta)$ have many local maximizers, we chose $M$ such that~$r(M)=1$ and then picked the elements of $D$ so that they were roughly placed near a circle drawn between $\bfovX{\eix{\theta} M}$ and the unit circle, biased towards the latter; see \cref{fig:hybrid_ex_fov} for a visualization. \cref{fig:hybrid_ex_htheta} shows how this choice of~$D$ indeed causes~$h(\theta)$ to have many local maximizers, while the randomly chosen $\eix{\theta}$ scalar means that the unique global maximizer may occur anywhere. The running times of our three algorithms on these $T_{n,\mu}$ examples are shown in \cref{fig:hybrid_comp}. Once again, we see that the running time of \cref{alg1} remains fairly constant across all the values of $\mu$, while the running time of \cref{alg2} is much faster for small values of $\mu$ but then blows up as $\mu \to 1$. Most importantly, \cref{fig:hybrid_comp} verifies that our hybrid algorithm indeed remains efficient for all values of~$\mu$ since it automatically detects when to switch from the cutting-plane approach to the level-set approach. In fact, our hybrid algorithm even becomes more efficient than \cref{alg1} for $\mu$ close to one. This is because when it switches to the level-set approach, \cref{alg2} often provided such good starting points that only one eigenvalue computation with~$R_\gamma - \lambda S$ was needed. In contrast, \cref{alg1} always required two eigenvalue computations with $R_\gamma - \lambda S$ on our $T_{n,\mu}$ test problems. \begin{figure} \centering \subfloat[$n=400$, $\mu = 0.4$.]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{T_400_4_fov}} \label{fig:hybrid_ex_fov} } \subfloat[$n=400$, $\mu = 0.99$]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=1.3cm 0cm 1.3cm 0cm,clip]{T_400_10_htheta}} \label{fig:hybrid_ex_htheta} } \caption{ Left: the field of values, eigenvalues, and the osculating circle at the unique point attaining the numerical radius are shown for one instance of $T_{n,\mu}$; to read the plot, see the caption of \cref{fig:demo}. Right: a plot of $h(\theta)$ for another instance of $T_{n,\mu}$. } \label{fig:hybrid_exs} \end{figure} \begin{figure} \centering \subfloat[$n=400$.]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=0.06cm 0cm 0.25cm 0cm,clip]{hybrid_comparison_400}} \label{fig:hybrid_400} } \subfloat[$n=800$]{ \resizebox*{7.0cm}{!}{\includegraphics[trim=0.06cm 0cm 0.25cm 0cm,clip]{hybrid_comparison_800}} \label{fig:hybrid_800.} } \caption{ Running times (in minutes, $\log_{10}$ scale) of the algorithms on $T_{n,\mu}$ with different normalized curvatures. } \label{fig:hybrid_comp} \end{figure} \section{Conclusion} \label{sec:conclusion} Via our new understanding of the local convergence rate of Uhlig's method, as well as how its overall cost blows up for disk matrices, we have precisely explained why Uhlig's method is sometimes much faster or much slower than the level-set method of Mengi and Overton. Moreover, this analysis has motivated our new hybrid algorithm that automatically switches between cutting-plane and level-set techniques in order to remain efficient across all numerical radius problems. Along the way, we have also identified inefficiencies in the earlier level-set and cutting-plane algorithms and addressed them via our improved versions of these two methodologies. \vspace{1cm} \small \bibliographystyle{alpha}
{ "timestamp": "2021-10-28T02:26:16", "yymm": "2002", "arxiv_id": "2002.00080", "language": "en", "url": "https://arxiv.org/abs/2002.00080", "abstract": "The main two algorithms for computing the numerical radius are the level-set method of Mengi and Overton and the cutting-plane method of Uhlig. Via new analyses, we explain why the cutting-plane approach is sometimes much faster or much slower than the level-set one and then propose a new hybrid algorithm that remains efficient in all cases. For matrices whose fields of values are a circular disk centered at the origin, we show that the cost of Uhlig's method blows up with respect to the desired relative accuracy. More generally, we also analyze the local behavior of Uhlig's cutting procedure at outermost points in the field of values, showing that it often has a fast Q-linear rate of convergence and is Q-superlinear at corners. Finally, we identify and address inefficiencies in both the level-set and cutting-plane approaches and propose refined versions of these techniques.", "subjects": "Numerical Analysis (math.NA); Optimization and Control (math.OC)", "title": "Convergence rate analysis and improved iterations for numerical radius computation", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.97997655637136, "lm_q2_score": 0.8175744695262775, "lm_q1q2_score": 0.8012038132235029 }
https://arxiv.org/abs/1707.01883
Hermann Hankel's "On the general theory of motion of fluids", an essay including an English translation of the complete Preisschrift from 1861
The present is a companion paper to "A contemporary look at Hermann Hankel's 1861 pioneering work on Lagrangian fluid dynamics" by Frisch, Grimberg and Villone (2017). Here we present the English translation of the 1861 prize manuscript from Göttingen University "Zur allgemeinen Theorie der Bewegung der Flüssigkeiten" (On the general theory of the motion of the fluids) of Hermann Hankel (1839-1873), which was originally submitted in Latin and then translated into German by the Author for publication. We also provide the English translation of two important reports on the manuscript, one written by Bernhard Riemann and the other by Wilhelm Eduard Weber, during the assessment process for the prize. Finally we give a short biography of Hermann Hankel with his complete bibliography.
\section{Introductory notes} \label{s:intro} \deffootnotemark{\textsuperscript{\thefootnotemark}}\deffootnote{2em}{1.6em}{${}^\thefootnotemark$\hspace{0.0319cm}\enskip} \setcounter{footnote}{0} Here we present, with some supplementary documents, a full translation from German of the winning essay, originally written in Latin, by Hermann Hankel, in response to the ``extraordinary mathematical prize''\footnote{The ordinary prize from the Philosophical Faculty of the University of G\"ottingen was a philosophical one; extraordinary prizes, however, could be launched in another discipline.} launched on 4th June 1860 by the Philosophical Faculty of the University of G\"ottingen with a deadline of end of March 1861. By request of the Prize Committee to Hankel, the essay was then revised by him and finally published in 1861, in German, as a \textit{Preisschrift}.\footnote{Hankel, \hyperlink{Hankel}{1861.} Hankel asked for permission to have his revised \textit{Preisschrift} published in German.} The Latin manuscript has been very probably returned to the Author and now appears to be irremediably lost.\footnote{This information has been conveyed to us through the G\"ottingen Archive. An attempt by us to obtain the Latin document from a descendant of Hankel was unsuccessful.} The G\"ottingen University Library possesses two copies of the 1861 German original edition of the \textit{Preisschrift}. Hankel participated in this prize competition shortly after his arrival in the Spring 1860 in G\"ottingen as a 21-years old student in Mathematics, from the University of Leipzig. The formulation of the prize (see section~\ref{s:Preisschrift}) highlighted the problem of the equations of fluid motion in Lagrangian coordinates and was presented in memory of Peter Gustav Lejeune-Dirichlet. A \textit{post mortem} paper of his had outlined the advantages of the use of the Lagrangian approach, compared to the Eulerian one, for the description of the fluid motion.\footnote{Lejeune-Dirichlet, \hyperlink{Dirichlet}{1860}; for more details, see the companion paper: Frisch, Grimberg and Villone, \hyperlink{FGV}{2017}.} During the decision process for the winning essay, there was an exchange of German-written letters among the committee members.\footnote{G\"ottingen University Archive, \hyperlink{G\"ottingen University Archive}{1860/1861}.} Notably, among all of these letters, the ones from Wilhelm Eduard Weber (1804--1891) and from Bernhard Riemann (1826--1866) have been decisive for the evaluation of the essay. Hankel had been the only one to submit an {essay, and the discussion among the committee members was devoted to deciding whether or not} the essay written by Hankel deserved to win the prize. For completeness, we provide an English translation of two of these letters, one signed by Bernhard Riemann, the other by Wilhelm Eduard Weber. As is thoroughly discussed in the companion paper \mbox{(Frisch, Grimberg and Villone, \hyperlink{FGV}{2017}),} Hankel's \textit {Preisschrift} reveals indeed a truly deep understanding of Lagrangian fluid mechanics, with innovative use of variational methods and differential geometry. Until these days, this innovative work of Hankel has remained apparently poorly known among scholars, with some exceptions. For details and references, see the companion paper. The paper is organised as follows. In section~\ref{s:Preisschrift}, we present the translation of the \textit{Preis-schrift}; this section begins with a preface signed by Hankel, containing the stated prize question together with the decision by the committee members. For our translation we used the digitized copy of the \textit{Preisschrift} from the HathyTrust Digital Library (indicated as a link in the reference). It has been verified by the G\"ottingen University Library that the text that we used is indeed the digitized copy of the 1861 original printed copy. Section~\ref {s:letters} contains the translation of two written judgements on Hankel's essay in the procedure of assessment of the prize, one by Weber and the other by Riemann. In section~\ref{s:HHpapers} we provide some biographical notes, together with a full publication list of Hankel. Let us elucidate our conventions for author/translator footnotes, comments and equation numbering. Footnotes by Hankel are denoted by an ``A.'' followed by a number (A stands for Author), enclosed in square brackets; translator footnotes are treated identically except that the letter ``A'' is replaced with ``T'' (standing for translator). For both author and translator footnotes, we apply a single number count, e.g., [T.1], [A.2], [A.3]. Very short translator comments are added directly in the text, and such comments are surrounded by square brackets. Only a few equations have been numbered by Hankel (denoted by numbers in round brackets). To be able to refer to all equations, especially relevant for the companion paper, we have added additional equation numbers in the format [p.n], which means the nth equation of \S p. Finally, we note that the abbreviations ``S.'' and ``Bd.'', which occur in the {\it Preisschrift}, refer respectively to the German words for ``page'' and ``volume''. \deffootnotemark{\textsuperscript{[T.\thefootnotemark]}}\deffootnote{2em}{1.6em}{[T.\thefootnotemark]\enskip} \setlength{\footnotesep}{0.24cm} \section{Hankel's Preisschrift translation} \label{s:Preisschrift} \noindent About the prize question set by the philosophical Faculty of Georgia Augusta on 4th June 1860\,:\,\footnote{The prize question is in Latin in the {\it Preisschrift}\,:\,``Aequationes generales motui fluidorum determinando inservientes duobus modis exhiberi possunt, quorum alter Eulero, alter Lagrangio debetur. Lagrangiani modi utilitates adhuc fere penitus neglecti clarissimus Dirichlet indicavit in commentatione postuma 'de problemate quodam hydrodynamico' inscripta; sed ab explicatione earum uberiore morbo supremo impeditus esse videtur. Itaque postulat ordo theoriam motus fluidorum aequationibus Lagrangianis superstructam eamque eo saltem perductam, ut leges motus rotatorii a clarissimo Helmholtz alio modo erutae inde redundent.'' At that time it was common to state prizes in Latin and to have the submitted essays in the same language.} \\\\ \indent\indent The most useful equations for determining fluid motion may be presented in two ways, one of which is Eulerian, the other one is Lagrangian. The illustrious Dirichlet pointed out in the posthumous unpublished paper ``On a problem of hydrodynamics'' the almost completely overlooked advantages of the Lagrangian way, but he was prevented from unfolding this way further by a fatal illness. So, this institution asks for a theory of fluid motion based on the equations of Lagrange, yielding, at least, the laws of vortex motion already derived in another way by the illustrious Helmholtz.\\\\ \noindent Decision of the philosophical Faculty on the present manuscript: \\\\ \indent\indent The extraordinary mathematical-physical prize question about the derivation of laws of fluid motion and, in particular, of vortical motion, described by the so-called Lagrangian equations, was answered by an essay carrying the motto: \,\,{\em The more signs express relationships in nature, the more useful they are.}\footnote{This motto is in Latin in the \textit{Preisschrift}: ``Tanto utiliores sunt notae, quanto magis exprimunt rerum relationes''. At that time it was common that an Author submitting his work for a prize signed it anonymously with a motto.}\hspace{0.1cm} This manuscript gives commendable evidence of the Author's diligence, of his knowledge and ability in using the methods of computation recently developed by contemporary mathematicians.\hspace{0.4cm}In particular $\S\,6$\footnotemark \footnotetext{A number such as ``$\S\,6$'' refers to a section number in the (lost) Latin version of the manuscript. Numbers are different in the revised German translation.} contains an elegant method to establish the equations of motion for a flow in a fully arbitrary coordinate system, from a point of view which is commonly referred to as Lagrangian.\hspace{0.4cm}However, when developing the general laws of vortex motion, the Lagrangian approach is unnecessarily left aside, and, as a consequence, the various laws have to be found by quite independent means.\hspace{0.4cm}Also the relation between the vortex laws and the investigations of Clebsch, reported in $\S.\,14.\,15 $, is omitted by the Author.\hspace{0.4cm}Nonetheless, as his derivation actually begins from the Lagrangian equations, one may consider the prize-question as fully answered by this manuscript.\hspace{0.4cm}Amongst the many good things to be found in this essay, the evoked incompleteness and some mistakes due to rushing which are easy to improve, do not prevent this Faculty from assigning the prize to the manuscript, but the Author would be obliged to submit a revised copy of the manuscript, improved according to the suggestions made above before it goes into print.\\ \noindent \centerline {\rule{5cm}{0.2mm}} \\\\ \indent\indent On my request, I have been permitted by the philosophical Faculty to have the manuscript, originally submitted in Latin, printed in German. --- \\\\ \indent\indent The above mentioned $\S.\,6$ coincides with $\S.\,5$ of the present essay. The $\S.\S\,14. 15$ included what is now in the note of S.\,45 of the text [now page~\pageref{footnoteClebsch}];\,\,these $\S.\S.$ were left out, on the one hand, because of lack of space, and on the other hand because, in the present view,\,\,these $\S.\S.$ are not anymore connected to the rest of the essay.\\ \indent\indent Leipzig,\hspace{0.1cm} September 1861.\\ $\phantom{C}$ \hfill Hermann Hankel. \quad\quad \vspace{2cm} \centerline{\fett{$\S.\,1.$} } \vspace{0.3cm} \noindent The conditions and forces which underlie most of the natural phenomena are so complicated and various, that it is rarely possible to take them into account by fully analytical means.{\hspace{0.4cm}Therefore, one should, for the time being, discard those forces and properties which evidently have little impact on the motion or the changes; this is done by just retaining the forces that are of essential and of fundamental importance.\hspace{0.4cm} Only then, once this first approximation has been made, one may reconsider the previously disregarded forces and properties, and modify the underlying hypotheses.\hspace{0.4cm}In the case of the general theory of motion of liquid fluids, it seems therefore advisable to take into account solely the continuity [of the fluid] and the constancy of volume, and to disregard both viscosity and internal friction which are also present as suggested by experience; it is advisable to take into account just the continuity and the elasticity, and consider the pressure determined as a function of the density. \hspace{0.1cm} Even if, in specific cases, the analytical methods have improved and may apply as well to more realistic hypotheses, these hypotheses are not yet suitable for a fundamental general theory.\\ \indent\indent The hypotheses of the general hydrodynamical equations, as given by \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r,} are the following. The fluid is considered as being composed of an aggregate of molecules, which are so small that one can find an infinitely large number of them in an arbitrarily small space.\hspace{0.4cm}Therefore, the fluid is considered as divided into infinitely small parts of dimensionality of the first order; each of these parts is filled with an infinite number of molecules, whose sizes have to be considered as infinitely small quantities of the second order.\hspace{0.4cm}These molecules fill space continuously and move without friction against each other.\\ \indent \indent The flow can be set in motion either by accelerating forces from the individual molecules or from external pressure forces.\footnotemark \footnotetext{Literally, Hankel wrote ``that emanates from the molecules''.} Considering the nature of fluids, one easily comes to the conclusion, also confirmed by experience, that the pressure on each fluid element of the external surface acts normally and proportionally to the size of that element. In order to have also a clear definition of the pressure at a given point within the fluid, let us think of an element of an arbitrary surface through this point: the pressure will be normal to this surface element and proportional to its size, but independent on its direction.\hspace{0.4cm}The difference between liquid and elastic flows\footnote {Nowadays, in this context, \textit {liquid} flows would be called \textit{incompressible} flows.}{ is then that, for the former, the density is a constant and, for the latter, the density depends on the pressure, and, conversely, the pressure depends on the density in a special way.\\ \indent\indent We shall not insist here on a detailed discussion of these properties as they are usually discussed in the better textbooks on mechanics.\\ \indent \indent \deffootnotemark{\textsuperscript{[A.\thefootnotemark]}} \deffootnote{2em}{1.6em}{[A.\thefootnotemark]\enskip} In order to study the above properties analytically, two methods have been so far applied, both owed to \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r.} \hspace{0.4cm}The first method\footnote {First published in the essay: Principes g\'en\'eraux du mouvement des fluides\,\, (Hist.\ de l'Acad.\ de Berlin, ann\'ee, \hyperlink{Euler226}{1755}).} considers the velocity of each point of the fluid as a function of position and time. If $u,v,w$ are the velocity components in the orthogonal coordinates $x,y,z$, then $u,v,w$ are functions of position $x,y,z$ and time $t$.\hspace{0.4cm}The velocity of the fluid in a given point is thus the velocity of the fluid particles flowing through that point.\hspace{0.4cm}This method was exclusively used for the study of motion of fluids until \mbox{D\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! t}\footnotemark \footnotetext{Untersuchungen \"uber ein Problem der Hydrodynamik.\,\, (Crelle's Journal, Bd.\ 55, [actually it is Bd. 58], S.181. [\hyperlink{Dirichlet2}{1861}]).} observed that this method had necessarily the drawback that the absolute space filled by the flow in general changes over time and, as a consequence, the coordinates $x,y,z$ are not entirely independent variables. The method just discussed seems appropriate if the flow is always filling the same space, i.e., in the case when the flow is filling the infinite space, or when the motion is stationary.\footnote{D\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! t used for this case the first \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r}ian method: Ueber einige F\"alle, in denen sich die Bewegung eines festen K\"orpers in einem incompressibelen fl\"ussigen Medium theoretisch bestimmen l\"asst [Ueber die Bewegung eines festen K\"orpers in einem incompressibeln fl\"ussigen Medium]. (Berichte der Berliner Akademie, \hyperlink{Dirichlet1}{1852}, S.12.).\,\,\, Also \mbox{R\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\!} used the first \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! rian} form in his {\em Essai}: Ueber die Fortpflanzung ebener Luftwellen von endlicher Schwingungsweite. (Bd. VIII d. Abhdlg. der G\"ott. Soc. \hyperlink{Riemann}{1860}).}\hspace{0.1cm}\\ \indent \indent The second, ingenious method of \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r}\footnote {De principiis motus fluidorum (Novi comm. acad. sc. Petropolitanae. Bd. XIV. Theil I. pro anno 1759 [\hyperlink{Euler1759}{1770}, in German].) im 6. Capitel: De motu fluidorum ex statu initiali definiendo. S.\,358.}} considers the coordinates $x,y,z$ of a flow particle, in any reference system, as a function of time $t$ and of its position $a,b,c$ at initial time $t=0$.\hspace{0.4cm}This method, by which the same fluid particle is followed during its motion, was reproduced, indeed in a slightly more elegant way, \deffootnotemark{\textsuperscript{[A.\thefootnotemark]}} \deffootnote{2em}{1.6em}{[A.\thefootnotemark]\enskip} by \mbox{L\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! g\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! g\hspace{0.036cm}} %{$\phantom{!}$\! e},\footnote{M\'ecanique analytique, \'ed III. par Bertrand. Bd.\,II. S. 250--261.\,\, The first edition of the M\'ecanique is of the year \hyperlink{Lagrange1}{1788}.} without giving any reference.\hspace{0.1cm} Since it appears that nowadays \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r}'s work is rarely read in detail, this method is considered due to \mbox{L\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! g\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! g\hspace{0.036cm}} %{$\phantom{!}$\! e.} \hspace{0.4cm}However, the method was already present in its full completeness in E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r's work, 29 years before.\hspace{0.4cm}I owe this interesting, historical note to my honoured Professor B. \mbox{R\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! n.}\hspace{0.4cm} According to this method, one has thus \begin{equation*} x=\varphi_1 (a,b,c,t), \qquad y=\varphi_2 (a,b,c,t), \qquad z=\varphi_3 (a,b,c,t), \tag*{[1.1]} \end{equation*} where $\varphi_1,\varphi_2,\varphi_3 $ are continuous functions of $a,b,c,t$.\,\, We have at initial time $t=0$ that \begin{equation*} \tag*{[1.2]} x=a, \hspace{2cm}y=b, \hspace{2cm}z=c \end{equation*} and, thus, the following conditions are valid for $\varphi_1,\,\varphi_2,\,\varphi_3$\,: \begin{equation*} \tag*{[1.3]} a=\varphi_1 (a,b,c,0), \qquad b=\varphi_2 (a,b,c,0), \qquad c=\varphi_3 (a,b,c,0) . \end{equation*} Obviously, one can take values of $t$ so small, that, according to Taylor's theorem, $x,y,z$ can be expanded in powers of $t$. Since at time $t=0$ we have $x=a,\,\, y=b,\,\, z=c$, it follows evidently that \begin{align*} \tag*{[1.4]} \begin{aligned} x\,&= \,a\,+\, A_1t\,+\,A_2 t^2\,+\, \ldots \\ y\,&= \,b\,+\, B_1t\,+\,B_2 t^2\,+ \hspace{0.09cm} \ldots \\ z\,&= \,c\,+\, C_1t\,+\,C_2 t^2\,+ \hspace{0.11cm} \ldots \end{aligned} \end{align*} From these equations, one easily finds that at time $t=0$: \begin{mynewequation} \tag*{[1.5]} \begin{aligned} \frac{dx}{da} &= 1, \,\,\, \,\frac{dx}{db} = 0, \,\,\,\,\frac{dx}{dc} = 0 \\ \frac{dy}{da} &= 0, \,\,\,\,\frac{dy}{db} = 1, \,\,\,\,\frac{dy}{dc} = 0 \\ \frac{dz}{da} &= 0, \,\, \,\,\frac{dz}{db} = 0,\,\, \,\,\frac{dz}{dc} = 1 \end{aligned} \end{mynewequation} \indent % In the following, we will try to give a presentation of the general theory of hydrodynamics based on the second method. It will turn out that the second [Lagrangian] form also merits to be preferred over the first one in some cases, because the fundamental equations of the former are more closely connected to the customary forms in mechanics.}\\[1cm] \centerline{\fett{$\S.\,2.$}}\\[0.3cm] \indent\indent The two infinitely near particles\\ \centerline {$a,\,\,b,\,\,c$} and\\ \centerline {$a+{\rm d}a, \,\,b+{\rm d}b, \,\,c+{\rm d}c$}\\\\ after some time $t$, will be at the two points\\\\ \centerline{$x,\,\, y, \,\,z$} \noindent and\\ \centerline {$x+{\rm d}x,\,\, y+{\rm d}y,\,\, z+{\rm d}z$,}\\\\ where one needs to put \begin{mynewequation} \tag*{[2.1]} \begin{aligned} {\rm d}x &=\frac{dx}{da} {\rm d}a + \frac{dx}{db} {\rm d}b + \frac{dx}{dc} {\rm d}c\\ {\rm d}y &=\frac{dy}{da} {\rm d}a + \frac{dy}{db} {\rm d}b + \frac{dy}{dc} {\rm d}c\\ {\rm d}z &=\frac{dz}{da} {\rm d}a + \frac{dz}{db} {\rm d}b + \frac{dz}{dc} {\rm d}c \end{aligned} \end{mynewequation} \noindent Let us think of $a,\,b,\, c$ as linear --- in general non-orthogonal --- coordinates; then, we have that ${\rm d}a,\, {\rm d}b,\, {\rm d}c$ are the coordinates of a point with respect to a congruent system of coordinates $S_0$, whose origin is at $a,\,b,\, c$.\hspace{0.4cm}As $x,\,y,\, z$ refer to the same coordinate system as $a,\,b,\, c$ do, then ${\rm d}x,\, {\rm d}y,\, {\rm d}z$ are the analogous coordinates with respect to a coordinate system, whose origin is at $x,\,y,\, z$.\hspace{0.4cm}Now, let us think of an infinitely small surface \begin{equation*} \tag*{[2.2]} F({\rm d}a,\, {\rm d}b,\, {\rm d}c) = 0 \end{equation*} with reference to $S_0$, then, at time $t$, the surface will have been moved into another surface \begin{equation*} \tag*{[2.3]} F({\rm d}x,\, {\rm d}y,\, {\rm d}z) = 0 \end{equation*} \noindent {with reference to $S$, or, \\ \begin{equation*} \tag*{[2.4]} F\Big(\frac{dx}{da} {\rm d}a + \frac{dx}{db} {\rm d}b + \frac{dx}{dc} {\rm d}c, \frac{dy}{da} {\rm d}a + \frac{dy}{db} {\rm d}b + \frac{dy}{dc} {\rm d}c, \frac{dz}{da} {\rm d}a + \frac{dz}{db} {\rm d}b + \frac{dz}{dc} {\rm d}c \Big)=0 \end{equation*} We see from this that an infinitely small algebraic surface of degree $n$ always remains of the same degree.\\\\ \indent \indent Hence, at any time, infinitely close points on a plane, always stay on a \mbox{plane; and since a} straight line may be thought of as an intersection of two planes, infinitely close points on a line at a certain time always stay on a line.\\\\ \indent\indent The points within an infinitely small ellipsoid will always stay in such an ellipsoid, because a closed surface cannot transform into a non-closed surface and, with the exception of the ellipsoid, all the surfaces of second degree are not closed.\hspace{0.4cm} The section of a plane and of an infinitely small ellipsoid will thus have to remain always the same.\hspace{0.4cm}Since that section always constitutes an ellipse, so an infinitely small ellipse will always remain the same. \\\\ \indent\indent The four points: \begin{mynewequation} \notag \begin{aligned} a\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,b\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,c\,\,\,\,\,\,\,\,\,\\ a+m\,{\rm d}a\,\,\,\,\,\,\,\,\,\,\,b + n\,\,{\rm d}b\,\,\,\,\,\,\,\,c + p\,\,{\rm d}c\\ \,\,\,\,\,\, a+ m'\,{\rm d}a\,\,\,\,\,\,\,\,\,b + n'\,{\rm d}b\,\,\,\,\,\,\, c + p'\,{\rm d}c\\ a+m''{\rm d}a \,\,\,\,\,\,\,\,b + n''{\rm d}b\,\,\,\,\,\,\, c + p''{\rm d}c\\ \end{aligned} \end{mynewequation} where $m,\,n,\,p,\,m',\,n',\,p',\,m'',\,n'',\,p''$ are finite numbers, at time $t$ will be at the position\\ \\ \centerline{\em \,\,\,\,\,\,x\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,y\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,z} \begin{tinyequation} \nonumber \begin {aligned} &\!\!x + \frac{dx}{da} m\,\,{\rm d}a + \frac{dx}{db} n\,\,{\rm d}b + \frac{dx}{dc} p\,\,{\rm d}c,\,\,\,\,\,\,\,\,\,\,y + \frac{dy}{da} m\,\,{\rm d}a + \frac{dy}{db} n\,\,{\rm d}b + \frac{dy}{dc} p\,\,{\rm d}c, \,\,\,\, z + \frac{dz}{da} m\,\,{\rm d}a + \frac{dz}{db} n\,\,{\rm d}b + \frac{dz}{dc} p\,{\rm d}c \\\\ \!\!&\!\!x+\frac{dx}{da} m'{\rm d}a + \frac{dx}{db} n'{\rm d}b + \frac{dx}{dc} p'{\rm d}c,\,\,\,\,\,\,\,\,\,\,y + \frac{dy}{da} m'{\rm d}a + \frac{dy}{db} n'{\rm d}b + \frac{dy}{dc} p'{\rm d}c, \,\,\,\,\,\, z + \frac{dz}{da} m'{\rm d}a + \frac{dz}{db} n'{\rm d}b + \frac{dz}{dc} p'{\rm d}c\,\\\\ \!\!&\!\!x+\frac{dx}{da} m''{\rm d}a + \frac{dx}{db} n''{\rm d}b + \frac{dx}{dc} p''{\rm d}c,\,\, y + \frac{dy}{da} m''{\rm d}a + \frac{dy}{db} n''{\rm d}b + \frac{dy}{dc} p''{\rm d}c, \,\,\, z + \frac{dz}{da} m''{\rm d}a + \frac{dz}{db} n''{\rm d}b + \frac{dz}{dc} p''{\rm d}c \end {aligned} \end{tinyequation} at time $t$. \indent\indent The volume of the tetrahedron $T_0$ whose vertices are those points at time $t = 0$, is expressed by the determinant\,:\\ \begin{equation*} \tag*{[2.5]} 6T_0= \begin{vmatrix} m& n & p\\m'& n' & p'\\m''&n''&p''\end{vmatrix} {\rm d}a\,{\rm d}b\,{\rm d}c \end{equation*} \indent\indent The volume of the tetrahedron $T$ whose vertices at time $t$ are formed by the same particles, is\,: \begin{smallequation} \tag*{[2.6]} 6T= \begin{vmatrix} \frac{dx}{da} m\,{\rm d}a + \frac{dx}{db} n\,{\rm d}b + \frac{dx}{dc} p\,{\rm d}c,\,\,\, &\frac{dy}{da} m\,{\rm d}a + \frac{dy}{db} n\,{\rm d}b + \frac{dy}{dc} p\,{\rm d}c,\,\,\, &\frac{dz}{da} m\,{\rm d}a + \frac{dz}{db} n\,{\rm d}b + \frac{dz}{dc} p\,{\rm d}c \\\\ \frac{dx}{da}m'{\rm d}a + \frac{dx}{db}n'{\rm d}b + \frac{dx}{dc}p'{\rm d}c,\,\,\, &\frac{dy}{da} m'{\rm d}a + \frac{dy}{db}n'{\rm d}b + \frac{dy}{dc} p'{\rm d}c, & \frac{dz}{da} m'{\rm d}a + \frac{dz}{db} n'{\rm d}b + \frac{dz}{dc} p'{\rm d}c\\\\ \frac{dx}{da}m''{\rm d}a + \frac{dx}{db} n'' {\rm d}b + \frac{dx}{dc} p'' {\rm d}c,\,\,\, &\frac{dy}{da} m'' {\rm d}a + \frac{dy}{db} n'' {\rm d}b + \frac{dy}{dc} p'' {\rm d}c, & \frac{dz}{da} m'' {\rm d}a + \frac{dz}{db} n'' {\rm d}b + \frac{dz}{dc} p'' {\rm d}c \end{vmatrix} \end{smallequation} \indent\indent By known theorems, this determinant can however be written as a product of two determinants\,: \begin{equation} \tag*{[2.7]} 6T= \begin{vmatrix} \frac{dx}{da}&\frac{dx}{db}&\frac{dx}{dc} \\\\ \frac{dy}{da}&\frac{dy}{db}&\frac{dy}{dc} \\\\ \frac{dz}{da}&\frac{dz}{db}&\frac{dz}{dc} \end{vmatrix} \begin{vmatrix} m& n & p\\\\m'& n' & p'\\\\m''&n''&p''\end{vmatrix} {\rm d}a\,{\rm d}b\,{\rm d}c \end{equation} or \begin{mynewequation} \tag*{[2.8]} \text{\small $T=T_0$} \begin{vmatrix} \frac{dx}{da}\,\,&\frac{dx}{db}\,\,&\frac{dx}{dc} \\\\ \frac{dy}{da}\,\,&\frac{dy}{db}\,\,&\frac{dy}{dc} \\\\ \frac{dz}{da}\,\,&\frac{dz}{db}\,\,&\frac{dz}{dc} \end{vmatrix} \end{mynewequation} \indent\indent It results from the preceding considerations, that all particles which at time $t=0$ are in the tetrahedron $T_0$, will be also in the tetrahedron $T$ at time $t$.\hspace{0.4cm}Let $\varrho_0$ be the mean density of the tetrahedron $T_0$ at time $t=0$, and $\varrho$ the mean density in $T$ at time $t$, so one has $T:T_0=\varrho_0:\varrho$ and hence \begin{mynewequation} \tag*{(1), [2.9]} \begin{vmatrix} \frac{dx}{da}&\frac{dx}{db}&\frac{dx}{dc} \\\\ \frac{dy}{da}&\frac{dy}{db}&\frac{dy}{dc} \\\\ \frac{dz}{da}&\frac{dz}{db}&\frac{dz}{dc} \end{vmatrix} \text{\small $=$} \,\frac{\varrho_0}{\varrho} \,. \end{mynewequation} If the density is constant, the fluid is a liquid flow and thus \deffootnotemark{\textsuperscript{[T.\thefootnotemark]}} \deffootnote{2em}{1.6em}{[T.\thefootnotemark]\enskip} \begin{mynewequation} \tag*{(2), [2.10]} \begin{vmatrix} \frac{dx}{da}&\frac{dx}{db}&\frac{dx}{dc} \\\\ \frac{dy}{da}&\frac{dy}{db}&\frac{dy}{dc} \\\\ \frac{dz}{da}&\frac{dz}{db}&\frac{dz}{dc} \end{vmatrix} \text{\small $=$} \,1. \end{mynewequation} \indent\indent One can reasonably refer to these equations as the density equations, more particularly the last one as the equation of the constancy of the volume.\\\\ \indent\indent The values of the functional determinant of $x, y, z$ with respect to $a,b,c$ may be used to develop a set of relationships, which are often needed to pass from the \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! rian representation to the other [i.e., to the Lagrangian representation; or reciprocally].} Indeed, if one solves the system of equations\,: \begin{mynewequation} \tag*{[2.11]} \begin{aligned} {\rm d}x=\frac{dx}{da} {\rm d}a + \frac{dx}{db} {\rm d}b + \frac{dx}{dc} {\rm d}c\\ {\rm d}y=\frac{dy}{da} {\rm d}a + \frac{dy}{db} {\rm d}b + \frac{dy}{dc} {\rm d}c\\ {\rm d}z=\frac{dz}{da} {\rm d}a + \frac{dz}{db} {\rm d}b + \frac{dz}{dc} {\rm d}c \end{aligned} \end{mynewequation} one has\,: \begin{mynewequation} \tag*{[2.12]} \begin{aligned} \Big(\frac{dy}{db} \frac{dz}{dc} - \frac{dy}{dc}\frac{dz}{db}\Big) {\rm d}x + \Big(\frac{dz}{db} \frac{dx}{dc} - \frac{dz}{dc}\frac{dx}{db}\Big) {\rm d}y + \Big(\frac{dx}{db} \frac{dy}{dc} - \frac{dx}{dc}\frac{dy}{db}\Big) {\rm d}z = \frac{\varrho_0}{\varrho} {\rm d}a\\\\ \Big(\frac{dy}{dc} \frac{dz}{da} - \frac{dy}{da}\frac{dz}{dc}\Big) {\rm d}x + \Big(\frac{dz}{dc} \frac{dx}{da} - \frac{dz}{da}\frac{dx}{dc}\Big) {\rm d}y + \Big(\frac{dx}{dc} \frac{dy}{da} - \frac{dx}{da}\frac{dy}{dc}\Big) {\rm d}z = \frac{\varrho_0}{\varrho} {\rm d}b\\\\ \Big(\frac{dy}{da} \frac{dz}{db} - \frac{dy}{db}\frac{dz}{da}\Big) {\rm d}x + \Big(\frac{dz}{da} \frac{dx}{db} - \frac{dz}{db}\frac{dx}{da}\Big) {\rm d}y + \Big(\frac{dx}{da} \frac{dy}{db} - \frac{dx}{db}\frac{dy}{da}\Big) {\rm d}z = \frac{\varrho_0}{\varrho} {\rm d}c \end{aligned} \end{mynewequation} where we have substituted the value $\varrho_0/\varrho$ for the functional determinant. \hspace{0.4cm}The comparison of these equations with\,: \begin{mynewequation} \tag*{[2.13]} \begin{aligned} \frac{da}{dx}\,{\rm d}x + \frac{da}{dy}\,{\rm d}y + \frac{da}{dz}\,{\rm d}z = {\rm d}a\\\\ \frac{db}{dx}\,{\rm d}x + \frac{db}{dy}\,{\rm d}y + \frac{db}{dz}\,{\rm d}z = {\rm d}b\\\\ \frac{dc}{dx}\, {\rm d}x + \frac{dc}{dy}\,{\rm d}y + \frac{dc}{dz}\,{\rm d}z = {\rm d}c \end{aligned} \end{mynewequation} gives this equation system\,: \begin{mynewequation}\tag*{(3), [2.14]} \left. \begin{aligned} \frac{\varrho_0}{\varrho}\frac{da}{dx} & = \frac{dy}{db} \frac{dz}{dc} - \frac{dy}{dc}\frac{dz}{db},\, \frac{\varrho_0}{\varrho}\frac{da}{dy} = \frac{dz}{db} \frac{dx}{dc} - \frac{dz}{dc}\frac{dx}{db}, \, \frac{\varrho_0}{\varrho}\frac{da}{dz} = \frac{dx}{db} \frac{dy}{dc} - \frac{dx}{dc}\frac{dy}{db} \\ \frac{\varrho_0}{\varrho}\frac{db}{dx} & = \frac{dy}{dc} \frac{dz}{da} - \frac{dy}{da}\frac{dz}{dc},\, \frac{\varrho_0}{\varrho}\frac{db}{dy} = \frac{dz}{dc} \frac{dx}{da} - \frac{dz}{da}\frac{dx}{dc}, \, \frac{\varrho_0}{\varrho}\frac{db}{dz} = \frac{dx}{dc} \frac{dy}{da} - \frac{dx}{da}\frac{dy}{dc} \\ \frac{\varrho_0}{\varrho}\frac{dc}{dx} & = \frac{dy}{da} \frac{dz}{db} - \frac{dy}{db}\frac{dz}{da},\, \frac{\varrho_0}{\varrho}\frac{dc}{dy} = \frac{dz}{da} \frac{dx}{db} - \frac{dz}{db}\frac{dx}{da}, \, \frac{\varrho_0}{\varrho}\frac{dc}{dz} = \frac{dx}{da} \frac{dy}{db} - \frac{dx}{db}\frac{dy}{da} \end{aligned} \qquad \right\} \end{mynewequation} \indent\indent If the equation (1) is differentiated with respect to one of the independent variables $a,b,c$, and these differentiations are indicated with $\delta$, one obtains \begin{align} \Big(\frac{dy}{db} \frac{dz}{dc} - \frac{dy}{dc}\frac{dz}{db}\Big) \frac{d \delta x}{da} + \Big(\frac{dy}{dc} \frac{dz}{da} - \frac{dy}{da}\frac{dz}{dc}\Big) \frac{d \delta x}{db} + \Big(\frac{dy}{da} \frac{dz}{db} - \frac{dy}{db}\frac{dz}{da}\Big) \frac{d \delta x}{dc} \nonumber \\ \nonumber \\ + \Big(\frac{dz}{db} \frac{dx}{dc} - \frac{dz}{dc}\frac{dx}{db}\Big) \frac{d \delta y}{da} + \Big(\frac{dz}{dc} \frac{dx}{da} - \frac{dz}{da}\frac{dx}{dc}\Big) \frac {d \delta y}{db} + \Big(\frac{dz}{da} \frac{dx}{db} - \frac{dz}{db}\frac{dx}{da}\Big) \frac {d \delta y}{dc} \nonumber \\ \nonumber \\ + \Big( \frac{dx}{db} \frac{dy}{dc} - \frac{dx}{dc}\frac{dy}{db}\Big) \frac{d \delta z}{da} + \Big(\frac{dx}{dc} \frac{dy}{da} - \frac{dx}{da}\frac{dy}{da}\Big) \frac{d \delta z}{db} + \Big(\frac{dx}{da} \frac{dy}{db} - \frac{dx}{db}\frac{dy}{da}\Big) \frac{d \delta z}{dc}\,\,\,\,\,\,\,\hspace{-0.4cm} \nonumber \\ \nonumber \\ = -\frac{\varrho_0}{\varrho} \frac{\delta \varrho}{\varrho}\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, \tag*{[2.15]} \end{align} and, by equations~(3): \begin{equation*} \tag*{[2.16]} \frac{d \delta x}{da}\frac{da}{dx} + \frac{d \delta x}{db}\frac{db}{dx} + \frac{d \delta x}{dc}\frac{dc}{dx} + \frac{d \delta y}{da}\frac{da}{dy} + \frac{d \delta y}{db}\frac{db}{dy} + \frac{d \delta y}{dc}\frac{dc}{dy} + \frac{d \delta z}{da}\frac{da}{dz} + \frac{d \delta z}{db}\frac{db}{dz} + \frac{d \delta z}{dc}\frac{dc}{dz} + \frac{\delta \varrho}{\varrho}= 0 \end{equation*} or \begin{equation*}\tag*{(4), [2.17]} \frac{d \delta x}{dx} + \frac{d \delta y}{dy}+ \frac{d \delta z}{dz}+ \frac{\delta \varrho}{\varrho}= 0 \end{equation*} \indent If by $\delta$, one understands the differentiation with respect to $t$, one has: \begin{equation*} \tag*{[2.18]} \varrho \Big(\frac{d \frac{dx}{dt}}{dx} + \frac{d \frac{dy}{dt}}{dy}+ \frac{d \frac{dz}{dt}}{dz}\Big) + \frac{{ \rm d} \varrho} {{\rm d} t} = 0 \end{equation*} Since $t$ appears not only explicitly in $\varrho$, but also implicitly in $\varrho$ through its dependence on $x$,$y$,$z$, one has \begin{equation*} \tag*{[2.19]} \frac{{ \rm d} \varrho} {{\rm d} t} = \frac{d\varrho}{dt} + \frac{d\varrho}{dx} \frac{dx}{dt} + \frac{d\varrho}{dy} \frac{dy}{dt} + \frac{d\varrho}{dz} \frac{dz}{dt} \end{equation*} If one sets \begin{equation*} \tag*{[2.20]} u = \frac{dx}{dt }, \,\, v= \frac{dy}{dt }, \,\, w = \frac{dx}{dt } \end{equation*} one thus has\,: \begin{equation*} \tag*{[2.21]} \varrho\Big( \frac{du}{dx}+ \frac{dv}{dy} + \frac{dw}{dz}\Big) + { \frac{d\varrho}{dt} } + \frac{d\varrho}{dx} u + \frac{d\varrho}{dy} v + \frac{d\varrho}{dz} w = 0 \end{equation*} or \begin{equation*}\tag*{(5), [2.22]} \frac{d\varrho}{dt} +\frac {d (\varrho u)}{dx} + \frac {d (\varrho v)}{dy} +\frac {d (\varrho w)}{dz} = 0 \end{equation*} This is the form in which \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r} first presented the density equation.\hspace{0.4cm}For the case when $\varrho$ is constant in time, in this first form of dependence, one obtains as the constancy-of-volume equation\,: \deffootnotemark{\textsuperscript{[A.\thefootnotemark]}} \deffootnote{2em}{1.6em}{[A.\thefootnotemark]\enskip} \begin{equation*}\tag*{(6), [2.23]} \frac{du}{dx} + \frac{dv}{dy} + \frac{dw}{dz} = 0 \end{equation*} L\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! g\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! g\hspace{0.036cm}} %{$\phantom{!}$\! e\footnote{M\'ecanique analytique, \,\, Bd.\ I., S. 179--183, Bd.\ II, S. 257--261. [First edition, \hyperlink{Lagrange1}{1788}].} treats the relation of these equations with (2) quite extensively; but this connection seems to be a special case of% \deffootnotemark{\textsuperscript{[A.\thefootnotemark]}} \deffootnote{2em}{1.6em}{[A.\thefootnotemark]\enskip}% a theorem by \mbox{J\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! b\hspace{0.036cm}} %{$\phantom{!}$\! i}},\footnote{C.\ G.\ J.\ Jacobi, Theoria novi multiplicatoris systemati aequationum differentialium vulgarium applicandi. \,\,\, Crelle's Journal, Bd.\ 27. S.\ 209 [\hyperlink{Jacobi}{1844}].} which for three variables is [actually] fully included in equations (1) and (5).\\ \indent\indent If an integral: \begin{equation*} \iiint f(x,y,z) \,\,\varrho \,\, {\rm d}x\,\, {\rm d}y\,\, {\rm d}z \end{equation*} which is extended over the whole fluid mass, is transformed into an integral over $a,\,b,\,c$, one has: \begin{equation*} \tag*{[2.24]} \iiint f(x,y,z) \,\,\varrho \,\, {\rm d}x\,\, {\rm d}y\,\, {\rm d}z = \iiint f(a,b,c) \,\,\varrho\,\, {\rm d}a\,\, {\rm d}b\,\, {\rm d}c \begin{vmatrix} \frac{dx}{da}\,\,&\frac{dx}{db}\,\,&\frac{dx}{dc} \\\\ \frac{dy}{da}\,\,&\frac{dy}{db}\,\,&\frac{dy}{dc} \\\\ \frac{dz}{da}\,\,&\frac{dz}{db}\,\,&\frac{dz}{dc} \end{vmatrix} \end{equation*} where also the second integral has to be extended over all particles, and $f(a,\,b,\,c)$ is the function into which $f(x,\,y,\,z) $ transforms by substituting $a,\,b,\,c$ for $x,\,y,\,z$.\hspace{0.4cm} Thus, from the density equation (1) follows\,: \begin{equation*} \tag*{[2.25]} \iiint f(x,y,z) \,\,\varrho \,\, {\rm d}x\,\, {\rm d}y\,\, {\rm d}z = \iiint f(a,b,c) \,\,\varrho_0\,\, {\rm d}a\,\, {\rm d}b\,\, {\rm d}c \end{equation*} --- an important transformation. \\[1cm] \centerline{\fett{$\S.\,3.$}}\\[0.3cm] \indent\indent Despite the complete analogy of these equations for the equilibrium and motion of liquid fluids with the equations for elastic fluids, \deffootnotemark{\textsuperscript{[T.\thefootnotemark]}} \deffootnote{2em}{1.6em}{[T.\thefootnotemark]\enskip}% there is still an essential difference with regard to their derivation. \hspace{0.4cm}For this reason, these cases have to be treated separately.\\\\ \indent\indent Before developing the equations of motion, it will be convenient to study more in detail the equilibrium conditions, at first for liquid fluids.\\\\ \indent\indent If $X, Y, Z$ are the accelerating forces in the direction of the coordinates axes, acting on the point $ x,\, y,\, z$, then it follows easily from the principle of virtual velocities that\,: \begin{equation*} \tag*{[3.1]} \iiint \big[\varrho\,\big(X \delta x + Y\delta y + Z \delta z \big) + p \delta L \big]\,\, {\rm d}x\,\,{\rm d}y\,\,{\rm d}z\,=\,0 \end{equation*} in which $L=0$ gives the density equation for incompressible fluids; $\delta L$ is its relative variation corresponding to the variations of coordinates $\delta x, \delta y, \delta z$, and $p$ is a not yet determined quantity. \,\,The integral has to be extended over all parts of the continuous flow.\hspace{0.4cm}From (4) in $\S.\,2$, since $\delta \varrho=0$, we have \begin{equation*} \tag*{[3.2]} \delta L = \frac{d \delta x}{dx} + \frac{d \delta y}{dy} + \frac{d \delta z}{dz} \end{equation*} And thus the previous integral becomes: \begin{equation} \tag*{(1), [3.3]} \iiint \Big[ \varrho \Big( X\delta x + Y \delta y + Z \delta z \Big) + p\Big(\frac{d \delta x}{dx} + \frac{d \delta y} {dy} +\frac{d \delta z}{dz} \Big ) \Big] {\rm d}x\,{\rm d}y\,{\rm d}z = 0 \end{equation} Integrating by parts, one finds\,: \begin{align} \tag*{[3.4]} \begin{aligned} \iiint p \frac{d\delta x}{dx} {\rm d}x\,{\rm d}y\,{\rm d}z = \iint p\, \delta x\,{\rm d}y\,{\rm d}z - \iiint \delta x \frac{dp}{dx}\, {\rm d}x\, {\rm d}y\, {\rm d}z \nonumber \\ \iiint p \frac{d\delta y}{dy} {\rm d}x\,{\rm d}y\,{\rm d}z = \iint p\, \delta y\,{\rm d}z\,{\rm d}x - \iiint \delta y \frac{dp}{dy}\, {\rm d}x\, {\rm d}y\, {\rm d}z \\ \iiint p \frac{d\delta z}{dz} {\rm d}x\,{\rm d}y\,{\rm d}z = \iint p\, \delta z\,{\rm d}x\,{\rm d}y - \iiint \delta z \frac{dp}{dz}\, {\rm d}x\, {\rm d}y\, {\rm d}z \nonumber \end{aligned} \end{align} where the double integrals extend over the surface of the fluid mass.\hspace{0.4cm} Thus, one has for the equation of the principle of virtual velocity \,: \begin{smallequation} \tag*{[3.5]} 0=\iiint\!\!\Big [\!\big(\varrho X \!-\! \frac{dp}{dx} \big) \delta x + \big(\varrho Y - \frac{dp}{dy} \big) \delta y + \big(\varrho Z - \frac{dp}{dz} \big) \delta z \! \Big] {\rm d}x {\rm d}y {\rm d}z + \!\int\! p ({\rm d}x \cos \alpha + {\rm d}y \cos \beta + {\rm d}z \cos \gamma) {\rm d}\omega \end{smallequation} where ${\rm d}\omega$ is an element of the external surface, and $\alpha$,\,$\beta$,\,$\gamma$ indicate the angles between the normal to the element $d\omega$ and the coordinates axes.\hspace{0.4cm}From these equations, it follows that the equilibrium conditions are\,: \begin{equation} \tag* {(2), [3.6]} \int p (\delta x \,\cos \alpha + \delta y\, \cos \beta + \delta z\, \cos \gamma) {\rm d} \omega =0 \end{equation} \begin{equation} \tag* {(3), [3.7]} \frac{dp}{dx} = \varrho X, \,\,\, \frac{dp}{dy} = \varrho Y, \,\,\, \frac{dp}{dz} = \varrho Z \end{equation} The last three equations require that the components $X,Y,Z$ be the differential quotients of an arbitrary function $V$ with respect to $x,y,z$; thus \begin{equation*} \tag*{[3.8]} X=\frac{dV}{dx}, \,\, Y=\frac{dV}{dy},\,\,Z=\frac{dV}{dz} \end{equation*} so that one has\,: \begin{equation*} \tag*{[3.9]} p = \varrho V + c \end{equation*} where $p$ is determined up to an arbitrary constant $c$.\\\\ \indent\indent Instead of equations $(3)$, one can also write\,: \begin{mynewequation} \tag*{[3.10]} \begin{aligned} (p_{x + {\rm d}x}- p_x)\,{\rm d}y\, {\rm d}z-\varrho X\, {\rm d}x\, {\rm d}y\, {\rm d}z=0\\ (p_{y + {\rm d}y}- p_y)\,{\rm d}z\, {\rm d}x-\varrho Y\, {\rm d}x\, {\rm d}y\, {\rm d}z=0\\ (p_{z + {\rm d}z}-p_z)\, {\rm d}x\, {\rm d}y-\varrho Z\,\,{\rm d}x\, {\rm d}y\, {\rm d}z=0\\ \end{aligned} \end{mynewequation} from which it is apparent that $p$ is the pressure at the point $x,y,z$, which acts against the given accelerating forces.\hspace{0.4cm}This pressure is determined up to an additive constant by $p=\varrho V + c$\,; it is fully determined if its value is given at any point.\hspace{0.4cm} Suppose now that in addition to the acceleration of the fluid particles there are also pressure forces acting on the external surface. We then find as an equilibrium condition that at each point of the external surface, these pressure forces must be equal and opposite to the pressure $p = \varrho V + c$.\\\\ \indent\indent In equation (2), the variations $\delta x, \delta y, \delta z $ depend on certain conditions which result from the nature of the walls. If in special cases the actual meaning of equation (2) is specified more precisely, then its mechanical necessity becomes manifest. Here we want to discuss just a few such cases. \\\\ \indent\indent Let us assume that a part of the external surface be free and that the same forces act on all parts of it. One can set the pressure to zero in the points of that free surface, so that $p$, the difference between the pressure in a certain point and the pressure in a point of the free surface, be exactly determined in all other remaining points of the fluid.\\\\ \indent\indent However, since $ \delta x, \delta y, \delta z$ are evidently arbitrary in a free surface, it follows that, if $(2)$ has to be satisfied, we must have $p=0$.\\\\ \indent\indent For fluid parts that are lying on a fixed wall, it is evident that no motion normal to the wall surface can take place.\hspace{0.4cm}The normal component of the motion is obviously\,: \begin{equation*} \delta x\,\cos \alpha + \delta y \,\cos \beta + \delta z\,\cos \gamma \end{equation*} and, since, this must vanish, equation (2) is indeed satisfied.\\\\ \indent\indent In the points which are on a moving wall, one can set \begin{equation*} \tag*{[3.11]} \delta x = \delta {x'} + \delta\xi,\,\, \delta y = \delta {y'} + \delta \eta, \,\,\delta z = \delta {z'} + \delta \zeta \end{equation*} where $\delta {x'}, \,\,\delta {y'} ,\,\,\delta {z'}$ are the motions relative to the wall and $\delta\xi,\,\, \delta \eta,\,\,\delta\chi$ are the motions of the fluid particles simultaneously in motion with the wall.\hspace{0.4cm}Therefore, one has instead of equation (2) \begin{equation*} \tag*{[3.12]} 0=\int p (\delta x' \,\, \cos\alpha + \delta y'\,\, \cos\beta + \delta z'\,\, \cos \gamma) {\rm d} \omega+ \int p (\delta\xi\,\, \cos\alpha + \delta \eta\,\, \cos\beta + \delta \zeta\,\, \cos \gamma) {\rm d} \omega \end{equation*} but, since no motion may happen against the wall, one has\,: \begin{equation*} \tag*{[3.13]} \delta x' \,\, \cos\alpha + \delta y'\,\, \cos\beta+ \delta z'\,\, \cos\gamma=0 \end{equation*} and therefore \begin{equation*} \tag*{[3.14]} \int p (\delta \xi \,\,\cos \alpha + \delta \eta\,\, \cos \beta + \delta \zeta \,\, \cos \gamma) {\rm d} \omega = 0 \end{equation*} which amounts to setting \begin{equation*} \tag*{[3.15]} \int \big (p_x \delta \xi + p_y \delta \eta + p_z \delta \zeta \,) {\rm d} \omega =0 \end{equation*} provided that $p_x, p_y, p_z$ are the components of the pressure with respect to the coordinates axes.\hspace{0.4cm}However this integral is the equilibrium condition of a body, on which act external surface forces $p_x, p_y, p_z$, where $ \delta \xi, \delta \eta, \delta \zeta$ indicate the variations, that the body can have, under the given circumstances.\hspace{0.4cm}Actually, also in this case, equation (2) is needed by the nature of things.% \deffootnotemark{\textsuperscript{[A.\thefootnotemark]}}\deffootnote{2em}{1.6em}{[A.\thefootnotemark]\enskip}% \footnote {Cf.\ M\'ec.\ analy.\ Bd. I, S.\ 193--201. [First edition, \hyperlink{Lagrange1}{1788}].} \\\\ \indent\indent We now discuss the case of elastic fluids; here the relation $L=0$ is not valid anymore.\hspace{0.4cm}In that case one has to consider also the forces due to elasticity of the fluid in addition to the accelerating and external pressure forces.\hspace{0.4cm}Let $p$ be the pressure at a point $x, y, z$, then this tends to reduce the volume of the element ${\rm d}x,\, {\rm d}y,\, {\rm d}z$\,; the momentum by this force is thus $p \delta ({\rm d}x \, {\rm d}y\, {\rm d}z)$; in order to get a different expression for $ \delta ({\rm d}x \, {\rm d}y \,{\rm d}z)$, we note that $\rho\,{\rm d}x\,{\rm d}y\,{\rm d}z$, being the mass of an element, is always the same; thus we have $\delta (\rho\,{\rm d}x\, {\rm d}y\,{\rm d}z)=0$; herefrom it follows that\: \begin{equation*} \tag*{[3.16]} \varrho \delta ( {\rm d} x \,\, {\rm d}y\,\, {\rm d}z) + {\rm d}x\,\, {\rm d}y\,\, {\rm d}z\,\, \delta\varrho=0 \end{equation*} and thus \begin{equation*} \tag*{[3.17]} \delta ( {\rm d}x\,\, {\rm d}y\,\, {\rm d}z) = - \frac{\delta \varrho}{\varrho} {\rm d}x\,\, {\rm d}y\,\, {\rm d} z \end{equation*} or, by equation $(4)$ in \S$.2$, \begin{equation*} \tag*{[3.18]} \delta ( {\rm d}x\,\, {\rm d}y\,\, {\rm d}z) \,\,=\,\,\Big(\frac{d \delta x}{dx}\,\,+ \frac{d \delta y}{dy}\,\,+\frac{d \delta z}{dz}\,\,\Big)\,\ {\rm d}x\,\,{\rm d}y\,\, {\rm d}z; \end{equation*} Therefore the equilibrium conditions are\,: \begin{equation*} \tag*{[3.19]} \iiint \Big[ \varrho \Big( X \delta x\,\,+\,\,Y\delta y\,\,+ Z\delta z \Big) \,\, + p \Big(\frac{d \delta x}{dx}\,\,+ \frac{d \delta y}{dy}\,\,+\frac{d \delta z}{dz}\,\,\Big) \Big]\,\ {\rm d}x\,\, {\rm d}y\,\, {\rm d}z = 0 \end{equation*} \indent\indent Since this equation is identical with $(1)$, the equilibrium equations for elastic and liquid flows are formally the same.\hspace{0.4cm} Also here, in accordance with equation $(2)$, we have\,: \begin{equation} \tag*{[3.20]} \frac{dp}{dx} = \varrho X, \,\,\frac{dp}{dy} = \varrho Y,\,\,\frac{dp}{dz} = \varrho Z \end{equation} For elastic flows, $\rho$ is a given function of $p$, say, $\varrho = \varphi (p)$.\hspace{0.4cm} Let us put \begin{equation*} \tag*{[3.21]} f(p) = \int \frac{{\rm d}p}{\varphi(p)} \end{equation*} from which it follows obviously that \begin{math} \displaystyle \frac{1}{\varrho} \frac{dp}{dx} = \frac{1}{\varphi(p)} \frac{dp}{dx} = \frac{df(p)}{dp} \frac{dp}{dx} = \frac{df(p)}{dx}, \end{math} therefore, the three equations for the equilibrium condition become\,: \begin{equation*} \tag*{[3.22]} \frac{df(p)}{dx} = X, \,\,\, \frac{df(p)}{dy} = Y, \,\,\, \frac{df(p)}{dx} = Z\,, \end{equation*} so that, also for elastic fluids in equilibrium, $X,Y,Z$ have to be partial differential quotients of the same function with respect to $x,y,z$.\hspace{0.4cm}As $\varphi(p)$, and consequently also $f(p)$ is known, $p$ can always be expressed through $X,Y,Z$.\\[1cm] \centerline{\fett{$\S.\,4.$}}\\[0.3cm] \indent\indent As follows from the considerations of $\S.\,3$, the principle of virtual velocities and lost forces for the motion of liquid and elastic fluids, implies\,: \tagsleft@true \begin{smallequation} \notag \,\,\,\,0=\iiint \biggl\{ \varrho \biggl[ \Big( X - \frac{d^2 x}{d t^2} \Big) \delta x + \Big( Y - \frac{d^2 y}{d t^2}\Big) \delta y + \Big(Z - \frac{d^2 z}{d t^2}\Big) \delta z \biggr] + p \biggl[ \frac{d \delta x}{dx} + \frac{d \delta y}{dy} + \frac{d \delta z}{dz} \biggr] \biggr\} {\rm d}x {\rm d}y {\rm d}z \end{smallequation} \tagsleft@false \vskip-1.15cm\begin{align}\tag*{[4.1]} \end{align} \vskip-1.07cm \mbox{\!\!\!\!\!\!(1)}\vskip0.8cm \noindent from which follows firstly equation $(2)$ of $\S.\,3$\,: \tagsleft@true \begin{align} \tag{2} \int p \big(\delta x \cos \alpha + \delta y \cos \beta + \delta z \cos \gamma \big) {\rm d} \omega = 0, \end{align} \tagsleft@false \vskip-1.5cm\begin{align}\tag*{[4.2]} \end{align} which concerns only the external surface; secondly, we have for the fundamental equations of liquid or elastic fluids, \tagsleft@true \begin{align}\tag{3} \varrho \Big ( \frac{d^2 x}{d t^2} - X \Big) + \frac{dp}{dx} =0,\,\, \varrho \Big ( \frac{d^2 y}{d t^2} - Y \Big) + \frac{dp}{dy} =0, \,\,\varrho \Big ( \frac{d^2 z}{d t^2} - Z \Big) + \frac{dp}{dz} =0 \,, \end{align} \tagsleft@false \vskip-1.5cm\begin{align}\tag*{[4.3]} \end{align} \noindent where $p$ indicates the pressure in each point. \\\\ \indent \indent According to the first \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! ian} method, the components $u,\,v,\,\omega$ are considered as function of time and space $x,\,y,\,z$. Therefore, \begin{mynewequation}\tag*{[4.4]} \begin{aligned} \frac{d^2 x}{dt^2} &=\frac{du}{dt} + \frac{du}{dx} u + \frac{du}{dy} v + \frac{du}{dz} w\,\,\,\\ \frac{d^2 y}{dt^2} &=\frac{dv}{dt} + \frac{dv}{dx} u + \frac{dv}{dy} v + \frac{dv}{dz} w\,\,\,\,\\ \frac{d^2 z}{dt^2} &=\frac{dw}{dt} + \frac{dw}{dx} u + \frac{dw}{dy} v + \frac{dw}{dz} w \end{aligned} \end{mynewequation} so that the fundamental equations in this form are the following\,: \begin{mynewequation}\tag*{(4), [4.5]} \left. \begin{aligned} \frac{du}{dt}\, +\, \frac{du}{dx} u + \frac{du}{dy} v + \frac{du}{dz} w - X + \frac{1}{\varrho}\frac{dp}{dx}=0\\\\ \frac{dv}{dt}\, +\,\, \frac{dv}{dx} u + \frac{dv}{dy} v + \frac{dv}{dz} w - Y + \frac{1}{\varrho}\frac{dp}{dy}=0\\\\ \frac{dw}{dt}\, + \frac{dw}{dx} u + \frac{dw}{dy} v + \frac{dw}{dz} w - Z + \frac{1}{\varrho}\frac{dp}{dz}=0 \end{aligned} \qquad \right\} \end{mynewequation} which may be also written in such a way that each equation is obtained from the other by cyclic permutation, namely\,: \begin{mynewequation}\tag*{(5), [4.6]} \left. \begin{aligned} \frac{du}{dt}\, +\, \frac{du}{dx} u + \frac{du}{dy} v + \frac{du}{dz} w - X + \frac{1}{\varrho}\frac{dp}{dx}=0\\\\ \frac{dv}{dt}\, +\,\, \frac{dv}{dy} v + \frac{dv}{dz} w + \frac{dv}{dx} u - Y + \frac{1}{\varrho}\frac{dp}{dy}=0\\\\ \frac{dw}{dt}\, + \frac{dw}{dz} w + \frac{dw}{dx} u + \frac{dw}{dy} v - Z + \frac{1}{\varrho}\frac{dp}{dz}=0 \end{aligned} \qquad \right\} \end{mynewequation} In addition to these formulae, there is also the density equation~(5) of the $\S.\,2$: \begin{equation*} \tag*{[4.7]} \frac{d\varrho}{dt} + \frac{d(\varrho u )}{dx} + \frac{d(\varrho v )}{dy} + \frac{d(\varrho w )}{dz} =0 \end{equation*} or, in particular, for liquid flows \begin{equation*} \tag*{[4.8]} \frac{du}{dx} + \frac{dv}{dy} + \frac{dw}{dz}=0 \end{equation*} We see that these four equations are sufficient to determine the four unknowns $u,v,w$ and $p$ as functions of $x,y,z$ and $t$;\, $\varrho$ is either a known function of $p$ or a constant.\\\\ \indent\indent By the second E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! erian representation [Lagrangian representation], equations~$(3)$ are respectively multiplied by $$ \frac{dx}{da},\,\,\frac{dy}{da},\,\,\frac{dz}{da}$$ and summed, then, similarly, by $$ \frac{dx}{db},\,\,\frac{dy}{db},\,\,\frac{dz}{db}$$ and $$ \frac{dx}{dc},\,\,\frac{dy}{dc},\,\,\frac{dz}{dc}$$ \noindent Thus, one obtains for the three fundamental equations: \begin{equation*}\tag*{(6), [4.9]} \left. \begin{aligned} \Big( \frac{d^2 x}{d t^2} - X \Big)\frac{dx}{da} + \Big( \frac{d^2 y}{d t^2} - Y \Big) \frac{dy}{da}+\Big( \frac{d^2 z}{d t^2} - Z \Big) \frac{dz}{da} + \frac{1}{\varrho} \frac{dp}{da} =0\\\ \Big( \frac{d^2 x}{d t^2} - X \Big)\frac{dx}{db} + \Big( \frac{d^2 y}{d t^2} - Y \Big) \frac{dy}{db}+\Big( \frac{d^2 z}{d t^2} - Z \Big) \frac{dz}{db} + \frac{1}{\varrho} \frac{dp}{db} =0\\\ \Big( \frac{d^2 x}{d t^2} - X \Big)\frac{dx}{dc} + \Big( \frac{d^2 y}{d t^2} - Y \Big) \frac{dy}{dc}+\Big( \frac{d^2 z}{d t^2} - Z \Big) \frac{dz}{dc} + \frac{1}{\varrho} \frac{dp}{dc} =0 \end{aligned} \qquad \right\} \end{equation*} and, in addition, there is the density equation (1) of $\S.\,2$ \begin{myequation} \tag*{[4.10]} \begin{vmatrix} \frac{dx}{da}&\frac{dx}{db}&\frac{dx}{dc} \\\\ \frac{dy}{da}&\frac{dy}{db}&\frac{dy}{dc} \\\\ \frac{dz}{da}&\frac{dz}{db}&\frac{dz}{dc} \end{vmatrix} =\frac{\varrho_0}{\varrho} \end{myequation} From these four equations $x,y,z$ and $p$ are found as functions of the initial location $a,b,c$ and time $t$.\\ \indent\indent Evidently, the solutions of these partial differential equations must contain arbitrary functions, which have to be determined from initial conditions and are in accordance with the nature of the walls and the flow boundaries.\\ \indent\indent These last equations, which are usually called after \mbox{L\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! g\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! g\hspace{0.036cm}} %{$\phantom{!}$\! e,} significantly simplify their form by setting\,: \begin{equation*} \tag*{[4.11]} X= \frac{dV}{dx}, \,\,Y=\frac{dV}{dy}, \,\, Z=\frac{dV}{dz} \end{equation*} so they become\,: \begin{equation*}\tag*{(7), [4.12]} \left. \begin{aligned} \frac{d^2 x}{d t^2}\frac{dx}{da} + \frac{d^2 y}{d t^2} \frac{dy}{da}+\frac{d^2 z}{d t^2} \frac{dz}{da} - \frac{dV}{da} + \frac{1}{\varrho}\frac{dp}{da}=0\\\\ \frac{d^2 x}{d t^2}\frac{dx}{db} + \frac{d^2 y}{d t^2} \frac{dy}{db}+\frac{d^2 z}{d t^2} \frac{dz}{db} - \frac{dV}{db} + \frac{1}{\varrho}\frac{dp}{db}=0\\\\ \frac{d^2 x}{d t^2}\frac{dx}{dc} + \frac{d^2 y}{d t^2} \frac{dy}{dc}+\frac{d^2 z}{d t^2} \frac{dz}{dc} - \frac{dV}{dc} + \frac{1}{\varrho}\frac{dp}{dc}=0 \end{aligned} \qquad \right\} \end{equation*} We will limit ourselves to this assumption about $X,Y,Z$ which, apart from the boundary conditions, \noindent coincides with the one that necessarily has to hold in the equilibrium state of the fluid. \\ \centerline{\fett{$\S.\,5.$}}\\[0.3cm] \indent\indent Partially integrating the last three terms of equation (1) of $\S.\,4$, ignoring the boundary contributions to the double integrals and using as already stated \begin{equation*} \tag*{[5.1]} X= \frac{dV}{dx}, \,\,Y=\frac{dV}{dy}, \,\, Z=\frac{dV}{dz} \,, \end{equation*} one obtains the following equation\,: \begin{equation*} \tag*{[5.2]} 0 = \iiint \varrho\,{\rm d}x\,{\rm d}y\, {\rm d}z\ \bigg\{\Big(\frac{d^2 x}{d t^2} - \frac{dV}{dx} + \frac{1}{\varrho}\frac{dp}{dx}\Big) \delta x + \Big(\frac{d^2 y}{d t^2} - \frac{dV}{dy} + \frac{1}{\varrho}\frac{dp}{dy}\Big) \delta y + \Big(\frac{d^2 z}{d t^2} - \frac{dV}{dz} + \frac{1}{\varrho}\frac{dp}{dz}\Big) \delta z \bigg\} \end{equation*} If one puts, as in $\S.\,3$, $\varrho= \varphi (p )$ and: \begin{equation*} \tag*{[5.3]} f(p) =\int\frac{ {\rm d}p}{\varphi(p)} \end{equation*} one has, \begin{smallequation} \tag*{[5.4]} \!\!\!\!0 = \iiint \varrho_0 \, {\rm d}a\, {\rm d}b\, {\rm d}c \bigg\{\Big(\frac{d^2 x}{d t^2} - \frac{dV}{dx} + \frac{df(p)}{dx}\Big) \delta x + \Big(\frac{d^2 y}{d t^2} - \frac{dV}{dy} + \frac{df(p)}{dy}\Big) \delta y + \Big(\frac{d^2 z}{d t^2} - \frac{dV}{dz} + \frac{df(p)}{dz}\Big) \delta z \bigg\} \,, \end{smallequation} provided that the transformation of the integral is made according to \S.\,2. Now, if one sets \begin{equation*} \tag*{[5.5]} \label{pointerOmega} V - f(p) = \Omega, \end{equation*} one can write\,: \begin{equation*} \tag*{[5.6]} 0 = \iiint \varrho_0 \, {\rm d}a\, {\rm d}b\, {\rm d}c \bigg[ \frac{d^2 x}{d t^2} \delta x + \frac{d^2 y}{d t^2} \delta y + \frac{d^2 z}{d t^2} \delta z -\delta \Omega \bigg] \end{equation*} If one now integrates under the triple integral with respect to the variable $t$ which is independent of $a,b,c$, one obtains \begin{equation} \tag*{[5.7]} \begin{aligned} \int \frac{d^2 x}{dt^2} \delta x\, {\rm d}t = \Big[\frac{dx}{dt} \delta x\Big] - \int \frac{dx}{dt} \frac{d \delta x}{dt} {\rm d} t\\\\ \int\frac{d^2 y}{dt^2} \delta y\, {\rm d}t= \Big[\frac{dy}{dt} \delta y\Big] - \int\frac{dy}{dt}\frac{d \delta y}{dt} {\rm d }t\\\\ \int\frac{d^2 z}{dt^2} \delta z\, {\rm d}t = \Big[\frac{dz}{dt} \delta z\Big] - \int\frac{dz}{dt}\frac{d \delta z}{dt} {\rm d} t \end{aligned} \end{equation} \noindent Dropping the triple integrals, one has the equation\,: \begin{equation*} \tag*{[5.8]} 0 = \iiint \varrho_0 {\rm d}a\, {\rm d}b\, {\rm d}c \int {\rm d}t \bigg\{\frac{dx}{dt} \delta \frac{dx}{dt} + \frac{dy}{dt} \delta \frac{dy}{dt} +\frac{dz}{dt} \delta \frac{dz}{dt} + \delta \Omega\bigg\} \end{equation*} which coincides with the following\,:\hypertarget{eq1}{} \begin{equation} \tag*{(1), [5.9]} 0 = \delta \iiint \varrho_0\, {\rm d}a\, {\rm d}b\, {\rm d}c \int {\rm d}t \Big [\Big(\frac{ds}{dt}\Big)^2 + 2 \Omega \Big] \,. \end{equation} The first three hydrodynamical fundamental equations are satisfied. Hence \begin{equation*} \notag \iiint \varrho_0\, {\rm d}a\, {\rm d}b\, {\rm d}c \int {\rm d}t \Big [\Big(\frac{ds}{dt}\Big)^2 + 2 \Omega \Big] \end{equation*} disappears; conversely, if the first variation of this integral with respect to $x,y,z$ is set to zero, then one obtains these first three equations. \\\\ \indent\indent This theorem, which can be easily considered, in view of the meaning of $\displaystyle \frac{ds}{dt}$ and $\Omega$, as a mechanical principle, has a certain analogy with the principle of least action. For us it possesses an analytical importance, since it gives an extremely simple tool for transforming the hydrodynamical equations. \hspace{0.4cm} Indeed, in order to introduce in these equations new coordinates, instead of $x,y,z$, it is only necessary to write the arc element \begin{equation*} \tag*{[5.10]} {\rm d}s^2= {\rm d}x^2 + {\rm d}y^2 + {\rm d}z^2 \end{equation*} in terms of the new coordinates and, then, apply the simple operation of variation, using the integral in the new coordinates\,: \begin{equation*} \notag \iiint \varrho_0\,{\rm d} a\, {\rm d} b\, {\rm d}c \int {\rm d} t \Big[\Big(\frac{ds}{dt}\Big)^2 + 2 \Omega \Big] \end{equation*} By setting the coefficients of the three variations to zero, we thereby obtain three equations in a similar form as equations (3) in $\S.\,4$\,; in order to write them into the first or second E\,u\,l\,e\,r\,ian form, one has to apply analogous procedures as was done in $\S\,4$.\\ \indent\indent If the new coordinates\,: \begin{equation*} \tag*{[5.11]} \varrho_1=f_1(x,\,\,y,\,\,z), \,\,\,\,\varrho_2=f_2(x,\,\,y,\,\,z),\,\,\,\,\varrho_3=f_3(x,\,\,y,\,\,z) \end{equation*} are used instead of $x,\,y,\,z$, one obtains: \begin{mynewequation} \tag*{[5.12]} \begin{aligned} {\rm d}x = \frac{dx}{d \varrho_1}{\rm d} \varrho_1 + \frac{dx}{d\varrho_2}{\rm d} \varrho_2 + \frac{dx}{d \varrho_3} {\rm d} \varrho_3\\ {\rm d}y = \frac{dy}{d \varrho_1}{\rm d} \varrho_1 + \frac{dy}{d\varrho_2}{\rm d} \varrho_2 + \frac{dy}{d \varrho_3} {\rm d} \varrho_3\\ {\rm d}z = \frac{dz}{d \varrho_1}{\rm d} \varrho_1 + \frac{dz}{d\varrho_2}{\rm d} \varrho_2 + \frac{dz}{d \varrho_3} {\rm d} \varrho_3\\ \end{aligned} \end{mynewequation} therefore \begin{equation*} \tag*{[5.13]} {\rm d}s^2 = {\rm d}x^2+{\rm d}y^2+{\rm d}z^2=N_1 {\rm d} \varrho_1^2 + N_2 {\rm d} \varrho_2^2 +N_3 {\rm d} \varrho_3^2+ 2 n_3 {\rm d} \varrho_1 {\rm d}\varrho_2 + 2n_1 {\rm d} \varrho_2 {\rm d} \varrho_3 + 2n_2 {\rm d} \varrho_3 {\rm d}\varrho_1 \end{equation*} where \begin{align} \tag*{[5.14]} \begin{aligned} N_1= \Big(\frac{dx}{d \varrho_1}\Big)^2 + \Big(\frac{dy}{d \varrho_1}\Big)^2 + \Big(\frac{dz}{d \varrho_1}\Big)^2\\ N_2= \Big(\frac{dx}{d \varrho_2}\Big) ^2 + \Big(\frac{dy}{d \varrho_2}\Big)^2 + \Big(\frac{dz}{d \varrho_2}\Big)^2\\ N_3= \Big(\frac{dx}{d \varrho_3}\Big) ^2 + \Big(\frac{dy}{d \varrho_3}\Big)^2 + \Big(\frac{dz}{d \varrho_3}\Big)^2 \end{aligned} \\[0.3cm] \tag*{[5.15]} \begin{aligned} n_3= \frac{dx}{d \varrho_1} \frac{dx}{d \varrho_2} + \frac{dy}{d \varrho_1} \frac{dy}{d \varrho_2} + \frac{dz}{d \varrho_1} \frac{dz}{d \varrho_2}\\\\ n_1= \frac{dx}{d \varrho_2} \frac{dx}{d \varrho_3} + \frac{dy}{d \varrho_2} \frac{dy}{d \varrho_3} + \frac{dz}{d \varrho_2} \frac{dz}{d \varrho_3}\\\\ n_2= \frac{dx}{d \varrho_3} \frac{dx}{d \varrho_1} + \frac{dy}{d \varrho_3} \frac{dy}{d \varrho_1} + \frac{dz}{d \varrho_3} \frac{dz}{d \varrho_1} \end{aligned} \end{align} \indent\indent Once $N_1, N_2, N_3,n_1,n_2,n_3$ are expressed in terms of the new variables $\varrho_1, \varrho_2, \varrho_3$, one has to vary the integral \begin{smallequation} \notag \iint\!\!\varrho_0 \, {\rm d}a\, {\rm d}b\, {\rm d}c \int\!\! {\rm d}t \Big[N_1 \Big(\frac{d \varrho_1}{dt} \Big)^2\! + N_2 \Big(\frac{d \varrho_2}{dt} \Big)^2 \!+N_3 \Big(\frac{d \varrho_3}{dt} \Big)^2\! + 2 n_3 \frac{d\varrho_1}{dt} \frac{d \varrho_2}{dt} + 2 n_1 \frac{d\varrho_2}{dt} \frac{d \varrho_3}{dt} + 2 n_2 \frac{d\varrho_3}{dt} \frac{d \varrho_1}{dt} + 2 \Omega \Big] \end{smallequation} with respect to these new variables. Then, after integration by parts with respect to $t$, one removes from the quadruple integral the quantities in which appear the time derivatives of the variations $\delta \varrho_1, \delta \varrho_2, \delta \varrho_3$. Then one has to set the coefficients of $\delta \varrho_1, \delta \varrho_2, \delta \varrho_3$ equal to zero.\hspace{0.4cm} After that one obtains the first three hydrodynamical fundamental equations, which are completed by a fourth one, the density equation.\hspace{0.6cm} In order to express also the density equation in the new coordinates, we notice that \noindent the volume element ${\rm d}x\,{\rm d}y\,{\rm d}z$ may be expressed in such coordinates very easily, namely\,: \begin{equation}\tag*{[5.16]} \begin{aligned} {\rm d}x\,{\rm d}y\,{\rm d}z= {\rm d}\varrho_1\, {\rm d}\varrho_2\, {\rm d}\varrho_3 \begin{vmatrix} \frac{dx}{d\varrho_1}&\frac{dx}{d\varrho_2}&\frac{dx}{d\varrho_3} \\\\ \frac{dy}{d\varrho_1}&\frac{dy}{d\varrho_2}&\frac{dy}{d\varrho_3} \\\\ \frac{dz}{d\varrho_3}&\frac{dz}{d\varrho_3}&\frac{dz}{d\varrho_3} \end{vmatrix} \end{aligned} \end{equation} herefrom follows \begin{equation} \notag ({\rm d} x\,{\rm d}y\,{\rm d}z)^2= ({\rm d}\varrho_1\,{\rm d}\varrho_2\,{\rm d}\varrho_3)^2 {\bf { \times}} \\[-0.5cm] \end{equation} \begin{equation}\tag*{[5.17]} \begin{vmatrix} (\frac{dx}{d\varrho_1})^2 +(\frac{dy}{d\varrho_1})^2+(\frac{dz}{d\varrho_1})^2,& \frac{dx}{d\varrho_1} \frac{dx}{d\varrho_2} + \frac{dy}{d\varrho_1} \frac{dy}{d\varrho_2}+\frac{dz}{d\varrho_1} \frac{dz}{d\varrho_2},& \frac{dx}{d\varrho_3} \frac{dx}{d \varrho_1} + \frac{dy}{d\varrho_3} \frac{dy}{d \varrho_1}+\frac{dz}{d\varrho_3}\frac{dz}{d\varrho_1}\\\\ \frac{dx}{d\varrho_1 }\frac{dx}{d\varrho_2} + \frac{dy}{d\varrho_1}\frac{dy}{d\varrho_2}+\frac{dz}{d\varrho_1} \frac{dz}{d\varrho_2},& (\frac{dx}{d\varrho_2})^2 +(\frac{dy}{d\varrho_2})^2+(\frac{dz}{d\varrho_2})^2,& \frac{dx}{d\varrho_2} \frac{dx}{d\varrho_3} + \frac{dy}{d\varrho_2} \frac{dy}{d\varrho_3}+\frac{dz}{d\varrho_2} \frac{dz}{d\varrho_3}\\\\ \frac{dx}{d\varrho_3} \frac{dx}{d\varrho_1} + \frac{dy}{d\varrho_3} \frac{dy}{d\varrho_1}+\frac{dz}{d\varrho_3} \frac{dz}{d\varrho_1},& \frac{dx}{d\varrho_2} \frac{dx}{d\varrho_3} + \frac{dy}{d\varrho_2} \frac{dy}{d\varrho_3}+\frac{dz}{d\varrho_2} \frac{dz}{d\varrho_3},& (\frac{dx}{d\varrho_3})^2 +(\frac{dy}{d\varrho_3})^2+(\frac{dz}{d\varrho_3})^2& \end{vmatrix} \end{equation} or, with the notation as defined above\,: \begin{mynewequation} \tag*{[5.18]} ({\rm d}x\,{\rm d}y\,{\rm d}z)^2= ({\rm d}\varrho_1\,{\rm d}\varrho_2\,{\rm d}\varrho_3)^2 \begin{vmatrix} N_1&n_3&n_2\\\\ n_3&N_2&n_1\\\\ n_2&n_1&N_3 \end{vmatrix} \end{mynewequation} \indent\indent Let us denote the values of $\varrho_1, \varrho_2,\varrho_3, N_1, N_2, N_3, n_1,n_2, n_3 $ at time $t=0$ with $ \varrho_1^0, \varrho_2^0,\varrho_3^0, N_1^0, N_2^0, N_3^0, n_1^0,n_2^0, n_3^0 $, then we get from the previous equation for $t=0$ \begin{mynewequation} \tag*{[5.19]} ({\rm d}a\,{\rm d}b\,{\rm d}c)^2= ({\rm d}\varrho_1^0\,{\rm d}\varrho_2^0\,{\rm d}\varrho_3^0)^2 \begin{vmatrix} N_1^0&n_3^0&n_2^0\\\\ n_3^0&N_2^0&n_1^0\\\\ n_2^0&n_1^0&N_3^0 \end{vmatrix} \end{mynewequation} One also has to think of the ensuing value of ${\rm d}a\,{\rm d}b\,{\rm d}c$ as substituted into the integral to be varied. But now the density equation is, for the general case\,: \begin{equation*} \tag*{[5.20]} \frac{ {\rm d}x\,{\rm d}y\,{\rm d}z}{{\rm d}a\,{\rm d}b\,{\rm d}c} = \frac{\varrho_0}{\varrho} \,. \end{equation*} Then, dividing $({\rm d}x\,\,{\rm d}y\,\,{\rm d}z)^2$ by $({\rm d}a\,\,{\rm d}b\,\,{\rm d}c)^2$, one obtains the density equation in the new variables\,: \begin{mynewequation} \tag*{[5.21]} \Big(\frac{{\rm d} \varrho_1}{{\rm d}\varrho_1^0} \frac{{\rm d} \varrho_2}{{\rm d}\varrho_2^0} \frac{{\rm d} \varrho_3}{{\rm d} \varrho_3^0} \Big)^2\,\,\,\Big(\frac{\varrho}{\varrho_0}\Big)^2 = \begin{vmatrix} N_1^0& n_3^0&n_2^0\\ n_3^0&N_2^0&n_1^0\\ n_2^0&n_1^0&N_3^0 \end{vmatrix} : \begin{vmatrix} N_1& n_3&n_2\\ n_3&N_2&n_1\\ n_2&n_1&N_3 \end{vmatrix} \end{mynewequation} or, as it is well-known \begin{mynewequation} \tag*{[5.22]} \text{\small ${\rm d}\varrho_1\,{\rm d}\varrho_2\,{\rm d}\varrho_3 = {\rm d}\varrho_1^0\,{\rm d}\varrho_2^0\,{\rm d}\varrho_3^0$} \begin{vmatrix} \frac{d\varrho_1}{d\varrho_1^0} &\frac{d\varrho_1}{d\varrho_2^0} &\frac{d\varrho_1}{d\varrho_3^0} \\\\ \frac{d\varrho_2}{d\varrho_1^0} &\frac{d\varrho_2}{d\varrho_2^0} &\frac{d\varrho_2}{d\varrho_3^0} \\\\ \frac{d\varrho_3}{d\varrho_1^0} &\frac{d\varrho_3}{d\varrho_2^0} &\frac{d\varrho_3}{d\varrho_3^0} \end{vmatrix} \,, \end{mynewequation} one finally obtains\,: \begin{mynewequation}\tag*{[5.23]} \begin{vmatrix} \frac{d\varrho_1}{d\varrho_1^0} &\frac{d\varrho_1}{d\varrho_2^0} &\frac{d\varrho_1}{d\varrho_3^0} \\\\ \frac{d\varrho_2}{d\varrho_1^0} &\frac{d\varrho_2}{d\varrho_2^0} &\frac{d\varrho_2}{d\varrho_3^0} \\\\ \frac{d\varrho_3}{d\varrho_1^0} &\frac{d\varrho_3}{d\varrho_2^0} &\frac{d\varrho_3}{d\varrho_3^0} \end{vmatrix} \cdot \text{\Large $\frac{\varrho}{\varrho_0}$} ={\boldsymbol {\surd }}\begin{vmatrix} N_1^0& n_3^0&n_2^0\\ n_3^0&N_2^0&n_1^0\\ n_2^0&n_1^0&N_3^0 \end{vmatrix} : \begin{vmatrix} N_1& n_3&n_2\\ n_3&N_2^0&n_1\\ n_2&n_1&N_3 \end{vmatrix} \end{mynewequation} \indent\indent All quantities appearing in this transformed density equation are already known through the transformation of the arc element.\hspace{0.4cm}Here, the problem of the transformation of the four hydrodynamical equations in an arbitrary coordinate system is reduced to the problem of the transformation of the arc element.\\ \indent\indent It is obvious that the applicability of this procedure does not depend on the number of variables and that, in the same way, by variation of the integral with respect to $x_1,x_2,...,x_n$: \begin{equation*} \notag \iiint \varrho_0\, {\rm d}a_1 {\rm d}a_2\,...\,{\rm d}a_n \int {\rm d}t \biggl\{ \Big(\frac{ds}{dt}\Big)^2 + 2 \Omega\biggr\} \end{equation*} one obtains \begin{mynewequation}\tag*{[5.24]} \left. \begin{aligned} \frac{d^2 x_1}{d t^2} \frac{dx_1}{da_1} + \frac{d^2 x_2}{d t^2} \frac{dx_2}{da_1}+...+ \frac{d^2 x_n}{d t^2} \frac{dx_n}{da_1}- \frac{dV}{da_1}+ \frac{1}{\varrho}\, \frac{dp}{da_1} =0\\ \,\,.\,\,\,\,.\,\,\,\,.\,\,\,\,.\,\,\,\,.\,\,\,\,.\,\,\,\,.\,\,\,\,.\,\,\,\,.\,\qquad \qquad \qquad \qquad \qquad \\ \frac{d^2 x_1}{d t^2} \frac{dx_1}{da_n} + \frac{d^2 x_2}{d t^2} \frac{dx_2}{da_n}+...+ \frac{d^2 x_n}{d t^2} \frac{dx_n}{da_n}- \frac{dV}{da_n}+ \frac{1}{\varrho} \frac{dp}{da_n} =0\\ \end{aligned} \qquad \right\} \end{mynewequation} \noindent where $a_1, a_2, .\,. \,a_n$ are the values of $x_1, x_2,.\,. \,x_n$ at time $t=0$.\hspace{0.4cm}Then, the transformation of these equations happens exactly in the same way.\hspace{0.4cm}For the sake of brevity, we here limit ourselves to three variables.\\ \indent\indent This transformation happens to be very simple when the new variables form an orthogonal system, a case, which apart from this simplification, is also very interesting since all frequently used coordinate systems are included therein.\\ \indent\indent The points where $\varrho_1$ takes a given value will in general form a surface, whose equation with respect to the axes of $x,\,y,\,z$ is \begin{equation*} \tag*{[5.25]} \varrho_1 = f_1 (x,\, y,\,z) \end{equation*} The cosines of the angles formed by the normal to the point $x, \,y, \,z$ of this surface and the coordinates axes, are\,: \begin{equation*} \tag*{[5.26]} \frac{1}{\Delta_1} \,\frac{d \varrho_1}{dx}, \,\,\,\,\,\,\,\frac{1}{\Delta_1}\,\frac{d \varrho_1}{dy}, \,\,\,\,\,\,\frac{1}{\Delta_1}\, \frac{d \varrho_1}{dz}, \,\,\,\,\,\,\,\,\,\Delta_1^2= \Big(\frac{d\varrho_1}{dx}\Big)^2 + \Big(\frac{d\varrho_1}{dy}\Big)^2 +\Big(\frac{d\varrho_1}{dz}\Big)^2 \end{equation*} The analogous cosines for the normal to the surface \begin{equation*} \tag*{[5.27]} \varrho_2 = f_2 (x,\, y,\, z) \end{equation*} are\,: \begin{equation*} \tag*{[5.28]} \frac{1}{\Delta_2} \,\frac{d \varrho_2}{dx}, \frac{1}{\Delta_2}\,\frac{d \varrho_2}{dy}, \frac{1}{\Delta_2}\, \frac{d \varrho_2}{dz}, \,\,\,\,\,\,\,\,\,\Delta_2^2= \Big(\frac{d\varrho_2}{dx}\Big)^2 + \Big(\frac{d\varrho_2}{dy}\Big)^2 +\Big(\frac{d\varrho_2}{dz}\Big)^2; \end{equation*} for the surface \begin{equation*} \tag*{[5.29]} \varrho_3 = f_3 (x, y, z) \end{equation*} the analogous cosines will be\,: \begin{equation*} \tag*{[5.30]} \frac{1}{\Delta_3} \,\frac{d \varrho_3}{dx}, \frac{1}{\Delta_3}\,\frac{d \varrho_3}{dy}, \frac{1}{\Delta_3}\, \frac{d \varrho_3}{dz}, \,\,\,\,\,\,\,\,\,\Delta_3^2= \Big(\frac{d\varrho_3}{dx}\Big)^2 + \Big(\frac{d\varrho_3}{dy}\Big)^2 +\Big(\frac{d\varrho_3}{dz}\Big)^2 \end{equation*} \indent\indent Suppose now that $\varrho_1, \, \varrho_2, \, \varrho_3$ form an orthogonal system; then, in the point of intersection the normals to the three surfaces $\varrho_1, \, \varrho_2, \, \varrho_3$ are mutually orthogonal to each other. The conditions that the axes of $x,\,y,\,z$ form an orthogonal system, when referring their positions to the normals of the surfaces $\varrho_1, \varrho_2,\,\varrho_3$, will be the following\,: \begin{mynewequation} \tag*{[5.31]} \left. \begin{aligned} \frac{1}{\Delta_1^2} \Big(\frac{d \varrho_1}{dx}\Big)^2 + \frac{1}{\Delta_2^2} \Big(\frac{d \varrho_2}{dx}\Big)^2 + \frac{1}{\Delta_3^2} \Big(\frac{d \varrho_3}{dx}\Big)^2 =1\\\\ \frac{1}{\Delta_1^2} \Big(\frac{d \varrho_1}{dy}\Big)^2 + \frac{1}{\Delta_2^2} \Big(\frac{d \varrho_2}{dy}\Big)^2 + \frac{1}{\Delta_3^2} \Big(\frac{d \varrho_3}{dy}\Big)^2 =1\\\\ \frac{1}{\Delta_1^2} \Big(\frac{d \varrho_1}{dz}\Big)^2 + \frac{1}{\Delta_2^2} \Big(\frac{d \varrho_2}{dz}\Big)^2 + \frac{1}{\Delta_3^2} \Big(\frac{d \varrho_3}{dz}\Big)^2 =1 \end{aligned} \qquad\right\} \end{mynewequation} \begin{mynewequation} \tag*{[5.32]} \left. \begin{aligned} \frac{1}{\Delta_1^2} \frac{d \varrho_1}{dx} \frac{d \varrho_1}{dy} + \frac{1}{\Delta_2^2} \frac{d \varrho_2}{dx} \frac{d \varrho_2}{dy}+ \frac{1}{\Delta_3^2} \frac{d \varrho_3}{dx} \frac{d \varrho_3}{dy} =0\\\\ \frac{1}{\Delta_1^2} \frac{d \varrho_1}{dy} \frac{d \varrho_1}{dz} + \frac{1}{\Delta_2^2} \frac{d \varrho_2}{dy} \frac{d \varrho_2}{dz}+ \frac{1}{\Delta_3^2} \frac{d \varrho_3}{dy} \frac{d \varrho_3}{dz} =0\\\\ \frac{1}{\Delta_1^2} \frac{d \varrho_1}{dz} \frac{d \varrho_1}{dx} + \frac{1}{\Delta_2^2} \frac{d \varrho_2}{dz} \frac{d \varrho_2}{dx}+ \frac{1}{\Delta_3^2} \frac{d \varrho_3}{dz} \frac{d \varrho_3}{dx} =0 \end{aligned} \qquad \right\} \end{mynewequation} \indent \indent Now we have: \begin{mynewequation} \tag*{[5.33]} \begin{aligned} \frac{\rm {d} \varrho_1}{\Delta_1} =\frac{1}{\Delta_1} \frac{d \varrho_1}{dx} {\rm {d}x} + \frac{1}{\Delta_1} \frac{d \varrho_1}{dy} {\rm {d}y} + \frac{1}{\Delta_1} \frac{d \varrho_1}{dz} {\rm {d}z} \\\\ \frac{\rm {d} \varrho_2}{\Delta_2} =\frac{1}{\Delta_2} \frac{d \varrho_2}{dx} {\rm {d}x} + \frac{1}{\Delta_2} \frac{d \varrho_2}{dy} {\rm {d}y} + \frac{1}{\Delta_2} \frac{d \varrho_2}{dz} {\rm {d}z} \\\\ \frac{\rm {d} \varrho_3}{\Delta_3} =\frac{1}{\Delta_3} \frac{d \varrho_3}{dx} {\rm {d}x} + \frac{1}{\Delta_3} \frac{d \varrho_3}{dy} {\rm {d}y} + \frac{1}{\Delta_2} \frac{d \varrho_3}{dz}{\rm {d}z} \end{aligned} \end{mynewequation} \indent\indent By squaring and adding these equations, using the given relations, one obtains\,: \begin{mynewequation} \tag*{[5.34]} \Big(\frac{\rm {d} \varrho_1}{\Delta_1}\Big)^2 + \Big(\frac{\rm {d} \varrho_2}{\Delta_2}\Big)^2 + \Big(\frac{\rm {d} \varrho_3}{\Delta_3}\Big)^2 \text{\small $={\rm d}x^2+{\rm d}y^2+{\rm d}z^2= {\rm d}s^2$} \end{mynewequation} The comparison of this expression with the general one yields\,: \begin{mynewequation} \tag*{[5.35]} N_1 = \frac{1}{A_1^2},\,\,\,N_2 = \frac{1}{A_2^2},\,\,\,N_3 = \frac{1}{A_3^2},\,\,\,n_1=n_2=n_3=0 \end{mynewequation} If one sets\,: \begin{mynewequation} \tag*{[5.36]} \begin{aligned} N_1=\Big(\frac{dx}{d\varrho_1}\Big)^2 + \Big(\frac{dy}{d\varrho_1}\Big)^2 + \Big(\frac{dz}{d\varrho_1}\Big)^2= \frac{1}{(\frac{d \varrho_1}{dx})^2 + (\frac{d \varrho_1}{dy})^2 + (\frac{d \varrho_1}{dz})^2 }\\ N_2=\Big(\frac{dx}{d\varrho_2}\Big)^2 + \Big(\frac{dy}{d\varrho_2}\Big)^2 + \Big(\frac{dz}{d\varrho_3}\Big)^2= \frac{1}{(\frac{d \varrho_2}{dx})^2 + (\frac{d \varrho_2}{dy})^2 + (\frac{d \varrho_2}{dz})^2 }\\ N_3=\Big(\frac{dx}{d\varrho_3}\Big)^2 + \Big(\frac{dy}{d\varrho_3}\Big)^2 + \Big(\frac{dz}{d\varrho_3}\Big)^2= \frac{1}{(\frac{d \varrho_3}{dx})^2 + (\frac{d \varrho_3}{dy})^2 + (\frac{d \varrho_3}{dz})^2 } \end{aligned} \end{mynewequation} so the equation for the density becomes\,: \begin{myequation} \tag*{(2), [5.37]} \begin{vmatrix} \frac{d\varrho_1}{d\varrho_1^0}\,\,\,\, &\frac{d\varrho_1}{d\varrho_2^0}\,\,\,\, &\frac{d\varrho_1}{d\varrho_3^0} \\\\ \frac{d\varrho_2}{d\varrho_1^0}\,\,\,\, &\frac{d\varrho_2}{d\varrho_2^0}\,\,\,\, &\frac{d\varrho_2}{d\varrho_3^0} \\\\ \frac{d\varrho_3}{d\varrho_1^0}\,\,\,\, &\frac{d\varrho_3}{d\varrho_2^0}\,\,\,\, &\frac{d\varrho_3}{d\varrho_3^0} \end{vmatrix} \cdot \frac{\varrho}{\varrho_0} =\sqrt {\frac{ N_1^0\, N_2^0\, N_3^0}{N_1\, N_2\, N_3}} \end{myequation} and the equation [``expression'' is here meant] to be varied is\,: \begin{mynewequation} \notag \int {\rm d} t \int {\rm d} T \left \{ N_1 \left(\frac{d\varrho_1}{dt} \right)^2 + N_2 \left(\frac{d \varrho_2}{dt} \right)^2 + N_3 \left(\frac{d \varrho_3}{dt} \right)^2 + 2 \Omega \right \} \end{mynewequation} where ${\rm d}T$ indicates the new element of mass $\varrho_0\,{\rm d}a\,{\rm d}b\,{\rm d}c$ written in the new coordinates.\hspace{0.4cm}The part dependent on $\delta\hspace{-0.03cm}\varrho_1$ of the variation of this integral is\,: \begin{smallequation} \notag \int {\rm d} t \int {\rm d} T \left \{ 2 N_1 \frac{d \varrho_1}{dt} \frac{d \delta\hspace{-0.03cm}\varrho_1}{dt} + \left( \frac{d \varrho_1}{dt} \right)^2 \frac{d N_1}{d \varrho_1} \delta\hspace{-0.03cm}\varrho_1 + \left( \frac{d \varrho_2}{dt} \right)^2 \frac{d N_2}{d \varrho_1} \delta\hspace{-0.03cm}\varrho_1 + \left( \frac{d \varrho_3}{dt} \right)^2 \frac{d N_3}{d \varrho_1} \delta\hspace{-0.03cm}\varrho_1 + 2 \frac{d \Omega}{d \varrho_1} \delta\hspace{-0.03cm}\varrho_1 \right \} \,. \end{smallequation} When the first member of this expression is integrated by parts in $t$, all members have the factor $\delta\hspace{-0.03cm}\varrho_1$. After it is set to zero, one obtains\,: \begin{myequation} \tag*{[5.38]} {\text{\footnotesize $2$}} \frac{d \Omega}{d\varrho_1} = \text{\footnotesize $2$} \frac{d \left(N_1 \frac{d \varrho_1}{dt}\right)} {dt} -\left( \frac{d \varrho_1}{dt} \right)^2 \frac{dN_1}{d \varrho_1} -\left( \frac{d \varrho_2}{dt} \right)^2 \frac{dN_2}{d \varrho_1} -\left( \frac{d \varrho_3}{dt} \right)^2 \frac{dN_3}{d \varrho_1} \,, \end{myequation} similarly, one has\,: \begin{equation*} \tag*{(3), [5.39]} 2\frac{d \Omega}{d\varrho_2} = 2 \frac{d \left(N_2 \frac{d \varrho_2}{dt}\right)} {dt} -\left( \frac{d \varrho_1}{dt} \right)^2 \frac{dN_1}{d \varrho_2} -\left( \frac{d \varrho_2}{dt} \right)^2 \frac{dN_2}{d \varrho_2} -\left( \frac{d \varrho_3}{dt} \right)^2 \frac{dN_3}{d \varrho_2} \end{equation*} \begin{equation*} \tag*{[5.40]} 2\frac{d \Omega}{d\varrho_3} = 2 \frac{d \left(N_3 \frac{d \varrho_3}{dt}\right)} {dt} -\left( \frac{d \varrho_1}{dt} \right)^2 \frac{dN_1}{d \varrho_3} -\left( \frac{d \varrho_2}{dt} \right)^2 \frac{dN_2}{d \varrho_3} -\left( \frac{d \varrho_3}{dt} \right)^2 \frac{dN_3}{d \varrho_3} \end{equation*} These are equations which are built analogously to (3) of \S\,4\,; in order for these equations to take the second \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! rian} form, the so-called \mbox{L\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! g\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! g\hspace{0.036cm}} %{$\phantom{!}$\! ian} form, we multiply in turn the previous equations by\,: $$ \frac{d \varrho_1}{d \varrho_1^0}, \,\,\,\,\,\, \frac{d \varrho_2}{d \varrho_1^0}, \,\,\,\,\,\, \frac{d \varrho_3}{d \varrho_1^0}, $$ and we add them; then we multiply by $$ \frac{d \varrho_1}{d \varrho_2^0}, \,\,\,\,\,\, \frac{d \varrho_2}{d \varrho_2^0}, \,\,\,\,\,\, \frac{d \varrho_3}{d \varrho_2^0}, $$ and we add them as well; finally, we multiply by $$ \frac{d \varrho_1}{d \varrho_3^0}, \,\,\,\,\,\, \frac{d \varrho_2}{d \varrho_3^0}, \,\,\,\,\,\, \frac{d \varrho_3}{d \varrho_3^0}, $$ and we add them too.\hspace{0.4cm}In this way, we get the following equations\,: \begin{mynewequation} \tag* {(4), [5.41]} \left. \begin{aligned} 2 \frac{d \Omega}{d {\varrho_1}^0} = 2 \frac{d\left(N_1 \frac{d\varrho_1}{dt} \right)}{dt} \frac{d \varrho_1}{d \varrho_1^0} + 2 \frac{d\left(N_2 \frac{d\varrho_2}{dt} \right)}{dt} \frac{d \varrho_2}{d \varrho_1^0} + 2 \frac{d\left(N_3 \frac{d\varrho_3}{dt} \right)}{dt} \frac{d \varrho_3}{d \varrho_1^0} \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, -\left( \frac{d \varrho_1}{dt} \right)^2 \frac{dN_1}{d \varrho_1^0} -\left( \frac{d \varrho_2}{dt} \right)^2 \frac{dN_2}{d \varrho_1^0} -\left( \frac{d \varrho_3}{dt} \right)^2 \frac{dN_3}{d \varrho_1^0} \\\\ 2 \frac{d \Omega}{d {\varrho_2}^0} = 2 \frac{d\left(N_1 \frac{d\varrho_1}{dt} \right)}{dt} \frac{d \varrho_1}{d \varrho_2^0} + 2 \frac{d\left(N_2 \frac{d\varrho_2}{dt} \right)}{dt} \frac{d \varrho_2}{d \varrho_2^0} + 2 \frac{d\left(N_3 \frac{d\varrho_3}{dt} \right)}{dt} \frac{d \varrho_3}{d \varrho_2^0} \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, -\left( \frac{d \varrho_1}{dt} \right)^2 \frac{dN_1}{d \varrho_2^0} -\left( \frac{d \varrho_2}{dt} \right)^2 \frac{dN_2}{d \varrho_2^0} -\left( \frac{d \varrho_3}{dt} \right)^2 \frac{dN_3}{d \varrho_2^0} \\\\ 2 \frac{d \Omega}{d {\varrho_3}^0} = 2 \frac{d\left(N_1 \frac{d\varrho_1}{dt} \right)}{dt} \frac{d \varrho_1}{d \varrho_3^0} + 2 \frac{d\left(N_2 \frac{d\varrho_2}{dt} \right)}{dt} \frac{d \varrho_2}{d \varrho_3^0} + 2 \frac{d\left(N_3 \frac{d\varrho_3}{dt} \right)}{dt} \frac{d \varrho_3}{d \varrho_3^0} \\ \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, -\left( \frac{d \varrho_1}{dt} \right)^2 \frac{dN_1}{d \varrho_3^0} -\left( \frac{d \varrho_2}{dt} \right)^2 \frac{dN_2}{d \varrho_3^0} -\left( \frac{d \varrho_3}{dt} \right)^2 \frac{dN_3}{d \varrho_3^0} \end{aligned} \qquad \right\} \end{mynewequation} \indent\indent A very elegant example of an orthogonal system are the elliptical coordinates $\varrho_1, \varrho_2,\varrho_3$ which can be defined as the roots of the equation with respect to $\varepsilon$\,: \begin{equation*} \tag*{[5.42]} \frac{x^2}{\alpha ^2 - \varepsilon^2} + \frac{y^2}{\beta^2 - \varepsilon^2} + \frac{z^2}{\gamma^2 - \varepsilon^2} =1 \end{equation*} and so taken that one has\,: \begin{equation*} \tag*{[5.43]} \alpha > \varrho_1 > \beta > \varrho_2>\gamma >\varrho_3 >0 \end{equation*} \noindent From the identity obtained after partial fraction decomposition\,: \begin{align} \frac {\left( \varepsilon^2 - \varrho_1^2 \right) \left( \varepsilon^2 - \varrho_2^2 \right) \left( \varepsilon^2 - \varrho_3^2 \right)} {\left( \varepsilon^2 - \alpha^2 \right) \left( \varepsilon^2 - \beta^2 \right) \left( \varepsilon^2 - \gamma^2\right)} = 1 - \frac{\left(\alpha^2 - \varrho_1^2 \right) \left(\alpha^2 - \varrho_2^2 \right) \left(\alpha^2 - \varrho_3^2 \right) }{ \left (\alpha^2 - \beta^2 \right)\left(\alpha^2 - \gamma^2 \right)} \cdot \frac{1}{\alpha^2 - \varepsilon^2} \hspace{2cm} \nonumber \\ - \frac{\left(\beta^2 - \varrho_1^2 \right) \left(\beta^2 - \varrho_2^2 \right) \left(\beta^2 - \varrho_3^2 \right) }{ \left (\beta^2 - \alpha^2 \right)\left(\beta^2 - \gamma^2 \right)} \frac{1}{\beta^2 - \varepsilon^2} - \frac{\left(\gamma^2 - \varrho_1^2 \right) \left(\gamma^2 - \varrho_2^2 \right) \left(\gamma^2 - \varrho_3^2 \right) }{ \left (\gamma^2 - \beta^2 \right)\left(\gamma^2 - \alpha^2 \right)} \frac{1}{\gamma^2 - \varepsilon^2} \,. \tag*{[5.44]} \end{align} When $\varepsilon$ is put equal to one of the roots $\varrho_1,\varrho_2,\varrho_3$ of the equation\,: \begin{equation*} \tag*{[5.45]} \frac{x^2}{\alpha^2 - \varepsilon^2} + \frac{y^2}{\beta^2 - \varepsilon^2} + \frac{z^2}{\gamma^2 -\varepsilon^2}=1 \end{equation*} one obtains\,: \begin{align} \frac{\left(\alpha^2 - \varrho_1^2 \right) \left(\alpha^2 - \varrho_2^2 \right) \left(\alpha^2 - \varrho_3^2 \right) }{ \left (\alpha^2 - \beta^2 \right)\left(\alpha^2 - \gamma^2 \right)} \frac{1}{\alpha^2 - \varepsilon^2} + \frac{\left(\beta^2 - \varrho_1^2 \right) \left(\beta^2 - \varrho_2^2 \right) \left(\beta^2 - \varrho_3^2 \right) }{ \left (\beta^2 - \alpha^2 \right)\left(\beta^2 - \gamma^2 \right)} \frac{1}{\beta^2 - \varepsilon^2} \nonumber \\ + \frac{\left(\gamma^2 - \varrho_1^2 \right) \left(\gamma^2 - \varrho_2^2 \right) \left(\gamma^2 - \varrho_3^2 \right) }{ \left (\gamma^2 - \beta^2 \right)\left(\gamma^2 - \alpha^2 \right)} \frac{1}{\gamma^2 - \varepsilon^2} =1 \tag*{[5.46]} \end{align} which compared with the previous equations gives the relations\,: \begin{align} \tag*{[5.47]} \begin{aligned} x^2 = \frac {\left( \alpha^2 - \varrho_1^2 \right) \left( \alpha^2 - \varrho_2^2 \right) \left( \alpha^2 - \varrho_3^2 \right)} {\left( \alpha^2 - \beta^2 \right) \left( \alpha^2 - \gamma^2 \right)} \\ \\ y^2 = \frac {\left( \beta^2 - \varrho_1^2 \right) \left( \beta^2 - \varrho_2^2 \right) \left( \beta^2 - \varrho_3^2 \right)} {\left( \beta^2 - \alpha^2 \right) \left( \beta^2 - \gamma^2 \right)} \\ \\ z^2 = \frac {\left( \gamma^2 - \varrho_1^2 \right) \left( \gamma^2 - \varrho_2^2 \right) \left( \gamma^2 - \varrho_3^2 \right)} {\left( \gamma^2 - \alpha^2 \right) \left( \gamma^2 - \beta^2 \right)} \end{aligned} \end{align} so that, if $\varepsilon$ is an arbitrary variable, one has\,: \begin{equation*} \tag*{[5.48]} \frac {\left( \varepsilon^2 - \varrho_1^2 \right) \left( \varepsilon^2 - \varrho_2^2 \right) \left( \varepsilon^2 - \varrho_3^2 \right)} {\left( \varepsilon^2 - \alpha^2 \right) \left( \varepsilon^2 - \beta^2 \right) \left( \varepsilon^2 - \gamma^2\right)} = 1 - \frac{x^2}{\alpha^2 - \varepsilon^2} - \frac{y^2}{\beta^2 - \varepsilon^2} - \frac{z^2}{\gamma^2 -\varepsilon^2} \end{equation*} When this equation is differentiated with respect to $\varepsilon$ and after setting $\varepsilon=\varrho_1, \varrho_2, \varrho_3$, one finds\,: \begin{mynewequation} \tag*{[5.49]} \begin{aligned} - \frac{\left(\varrho_1^2 - \varrho_2^2 \right) \left(\varrho_1^2 - \varrho_3^2 \right)}{ \left (\varrho_1^2 - \alpha^2 \right)\left(\varrho_1^2 - \beta^2 \right)\left(\varrho_1^2 - \gamma^2 \right)} =\left( \frac{x}{\alpha^2 - \varrho_1^2}\right)^2 + \left( \frac{y}{\beta^2 - \varrho_1^2}\right)^2 + \left( \frac{z}{\gamma^2 - \varrho_1^2}\right)^2 \\ - \frac{\left(\varrho_2^2 - \varrho_1^2 \right) \left(\varrho_2^2 - \varrho_3^2 \right)}{ \left (\varrho_2^2 - \alpha^2 \right)\left(\varrho_2^2 - \beta^2 \right)\left(\varrho_2^2 - \gamma^2 \right)} =\left( \frac{x}{\alpha^2 - \varrho_2^2}\right)^2 + \left( \frac{y}{\beta^2 - \varrho_2^2}\right)^2 + \left( \frac{z}{\gamma^2 - \varrho_2^2}\right)^2 \\ - \frac{\left(\varrho_3^2 - \varrho_1^2 \right) \left(\varrho_3^2 - \varrho_2^2 \right)}{ \left (\varrho_3^2 - \alpha^2 \right)\left(\varrho_3^2 - \beta^2 \right)\left(\varrho_3^2 - \gamma^2 \right)} =\left( \frac{x}{\alpha^2 - \varrho_3^2}\right)^2 + \left( \frac{y}{\beta^2 - \varrho_3^2}\right)^2 + \left( \frac{z}{\gamma^2 - \varrho_3^2}\right)^2 \\ \end{aligned} \end{mynewequation} Logarithmical differentiation of the equations by which $x^2,y^2, z^2$ are represented as functions of $\varrho_1^2,\varrho_2^2, \varrho_3^2$ gives\,: \begin{mynewequation} \tag*{[5.50]} \begin{aligned} - \frac{dx}{d \varrho_1} = \frac{x \varrho_1}{\alpha^2 - \varrho_1^2}, \,\,\, - \frac{dx}{d \varrho_2} = \frac{x \varrho_2}{\alpha^2 - \varrho_2^2}, \,\,\, -\frac{dx}{d \varrho_3} = \frac{x \varrho_3}{\alpha^2 - \varrho_3^2}\\ - \frac{dy}{d \varrho_1} = \frac{y \varrho_1}{\beta^2 - \varrho_1^2}, \,\,\, - \frac{dy}{d \varrho_2} = \frac{y \varrho_2}{\beta^2 - \varrho_2^2}, \,\,\, -\frac{dy}{d \varrho_3} = \frac{y \varrho_3}{\beta^2 - \varrho_3^2}\\ \frac{dz}{d \varrho_1} = \frac{z \varrho_1}{\gamma^2 - \varrho_1^2}, \,\,\, - \frac{dz}{d \varrho_2} = \frac{z \varrho_2}{\gamma^2 - \varrho_2^2}, \,\,\, -\frac{dz}{d \varrho_3} = \frac{z \varrho_3}{\gamma^2 - \varrho_3^2} \end{aligned} \end{mynewequation} so that one has\,: \begin{mynewequation} \tag*{[5.51]} \begin{aligned} N_1 = \varrho_1^2 \left\{ \left( \frac{x}{\alpha^2 - \varrho_1^2}\right)^2 + \left( \frac{y}{\beta^2 - \varrho_1^2}\right)^2 + \left( \frac{z}{\gamma^2 - \varrho_1^2}\right)^2 \right \}\\ N_2 = \varrho_2^2 \left \{ \left( \frac{x}{\alpha^2 - \varrho_2^2}\right)^2 + \left( \frac{y}{\beta^2 - \varrho_2^2}\right)^2 + \left( \frac{z}{\gamma^2 - \varrho_2^2}\right)^2 \right \}\\ N_3 = \varrho_3^2 \left \{ \left( \frac{x}{\alpha^2 - \varrho_3^2}\right)^2 + \left( \frac{y}{\beta^2 - \varrho_3^2}\right)^2 + \left( \frac{z}{\gamma^2 - \varrho_3^2}\right)^2 \right \} \end{aligned} \end{mynewequation} From these relations, one can therefore write\,: \begin{mynewequation} \tag*{[5.52]} \begin{aligned} N_1 = -\varrho_1^2 \frac{\left(\varrho_1^2 - \varrho_2^2 \right) \left(\varrho_1^2 - \varrho_3^2 \right)}{ \left (\varrho_1^2 - \alpha^2 \right)\left(\varrho_1^2 - \beta^2 \right)\left(\varrho_1^2 - \gamma^2 \right)}\\\\ N_2 = -\varrho_2^2 \frac{\left(\varrho_2^2 - \varrho_1^2 \right) \left(\varrho_2^2 - \varrho_3^2 \right)}{ \left (\varrho_2^2 - \alpha^2 \right)\left(\varrho_2^2 - \beta^2 \right)\left(\varrho_2^2 - \gamma^2 \right)}\\\\ N_3 = - \varrho_3^2 \frac{\left(\varrho_3^2 - \varrho_1^2 \right) \left(\varrho_3^2 - \varrho_2^2 \right)}{ \left (\varrho_3^2 - \alpha^2 \right)\left(\varrho_3^2 - \beta^2 \right)\left(\varrho_3^2 - \gamma^2 \right)} \end{aligned} \end{mynewequation} \indent\indent These three quantities are recognised as positive because of \begin{mynewequation} \tag*{[5.53]} \alpha > \varrho_1 > \beta > \varrho_2>\gamma >\varrho_3 >0 \end{mynewequation} have only to be substituted into equations~(4) to obtain the hydrodynamical fundamental equations for elliptical coordinates.\\ \indent\indent The polar coordinate system $r, \theta, \varphi$, which is determined by\,: \begin{mynewequation} \tag*{[5.54]} x = r \cos \theta, \, \, y = r \sin \theta \cos \varphi, \, \, z= r \sin \theta \sin \varphi \end{mynewequation} is orthogonal\,;\hspace{0.4cm}in fact one has\,: \begin{mynewequation} \tag*{[5.55]} {\rm d}s^2= {\rm d}x^2 + {\rm d} y^2 + {\rm d} z^2 = {\rm d} r^2 + r^2 {\rm d} \theta^2 + r^2 \sin^2 \theta {\rm d} \varphi \end{mynewequation} so that one has \begin{mynewequation} \tag*{[5.56]} N_1 = 1 , \,\,\,N_2 = r^2,\,\, N_3 = r^2 \sin^2 \theta . \end{mynewequation} \noindent The density equation (2) becomes\,: \begin{myequation} \tag*{[5.57]} \frac{\varrho}{\varrho_0} \begin{vmatrix} \frac{dr}{dr_0}\,\,\,\,&\frac{d\theta}{dr_0}\,\,\,\,&\frac{d\varphi}{d r_0} \\\\ \frac{dr}{d\theta_0}\,\,\,\,&\frac{d \theta}{d\theta_0}\,\,\,\,&\frac{d \varphi}{d\theta_0} \\\\ \frac{dr}{d\varphi_0}\,\,\,\,&\frac{d \theta}{d\varphi_0}\,\,\,\,&\frac{d \varphi}{d\varphi_0} \end{vmatrix} = \frac{r_0^2 \sin \theta_0}{r^2 \sin \theta} \end{myequation} In this case, by setting \begin{align} \tag*{[5.58]} \begin{aligned} \Phi_1&= \frac{d^2 r}{dt^2} - r \left( \frac{d \theta}{dt}\right)^2 - r \sin ^2 \theta \left(\frac{d \varphi}{dt} \right )^2 \\ \Phi_2&= \frac{d \left( r^2 \frac{ d\theta}{dt}\right)}{dt} - \left( \frac{d \varphi}{dt} \right )^2 r^2 \sin \theta \cos \theta \, \, \, \,\, \\ \Phi_3 &= \frac{d \left(r^2 \sin^2 \theta \frac{d \varphi}{dt}\right)}{dt} \end{aligned} \end{align} equations~(3) become very simply\,: \begin{equation*} \tag*{[5.59]} \frac{d \Omega}{dr} = \Phi_1, \,\,\, \frac{d \Omega}{d \theta} = \Phi_2, \,\,\,\frac{d \Omega}{d \varphi} = \Phi_3, \,\,\, \end{equation*} and equations~(4) become\,: \begin{mynewequation} \tag*{[5.60]} \begin{aligned} \Phi_1 \frac{dr}{dr_0} + \Phi_2 \frac{d \theta}{d r_0} + \Phi_3 \frac{d \varphi}{dr_0} - \frac{d \Omega}{d r_0}\,\,\,=0\\ \Phi_1\,\, \frac{dr}{d\theta_0} + \Phi_2 \frac{d \theta}{d\theta_0} + \Phi_3 \frac{d \varphi}{d\theta_0} - \frac{d \Omega}{d \theta_0}\hspace{0.09cm}=0\\ \Phi_1 \frac{dr}{d \varphi_0} + \Phi_2 \frac{d \theta_0}{d \varphi_0} + \Phi_3 \frac{d \varphi}{d\varphi_0} - \frac{d \Omega}{d \varphi_0}=0\\ \end{aligned} \end{mynewequation} \indent\indent The same transformation can be directly performed in the following way.\hspace{0,4cm} One observes that\,: \begin{smallequation} \tag*{[5.61]} \begin{aligned} \!\!\frac{d^2 x}{dt^2} &= \cos \theta \frac{d^2 r }{d t^2} - r \sin \theta \frac{d^2 \theta}{d t^2} - r \cos \theta \left(\frac{d \theta}{dt} \right )^2 - 2 \sin \theta \frac{dr}{dt } \frac{d \theta}{dt}\,\,\hspace{5.6cm} \\ \\ \!\!\frac{d^2 y}{dt^2} &= \sin \theta \cos \varphi \frac{d^2 r}{d t^2} + r \cos \theta \cos \varphi \frac{d^2 \theta}{d t^2} - r \sin \theta \sin \varphi \frac{d^2 \varphi}{d t^2} - r \sin \theta \sin \varphi \frac{d^2 \varphi }{d t^2} - r \sin \theta \cos \varphi \left(\frac{d \theta}{dt} \right )^2 \\ &-r \sin \theta \cos \varphi \left(\frac{d \varphi}{dt} \right )^2 + 2 \cos \theta \cos \varphi \frac{dr}{dt } \frac{d \theta}{dt} - 2 \sin \theta \sin \varphi \frac{dr}{dt } \frac{d \varphi}{dt} - 2 r \cos \theta \sin \varphi \frac{d \theta}{dt } \frac{d \varphi}{dt}\hspace{1.5cm} \\ \\ \!\!\frac{d^2 z}{dt^2} &=\sin \theta \sin \varphi \frac{d^2 r }{d t^2} + r \cos \theta \sin \varphi \frac{d^2 \theta }{d t^2} + r \sin \theta \cos \varphi \frac{d^2 \varphi}{d t^2} - r \sin \theta \sin \varphi \left(\frac{d \theta}{dt} \right )^2\,\hspace{2.7cm} \\ &- r \sin \theta \sin \varphi \left(\frac{d \varphi}{dt} \right )^2 + 2 \cos \theta \sin \varphi \frac{dr}{dt } \frac{d \theta}{dt} + 2 \sin \theta \cos \varphi \frac{dr}{dt } \frac{d \varphi}{dt} + 2 r \cos \theta \cos \varphi \frac{d \theta}{dt } \frac{d \varphi}{dt} \hspace{1.3cm} \end{aligned} \end{smallequation} Furthermore one has\,: \begin{align} \tag*{[5.62]} \begin{aligned} \frac{dx}{da} &= \cos \theta \frac{dr}{da} - r \sin \theta \frac{d \theta}{da}\hspace{4.2cm} \\ \\ \frac{dy}{da} &= \sin \theta \cos \varphi \frac{dr}{da } + r \cos \theta \cos \varphi \frac{d \theta}{da } - r \sin \theta \sin \varphi \frac{d \varphi}{da } \\ \\ \frac{dz}{da} &= \sin \theta \sin \varphi \frac{dr}{da } + r \cos \theta \sin \varphi \frac{d \theta}{da } + r \sin \theta \sin \varphi \frac{d \varphi}{da }\,; \end{aligned} \end{align} then the equation\,: \begin{equation*} \tag*{[5.63]} \frac{d^2 x}{dt^2} \frac{dx}{da} + \frac{d^2 y}{dt^2} \frac{dy}{da} + \frac{d^2 z}{dt^2} \frac{dz}{da} - \frac{d \Omega}{da}=0 \end{equation*} becomes\,: \begin{equation*} \tag*{[5.64]} \Phi_1 \frac{dr}{da} + \Phi_2 \frac{d \theta}{da} + \Phi_3 \frac{d \varphi}{da} - \frac{d \Omega}{da}=0 \end{equation*} where $\Phi_1,\Phi_2,\Phi_3$ have the meaning as written above.\hspace{0.4cm}In addition to this equation, there are two others\,: \begin{equation*} \tag*{[5.65]} \begin{aligned} \Phi_1\frac{dr}{db} + \Phi_2 \frac{d \theta}{db} + \Phi_3 \frac{d \varphi}{db} - \frac{d \Omega}{db}=0\\\\ \Phi_1\frac{dr}{dc} + \Phi_2 \frac{d \theta}{dc} + \Phi_3 \frac{d \varphi}{dc} - \frac{d \Omega}{dc}=0 \end{aligned} \end{equation*} where $a, b, c$ depend on $r_0, \theta_0, \varphi_0$ through the relations\,: \begin{equation*} \tag*{[5.66]} a=r_0 \cos \theta_0,\,\, b= r_0 \sin \theta_0\cos \varphi_0,\,\, c=r_0 \sin \theta_0 \sin \varphi_0 \end{equation*} The change of variables from $a,\,b,\,c$ to $r_0, \,\theta_0,\,\varphi_0$ into the hydrodynamical equations can be easily carried out by multiplying with the appropriate factors and adding the equations. Then, one arrives at the above formulae in a different way. \\ \indent\indent The transformation of the fundamental equations into cylindrical coordinates is extremely simple.\hspace{0.4cm}Namely, if one sets\,: \begin{equation*} \tag*{[5.67]} x = r \cos \theta, \,\,y = r \sin \theta, \,\,z=z \end{equation*} then one has\,: \begin{equation*} \tag*{[5.68]} {\rm d}s^2= {\rm d}x^2 + {\rm d}y^2 + {\rm d}z^2= {\rm d}r^2 + r^2 {\rm d} \theta^2 + {\rm d}z^2 \end{equation*} so that \begin{equation*} \tag*{[5.69]} N_1 = 1,\,\, N_2= r^2,\,\, N_3=1 \end{equation*} hence\,: \begin{equation*} \tag*{[5.70]} \begin{aligned} \frac{d^2 r}{dt^2} - r \left(\frac{d \theta}{dt} \right)^2 &= \frac{d \Omega}{dr}\\ \frac{d \left( r^2 \frac{d \theta}{dt} \right) }{dt} &= \frac{d \Omega}{d \theta}\\ \frac{d^2 z}{dt^2} &= \frac{d \Omega}{dz} \end{aligned} \end{equation*} \indent\indent In case $r,\, \theta,\, z$ need to be expressed as functions of the initial values $r_0, \,\, \theta_0, \,\, z_0$, one obtains the equations\,: \begin{equation*} \tag*{[5.71]} \begin{aligned} \left( \frac {d^2 r}{dt^2} - r \left( \frac{d \theta}{dt} \right)^2 \right) \frac{dr}{dr_0} + \frac{d \left(r^2 \frac{d \theta}{dt}\right)}{dt} \frac{d \theta}{d r_0} + \frac{d^2 z}{dt^2} \frac{dz}{d r_0} &= \frac{d \Omega}{dr_0}\\\\ \left( \frac {d^2 r}{dt^2} - r \left( \frac{d \theta}{dt} \right)^2 \right) \frac{dr}{d \theta_0} + \frac{d \left(r^2 \frac{d \theta}{dt}\right)}{dt} \frac{d \theta}{d \theta_0} + \frac{d^2 z}{dt^2} \frac{dz}{d \theta_0} &= \frac{d \Omega}{d \theta_0} \\\\ \left( \frac {d^2 r}{dt^2} - r \left( \frac{d \theta}{dt} \right)^2 \right) \frac{dr}{dz_0} + \frac{d \left(r^2 \frac{d \theta}{dt}\right)}{dt} \frac{d \theta}{d z_0} + \frac{d^2 z}{dt^2} \frac{dz}{d z_0} &= \frac{d \Omega}{dz_0} \end{aligned} \end{equation*} The density equation becomes\,: \begin{mynewequation}\tag*{[5.72]} \left| \begin{aligned} \frac{dr}{dr_0}\,\,\,\,\,\,&\frac{dr}{d\theta_0}&\frac{dr}{d\varphi_0} \\\\ \frac{d\theta}{dr_0}\,\,\,\,\,&\frac{d\theta}{d\theta_0}&\frac{d\theta}{d\varphi_0} \\\\ \frac{d\varphi}{dr_0}\,\,\,\,\,&\frac{d\varphi}{d\theta_0}&\frac{d\varphi}{d\varphi_0} \end{aligned} \right| = \text{\Large $\frac{r_0}{r}$} \end{mynewequation} If we take the initial conditions and the accelerating forces to be symmetric with respect to the $z$ axis, these equations become advantageous\,; then we have $\frac{d \Omega}{d \theta} = 0$, therefore $\frac{d \left( r^2 \frac{d \theta}{dt}\right)}{dt}=0,$ and thus\,: \begin{equation*} \tag*{[5.73]} \frac{d \theta}{dt}=\frac{H}{r^2}, \end{equation*} where $H$ is a time-independent constant which has always the same value for a certain particle, but varies from particle to particle and has to be determined by the initial conditions.\hspace{0.4cm}Since $\frac{d \theta}{dt}$ is the rotational velocity of a particle around the $z$ axis, the rotational velocity of one and the same particle around the symmetry axis is inversely proportional to the relative distance squared of the particle to the axis.\hspace{0.4cm}We see from this that no particle initially rotating ceases to rotate under the influence of forces generated by a potential and, conversely, no particle begins to rotate if it is not initially in rotation. \deffootnotemark{\textsuperscript{[A.\thefootnotemark]}} \deffootnote{2em}{1.6em}{[A.\thefootnotemark]\enskip} \hspace{0.4cm}This elegant theorem is thanks to \mbox{S\hspace{0.036cm}} %{$\phantom{!}$\! v\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! b\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! g,} {who obtained it in cylindrical coordinates from the first \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! ian} equations which, actually, are for this purpose slightly less convenient than the second \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! ian} equations \footnote{On fluides r\"orelse. Kongl. Vetenskaps-Academiens Handlingar f\"or {\aa}r 1839, S.\,139. Stockholm. Also cf: Sur le mouvement des fluides.\,\, Crelle's Journal Bd. 24, S.\,157, [\hyperlink{Svanberg}{1842}].} \vspace{0.6cm} \centerline{\fett{$\S.\,6.$}} \vspace{0.3cm} \indent\indent {Using the $\Omega$ function introduced above (cf. S.\,18), the hydrodynamical fundamental equations can be written as\,: \begin{equation*} \tag*{[6.1]} \begin{aligned} \frac{d u}{d t}\frac{dx}{da} + \frac{d v}{d t} \frac{dy}{da}+\frac{dw }{d t} \frac{dz}{da} - \frac{d \Omega}{da} =0\\\\ \frac{d u}{d t}\frac{dx}{db} + \frac{d v}{d t} \frac{dy}{db}+\frac{dw }{d t} \frac{dz}{db} - \frac{d \Omega}{db} =0\\\\\frac{d u}{d t}\frac{dx}{dc} + \frac{d v}{d t} \frac{dy}{dc}+\frac{dw }{d t} \frac{dz}{dc} - \frac{d \Omega}{dc} =0 \end{aligned} \end{equation*} \indent\indent From these equations, one can easily eliminate the function $\Omega$ and obtain equations which represent all possible fluid motions under the influence of potential forces.\hspace{0.4cm}The elimination is easily done through differentiations with respect to $a,b,c$ and subtractions from which one obtains\,: \begin{equation*} \tag*{[6.2]} \begin{aligned} \frac{d^2 u}{d t dc}\frac{dx}{db} - \frac{d^2 u}{d t db}\frac{dx}{dc} + \frac{d^2 v}{d t dc} \frac{dy}{db} -\frac{d^2 v}{d t db} \frac{dy}{dc} +\frac{d^2 w }{d t dc} \frac{dz}{db} - \frac{d^2 w }{d t db} \frac{dz}{dc} =0\\\\ \frac{d^2 u}{d t da}\frac{dx}{dc} - \frac{d^2 u}{d t dc}\frac{dx}{da} + \frac{d^2 v}{d t da} \frac{dy}{dc} -\frac{d^2 v}{d t dc} \frac{dy}{da} +\frac{d^2 w }{d t da} \frac{dz}{dc} - \frac{d^2 w }{d t dc} \frac{dz}{da} =0\\\\ \frac{d^2 u}{d t db}\frac{dx}{da} - \frac{d^2 u}{d t da}\frac{dx}{db} + \frac{d^2 v}{d t db} \frac{dy}{da} -\frac{d^2 v}{d t da} \frac{dy}{db} +\frac{d^2 w }{d t db} \frac{dz}{da} - \frac{d^2 w }{d t da} \frac{dz}{db} =0 \end{aligned} \end{equation*} One can readily integrate these equations with respect to time by writing each of the three differences in these equations as an exact time derivative.\hspace{0.4cm} Denoting the time-independent integration constants as $2A,2B,2C$, one finds\,: \begin{equation*}\tag*{(1), [6.3]} \left. \begin{aligned} \frac{du}{d c}\frac{dx}{db} - \frac{du}{d b}\frac{dx}{dc} + \frac{dv}{d c}\frac{dy}{db} - \frac{dv}{d b}\frac{dy}{dc} + \frac{dw}{d c}\frac{dz}{db} - \frac{dw}{d b}\frac{dz}{dc} =2A\\\\ \frac{du}{d a}\frac{dx}{dc} - \frac{du}{d c}\frac{dx}{da} + \frac{dv}{d a}\frac{dy}{dc} - \frac{dv}{d c}\frac{dy}{da} + \frac{dw}{d a}\frac{dz}{dc} - \frac{dw}{d c}\frac{dz}{da} =2B\\\\ \frac{du}{d b}\frac{dx}{da} - \frac{du}{d a}\frac{dx}{db} + \frac{dv}{d b}\frac{dy}{da} - \frac{dv}{d a}\frac{dy}{db} + \frac{dw}{d b}\frac{dz}{da} - \frac{dw}{d a}\frac{dz}{db} =2C \end{aligned} \qquad \right\} \end{equation*} The left sides of these integral equations can be seen as the difference of two differential quotients [a curl is meant]. \hspace{0.4cm}Namely, defining\,: \\ \begin{equation*} \tag*{(2), [6.4]} \left. \begin{aligned} \alpha = u \frac{dx}{da} + v \frac{dy}{da} + w \frac{dz}{da}\\\\ \beta = u \frac{dx}{db} + v \frac{dy}{db} + w \frac{dz}{db}\\\\ \gamma = u \frac{dx}{dc} + v \frac{dy}{dc} + w \frac{dz}{dc} \end{aligned} \qquad \right\} \end{equation*} one obtains, instead of $(1)$\,: \begin{equation*} \tag*{(3), [6.5]} \label{pointer2} \frac{d \beta}{dc} - \frac{d \gamma}{db} = 2A,\,\,\, \frac{d \gamma}{da} - \frac{d \alpha}{dc} = 2B, \,\,\,\frac{d \alpha}{db} - \frac{d \beta}{da} = 2C \end{equation*} These interesting relations are the analogues of the equations that \mbox{C\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! y}\footnote{% In an essay prized by the Paris Academy: M\'emoire sur la th\'eorie de la propagation des ondes \`a la surface d'un fluide pesant d'une profondeur infinie (M\'em. sav. \'etran.\ Bd.\ 1) [\hyperlink{Cauchy1}{1827}].} found already in 1816 [actually, already in 1815] for the first E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! ian dependence.\hspace{0.4cm}They attained their actual importance only when \mbox{H\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! z}\footnote{% Ueber Integrale der hydrodynamischen Gleichungen, welche den Wirbelbewegungen entsprechen, Crelle's Journal, Bd. 55, S.105, [\hyperlink{Helmholtz1}{1858}]. [\textit {English translation}: On Integrals of the hydrodynamic equations that correspond to vortex motions. Philos. Mag. {\bf 4}, Vol. 33, 485--511, [\hyperlink{Helmholtz1}{1868}].} realised their mechanical significance% \deffootnotemark{\textsuperscript{[T.\thefootnotemark]}}\deffootnote{2em}{1.6em}{[\hspace{0.01cm}T.\thefootnotemark]\enskip}% \footnote{It appears likely that Helmholtz was not aware of Cauchy's equations but stressed the importance of vortex dynamics.} [of these equations], and thereby laid the foundations for a peculiar treatment of hydrodynamics.\hspace{0.4cm}It will be our next task to investigate this significance using a method adapted to the second \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r} dependence; for this we first need to develop an appropriate theorem.\\\\ \centerline{\fett{$\S.\,7.$}}\\[0.3cm] \indent\indent For this purpose we start from the known theorem that if $\xi$ and $\eta$ are arbitrary continuous functions of $x$ and $y$, one has the relation\,: \begin{equation*} \tag*{[7.1]} \int \left( \xi {\rm d}x + \eta {\rm d}y \right) =\iint \left( \frac{d \xi}{dy} - \frac{d \eta}{dx}\right) {\rm d}x \, {\rm d}y \end{equation*} where the double integral has to be extended over all elements of a domain on the $xy$ plane, and the simple integral is over the boundary of the domain suitably oriented.% \deffootnotemark{\textsuperscript{[A.\thefootnotemark]}}\deffootnote{2em}{1.6em}{[A.\thefootnotemark]\enskip}% \footnote{Cf.\ B.\ Riemann, Lehrs\"atze aus der analysis situs f\"ur die Theorie der Integrale von zweigliedrigen vollst\"andigen Differentialien. Crelle's Journal Bd.\ 54, S.\ 105, [\hyperlink{Riemann1}{1857}].}\\ \indent\indent One can generalise this theorem in the following way\,: Let there be an arbitrary closed curve in space and consider the path integral over all elements of this curve\,: \begin{equation*} \notag \int \left( \xi\, {\rm d}x + \eta\, {\rm d}y + \zeta\, {\rm d}z \right) \end{equation*} where $\xi,\,\eta,\,\chi$ are arbitrary continuous functions of $x,y,z$.\hspace{0.4cm}Let us think of an arbitrary connected surface limited by this curve, so that one can consider on this whole surface $z$ as a function of $x$ and $y$ and set\,: \begin{equation*} \tag*{[7.2]} {\rm d}z = \frac{dz}{dx} {\rm d}x + \frac{dz}{dy} {\rm d}y, \end{equation*} from which the given integral transforms into\,: \begin{equation*} \notag \int \left\{ \left (\xi + \frac{dz}{dx} \zeta \right) {\rm d}x + \left (\eta + \frac{dz}{dy} \zeta \right) {\rm d}y \right \} \end{equation*} By the {stated theorem for [the case of] two independent variables, this integral becomes\,: \begin{equation*} \notag \iint \left\{ \frac{ d\left(\xi + \frac{dz}{dx} \zeta \right)}{dy} - \frac{\left(\eta + \frac{dz}{dy} \zeta \right)}{dx}\right\} {\rm d}x {\rm d}y \end{equation*} Since $z$ is a function of $x$ and $y$, one has\,: \begin{equation*} \tag*{[7.3]} \begin{aligned} \frac{ d\left(\xi + \frac{dz}{dx} \zeta \right)}{dy} &= \frac{d \xi}{dy} + \frac{d \xi}{dz} \frac{d z}{dy} + \frac{d \zeta}{dy} \frac{d z}{dx} + \frac{d \zeta}{dz} \frac{d z}{dx}\frac{d z}{dy} + \frac{d^2 z}{dx dy} \zeta\\ \frac{ d\left(\eta + \frac{dz}{dy} \zeta \right)}{dx} &= \frac{d \eta}{dx} + \frac{d \eta}{dz} \frac{d z}{dx} + \frac{d \zeta}{dx} \frac{d z}{dy} + \frac{d \zeta}{dz} \frac{d z}{dy}\frac{d z}{dx} + \frac{d^2 z}{dy dx} \zeta \end{aligned} \end{equation*} and so one has the equation\,: \begin{smallequation} \tag*{[7.4]} \int \left( \xi {\rm d}x + \eta {\rm d}y + \zeta {\rm d}z \right) = \iint \left \{ \left(\frac{d \xi} {dy} - \frac{d \eta}{dx} \right) + \left(\frac{d \zeta} {dy} - \frac{d \eta}{dz} \right)\frac{dz}{dx} + \left( \frac{d \xi}{dz} - \frac{d \zeta}{dx}\right) \frac{dz}{dy} \right\}{\rm d}x {\rm d}y, \end{smallequation} where now the double integral has to be extended over all the elements of the surface limited by the curve.\hspace{0.4cm}This otherwise arbitrary surface through the curve has just to satisfy the condition that the part embedded in the curve is not multiply-connected and that the curve forms a complete boundary to it.\hspace{0.4cm}Let $\lambda,\mu, \nu$ be the angles of the normal drawn to the surface with the coordinates axes, one has\,: \begin{equation*} \tag*{[7.5]} \frac{dz}{dx} = - \frac{\cos \lambda}{\cos \nu}, \,\,\, \frac{dz}{dy} = - \frac{\cos \mu}{\cos \nu}; \end{equation*} from which follows that one has \begin{equation*}\tag*{(1), [7.6]} \begin{aligned} &\hspace{4cm}\int \left( \xi {\rm d}x + \eta {\rm d}y +\zeta {\rm d}z \right)=\\ &\int \left \{ \left(\frac{d \eta} {dz} - \frac{d \zeta}{dy} \right)\cos \lambda + \left(\frac{d \zeta} {dx} - \frac{d \xi}{dz} \right) \cos \mu + \left( \frac{d \xi}{dy} - \frac{d \eta}{dx}\right) \cos \nu \right\} {\rm d} \sigma \end{aligned} \end{equation*} where $\displaystyle \frac{ {\rm d}x\,{\rm d}y}{\cos \nu} = {\rm d} \sigma$ denotes the surface element. \\ \indent\indent For our purposes, we can bring both of the above integrals into more convenient forms\,:\\\\ \indent\indent If one determines three angles $\lambda', \, \, \mu', \,\,\nu'$ so that\,: \begin{equation*} \tag*{[7.7]} \cos \lambda': \cos \mu': \cos \nu'=\left(\frac{d \eta} {dz} - \frac{d \zeta}{dy} \right) : \left(\frac{d \zeta} {dx} - \frac{d \xi}{dz} \right) : \left( \frac{d \xi}{dy} - \frac{d \eta}{dx} \right) \end{equation*} and, at the same time, assumes that they are the angles of a definite direction with the coordinates axes, so that\,: \begin{equation*} \tag*{[7.8]} \cos^2 \lambda' + \cos^2 \mu' + \cos^2 \nu' =1 \end{equation*} one finds\,: \begin{equation*} \tag*{[7.9]} 2 \Delta \cos \lambda' =\frac{d \eta} {dz} - \frac{d \zeta}{dy},\,\, 2 \Delta \cos \mu' = \frac{d \zeta} {dx} - \frac{d \xi}{dz}, \,\, 2 \Delta \cos \nu' = \frac{d \xi}{dy} - \frac{d \eta}{dx} \end{equation*} where\,: \begin{equation*} \tag*{[7.10]} 4 \Delta^2 = \left(\frac{d \eta} {dz} - \frac{d \zeta}{dy} \right) ^2 + \left(\frac{d \zeta} {dx} - \frac{d \xi}{dz} \right) ^2 + \left( \frac{d \xi}{dy} - \frac{d \eta}{dx} \right)^2 \end{equation*} The integral above, whose element is ${\rm d}\sigma$, now goes into \begin{equation*} \notag 2 \int \Delta \left(\cos \lambda \cos \lambda' + \cos \mu \cos \mu' + \cos \nu \cos \nu' \right) {\rm d}\sigma \end{equation*} and one obtains\,: \begin{equation*} \tag*{(2), [7.11]} \int \left\{\left( \frac{d \eta}{dz} - \frac{d \zeta}{dy} \right) \cos \lambda + \left( \frac{d \zeta}{dx} - \frac{d \xi}{dz} \right) \cos \mu + \left( \frac{d \xi}{dy} - \frac{d \eta}{dx} \right) \cos \nu \right \}{\rm d}\sigma = 2\int \Delta \cos \theta {\rm d}\sigma \,, \end{equation*} where $\theta$ is the angle between the directions determined by $\lambda, \mu, r$ and $\lambda',\mu', r' $ .\\ \indent\indent Indicating by $\varrho, \sigma, \tau$ the angles of the coordinate axes with the tangent to the curve at the point $x,y,z$, one has\,: \begin{equation*} \tag*{[7.12]} {\rm d}x= \cos \varrho {\rm d}s, \,\,\, {\rm d}y = \cos \sigma {\rm d}s, \, \, \, {\rm d}z= \cos \tau {\rm d}s, \end{equation*} where ${\rm d}s$ indicates the arc element.\hspace{0.4cm}Let now $ \varrho', \sigma', \tau'$ be the angles of a direction with the coordinate axes, so that\,: \begin{equation*} \tag*{[7.13]} \cos^2 \varrho' + \cos^2 \sigma' + \cos^2 \tau' = 1, \end{equation*} and furthermore [require that] \begin{equation*} \tag*{[7.14]} \cos \varrho': \cos \sigma': \cos \tau'= \xi:\eta:\zeta, \end{equation*} so one has\,: \begin{equation*} \tag*{[7.15]} U^2 = \xi^2 + \eta^2 + \zeta^2 \end{equation*} and \begin{equation*} \tag*{[7.16]} U \cos \rho'=\xi,\,\, U \cos \sigma' = \eta, \,\,U \cos \tau' = \zeta. \end{equation*} These values of $\xi,\,\eta, \,\zeta$ and ${\rm d}x$, ${\rm d}y$, ${\rm d}z$ substituted into the simple integral, give\,: \begin{equation*} \tag*{[7.17]} \int \left( \xi {\rm d} x + \eta {\rm d}y + \zeta {\rm d}z\right)\, = \int U (\cos \varrho \cos \varrho' +\cos \sigma \cos \sigma' + \cos \tau \cos \tau' ) {\rm d}s \end{equation*} or \begin{equation*}\tag*{(3), [7.18]} \int\left( \xi {\rm d} x + {\eta\rm d} y + \zeta {\rm d}z \right)\, =\,\int U \cos \theta' {\rm d}s \end{equation*} where $\theta'$ indicates the angle between the directions determined by $\varrho,\sigma,\tau$ and $\varrho',\sigma',\tau'$. \\[1cm] \centerline{\fett{$\S.\,8.$}}\\[0.3cm] \indent\indent After development of this lemma we come back to the task described at the end of $\S\,6$\,:\\ \indent\indent According to $(1)$ of $\S.\,7.$, one can write the equation\,: \begin{smallequation} \tag*{[8.1]} \int\left( \alpha {\rm d}a + \beta {\rm d}b + \gamma {\rm d}c \right)\, =\, \int \left \{ \left( \frac{d \beta}{dc} - \frac {d \gamma}{db} \right)\cos \lambda + \left( \frac{d \gamma}{da} - \frac {d \alpha}{dc} \right)\cos \mu + \left(\frac{d \alpha}{db} - \frac {d \beta}{da} \right) \cos \nu \right \} {\rm d} \sigma_0 \end{smallequation} where the first [l.h.s.] integral is over a closed curve, the second [r.h.s.] over an arbitrary simply-connected surface bounded by that curve, and where $\lambda, \mu, \nu$, are the angles of the normal to that surface with the coordinate axes, ${\rm d} \sigma_0$ is the surface element and $\alpha,\beta,\gamma $ are arbitrary functions of $ a,b,c $.\hspace{0.4cm}According to $(2)$ of the $\S\,7$, the second integral can be transformed, so that\,: \begin{equation*} \tag*{(1), [8.2]} \int\left( \alpha\,{\rm d}a + \beta\, {\rm d}b + \gamma \, {\rm d}c \right)\, =\, 2 \int \Delta_0 \cos \theta_0 {\rm d} \sigma_0 \,. \end{equation*} If one now takes $\alpha, \beta,\gamma $ to be the quantities defined in $(2)$ of $\S\,6$ and makes use of equations~(3) of $\S\,6$, one obtains\,: \begin{equation*} \tag*{[8.3]} \Delta_0 = \sqrt{A^2 + B^2 + C^2} \,. \end{equation*} As to $\theta_0$, it is the angle between the directions determined by $\lambda, \mu, \nu$ and $\lambda', \mu', \nu'$, the latter being determined by\,: \begin{equation*} \tag*{[8.4]} \Delta_0 \cos \lambda' = \Delta, \, \, \Delta_0 \cos \mu' = B, \, \, \Delta_0 \cos \nu' = C \,. \end{equation*} The integral on the right-hand side of the preceding equation $(1)$ does not depend on time, thus one may choose an arbitrary value of $t$ on its left-hand side; if one sets $t=0$, then $\alpha, \beta, \gamma $ go into the initial values of $u,\,v,\,w$ that we denote with $u_0, v_0, w_0$. One then obtains\,: \begin{equation*} \tag*{[8.5]} \int \left ( u_0 {\rm d}a + v_0 {\rm d}b + w_0 {\rm d}c \right) = 2 \int \Delta_0 \cos \theta_0 {\rm d} \sigma_0 \end{equation*} According to $(3)$ of the $\S.\,7.$, the first integral may be transformed and thus one finds\,: \begin{equation*} \tag*{(2)\,, [8.6]} \int U_0 \cos \theta'_0 {\rm d} s_0 = 2 \int \Delta_0 \cos \theta_0 {\rm d} \sigma_0 \end{equation*} where \begin{equation*} \tag*{[8.7]} U_0 = \sqrt{u_0 ^2 + v_0^2 +w_0^2} \end{equation*} is the initial velocity, $\theta'_0$ is the angle between $U_0$ and the element ${\rm d}s_0$ of the closed curve.\\ \indent\indent Let us assume, that the given closed curve is a circle with an infinitely small radius $r$ and the surface spanned by the curve is the disk of the circle. So it is clear that $\Delta_0 \cos \theta_0 $ for different points of the circle surface changes only by infinitely small quantities.\hspace{0.4cm} Then, we have $\int \Delta_0\, \cos \theta_0 \,{\rm d} \sigma =\pi r^2 \cdot \Delta_0 \cos \theta_0$.\hspace{0.4cm}The component of the initial velocities with respect to the tangent to this circle is $U_0 \cos \theta_0'$, a quantity, that for differents points of the circumference will take different values.\hspace{0.4cm}However, $U_0 \cos \theta_0'$ in each point may be considered as the sum of two velocities $T_0 + T_0'$, where $T_0'$ is the progressive motion projected to a tangent, which is common to all points of the circle and $T_0$ the tangential velocity by the rotation around the center of the infinitely small circle.\hspace{0.4cm}The progressive motion of all particles is the same\,: as a consequence its component $T'_0$ with respect to the tangent of the circle will be the same for two particles diametrically opposite, but of opposite signs, so that $\int T_0' {\rm d}s_0=0$, when one integrates over the whole circle.\hspace{0.4cm}The relative velocity of the particles will be the same except for infinitely small quantities provided the velocities are continous functions of the position, which we always suppose.\hspace{0.4cm}So, it follows that $\int T_0 {\rm d}s_0=T_0 \cdot \int {\rm d}s_0=2\pi r \cdot T_0$.\hspace{0.4cm}From the equation (2) it follows now $2\pi r \cdot T_0= 2 \cdot \pi r^2 \cdot \Delta_0 \cos \theta_0$ or\,: \begin{equation*} \tag*{[8.8]} \Delta_0 \cos \theta_0= \frac{T_0}{r} \end{equation*} Since $T_0$ is the tangential velocity, so $T_0\,: r$ is the rotational velocity around the nfinitely small distant center of the circle, or --- as one can say, in order to take into account, also the location and orientation of the circle --- is the rotational velocity around the normal to the surface of the circle taken as axis. Since $\theta_0$ is the angle between the normal to the surface element and the direction determined by the angles $\lambda',\mu',\nu'$, then $\Delta_0$ in a point becomes the rotational velocity around an axis passing by this point and oriented in the direction $\lambda',\mu',\nu'$.\\ \indent\indent Furthermore $A,\,\,B,\,\,C$ are the components of the rotational velocity of a particle $a,\,b,\,c$ around axes, which are parallel to the coordinates axes through the point $a,\,b,\,c$. \\[1cm] \centerline{\fett{$\S\,9.$}}\\[0.3cm] \indent\indent Analogously to $(2)$ of $\S.\,8$, one has at each time $t$\,: \begin{equation*} \tag*{(1), [9.1]} \int U \cos \theta' {\rm d} s = 2 \int \Delta \cos \theta {\rm d}\sigma \end{equation*} where now $ U \cos \theta' $ is the component of the velocity of a particle $x, y, z$ with respect to the tangent to the closed curve, over whose element ${\rm d}s$ the first integral has to be extended, and where $\Delta \cos \theta$ is the component of the rotational velocity expressed with respect to the normal to the element ${\rm d} \sigma$ of the surface, which is bounded by that curve.\\ \indent\indent According to $(3)$ of $\S.\,7$, one has \begin{equation*} \tag*{[9.2]} \int U \cos \theta' {\rm d}s = \int \left( u{\rm d}x + v {\rm d}y + w {\rm d}z \right) \end{equation*} but, since from $ (2) $ of $\S.\,6$ one easily infers that \begin{equation*} \tag*{[9.3]} \alpha {\rm d} a + \beta {\rm d} b + \gamma {\rm d} c = u {\rm d}x + v {\rm d}y + w {\rm d}z, \end{equation*} one has\,: \begin{equation*} \tag*{[9.4]} \int U \cos \theta' {\rm d}s = \int \left(\alpha {\rm d}a + \beta {\rm d}b + \gamma {\rm d}c \right) \end{equation*} The second integral is calculated according to $(1)$ of $\S.\,8$\,: \begin{equation*} \tag*{[9.5]} \int U \cos \theta' {\rm d}s = 2 \int \Delta_0 \cos \theta_0 {\rm d} \sigma_0 \,. \end{equation*} Hence, because of equation $(1)$, one obtains\,: \begin{equation*} \tag*{(2), [9.6]} \int \Delta \cos \theta {\rm d}\sigma = \int \Delta_0 \cos \theta_0 {\rm d} \sigma_0 \end{equation*} Herefrom follows that $\int \Delta \cos \theta' {\rm d}\sigma$ is constant with respect to time, provided the integral is always extended over a surface moving with the flow and consisting of the same particles. Following \mbox{H\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! z,} if one designates by \mbox{r\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! l} \,\mbox{i\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! s\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! y} the product of the rotational velocity around the normal to the surface as an axis, times the size of the surface element, one obtains the following result\,:\\ \indent\indent The integral of the rotational intensity over a surface always formed by the same particles remains unchanged in time. \\ \indent\indent Since this theorem is valid, irrespective of how small the surface may be, it is also valid for each single surface element\,: the rotational intensity of a surface element always stays the same.\hspace{0.4cm}Because such a particle cannot spread out infinitely, its rotational velocity cannot decrease infinitely. It follows herefrom that no particle once put into rotational motion can stop rotating\,; and on the other hand, one easily sees that no particle which at initial time is not rotating, may ever begin to rotate.\\ \indent\indent One must remark that these results are obtained under the assumption that the accelerating forces acting on the fluid are partial derivatives of a potential function. If, however, the accelerating forces do not possess this property, these theorems do not apply.\hspace{0.4cm}Herefrom one obtains a criterion to know whether accelerating forces acting on a fluid without pressure forces, have or do not have a potential.{ In the latter case, the rotational intensity of single particles is not conserved; in general new particles begin rotating, and rotating particles will lose this characteristic motion.\\ \indent\indent So far we have considered only rotating surface elements, but let us take into account a mass element which is contained in a cylinder whose axis is the rotation axis. Its constant mass is the product of its transverse section, its length and its density.\hspace{0.4cm}Since the product of the rotational velocity by the transverse section is constant, one sees that for each element the ratio of its rotational velocity to the product of the distance measured in the direction of its rotation axis by the density is constant.\label{pointer}\hspace{0.4cm} Therefore, if the fluid is liquid, i.e., its density considered as constant, the ratio of the rotational velocity to the length of the particle is constant.\\[1cm] \centerline{\fett{$\S.\,10.$}}\\[0.3cm] \indent\indent In his theory of rotational motion, \mbox{H\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! z} introduced this following important principle\,: instead of considering the whole rotating mass, one should fragment it into \mbox{v\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! x} \mbox{l\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! s.} Here, a vortex line is a line lying in the flow so that its direction will stay always parallel to the instantaneous rotation axis.} By v\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! x f\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! t, we understand the infinitely thin cylinder which, wrapped around the vortex line, includes the rotating particles.\\ \indent \indent Denoting by ${\rm d}a, {\rm d}b, {\rm d}c$ [the three components of]} an element of such a vortex line at time $t=0$, one obviously has\,: \begin{equation*} \tag*{(1), [10.1]} {\rm d}a : {\rm d}b :{\rm d}c = A\,:\,B\,:\,C \end{equation*} Let $\varphi$ and $\psi$ be functions of $a,\, b,\,c$, such that the vortex lines at time $t=0$ are obtained by setting these two functions to constant values. Then the following conditions must hold\,: \begin{equation*} \tag*{[10.2]} \begin{aligned} \frac{d \varphi}{da} {\rm d}a + \frac{d \varphi}{db} {\rm d}b + \frac{{\rm d}\varphi}{dc} {\rm d}c = 0\\\\ \frac{d \psi}{da} {\rm d}a + \frac{d \psi}{db} {\rm d}b + \frac{d \psi}{dc} {\rm d}c = 0\\ \end{aligned} \end{equation*} or, according to $(1)$\,: \begin{equation*} \tag*{[10.3]} \begin{aligned} \frac{d \varphi}{da} A + \frac{d \varphi}{db} B + \frac{d \varphi}{dc} C &= 0\\ \frac{d \psi}{da} A + \frac{d \psi}{db} B + \frac{d \psi}{dc} C &= 0 \,.\\ \end{aligned} \end{equation*} If $A,\,B,\,C$ were} known, one could find $\varphi$ and $\psi$ from these equations by integration. \hspace {0.4cm}One can easily observe that $\varphi, \,\psi$ must be such that, one may write\,: \begin{equation*} \tag*{(2), [10.4]} -2 A = \begin{vmatrix} \frac{d \varphi}{db}& \frac{d \varphi}{dc} \\\\ \frac{d \psi}{db}& \frac{d \psi}{dc} \end{vmatrix},\\ \,\,-2 B = \begin{vmatrix} \frac{d \varphi}{dc}& \frac{d \varphi}{da} \\\\ \frac{d \psi}{dc}& \frac{d \psi}{da} \end{vmatrix},\\ \,\,-2 C = \begin{vmatrix} \frac{d \varphi}{da}& \frac{d \varphi}{db} \\\\ \frac{d \psi}{da}& \frac{d \psi}{db} \end{vmatrix} \,.\\ \end{equation*} Indeed, from the preceding equations follows that $A, B, C$ are proportional to these determinants\,; however, by substitution of~(2) into~(3) of $\S.\,6$, the following relation is identically satisfied\,: \begin {equation*} \tag*{(3), [10.5]} \frac{dA}{da} + \frac{dB}{db} +\frac{dC}{dc} =0 \,. \end{equation*} Thus, $\varphi,\, \psi$ are always determined by integration in such a way that equations~(2) are satisfied.\\ \indent\indent From these considerations it follows that rotating particles may always be considered as arranged in vortex filaments. \hspace{0.4cm} When such a vortex filament moves along with the fluid, always the same particles will belong to it, as no particle belonging to the vortex filament may lose its rotational motion and, furthermore, each element parallel to the rotation axis of the vortex filament always remains parallel. Namely, a surface element, which, at some time, is parallel to the rotation axis, will have a vanishing rotational intensity with respect to the direction of its normal; since the rotational intensity always stays the same, the normal to the surface element always stays perpendicular to the rotation axis, because the particles themselves always maintain their rotational motion around the axis of the vortex filament.\hspace{0.4cm}Therefore, one obtains the equations of the vortex lines at time $t$, if one expresses the values of $a,\,b,\,c$ in $\varphi$ and $\psi$ through $x,\,y,\,z,\,t$.\hspace{0.4cm}The equations of the vortex lines at time $t$ must be such that $\displaystyle \frac{{\rm d} \varphi}{{\rm d}t} =0$ and $\displaystyle \frac{{\rm d}\psi}{{\rm d}t} =0$ are satisfied. Here one has to differentiate with respect to $t$, whether it appears explicitly or implicitly, through $x,\,y,\,z$. \hspace{0.4cm} In motion, the vortex filament will sometimes grow in density and size and sometimes decrease, namely in such a way that the length of a vortex-tube element multiplied by the density remains proportional to the rotational velocity. \mbox{(Cf.\ S.\,40 [now page~\pageref{pointer}].)}\\ \indent\indent The rotational intensity in each section of this vortex filament will remain unchanged over time.\hspace{0.2cm} Moreover, it is also the same for all sections. \hspace{0.4cm}Obviously we have $\int \Delta \cos \theta {{\rm d} \sigma}=0$ if the integral extends over a closed surface.\hspace{0.4cm} Namely, let us think of a closed curve drawn on this surface. In order to extend $\int \Delta \cos \theta {{\rm d}\sigma}$ over {the two portions of the surface separated by the curve, one has to integrate $\int U \cos \theta' {\rm d}s$ twice but in the opposite direction over this curve, so that one has $\int \Delta \cos \theta {{\rm d}\sigma}=0$.\hspace{0.4cm} Let this closed surface be formed by two sections of a vortex filament and the part of the filament surface between them, then $\int \Delta \cos \theta {{\rm d}\sigma}$ will disappear for the latter part, and so will the sum of the integrals for both sections. Indeed, since $\cos \theta$, in a section, is $+1$ [and] $-1$ in the other one, the rotational intensity with regard to the axis of the vortex filament is the same for each section. \\ \indent\indent The results of the preceding investigations about these particular rotational motions are essentially due to \mbox{H\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! z,} who, by restricting himself to liquid flows, derived them in a classical essay\footnote{Crelle's Journal, Bd.\ 55, S.\,33 and ff [\hyperlink{Helmholtz1}{1858}].} from the \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! }ian first form of the equations. \indent\indent A special case of the theorem on the constancy of the rotational velocity has been already given by \mbox{S\hspace{0.036cm}} %{$\phantom{!}$\! v\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! b\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! g\,:}\\ \indent\indent Namely, when all motions happen symmetrically around an axis and circular vortex filaments centered on the symmetry axis} are given in the flow with radius $r$ and infinitely small section $\omega$, such a vortex filament may be considered as a cylinder of height $2 \pi r$ and basis $\omega$ so that its content will be $ 2 \pi r \omega$.\hspace{0.4cm}Since the rotational motion of the particles is conserved, the content $2 \pi r \omega$ for each vortex filament must be constant in time.\hspace{0.4cm}Then $\Delta \omega r$ must be proportional to the rotational velocity $\Delta$.\hspace{0.4cm}Since $\Delta \omega$, as rotational intensity, is constant, so the rotational velocity $\Delta$ for each vortex ring at different times must be proportional to its radius.\\ \indent\indent This is the meaning of the formulae derived by \mbox{S\hspace{0.036cm}} %{$\phantom{!}$\! v\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! b\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! g} from the first \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r\,ian} form\footnote {Crelle's Journal Bd.\ 24. S.\ 159. Nro.\ 31 [\hyperlink{Svanberg}{1842}].} of the fundamental equations.\\[1cm] \centerline{\fett{$\S \,11.$}} \\[0.3cm] \indent\indent The concept of vortex lines gives rise to an interesting transformation of the hydrodynamical equations in their first form: Namely, by introducing as dependent variables, functions which are in a certain relation to $u,v,w$.\\ \indent\indent From equations (2) of $\S \,10$ follows indeed\,: \begin{equation*} \tag*{[11.1]} \begin{aligned} \frac{d\gamma}{db} - \frac{d\beta }{dc}= \frac{d\varphi}{db}\frac{d\psi}{dc} - \frac{d\varphi}{dc}\frac{d\psi}{db}\\\\ \frac{d\alpha}{dc} - \frac{d\gamma }{da}= \frac{d\varphi}{dc}\frac{d\psi}{da} - \frac{d\varphi}{da}\frac{d \psi}{dc}\\\\ \frac{d \beta}{da} - \frac{d\alpha }{db} = \frac{d\varphi}{da}\frac{d\psi}{db} - \frac{d\varphi}{db}\frac{d\psi}{dc} \end{aligned} \end{equation*} provided that $\varphi$ and $\psi$ are functions of $a,b,c$ that represent the vortex line, when set equal to constants. Herefrom follows that one can represent $\alpha, \beta, \gamma$ in the form\,: \begin{equation*} \tag*{(1), [11.2]} \alpha = \frac{dF}{d a} + \varphi \frac{d \psi}{d a}, \,\,\,\, \beta= \frac{dF}{d b} + \varphi \frac{d \psi}{d b}, \,\,\,\,\gamma = \frac{dF}{d c} + \varphi \frac{d \psi}{d c} \end{equation*} where in general $F$ will be function of $a,\,b,\,c,\,t$.\hspace{0.4cm}Herefrom one finds\,: \begin{equation*} \tag*{[11.3]} \begin{aligned} \alpha \frac{da}{dx} + \beta \frac{db}{dx} + \gamma \frac{dc}{dx} = \frac{dF}{dx} + \varphi \frac{d \psi}{dx}\\\\ \alpha \frac{da}{dy} + \beta \frac{db}{dy} + \gamma \frac{dc}{dy} = \frac{dF}{dy} + \varphi \frac{d \psi}{dy}\\\\ \alpha \frac{da}{dz} + \beta \frac{db}{dz} + \gamma \frac{dc}{dz} = \frac{dF}{dx} + \varphi \frac{d \psi}{dz} \end{aligned} \end{equation*} \hspace{0.6cm}By solving the system (2) of $\S\,6$ using the relations $(3)$ of $\S\,2$, one finds\,: \begin{equation*} \tag*{[11.4]} \begin{aligned} \alpha \frac{da}{dx} + \beta \frac{db}{dx} + \gamma \frac{dc}{dx} &= u\\\\ \alpha \frac{da}{dy} + \beta \frac{db}{dy} + \gamma \frac{dc}{dy} &= v\\\\ \alpha \frac{da}{dz} + \beta \frac{db}{dz} + \gamma \frac{dc}{dz} &= w \end{aligned} \end{equation*} so that one can always set\,: \begin{equation*} \tag*{(2), [11.5]} u = \frac{dF}{dx} + \varphi \frac{d \psi} {dx}, \,\,\, v= \frac{dF}{dy} + \varphi \frac{d \psi}{dy},\,\,\, w= \frac{dF}{dz} + \varphi \frac{d \psi}{dz} \end{equation*} where $\varphi= Const$ and $\psi=Const$ are the equations of the vortex lines.\hspace{0.4cm} If one replaces the quantities $a,\,b,\,c$ in $\varphi$ and $\psi$ in terms of $x,y,z$ [at time] $t$, one obtains\;: \begin{equation*} \tag*{[11.6]} \frac{{\rm d} \varphi}{{\rm d}t} = 0, \qquad \frac{{\rm d} \psi}{{\rm d}t}=0 \end{equation*} whereby differentiation has to be performed when $t$ appears explicitly as well as implicitly in $x,y,z$. Hence we have\,: \begin{equation*} \tag*{[11.7]} \begin{aligned} \frac{d \varphi}{dt} + \frac{d \varphi}{dx} u + \frac{d \varphi}{dy} v +\frac{d \varphi}{dz} w &=0\\\\ \frac{d \psi}{dt} + \frac{d \psi}{dx} u + \frac{d \psi}{dy} v +\frac{d \psi}{dz} w &=0 \ \end{aligned} \,. \end{equation*} Substituting the values of $u,\,v,\,w$ [taken from~(2)], we obtain\,: \begin{equation*}\tag*{(3), [11.8]} \left. \begin{aligned} \frac{d \varphi}{dt} + \frac{d \varphi}{dx} \left (\frac{dF}{dx}+ \varphi \frac{d \psi}{dx}\right) + \frac{d \varphi}{dy} \left (\frac{dF}{dy}+ \varphi \frac{d \psi}{dy}\right) + \frac{d \varphi}{dz} \left (\frac{dF}{dz}+ \varphi \frac{d \psi}{dz}\right) =0 \\\\ \frac{d \psi}{dt} + \frac{d \psi}{dx} \left (\frac{dF}{dx}+ \varphi \frac{d \psi}{dx}\right) + \frac{d \psi}{dy} \left (\frac{dF}{dy}+ \varphi \frac{d \psi}{dy}\right) + \frac{d \psi}{dz} \left (\frac{dF}{dz}+ \varphi \frac{d \psi}{dz}\right) =0 \end{aligned} \qquad \right\} \end{equation*} In addition to these relations there is also the density equation $(5)$ of $\S.\,2.$ \begin{equation*} \tag*{[11.9]} \frac{d \varrho}{dt} + \frac{d }{dx} \varrho \left (\frac{dF}{dx}+ \varphi \frac{d \psi}{dx}\right) + \frac{d }{dy} \varrho \left (\frac{dF}{dy}+ \varphi \frac{d \psi}{dy}\right) + \frac{d }{dz} \varrho \left (\frac{dF}{dz}+ \varphi \frac{d \psi}{dz}\right) =0 \end{equation*} We observe that, in general, these three equations are not sufficient to determine the four unknown functions $F,\varphi,\psi$ and $\varrho$. Only in the case of liquid fluids, where $\varrho$ is constant, are these sufficient, insofar as the latter expression transforms into\,: \begin{equation*} \tag*{(4), [11.10]} \frac{d}{dx} \left(\frac{dF}{dx}+ \varphi \frac{d \psi}{dx} \right) + \frac{d}{dy} \left(\frac{dF}{dy}+ \varphi \frac{d \psi}{dy} \right) + \frac{d}{dz} \left(\frac{dF}{dz}+ \varphi \frac{d \psi}{dz} \right) =0 \end{equation*} These are the transformed equations in the elegant form that \mbox{C\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! b\hspace{0.036cm}} %{$\phantom{!}$\! s\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! h} found by another way, without giving the meaning of $\varphi$ and $\psi$.\footnote{% Ueber die Integration der hydrodynamischen Gleichungen (Crelle's Journal Bd.\ 56, S.\ 1 [\hyperlink{Clebsch1}{1859}]). In this essay \mbox{C\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! b\hspace{0.036cm}} %{$\phantom{!}$\! s\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! h} starts from the first form of the \mbox{E\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! r} equation to perform the above transformation. On the one hand, the method applied by him is susceptible of a bigger simplification; on the other hand, one can give a procedure quite analogous to that used above to prove that $\displaystyle\frac{{\rm d} \varphi} {dt}= 0$ and $\displaystyle\frac{{\rm d} \psi} {dt}= 0.$\hspace{0.4cm} \mbox{C\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! b\hspace{0.036cm}} %{$\phantom{!}$\! s\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! h} showed (in Theorem 2) that the equations transformed by the introduction of $F, \varphi, \psi$ are the conditions for the variation of a quadruple integral to disappear; this is [somewhat] similar to what \mbox{C\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! b\hspace{0.036cm}} %{$\phantom{!}$\! s\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! h} did for steady-state flows in his paper\,: Ueber eine allgemeine Transformation der hydrodynamischen Gleichungen (Crelle's Journal Bd.\ 54, S.\ 301 [\hyperlink{Clebsch2}{1857}]). \hspace{0.4cm}In another form, this is the Theorem, I derived directly from the principle of the virtual velocities $\S.\,5$.\,eq.\ (\hyperlink{eq1}{1}).\hspace{0.4cm}One can easily show that from this result one arrives at the introduction of the functions $a$ for the case of the stationary motion in a simpler way than \mbox{C\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! b\hspace{0.036cm}} %{$\phantom{!}$\! s\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! h} does in the latter essays\,; for a lack of space, we are not giving a more accurate argumentation here. }\label{footnoteClebsch}\\[1cm] \centerline {\fett{$\S.\,12$}} \\[0.3cm] If one puts into the equations $(3)$ of $\S.\,6$, $t=0$, one finds\,: \begin{equation*} \tag*{[12.1]} \frac{d v_0}{dc} - \frac{d w_0}{db} = 2A, \,\,\,\, \frac{d w_0}{da} - \frac{d u_0}{dc} = 2B,\,\,\,\, \frac{d u_0}{db} - \frac{d v_0}{da} = 2 C \end{equation*} where $A,\,\,B,\,\,C$ are the rotational velocities with respect to the coordinates axes in the point $a,\,\,b,\,\,c$ at time $t=0$.\hspace{0.4cm}Quite similarly, at time $t$ we have\,: \begin{equation*} \tag*{(1), [12.2]} \frac{d v}{dz} - \frac{d w}{dy} = 2X, \,\, \frac{d w}{dx} - \frac{d u}{dz} = 2Y,\,\,\, \frac{du}{dy} - \frac{d v}{dz} = 2 Z \end{equation*} These are the equations cited at page 34 [now page~\pageref{pointer2}] found by \mbox{C\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! y,} where $X,\,Y,\,Z$ mean the rotational velocities with respect to the coordinates $x,\,\,y,\,\,z$ at time $t$ taken as axes.\hspace{0.4cm}If $A,\,B,\,C$ are zero overall, i.e.\ at the beginning of the motion, there are no rotating particles and thus $X,\,Y,\,Z$ will stay zero and hence, one can write \begin{equation*} \tag*{(2), [12.3]} u= \frac{dF}{dx}, \,\, v= \frac{dF}{dy}, \,\, w = \frac{dF}{dz}. \end{equation*} Restricting ourselves to liquid flows, by this substitution the density equation (6) of $\S.\,2.$ becomes\,: \begin{equation*} \tag*{(3), [12.4]} \frac{d^2 F}{dx^2} + \frac{d^2 F}{dy^2} + \frac{d^2 F}{dz^2} =0\,; \end{equation*} therefore, $F$ satisfies \mbox{L\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! p\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! e}'s differential equation for all fluid particles\,; for this reason the function $F$ has been designated by \mbox{H\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! z} as the \mbox{v\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! y} \mbox{p\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! a\hspace{0.036cm}} %{$\phantom{!}$\! l.}\hspace{0.4cm} From the equation~(3) derives an interesting analogy between the fluid motions in a simply connected space and the effects of magnetic masses\,: the velocities are equal and aligned to the forces exerted by a certain distribution of magnetic masses on the surface on a magnetic particle in the interior. \\ \indent \indent If the function $F$ is determined by~(3) by suitable boundary conditions on the surface, it is still necessary to determine the pressure $p$ in order to get a complete solution to the problem. The required equation for that is easily obtained, by using the values of $u, \, v, \, w$ from~(2) in~(4) of $\S.\,4$. Then, one obtains\,: \begin{equation*} \tag*{[12.5]} \left. \begin{aligned} \frac{d^2 F}{dt dx} + \frac{d^2 F}{ dx^2} \frac{dF}{dx }+ \frac{d^2 F}{ dx dy} \frac{dF}{dy } + \frac{d^2 F}{ dx dz} \frac{dF}{dz } - \frac{d V}{dx} + \frac{1}{\varrho} \frac{dp}{dz} = 0\\\\ \frac{d^2 F}{dt dy} + \frac{d^2 F}{ dx dy} \frac{dF}{dx }+ \frac{d^2 F}{ dy^2} \frac{dF}{dy } + \frac{d^2 F}{ dy dz} \frac{dF}{dz } - \frac{d V}{dy} + \frac{1}{\varrho} \frac{dp}{dy} = 0\\\\ \frac{d^2 F}{dt dz} + \frac{d^2 F}{ dz dx} \frac{dF}{dx }+ \frac{d^2 F}{ dz dy} \frac{dF}{dy } + \frac{d^2 F} {dz^2} \frac{dF}{dz } - \frac{d V}{dz} + \frac{1}{\varrho} \frac{dp}{dz}=0 \end{aligned} \qquad \right\} \end{equation*} Through multiplication with ${\rm d} x, {\rm d} y, {\rm d} z$ and summation, one gets\,: \begin{equation*} \tag*{[12.6]} {\rm d} \left(\frac{d F}{dt} \right) + \frac{1}{2} {\rm d}\left( \frac{dF}{dx}\right)^2 + \frac{1}{2} {\rm d} \left( \frac{dF}{dy}\right)^2 + \frac{1}{2} {\rm d} \left( \frac{dF}{dz}\right)^2 - {\rm d} \Omega = 0 \end{equation*} where $\Omega$ is the function defined on S.\,18 [eq.\,\ref{pointerOmega}, now on page~\pageref{pointerOmega}].\hspace{0.4cm} It follows herefrom by integration\,: \begin{equation*} \tag*{[12.7]} \frac{d F}{dt} +\frac{1}{2} \left[ \left( \frac{dF}{dx}\right)^2 + \left( \frac{dF}{dy}\right)^2 + \left( \frac{dF}{dz}\right)^2 \right] - \Omega =0 \end{equation*} where the additive integration constant, which is a function of $t$, can be included in $F$. \hspace{0.4cm} Apart from the unknown $p$ the known function $V$ is contained in $\Omega$, from which one can easily determine $p$ once $F$ is obtained. \\[0.5cm] \centerline{\fett{$\S.\,13.$}} \vspace{0.3cm} \indent\indent From the relations $(1)$ of $\S.\,12$, \begin{equation*} \tag*{(1), [13.1]} \frac{d v}{dz} - \frac{d w}{dy} = 2X, \,\,\frac{d w}{dx} - \frac{d u}{dz} =2Y,\,\,\, \frac{d u}{dy} - \frac{d v}{dz} = 2 Z \end{equation*} and from the density equation $(6)$ of $\S.\,2$ for liquid fluids \begin{equation*} \tag*{[13.2]} \frac{du}{dx} + \frac{dv}{dy} + \frac{d w}{dz} =0 \,, \end{equation*} $u,\,v,\,w$ {can be determined as functions of $X,\,Y,\,Z$. \hspace{0.4cm} Indeed, one finds easily from these equations\,: \begin{equation*} \tag*{[13.3]} \begin{aligned} \frac{d^2 u}{d x^2}\,\, + \frac{d^2 u}{d y^2}\, + \frac{d^2 u}{d z^2}\, = 2 \left( \frac{d Z}{dy}\,- \frac{d Y}{dz}\right)\\\\ \frac{d^2 v}{d x^2}\,\, + \frac{d^2 v}{d y^2}\, + \frac{d^2 v}{d z^2}\, = 2 \left( \frac{d X}{dz}\, - \frac{d Z}{dx}\right)\\\\ \frac{d^2 w}{d x^2} + \frac{d^2 w}{d y^2} + \frac{d^2 w}{d z^2}\, = 2 \left( \frac{d Y}{dx} - \frac{d X}{dy}\right) \end{aligned} \end{equation*} The integration of these partial differential equations follows from known theorems\,: $u,\,v,\,w$ appear as potential functions of fictitious, attracting masses, which are distributed through the fluid-filled space with the density, \begin{equation*} \tag*{[13.4]} -\frac{1}{2 \pi} \left( \frac{d Z}{dy} - \frac{d Y}{dz}\right), \,\, -\frac{1}{2 \pi} \left( \frac{d X}{dz} - \frac{d Z}{dx}\right),\,\, -\frac{1}{2 \pi} \left( \frac{d Y}{dz} - \frac{d X}{dy}\right) \,. \end{equation*} Denoting by $u_1,\, v_1, \, w_1$ the velocity components at point $x_1, \, y_1,\, z_1$, and by $r$ the distance of this point to $x,\,y\,z$, one has\,: \begin{equation*} \tag*{[13.5]} \begin{aligned} u_1 =\frac{1}{2 \pi} \iiint \left( \frac {dY}{dz} - \frac{dZ}{dy}\right) \frac{{\rm d}x {\rm d}y {\rm d}z}{r}\\\\ v_1 =\frac{1}{2 \pi} \iiint \left( \frac {dZ}{dx} - \frac{dX}{dz}\right) \frac{{\rm d}x {\rm d}y {\rm d}z}{r}\\\\ w_1 =\frac{1}{2 \pi} \iiint \left( \frac {dX}{dy} - \frac{dY}{dx}\right) \frac{{\rm d}x {\rm d}y {\rm d}z}{r} \end{aligned} \end{equation*} where the integrals have to be extended over all points of the continuous fluid.\hspace{0.4cm}These are not yet the most general values of $u_1, \, v_1, \, w_1$\,: since these values $u_1, \, v_1, \, w_1$ are determined, one can always add to them successively \begin{equation*} \notag \begin{aligned} \frac{dP_1}{dx_1},\,\,\,\frac{dP_1}{dy_1}, \,\,\,\frac{dP_1}{dz_1} \,, \end{aligned} \end{equation*} while equations~(1) are still satisfied as well as the density equation, provided that\,: \begin{equation*} \tag*{[13.6]} \frac{d^2 P}{dx^2} + \frac{d^2P}{dy^2} + \frac{d^2P}{dz^2} =0 \end{equation*} is assumed for all points of the fluid.\hspace{0.4cm}One can consider here $P$ as the potential function of attracting masses which are outside of the space filled with the fluid, and must be dermined so that the conditions for $u_1,\, v_1, \, w_1$ on the fluid's surface are satisfied.\\\\ \indent\indent The values found for $u_1,\, v_1, \, w_1$ in this way can be transformed through integration by parts.\hspace{0.4cm}Since\,: \begin{equation*} \tag*{[13.7]} r^2=\left(x -x_1 \right)^2 + \left(y - y_1 \right)^2+\left(z -z_1 \right)^2 \end{equation*} one finds\,: \begin{equation*} \tag*{[13.8]} \begin{aligned} \iiint \frac{dY}{dz} \frac{{\rm d}x\,{\rm d}y\, {\rm d}z}{r}= \iint Y \frac{{\rm d}x\, {\rm d}y}{r} + \iiint Y \frac {z-z_1}{r^3} {\rm d}x\, {\rm d}y\, {\rm d}z\\\\ \iiint \frac{dZ}{dy} \frac {{\rm d}x\, {\rm d}y\, {\rm d}z}{r}= \iint Z \frac{{\rm d}x\, {\rm d}z}{r} + \iiint Z \frac {y-y_1}{r^3} {\rm d}x\, {\rm d}y\, {\rm d}z \end{aligned} \end{equation*} and thus \begin{equation*}\tag*{(2), [13.9]} \left. \begin{aligned} u_1 &=\frac{dP_1}{dx_1} + \frac{1}{2\pi} \int \left( Y \cos \gamma - Z \cos \beta \right ) \frac{{\rm d} \omega}{r} + \frac{1}{2\pi} \iiint \left\{ Y(z-z_1) - Z (y - y_1) \right\}\frac{{\rm d}x\, {\rm d}y\, {\rm d}z }{r^3} \,\,\,\,\\\\ &\text{\qquad\qquad\qquad\qquad and in the same way\,:} \\\\ v_1 &=\frac{dP_1}{dy_1} + \frac{1}{2\pi} \int ( Z \cos \alpha - X \cos \gamma ) \frac{{\rm d} \omega}{r} + \frac{1}{2\pi} \iiint \left\{ Z(x-x_1) - X (z - z_1) \right\}\frac{{\rm d}x\, {\rm d}y\, {\rm d}z }{r^3}\,\,\,\, \\\\ w_1 &=\frac{dP_1}{dz_1} + \frac{1}{2\pi} \int ( X \cos \beta - X \cos \alpha ) \frac{{\rm d} \omega}{r} + \frac{1}{2\pi} \iiint \left\{ X(y-y_1) - Y (x - x_1) \right\} \frac{{\rm d}x\, {\rm d}y\, {\rm d}z }{r^3} \end{aligned} \right\} \end{equation*} where ${\rm d}\omega$ denotes a surface element and $\alpha,\beta, \gamma$ the angle formed by the normal to it with the coordinates axes.\\ \indent\indent These analytical formulas for the representation of $u_1,\,v_1,\,w_1$ in terms of $X,\,Y,\,Z$ allow for an interesting interpretation leading to a striking analogy between the effect of vortex filaments and that of electrical currents.\hspace{0.4cm}Namely, if we indicate the parts of $u_1,\,v_1,\,w_1$ which originate from the elements ${\rm d}x\, { \rm d}y\, {\rm d}z$ of the triple integrals with ${\rm d}u_1 { \rm d}v_1 {\rm d}w_1$ then one has\,: \begin{equation*} \tag*{[13.10]} \begin{aligned} {\rm d} u_1 = \frac{1}{2\pi} \left \{ Y(z - z_1, -Z(y-y_1)\right \} \frac{{\rm d}x\, {\rm d}y\, {\rm d}z }{r^3}\\\\ {\rm d} v_1 = \frac{1}{2\pi} \left \{ Z (x - x_1, -Z(z-z_1)\right \} \frac{{\rm d}x\, {\rm d}y\, {\rm d}z }{r^3}\\\\ {\rm d} w_1 = \frac{1}{2\pi} \left \{ X (y - y_1, -Y(x-x_1)\right \} \frac{{\rm d}x\, {\rm d}y\, {\rm d}z }{r^3} \end{aligned} \end{equation*} By known theorems, from these equations one derives the relations\,: \begin{align} \tag*{[13.11]} &(x - x_1)\, {\rm d} u_1 + (y - y_1) \,{\rm d} v_1 + (z - z_1)\, {\rm d} w_1=0, \\ \nonumber \\ &X {\rm d}u_1 + Y{\rm d} v_1 + Z {\rm d} w_1 =0,\hspace{6.8cm} \tag*{[13.12]}\\\nonumber\\ &{\rm d}u_1^2 + {\rm d}v_1^2 + {\rm d}w_1^2 = \left(\frac{{\rm d}x {\rm d}y {\rm d}z}{2 \pi r^3}\right)^2 \Big\{ (X^2 + Y^2 + Z^2) [(x-x_1)^2 + (y-y_1)^2 + (z - z_1)^2] \nonumber\\ &\quad\hspace{5cm} - [X(x-x_1)+ Y(y-y_1) + Z(z-z_{3})]^2 \Big\} \,. \tag*{[13.13]} \end{align} \noindent The first two equations show that the velocity with the components $ {\rm d}u_1, {\rm d}v_1, {\rm d}w_1$ \begin{equation*} \tag*{[13.14]} {\rm d} U_1 = \sqrt{{\rm d}u_1^2 + {\rm d}v_1^2 + {\rm d}w_1^2 } \end{equation*} is normal to the plane containing the point $x_1, y_1, z_1$ and the rotation axis of the fluid particles $x,\, y,\,z$.\,\,The previous equation can also be written\,: \begin{equation*} \tag*{[13.15]} {\rm d} U_1^2=\Big(\frac{{\rm d}x\, {\rm d}y\, {\rm d}z}{2 \pi r^2} \Delta \Big)^2 \left\{ 1-\Big[\frac{X}{\Delta} \frac{x-x_1}{r } +\frac{Y}{\Delta} \frac{y-y_1}{r } + \frac{Z}{\Delta} \frac{z-z_1}{r }\Big]^2 \right\} \end{equation*} where $\Delta$ means the rotational velocity.\hspace{0.4cm}Herefrom one finds\,: \begin{equation*} \tag*{[13.16]} {\rm d} U_1= \frac{{\rm d}x\,{\rm d}y\,{\rm d}z}{2 \pi r^2} \Delta \sin \varepsilon \end{equation*} where $\varepsilon$ indicates the angle between the rotation axis of the particle $x, \,y, \,z$ and the connecting line $r$ between this point and $x_1, y_1, z_1$.\hspace{0.4cm}Each rotating particle $x,\,y,\,z$ generates, then, into another particle of the fluid $x_1, y_1, z_1$ a velocity which is normal to the plane passing through the rotation axis of the particle $x,\,y,\,z$ and $x_1, y_1, z_1$. This velocity is directly proportional to the volume ${\rm d}x\, {\rm d}y\, {\rm d}z$ of the particle $x,\,y,\,z$, to its rotational velocity $\Delta$ and to the sinus of the angle $\varepsilon$ between the rotation axis of $x,\, y,\, z$ and the connecting line $r$ of both particles, is inversely proportional to the square of the distance $r$ between both particles.\\ \indent\indent However, according to A\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! p\hspace{0.036cm}} %{$\phantom{!}$\! \`e\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! e's Law, this is the same force that an electrical particle of intensity $\Delta$ located at $x,\,y,\,z$ in a current aligned parallely to the rotational axis, would exert on a little magnet located at $x_1,\,y_1,\,z_1$.\\ \indent\indent This highly remarkable analogy, whose discovery is due to \mbox{H\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! z,} is firstly of great importance for the theory of the vortex filaments in liquid fluids --- since equations~(2) are only applicable to such liquid fluids. Indeed, it allows to apply theorems developed in electrodynamics, with minor modifications, to hydrodynamics, making the visualisation significantly simpler. Furthermore, this analogy is also of a certain value for electrodynamics, since it allows to analyse the electrodynamical processes not based on the mutual elementary interactions between two particles --- as is it is usually done following \mbox{A\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! p\hspace{0.036cm}} %{$\phantom{!}$\! \`e\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! e's} procedure --- but by considering infinitely thin closed currents as a basis for the whole theory, in analogy with vortex filaments}. \\[1cm] \centerline{\fett{$\S.\,14.$}}\vspace{0.3cm} \indent\indent The principle of virtual velocities for the motion of liquid fluids\: \begin{equation*}\tag*{(1), [14.1]} \frac{d \delta x}{dx} + \frac{d \delta y}{dy} + \frac{d \delta z}{dz} = 0, \end{equation*} gives the next equation, following from (1) of $\S.\,4$\,: \begin{equation*} \tag*{[14.2]} \iint \Big[(\frac{d^2 x}{dt^2} -X) \delta x + (\frac{d^2 y}{dt^2} -Y) \delta y + (\frac{d^2 z}{dt^2} -Z) \delta z \Big] {\rm d}x\, {\rm d}y\,{\rm d}z =0 . \end{equation*} \noindent If one takes, instead of the virtual velocities, the actual velocities, one has to put, instead of $\delta x, \delta y, \delta z$\,: \begin{equation*} \tag*{[14.3]} {\rm d}x = \frac{dx}{dt}{{\rm d}t}, \,\,\, {\rm d}y=\frac{dy}{dt}{{\rm d}t}, \,\,\,{\rm d}z=\frac{dz}{dt}{{\rm d}t}, \end{equation*} and, since (1), turns into the continuity equation, when limiting oneself to liquid flow, one finds, by the usual assumptions about $X,Y,Z$\,: \begin{equation*} \tag*{[14.4]} \iiint \Big(\frac{d^2 x}{dt^2}\frac{d x}{dt} +\frac{d^2 y}{dt^2}\frac{d y}{dt} +\frac{d^2 z}{dt^2}\frac{d z}{dt}\Big ) {\rm d} x\,{\rm d} y\,{\rm d}z = \iiint \Big( \frac{dV}{dx} \frac{dx}{dt} + \frac{dV}{dy} \frac{dy}{dt} + \frac{dV}{dz}\frac{dz}{dt}\Big) {\rm d} x\,{\rm d} y\,{\rm d}z \end{equation*} From this equation, after integration in time, denoting by \begin{equation*} \tag*{[14.5]} K=\frac{1}{2}\iiint \Big\{\Big(\frac{dx}{dt}\Big)^2 + \Big(\frac{dy}{dt}\Big)^2 + \Big(\frac{dz}{dt}\Big)^2 \Big\}{\rm d} x\,{\rm d} y\,{\rm d}z \end{equation*} the living force \begin{equation*} \tag*{[14.6]} K = \text{{\em Const.}} + \iiint V {\rm d}x\,{\rm d}y\,{\rm d}z \end{equation*} wherein the constant is independent of $t$.\hspace{0.4cm}This equation of the living force has been presumably given for the first time by \mbox{L\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! j\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! u\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! e-D\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! t}.\footnote{Untersuchungen \"uber ein Problem der Hydrodynamik. Crelle's Journal, Bd. 58, S. 202, [\hyperlink{Dirichlet}{1860}].}\\ \indent\indent Consequently, the living force will be independent of time, provided that the integral $\iiint V {\rm d}x\,{\rm d}y\,{\rm d}z$ is constant in time, in other words, since $V$ for each point in space is independent of $t$.\hspace{0.4cm}The condition that the fluid always fills the same absolute space may be also expressed by the fact that the normal velocity components of the particles at the surface is zero.\hspace{0.4cm}One may show directly that, in this case, $K$ is constant. Namely, one has\,: \begin{equation*} \tag*{[14.7]} \frac{d K}{dt }= \iiint \Big(\frac{dV}{dx} u + \frac{dV}{dy}v + \frac{dV}{dz}w\Big) {\rm d } x\,\,{\rm d} y\,\,{\rm d}z \end{equation*} and because of \begin{equation*} \tag*{[14.8]} \begin{aligned} \iiint \frac{dV}{dx} u {\rm d}x\,{\rm d}y\,{\rm d}z &= \iint Vu{\rm d}y\,{\rm d}z - \iiint V \frac{du}{dx} {\rm d}x\,{\rm d}y\,{\rm d}z \\\ \iiint \frac{dV}{dy} v {\rm d}x\,{\rm d}y\,{\rm d}z &= \iint Vv{\rm d}z\,{\rm d}x - \iiint V \frac{dv}{dy} {\rm d}x\,{\rm d}y\,{\rm d}z\\\ \iiint \frac{dV}{dz} w {\rm d}x\,{\rm d}y\,{\rm d}z &= \iint Vw{\rm d}x\hspace{0.01cm}{\rm d}y - \iiint V \frac{dw}{dx}{\rm d}x\hspace{0.02cm}{\rm d}y\hspace{0.02cm}{\rm d}z \end{aligned} \end{equation*} \noindent one has\,: \begin{equation*} \tag*{[14.9]} \frac{d K}{dt} = \iint Vu {\rm d}y\,{\rm d}z+\iint Vv {\rm d}z\,{\rm d}x + \iint Vw{\rm d}x {\rm d}y - \iiint V\Big( \frac{du}{dx} + \frac{dv}{dy} +\frac{dw}{dz}\Big) {\rm d}x\, {\rm d}y \,{\rm d}z \end{equation*} The density equation cancels out the last integral and one finds that \begin{equation*} \tag*{[14.10]} \frac{d K}{dt} =\int V\,U_n\, {\rm d} \omega, \end{equation*} where $U_n$ denotes the normal velocity of a particle at the external surface, and where ${\rm d} \omega$ is the surface element.\hspace{0.4cm}If $U_n =0$ on the whole surface, we find indeed that $\displaystyle \frac{dK}{dt}=0$.\\\\ \indent\indent For every stationary motion the normal velocity component at the surface vanishes\,; the same happens when the fluid is surrounded by motionless rigid walls\,; this includes also the case when the liquid mass extends to infinity in all directions, since this amounts to having the flow contained in an infinitely large sphere.\hspace{0.4cm}In all these cases the living forces are constant in time.\\\\ \indent\indent When there is no rotational motion in the flow, according to $\S.\,12$, one can set\,: \begin{equation*} \tag*{[14.11]} u = \frac{dF}{dx}, \hspace{1cm} v = \frac{dF}{dy}, \hspace{1cm} w = \frac{dF}{dz} \end{equation*} so that we obtain \begin{equation*} \tag*{[14.12]} K= \frac{1}{2} \iiint \Big\{ \Big( \frac{dF}{dx} \Big)^2 + \Big( \frac{dF}{dy} \Big)^2 + \Big( \frac{dF}{dz} \Big)^2 \Big\} {\rm d}x\,\,{\rm d}y\,\,{\rm d}z \end{equation*} \indent\indent By a well-known theorem, one finds\,:% \deffootnotemark{\textsuperscript{[T.\thefootnotemark]}}\deffootnote{2em}{1.6em}{[T.\thefootnotemark]\enskip}% \footnote{The factor $F$ in front of the Laplacian in the right-most term was omitted by Hankel.} \begin{smallequation} \tag*{[14.13]} \iiint \Big\{ \Big(\frac{dF}{dx}\Big)^2 + \Big( \frac{dF}{dy}\Big)^2 + \Big( \frac{dF}{dz}\Big)^2 \Big\}{\rm d}x\,{\rm d}y\,{\rm d}z = -\int F\frac{dF}{dn} {\rm d} \omega - \iiint F \big( \frac{d^2 F}{d x^2} + \frac{d^2 F}{d y^2}+ \frac{d^2 F}{d z^2} \big) {\rm d}x\,{\rm d}y\,{\rm d}z \end{smallequation} But, by (3) of $\S.\,12$, the last integral vanishes, and thus one has\,: \begin{equation*} \tag*{[14.14]} K=- \int F\frac{dF}{dn} \, {\rm d} \omega \end{equation*} Furthermore, if the flow is in a stationary motion, then {\small $\displaystyle\frac{dF}{dn}$}, being the velocity component normal to the external surface, must vanish, and thus $K=0$ from which follows\,: \begin{equation*} \tag*{[14.15]} \frac{dF}{dx}=0, \hspace{1cm} \frac{dF}{dy} = 0, \hspace{1cm} \frac{dF}{dz} = 0 \end{equation*} so that there is no motion at all.\hspace{0.4cm} Thus, a motion driven by a velocity potential can never become stationary and conversely, a stationary motion will never have a velocity potential\;--- an interesting theorem discovered by \mbox{H\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! m\hspace{0.036cm}} %{$\phantom{!}$\! h\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! l\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! z.} \vspace{1.2cm} \\\ \centerline{C\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! r\hspace{0.036cm}} %{$\phantom{!}$\! e\hspace{0.036cm}} %{$\phantom{!}$\! c\hspace{0.036cm}} %{$\phantom{!}$\! t\hspace{0.036cm}} %{$\phantom{!}$\! i\hspace{0.036cm}} %{$\phantom{!}$\! o\hspace{0.036cm}} %{$\phantom{!}$\! n\hspace{0.036cm}} %{$\phantom{!}$\! s [already implemented] } \vspace{0.5cm} \noindent p.\,20.\,\, l.\,7 must be $\big(\frac{dx}{d \varrho_1} \big)^2$ instead of $\big(\frac{dx}{d \varrho_1} \big)^3$ \\ p.\,30.\,\, l.\,10 in front of $2r \cos \theta \cos \varphi \frac{d \theta}{dt} \frac{d \varphi}{d t}$ there must be a $+$ sign \vspace{0.8cm} \section{Letters exchanged for the prize assignment: judgements of Bernhard Riemann and Wilhelm Eduard Weber} \label{s:letters} \subsection*{{\bf Judgement on Hankel's manuscript by Bernhard Riemann}} Decision on the received manuscript on the mathematical prize question carrying the motto: \,\,{\em The more signs express relationships in nature, the more useful they are}:\footnote{This motto is in Latin in the \textit{Preisschrift}: ``Tanto utiliores sunt notae, quanto magis exprimunt rerum relationes''.} \indent\indent The manuscript gives commendable evidence of the Author's diligence, of his knowledge and ability in using the methods of computation recently developed by contemporary mathematicians.\hspace{0.4cm}This is particularly shown in $\S\,6$ of the manuscript, which contains an elegant method for building the equations of motion for a flow in a fully arbitrary coordinate system, from a point of view which is commonly referred to as Lagrangian. \hspace{0.1cm} However, when further developing the general laws of vortex motion, the Lagrangian approach is unnecessarily left aside, and, as a consequence, the various laws had to be found by other and quite independent means.\hspace{0.4cm} Also the relation between the equations of motion and the investigations of Clebsch, reported in $\S.\,14.\,15 $, is omitted by the Author.\hspace{0.4cm}Nonetheless, as his derivation actually begins from the Lagrangian equations, one may consider the requirement posed by the prize-question as fulfilled by this part of the manuscript (if one substitutes the wrong text of a given proof of a theorem used by the Author with the right one contained in his note and leaves out of consideration the mistakes due to rushing in $\S.\,9$).\footnote{The $\S\,9$ refers to the Latin manuscript.} \hspace{0.4cm} In the opinion of the referee, the imperfectness just evoked in handling this part of the question did not give any sufficient reason to deny the prize to the manuscript. However, the Author will have to consolidate this part of the manuscript by a reworking, after which the same would gain in shortness and uniformity. A more important reason against the assignment of the prize could be the incorrectnesses occuring in several places. These [incorrectnesses] do not get to the core of the argument, except [in] two [paragraphs] (\S 3 and \S 8) in which the obscurity of another writer may as well serve as an excuse for the Author. [These incorrectnesses] may still be passed over, if our highly esteemed Faculty would like to assign the prize to this manuscript, in view of all manner of good things it contains.\\ \begin{figure}[t] \includegraphics[width=\textwidth]{LettersRiemannWeber.pdf} \caption {Scanned image of the letters from Riemann and Weber, taken from the \textit{G\"ottingen University Archive}. The image is the result of merging two consecutive pages of the full text of the exchanged letters amongst the commission members for the prize assignment. Riemann's letter begins on the left page and ends halfway on the right page. Weber's letter follows after Riemann's letter.} \end{figure} \vskip-0.3cm \indent G\"ottingen, 7th Mai 1861 $\phantom{C}$ \hfill B.\ Riemann \quad \quad \vspace{0.7cm} \subsection*{\bf{Judgement on Hankel's manuscript by Wilhelm Eduard Weber}} With the Faculty approval, I have consulted Prof.\ Riemann, whose judgement I fully agree with, both for the launching of a pure mathematical prize and for the evaluation of the received manuscript. In any case, the work deserves praiseful recognition, and since it needs just a few corrections indicated by my colleague Riemann, thereby easy to do, in order to meet the task requirements, it seems to me that the prize assignment does not give rise to any concern. Nevertheless, the Author will have to consign again his revised work before it is going to print, according to the given suggestions. In my consideration, some incorrectnesses and hastinesses find a fair excuse, in the indeed sparsely proportionated time for such a task, given that only the autumn holiday could be used for the scope.\\ $\phantom{C}$ \hfill Wilhelm Weber \quad \quad \vspace{0.7cm} \section{Hankel's biographical notes and list of published papers} \label{s:HHpapers} \begin{figure}[t] \includegraphics[width=0.4\textwidth]{Hankel.jpg} \caption{Portrait of Hermann Hankel (from Wikimedia Commons).} \end{figure} \subsection*{\bf{Biographical notes about Hermann Hankel}} \deffootnotemark{\textsuperscript{[T.\thefootnotemark]}}\deffootnote{2em}{1.6em}{[T.\thefootnotemark]\enskip} Hermann Hankel was born in Halle, near Leipzig, on 14th February 1839. His father was Gottfried Wilhelm Hankel, a renowned physicist. Hermann Hankel was a brilliant student already in high school, with a particular interest in mathematics and its history. From 1857, he studied Mathematics at the Leipzig University under the mathematicians Scheibner, Moebius and Drobisch. Then, he continued his studies in G\"ottingen, where, arriving in April 1860, he could attend, among others, Riemann's lectures. In G\"ottingen he won in 1861 the extraordinary mathematical prize launched in June 1860 by the Faculty of G\"ottingen with an essay on the fluid motion theory to be elaborated in a Lagrangian framework. Also in 1861, he obtained his Doctor degree in Leipzig with the dissertation: ``\"Uber eine besondere Classe der symmetrischen Determinanten''.\footnote{``On a particular class of symmetric determinants''.} Then, in the autumn of the same year, he went to Berlin, where he could attend courses of Weierstrass and Kronecker. In 1862 he returned to Leipzig and, in 1863, at the same place, he habilitated as a \textit{Privatdozent} with a thesis on the Euler integrals with an unlimited variability of the argument. The writing of the habilitation thesis was probably firstly induced by the lectures of Riemann about functions of complex variables. In the spring 1867, he became extraordinary Professor at Leipzig University and, in the same year, ordinary Professor in Erlangen, then, in T\"ubingen in 1869. He was married to Marie Dippe, who much later became a very important Esperantist. During his life, Hankel was advisor for doctoral dissertations in mechanics, real functions and geometry. He died prematurely on 29th August 1873. Hermann Hankel is known for his Hankel functions, a type of cylindrical functions, Hankel transforms, integral transformations whose kernels are Bessel functions of the first kind, and Hankel matrices, with constant skew diagonals. Hankel was the first to recognise the significance of Grassmann's extension theory (``Ausdehnungslehre''). Hankel had a passion for research in history of mathematics and published meaningful writings also in this domain (his inaugural lesson in T\"ubingen was about the development of Mathematics in the last centuries). Curiously, his prized work on the fluid-dynamic theory in Lagrangian coordinates written as a student, is little known.\footnote{For Hankel's biography see: Cantor, \hyperlink{Cantor}{1879}; Crowe, \hyperlink{Crowe}{2008} Monna, \hyperlink{Monna}{1973}, von Zahn, \hyperlink{Zahn}{1874}.} \vspace{0.7cm} \subsection*{\bf{List of papers of Hermann Hankel}} \begin{enumerate} \item[1)] Hankel, Hermann. 1861. \textit {Zur allgemeinen Theorie der Bewegung der Fl\"ussigkeiten.} Eine von der philosophischen Facult\"at der Georgia Augusta am 4.\ Juni 1861 gekr\"onte Preisschrift, G\"ottingen. \HD{http://babel.hathitrust.org/cgi/pt?id=mdp.39015035826760;view=1up;seq=5}{Druck der Dieterichschen Univ.-Buchdruckerei. W.FR.Kaestner, G\"ottingen.} \item[2)] Hankel, Hermann. 1861. \textit {\"Uber eine besondere Classe der symmetrischen Determinanten.} \HD{https://books.google.fr/books/about/Ueber_eine_besondere_Classe_der_symmetri.html?id=vHZaAAAAcAAJ&redir_esc=y}{Inaugural-Dissertation zur Erlangung der philosophischen Doktorw\"urde an der Universit\"at Leipzig von Hermann Hankel.} \item[3)] Hankel, Hermann. 1862. \"Uber die Transformation von Reihen in Kettenbr\"uche. \HD{http://gdz.sub.uni-goettingen.de/pdfcache/PPN599415665_0007/PPN599415665_0007___LOG_0026.pdf} {\textit{Zeitschrift f\"ur Mathematik und Physik}, {\bf 7}, 338--343.} Also in \textit {Berichte \"uber die Verhandlungen der k\"oniglich s\"achsichen Gesellschaft der Wissenschaften zu Leipzig}, mathematisch-physische Classe {\bf 14}, 17-22, 1862. Verlag der S\"achsischen Akademie der Wissenschaften zu Leipzig. \item[4)] Hankel, Hermann, (signed as Hl.). 1863. Aufsatz \"uber \textit {Ein Beitrag zu den Untersuchungen \"uber die Bewegung eines fl\"ussigen gleichartigen Ellipsoides} by B.Riemann. \HD{https://play.google.com/books/reader?id=zt0EAAAAQAAJ&printsec=frontcover&output=reader&hl=en&pg=GBS.PA50}{\textit{Die Fortschritte der Physik} im Jahre 1861, {\bf 17}, 50--57.} \item[5)] Hankel, Hermann, (signed as Hl.). 1863. Aufsatz \"uber \textit {Zur allgemeinen Theorie der Bewegung der Fl\"ussigkeiten. Eine von der philosophischen Facult\"at der Georgia Augusta am 4.\ Juni 1861 gekr\"onte Preisschrift, G\"ottingen} by H.\ Hankel. \HD{https://play.google.com/books/reader?id=zt0EAAAAQAAJ&printsec=frontcover&output=reader&hl=en&pg=GBS.PA57} {\textit{Die Fortschritte der Physik} im Jahre 1861, {\bf 17}, 57--61.} \item[6)] Hankel, Hermann, (signed as Hl.). 1863. Aufsatz \"uber \textit {D\'eveloppements relatifs au \S 3 de recherches de Dirichlet sur un probl\`eme d'hydrodynamique} by F.\ Brioschi. \HD{https://play.google.com/books/reader?id=zt0EAAAAQAAJ&printsec=frontcover&output=reader&hl=en&pg=GBS.PA61}{\textit{Die Fortschritte der Physik} im Jahre 1861, {\bf 17}, 61--62.} \item[7)] Hankel, Hermann. 1863. \textit {Die Euler'schen integrale bei unbeschr\"ankter variabilit\"at des Argumentes}: \HD{https://ia902205.us.archive.org/28/items/dieeulerschenin00hankgoog/dieeulerschenin00hankgoog.pdf}{zur Habilitation in der Philosophischen Facult\"at der Universit\"at, Leipzig, Voss.} An extrait is published in \HD{http://gdz.sub.uni-goettingen.de/pdfcache/PPN599415665_0009/PPN599415665_0009___LOG_0006.pdf}{\textit{Zeitschrift f\"ur Mathematik und Physik}, 1864, {\bf 9}, 1--21.} \item[8)] Hankel, Hermann. 1864. Die Zerlegung algebraischer Functionen in Partialbr\"uche nach den Prinzipien der complexen Functionentheorie. \HD{http://gdz.sub.uni-goettingen.de/dms/load/img/?PID=PPN599415665_0009\%7CLOG_0025} {\textit{Zeitschrift f\"ur Mathematik und Physik} {\bf 9}, 425--433.} \item[9)] Hankel, Hermann. 1864. Mathematische Bestimmung des Horopters. \HD{http://gallica.bnf.fr/ark:/12148/bpt6k15207b/f600.image.r=}{\textit{Annalen der Physik und Chemie}, {\bf 122}, 575--588.} \item[10)] Hankel, Hermann. 1864. \textit{\"Uber die Vieldeutigkeit der Quadratur und rectification algebraischer Curven}. \HD{https://archive.org/stream/bub_gb_M6dWcNSvxJYC\#page/n0/mode/2up}{Eine Gratulationsschrift zur Feier des f\"unfzigjaehrigen Doctorjubilaeums des Herren August Ferdinand Moebius am 11 Dezember 1864.} \item[11)] Hankel, Hermann. 1867. Ein Beitrag zur Beurteilung der Naturwissenschaft des griechischen Altertum. \HD{https://babel.hathitrust.org/cgi/pt?id=hvd.hw2918;view=1up;seq=436}{\textit {Deutsche Vierteljahresschrift} {\bf 4}, 120-155.} \item[12)] Hankel, Hermann. 1867. \textit {Vorlesungen \"uber die complexen Zahlen und ihre Functionen} in \HD{https://books.google.it/books?id=MkttAAAAMAAJ&printsec=frontcover&hl=it&source=gbs_ge_summary_r&cad=0\#v=onepage&q&f=false}{zwei Theilen, Voss, Leipzig.} \item[13)] Hankel, Hermann. 1867. Darstellung symmetrischer Functionen durch die Potenzsummen. \HD{https://www.digizeitschriften.de/download/PPN243919689_0067/PPN243919689_0067___log8.pdf}{\textit {Journal f\"ur die reine und angewandte Physik}, {\bf 67}, 90--94.} \item[14)] Hankel, Hermann. 1868. Die Astrologie um 1600 mit besonderer R\"ucksicht auf das Verhaeltnis Keppler's und Wallenstein's. \HD{http://reader.digitale-sammlungen.de/de/fs1/object/display/bsb10612800_00295.html?contextType=scan&contextSort=score\%2Cdescending&contextRows=10&context=hankel}{\textit {Westermann Monatshefte}, {\bf 25}, 281--294. } \item[15)] Hankel, Hermann. 1869. \textit {Die Entwickelung der Mathematik in den letzten Jahrhunderte}. \HD{https://archive.org/details/bub_gb_TE3kAAAAMAAJ}{Ein Vortrag beim Eintritt in den akademischen Senat der Universita\"at T\"ubingen am 29. April 1869 gehalten, Fuessche, T\"ubingen. } \item[16)] Hankel, Hermann. 1869. Die Entdeckung der Gravitation -- und Pascal - Ein literarisches Bericht von Dr.\ Hermann Hankel. \HD{http://gdz.sub.uni-goettingen.de/pdfcache/PPN599415665_0014/PPN599415665_0014___LOG_0014.pdf}{\textit{Zeitschrift f\"ur Mathematik und Physik}, {\bf 14}, 165--173.} \item[17)] Hankel, Hermann. 1869. Beweis eines Hilfsatzes in der Theorie der bestimmten Integrale. \HD{http://gdz.sub.uni-goettingen.de/dms/load/img/?PID=PPN599415665_0014\%7CLOG_0030&physid=PHYS_0440}{\textit {Zeitschrift f\"ur Mathematik und Physik}, {\bf 14}, 436--437.} \item[18)] Hankel, Hermann. 1869. Die Cylinderfunctionen erster und zweiter Art. \HD{http://gdz.sub.uni-goettingen.de/pdfcache/PPN235181684_0001/PPN235181684_0001___LOG_0041.pdf}{\textit {Mathematische Annalen}, {\bf 1}, 467--501.} \item[19)] Hankel, Hermann. 1870. Die Entdeckung der Gravitation durch Newton. \HD{http://reader.digitale-sammlungen.de/de/fs1/object/goToPage/bsb10612803.html?pageNo=496}{\textit {Westermann Monatshefte}, {\bf 27}, 482--493.} \item[20)] Hankel, Hermann. 1872. Intorno al volume intitolato: \textit{Geschichte der mathematischen Wissenschaften. 1.\ Theil. Von den \"altesten Zeiten bis Ende des 16.\ Jahrhunderts} of H.Suter. Relazione del dottor Ermanno Hankel. \HD{http://gdz.sub.uni-goettingen.de/pdfcache/PPN599471603_0005/PPN599471603_0005___LOG_0022.pdf}{\textit {Bullettino di bibliografia e di storia delle scienze matematiche e fisiche}, {\bf 5}, 297--300.} \item[21)] Hankel, Hermann. 1874. \textit{Zur Geschichte der Mathematik in Althertum und im Mittelalter}, (published \textit{post-mortem}), \HD{http://gallica.bnf.fr/ark:/12148/bpt6k82883t}{Druck und Verlag von B.G.\ Teubner.} \item[22)] Hankel, Hermann. 1875. \textit{Die Elemente der projectivischen Geometrie in synthetischer Behandlung.} \HD{https://archive.org/stream/dieelementederp00hankgoog\#page/n7/mode/2up}{Vorlesungen von Dr. Hermann Hankel, Teubner, Leipzig.} \item[23)] Hankel, Hermann. 1875. Bestimmte Integrale mit Cylinderfunctionen. \HD{http://gdz.sub.uni-goettingen.de/pdfcache/PPN235181684_0008/PPN235181684_0008___LOG_0037.pdf}{\textit {Mathematische Annalen}, { \bf 8}, 453--470. } \item[24)] Hankel, Hermann. 1875. Die Fourier'schen Reihen und Integrale f\"ur Cylinderfunctionen. \HD{http://www.digizeitschriften.de/download/PPN235181684_0008/PPN235181684_0008___log38.pdf}{\textit {Mathematische Annalen}, {\bf 8}, 471--494.} \item[25)] Hankel, Hermann. 1882 (1870). Untersuchungen \"uber die unendlich oft oszillierenden und unstetigen Funktionen. Abdruck aus dem Gratulationsprogramm der T\"ubinger Universit\"at vom 6. M\"arz 1870. \HD{http://www.digizeitschriften.de/download/PPN235181684_0020/PPN235181684_0020___log14.pdf}{\textit { Mathematische Annalen}, {\bf 20}, 63--112.} \item[26)] Hankel, Hermann. 1818--1881, Lagrange Lehrsatz. In \HD{http://gdz.sub.uni-goettingen.de/pdfcache/PPN358976863/PPN358976863___LOG_0132.pdf} {\textit {Allgemeine Encyclop\"adie der Wissenschaften und K\"unste}, J.S.\ Ersch, J.G.\ Grube, 353--367.} \item[27)] Hankel, Hermann. 1818--1881. Gravitation. In \HD{http://gdz.sub.uni-goettingen.de/pdfcache/PPN358787696/PPN358787696___LOG_0301.pdf} {\textit {Allgemeine Encyclop\"adie der Wissenschaften und K\"unste}, J.S.\ Ersch, J.G.\ Grube, 313--355.} \item[28)] Hankel, Hermann. 1818--1881. Grenze. In \HD{http://gdz.sub.uni-goettingen.de/pdfcache/PPN358976863/PPN358976863___LOG_0132.pdf} {\textit {Allgemeine Encyclop\"adie der Wissenschaften und K\"unste} J.S.\ Ersch, J.G.\ Grube, 185--211.} \end{enumerate} \vspace{0.5cm} \noindent {\it Acknowledgements.} We are grateful to Uriel Frisch and tho the referees for useful remarks. We thank the Observatoire de la C\^ote d'Azur and the Laboratoire J.-L.~Lagrange for their hospitality. CR is supported by the DFG through the SFB-Transregio TRR33 ``The Dark Universe''.
{ "timestamp": "2017-09-19T02:14:27", "yymm": "1707", "arxiv_id": "1707.01883", "language": "en", "url": "https://arxiv.org/abs/1707.01883", "abstract": "The present is a companion paper to \"A contemporary look at Hermann Hankel's 1861 pioneering work on Lagrangian fluid dynamics\" by Frisch, Grimberg and Villone (2017). Here we present the English translation of the 1861 prize manuscript from Göttingen University \"Zur allgemeinen Theorie der Bewegung der Flüssigkeiten\" (On the general theory of the motion of the fluids) of Hermann Hankel (1839-1873), which was originally submitted in Latin and then translated into German by the Author for publication. We also provide the English translation of two important reports on the manuscript, one written by Bernhard Riemann and the other by Wilhelm Eduard Weber, during the assessment process for the prize. Finally we give a short biography of Hermann Hankel with his complete bibliography.", "subjects": "History and Overview (math.HO); Analysis of PDEs (math.AP); History and Philosophy of Physics (physics.hist-ph)", "title": "Hermann Hankel's \"On the general theory of motion of fluids\", an essay including an English translation of the complete Preisschrift from 1861", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9744347857055181, "lm_q2_score": 0.8221891392358015, "lm_q1q2_score": 0.8011696977006426 }
https://arxiv.org/abs/0805.4174
On a generalization of Christoffel words: epichristoffel words
Sturmian sequences are well-known as the ones having minimal complexity over a 2-letter alphabet. They are also the balanced sequences over a 2-letter alphabet and the sequences describing discrete lines. They are famous and have been extensively studied since the 18th century. One of the {extensions} of these sequences over a $k$-letter alphabet, with $k\geq 3$, are the episturmian sequences, which generalizes a construction of Sturmian sequences using the palindromic closure operation. There exists a finite version of the Sturmian sequences called the Christoffel words. They are known since the works of Christoffel and have interested many mathematicians. In this paper, we introduce a generalization of Christoffel words for an alphabet with 3 letters or more, using the episturmian morphisms. We call them the {\it epichristoffel words}. We define this new class of finite words and show how some of the properties of the Christoffel words can be generalized naturally or not for this class.
\section{Introduction} As far as we know, Sturmian sequences first appeared in the literature at the 18th century in the precursory works of the astronomer Bernoulli \cite{jb1772}. They later appeared in the 19th century in Christoffel \cite{ebc1875} and Markov \cite{am1882} works. The first deep study of these sequences is given in \cite{mh1938,mh1940} where the name {\it Sturmian sequence} appears for the first time. At the end of the 20th century and more recently, many mathematicians have been interested in those sequences, for instance \cite{ch1973,emc1974,kbs1976,tcb1993,bpr1994,gz1995,adl1997,jb2002}. Recent books also show this interest \cite{ml2002,pf2002,as2003,blrs2008} as well as a recent survey \cite{jb2007}. In this wide literature, we find different characterizations of the Sturmian sequences. In particular, they are the sequences over a 2-letter alphabet having the minimal complexity, they also are the balanced sequences over a 2-letter alphabet and they code discrete lines. These different characterizations show how the Sturmian sequences occur in different fields as number theory \cite{rm19852,rjs1991,rt2000,rt2001,bv2003,rjs2004,go2005}, discrete geometry, crystallography \cite{bt1986} and symbolic dynamics \cite{mh1938,mh1940,gah1944,mq1987}. Since the end of the 20th century, numerous generalizations of Sturmian sequences have been introduced for an alphabet with more than 2 letters. Among them, one natural generalization is called the {\it episturmian sequences} and is using the palindromic closure property of Sturmian sequences \cite{adl19972}. The first construction of episturmian sequences is due to \cite{djp2001}. Previously the first introduction and study of an episturmian sequence was that of the Tribonacci word \cite{gr1982} and an important class of episturmian sequences, now called the Arnoux-Rauzy sequences, had been considered in \cite{ar1991,rz2000}. More recently the whole class was extensively studied, for instance in \cite{jv2000,rz2000,djp2001,jp2002,jp2004,jj2005,ag2005,pv2007,gr20072,ag2006,bdldlz2008,gjp2007,glr2008}. For surveys about episturmian sequences, see for instance \cite{jb2007,gj2009}. The finite version of Sturmian sequences, called {\it Christoffel words}, has been also well studied \cite{ebc1875,ml2002,br2006,bdlr2007,kr2007}. It is known that any finite standard Sturmian word, that is the words obtained by standard Sturmian morphisms to a letter, is conjugate to a Christoffel word. A Christoffel word is then the smallest word, with respect to the lexicographic order, in the conjugacy class of a finite standard Sturmian word. Finite factors of the episturmian sequences appeared for instance in \cite{gjp2007}. The class of standard episturmian words is naturally defined as the set of finite words obtained by standard episturmian morphisms to letter, but no generalization of the Christoffel words have been introduced yet. In this paper, we introduce such a generalization that we naturally call the {\it epichristoffel words}. Note that it naturally appears that for each standard episturmian word, there exists a conjugate which is an epichristoffel word, and conversely. The paper is organized as follows. We first recall some basic definitions of combinatorics on words and we establish the notation used in this paper. We recall the definitions and some properties of the Sturmian sequences, the Christoffel words and the episturmian sequences. Then we introduce our new class of finite words: the {\it epichristoffel ones}. We prove how some of the properties of the Christoffel words can be generalized for an alphabet with more than $2$ letters. We then describe an algorithm which determines if a given $k$-tuple describes the occurrence numbers of letters in an epichristoffel word or not. If so, we show how to construct it. Finally, we prove the next theorem, which is a generalization of a result for Christoffel words \cite{dldl2006}, that characterizes epichristoffel conjugates. \\ \noindent {\bf Theorem } {\it Let $w$ be a finite primitive word different from a letter. Then the conjugates of $w$ are all factors of the same episturmian sequence if and only if $w$ is conjugate to an epichristoffel word. } \section{Definitions and notation} Throughout this paper, $\mathcal{A}$ denotes a finite alphabet containing $k$ letters $a_0, a_1, \ldots, a_{k-1}$. A {\it finite word} is an element of the free monoid $\mathcal{A} ^*$. If $w=w[0]w[1]\cdots w[n-1]$, with $w[i] \in \mathcal{A}$, then $w$ is said to be a finite word of {\it length} $n$ and we write $|w|=n$. By convention, the {\it empty word} is denoted $\varepsilon$ and its length is 0. We define $\mathcal{A} ^\omega$ the set of right infinite words, also called {\it sequences}, over the alphabet $\mathcal{A}$ and then, $\mathcal{A} ^\infty=\mathcal{A}^* \cup \mathcal{A}^\omega$ is the set of finite and infinite words. The number of occurrences of the letter $a_i$ in $w$ is denoted $|w|_{a_i}$. The {\it reversal} of the word $w=w[0]w[1]\cdots w[n-1]$ is $\widetilde{w}=w[n-1]w[n-2] \cdots w[0]$ and if $\widetilde{w}=w$, then $w$ is said to be a {\it palindrome}. A finite word $f$ is a {\it factor} of $w \in \mathcal{A}^\infty$ if $w=pfs$ for some $p \in \mathcal{A}^*, s \in \mathcal{A}^\infty$. If $p=\varepsilon$ (resp. $s=\varepsilon$), $f$ is called a {\it prefix} (resp. a {\it suffix}) of $w$. Let $u\in \mathcal{A}^*$ and $n\in \mathbb N$. We denote by $u^n$ the word $u$ repeated $n$ times and we called it a {\it $n$-th power word}. A factor $\alpha^k$ of the word $w$, with $\alpha \in \mathcal{A}$ and $k \in \mathbb N$ locally maximum, is called a {\it block of $\alpha$ of length $k$ in $w$}. Let $u, v$ be two palindromes, then $u$ is a {\it central factor} of $v$ if $v=wu\tilde w$ for some $w \in \mathcal{A}^*$. The {\it right palindromic closure} of $w \in \mathcal{A}^*$ is the shortest palindrome $u=w^{(+)}$ having $w$ as prefix. The {\it set of factors} of $w \in \mathcal{A}^\omega$ is denoted $F(w)$ and $F_n(w)=F(w)\cap \mathcal{A} ^n$ is the set of all factors of $w$ of length $n \in \mathbb N$. The complexity function is given by $P(n)=|F_n(w)|$ and is the number of distinct factors of $w$ of length $n \in \mathbb N$. Two words $w$ and $w'$ are said {\it equivalent} if they have the same set of factors: $F(w)=F(w')$. The {\it conjugacy class} $[w]$ of $w \in \mathcal{A} ^n$ is the set of all words $w[i]w[i+1]\cdots w[n-1]w[0]\cdots w[i-1]$, for $0 \leq i \leq n-1$. If $w$ is not the power of a shorter word, then $w$ is said to be {\it primitive} and has exactly $n$ conjugates. If $w$ is the smallest of its conjugacy class, relatively to some lexicographic order, then $w$ is called {\it a Lyndon word}. Let $w$ be an infinite word, then a factor $f$ of $w$ is {\it right} (resp. {\it left}) {\it special} in $w$ if there exist $a,b \in \mathcal{A}$, $a \neq b$, such that $fa, fb \in F(w)$ (resp. $af, bf \in F(w)$). A word $w$ over $\mathcal{A}$ is {\it balanced} if for all factors $u$ and $v$ of $w$ having the same length, for all letters $a \in \mathcal{A}$, one has $$\big ||u|_a-|v|_a\big |\leq 1.$$ If $w=pus \in \mathcal{A}^\infty$, with $p,u \in \mathcal{A}^*$ and $s \in \mathcal{A}^\infty$, then $p^{-1}w$ denotes the word $us$. Similarly, $ws^{-1}$ denotes the word $pu$. An integer $p\in \mathbb N$ is a {\it period} of the word $w=w[0]w[1]\cdots w[n-1]\in \mathcal{A}^*$ if $w[i]=w[i+p]$ for $0\leq i < n-p$. When $p=0$, the period is {\it trivial}. If $p$ is the smallest non trivial period of $w$, then the {\it fractionnary root} of $w$ is defined as the prefix $z_w$ of $w$ of length $p$. An infinite word $w\in \mathcal{A}^\omega$ is {\it periodic} (resp. {\it ultimately periodic}) if it can be written as $w=u^\omega$ (resp. $w=vu^\omega$), for some $u,v \in \mathcal{A}^*$. If $w$ is not ultimately periodic, then it is {\it aperiodic}. A {\it morphism} $f$ from $\mathcal{A}^*$ to $\mathcal{A}^*$ is a mapping from $\mathcal{A}^*$ to $\mathcal{A}^*$ such that for all words $u,v\in\mathcal{A}^*$, $f(uv)=f(u)f(v)$. A morphism extends naturally on infinite words. \section{Sturmian, Christoffel and episturmian words} Before introducing our generalization of Christoffel words, inspired by the definition of episturmian sequences, let us recall the definition of these well-known families and some of their properties. \subsection{Sturmian words and morphisms} One of the classical definitions of Sturmian sequences is the one given by Morse and Hedlund \cite{mh1940}: \\ \noindent {\bf Definition } Let $\rho$, called the intercept, and $\alpha$, called the slope, be two real numbers with $\alpha$ irrational such that $0\leq \alpha< 1$. For $n\geq 0$, let \begin{center} $s[n] = \left \{ \begin{tabular}{l} $a$ \textnormal{if} $\lfloor \alpha(n+1) +\rho \rfloor = \lfloor \alpha n +\rho \rfloor$,\\ $b$ \textnormal {otherwise}, \end{tabular} \right . $ \\ $s'[n] = \left \{ \begin{tabular}{l} $a$ \textnormal{if} $\lceil \alpha(n+1) +\rho \rceil = \lceil \alpha n +\rho \rceil$,\\ $b$ \textnormal {otherwise}. \end{tabular} \right . $\\ \end{center} Then the sequences $$s_{\alpha,\rho}=s[0] s[1]s[2]\cdots \quad \quad \textnormal{ and } \quad \quad s'_{\alpha, \rho}=s'[0]s'[1]s'[2]\cdots$$ are {\it Sturmian} and conversely, a Sturmian sequence can be written as $s_{\alpha, \rho}$ or $s'_{\alpha,\rho}$ for $\alpha$ irrational and $\rho \in \mathbb{R}$. \\ Sturmian sequences have several characterizations. For more details about this class of words, we refer the reader to the section in \cite{ml2002} devoted to Sturmian sequences.\\ \noindent {\bf Proposition } \cite{ch1973} A sequence $s$ is Sturmian if and only if for all $n \in \mathbb N$, $P(n)=n+1$.\\ \newpage \begin{theo}{\rm \cite{ml2002}} Let $s$ be a sequence. The following assertions are equivalent: \vspace{-0.2cm} \begin{itemize} \item [\rm i)] $s$ is Sturmian; \item [\rm ii)] $s$ is balanced and aperiodic. \end{itemize} \end{theo} \begin{defn}\cite{ml2002} A {\it morphism} $f$ is {\it Sturmian} if $f(s)$ is Sturmian for all Sturmian sequences~$s$. \end{defn} \subsection{Christoffel words} In discrete geometry, Christoffel words are defined as the discretization of a line having a rational slope, as introduced in \cite{bl1993}. In symbolic dynamics, they are defined by exchange of intervals \cite{mh1940} as follows.\\ \noindent {\bf Definition } Let $p$ and $q$ be positive relatively prime integers and $n=p+q$. Given an ordered $2$-letter alphabet $\{a < b\}$, the {\it Christoffel word w of slope $p/q$} over this alphabet is defined as $w=w[0]w[1]\cdots w[n-1]$, with $$w[i] = \left \{ \begin{tabular}{l} $a$ \textnormal{if} $ip \mod n > (i-1)p \mod n,$ \\ $b$ \textnormal{if} $ip \mod n < (i-1)p \mod n,$ \end{tabular} \right . $$ for $0\leq i\leq n-1$, where $k \mod n$ denotes the remainder of the Euclidean division of $k$ by $n$.\\ Notice that since $p$ and $q$ are relatively prime, a Christoffel word is always primitive. {Other important properties of Christoffel words will be recalled just before their generalizations in Section 4.} \subsection{Episturmian sequences and morphisms} One of the possible generalizations of Sturmian sequences for an alphabet with $3$ letters or more is the set of episturmian sequences. Let us first recall the definition of standard episturmian sequences as introduced initially by Droubay, Justin and Pirillo. \begin{defn} \cite{djp2001} \label{def1epis} A sequence $s$ is {\it standard episturmian} if it satisfies one of the following equivalent conditions. \vspace{-0.2cm} \begin{itemize} \item [\rm i)] For every prefix $u$ of $s$, $u^{(+)}$ is also a prefix of $s$. \item [\rm ii)] Every leftmost occurrence of a palindrome in $s$ is a central factor of a palindromic prefix of $s$. \item [\rm iii)] There exist a sequence $u_0=\varepsilon, u_1, u_2, \ldots$ of palindromes and a sequence $\Delta(s)=x[0]x[1]\cdots$, with $x[i] \in \mathcal{A}$, such that $u_n$ defined by $u_{n+1}=(u_nx[n])^{(+)}$, with $n \geq 0$, is a prefix of $s$. \end{itemize} \end{defn} \begin{defn} \cite{djp2001} \label{episst} A sequence $t$ is {\it episturmian} if $F(t)=F(s)$ for a standard episturmian sequence $s$. \end{defn} An equivalent definition is that a sequence $s \in A^\omega$ is {\it episturmian} if its set of factors is closed under reversal and $s$ has at most one right (or equivalently left) special factor for each length. \begin{nota} \textnormal{\cite{jj2005}} Let $w=w[0]w[1]\cdots w[n-1]$, with $w[i] \in \mathcal{A}$, and $u_0=\varepsilon$, \ldots, $u_{n}=(u_{n-1}w[n-1])^{(+)}$, the palindromic prefixes of $u_{n}$. Then \textnormal{Pal}$(w)$\index{$\textnormal{Pal}(w)$} denotes the word $u_{n}$. \end{nota} In Definition \ref{def1epis}, $\Delta(s)$ is called the {\it directive sequence} of the standard episturmian sequence~$s$. Since $\Delta(s)$ is the limit of its prefixes and $s$ is the limit of the $u_n$, it is natural to write $s=\hbox{\rm Pal}(\Delta(s))$. Let us recall from \cite {jj2005} a useful property of the operator $\hbox{\rm Pal}$. \begin{lem} \label{lemjj} {\rm \cite{jj2005}} Let $x \in \mathcal{A}$, $w \in \mathcal{A}^*$. If $|w|_x=0$, then $\hbox{\rm Pal}(wx)=\hbox{\rm Pal}(w)x\hbox{\rm Pal}(w)$. Otherwise, we write $w=w_1xw_2$ with $|w_2|_x=0$. The longest palindromic prefix of $\hbox{\rm Pal}(w)$ which is followed by $x$ in $\hbox{\rm Pal}(w)$ is $\hbox{\rm Pal}(w_1)$. Thus, $\hbox{\rm Pal}(wx)=\hbox{\rm Pal}(w)\hbox{\rm Pal}(w_1)^{-1}\hbox{\rm Pal}(w)$. \end{lem} \begin{defn} For $a, b \in \mathcal{A}$, we define the following endomorphisms of $\mathcal{A} ^*$: \vspace{-0.2cm} \begin{itemize} \item [\rm i)] $\psi_a(a)=\overline{\psi }_a(a)=a$; \item [\rm ii)] $\psi_a(x)=ax$, if $x \in \mathcal{A} \setminus \{a\}$; \item [\rm iii)] $\overline{\psi }_a(x)=xa$, if $x \in \mathcal{A} \setminus \{a\}$; \item [\rm iv)] $\theta_{ab}(a)=b$ , $\theta_{ab}(b)=a$, $\theta_{ab}(x)=x$, $x \in \mathcal{A} \setminus \{a,b\}$. \end{itemize} \end{defn} The endomorphisms $\psi$ and $\overline \psi$ can be naturally extended to a finite word $w=w[0]w[1]\cdots w[n-1]$. Then $\psi_w(a)=\psi_{w[0]}(\psi_{w[1]}(\cdots (\psi_{w[n-1]}(a))\cdots ))$ and $\overline \psi_w(a)=\overline \psi_{w[0]}(\overline \psi_{w[1]}(\cdots (\overline \psi_{w[n-1]}(a))\cdots ))$ , with $a \in \mathcal{A}$. Similarly to the Sturmian morphisms, we can define the episturmian morphisms as follows. \begin{defn}\cite{jp2002} The set $\mathscr{E}$ of {\it episturmian morphisms} is the monoid generated by the morphisms $\psi_a, \overline{\psi}_a, \theta_{ab}$ \index{$\psi_a$}\index{$\overline{\psi}_a$}\index{$ \theta_{ab}$} under composition. The set $\mathscr{S}$ of {\it standard episturmian morphisms} is the submonoid generated by the $\psi_a$ and $\theta_{ab}$; the set of {\it pure episturmian morphisms} is the submonoid generated by the $\psi_a$ and $\overline \psi_a$. \end{defn} As the Sturmian morphism, the episturmian ones have the following characteristic property: a morphism $f$ is {\it episturmian} if $f(s)$ is episturmian for any episturmian sequence~$s$. \section{Epichristoffel words} In this section, we generalize Christoffel words to a $k$-letter alphabet and we call this generalization {\it epichristoffel words}. Let us first recall some properties of Christoffel words that will be used to define their generalization. \begin{lem} {\rm \cite{bdl1997}} \label{LyndonChristo}A word $w$ is a Christoffel word if and only if $w$ is a balanced Lyndon word. \end{lem} The next proposition follows from S\'e\'ebold, Richomme, Kassel and Reutenauer works \cite{ps1996,ps1998,gr2007,kr2007} and is proved in \cite{wfc1999}. \begin{prop} \label{christoMorph} Christoffel words and their conjugates are exactly the words obtained by the application of Sturmian morphisms to a letter. \end{prop} {Lemma }\ref{LyndonChristo} {and Proposition} \ref{christoMorph} {have for consequence the following corollary.} \begin{cor} \label{agen} In the conjugation class of a Christoffel word, the Lyndon word is the Christoffel word. \end{cor} {Note that Corollary }\ref{agen} { is the result we will extend as a definition of epichristoffel words.} \begin{defn}\label{defEpiClass} A finite word $w \in \mathcal{A}^*$ belongs to an {\it epichristoffel class} if it is the image of a letter by an episturmian morphism. \end{defn} \begin{defn} A finite word $w \in \mathcal{A}^*$ is {\it epichristoffel}\index{mot!epichristoffel@\'epichristoffel} if it is the unique Lyndon word occurring in an epichristoffel class. \end{defn} In the sequel, a word in an epichristoffel class will be called {\it $c$-epichristoffel}, {for short.} {The following result insures that the epichristoffel classes are well-defined.} \begin{prop} \label{propguil} Let $w$ and $w'$ be conjugate finite words. Then $w=\phi(u)$ and $w'=\phi'(u')$, with $\phi, \phi' \in \{\psi_a, \overline{\psi}_a \}$, for $u, u' \in \mathcal{A}^*$, $a \in \mathcal{A}$ if and only if $u$ and $u'$ are conjugate. \end{prop} {\nl\it Proof.\ } \begin{itemize} \item [($\Longrightarrow$)] Without loss of generality, we can suppose that $\phi=\phi'=\psi_a$, since $\psi_a(w)=a\overline \psi_a(w) a^{-1}$ and so, $\psi_a(w)$ is conjugate to $\overline \psi_a(w)$ for any word $w$. Thus, we can write $w=a^{n_0}v[0]a^{n_1}v[1]\cdots a^{n_k}v[k]$, with $v[i] \neq a$ and $n_i >0$ for $0 \leq i \leq k$. Since $w=\psi_a(u)$, using injectivity of $\psi_a$, we have $u=a^{n_0-1}v[0]a^{n_1-1}v[1]\cdots a^{n_k-1}v[k]$. Since $w$ and $w'$ are conjugate, we can write $w'=a^{\alpha}v[i]a^{n_{i+1}}v[i+1]\cdots a^{n_{i-1}}v[i-1]a^\beta$, with $\alpha+\beta = n_i$ and $\alpha \geq 1$. Thus, $u'=a^{\alpha-1}v[i]a^{n_{i+1}-1}v[i+1]\cdots a^{n_{i-1}-1}v[i-1]a^\beta$. Comparing $u$ and $u'$, we conclude that $u$ is conjugate to $u'$. \item[($\Longleftarrow$)] If $u$ and $u'$ are conjugate, then there exist $v,t$ such that $u=vt$ and $u'=tv$. Applying respectively the morphisms $\phi$ and $\phi'$ over $u$ and $u'$, we obtain $\phi(u)=\phi(v)\phi(t)$ and $\phi'(u')=\phi'(t)\phi'(v)$. If $\phi=\phi'$ the result follows. Otherwise, let us suppose $\phi=\psi_a$ and $\phi'=\overline \psi_a$. Then we conclude using the fact that $\psi_a(u)=a\overline \psi_a(u)a^{-1}$: $$\psi_a(u)=a\overline \psi_a(v)a^{-1}a\overline \psi(t)a^{-1}= a\overline \psi_a(v)\overline \psi_a(t)a^{-1}.$$ \vspace{-1.5cm} {\flushright \rule{1ex}{1ex} \par\medskip} \end{itemize} The finite factors of episturmian sequences, also called finite Arnoux-Rauzy words, have already been studied. In \cite {jp2002}, the authors used a subclass of $c$-epichristoffel words without mentioning that it is a generalization of Christoffel words. In their paper, they denoted by $h_n$, the standard episturmian words, that is the words obtained by the application of standard episturmian morphisms to a letter. The $c$-epichristoffel words are exactly the set of all conjugates of the standard episturmian words and the smallest one in the conjugacy class is epichristoffel. Notice that they form a subclass of the Arnoux-Rauzy word, since they all are factor of episturmian sequences, but any factor of episturmian sequence is not necessarily obtained by an episturmian morphism to a letter. For instance, the word $abacab\underline{aabac}ababacabaabacaba\cdots$ contains the finite Arnoux-Rauzy word $aabac$ which is not $c$-epichristoffel. In \cite{jp2002}, the authors proved the $2$ following properties. \begin{prop} \textnormal{(\cite{jp2002}, prop. 2.8, prop. 2.12)} \label{propal1}Every standard episturmian word is primitive and can be written as the product of $2$ palindromic words. \end{prop} It is clear that any standard episturmian word is conjugate to an epichristoffel word. Proposition \ref{propguil} can be used to show the converse. Consequently, Proposition \ref{propal1} can be generalized for any $c$-epichristoffel word, {using the following lemma.} \begin{lem} \textnormal{(\cite{djp2001}, Lemma 3)} \label{lemcon} {The word $u\in \mathcal{A}^*$ is a palindrome if and only if $\psi_a(u)a$ and $a \overline \psi_a(u)$ are so, $a \in \mathcal{A}$.} \end{lem} \begin{prop} \label{propal} Every $c$-epichristoffel word is primitive and can be written as the product of $2$ palindromic words. \end{prop} {\nl\it Proof.\ } {By induction over the number of morphisms. For a single morphism applied over a letter, we get $w=ab$, with $a, b \in \mathcal{A}$ and $a\neq b$, which is the product of two palindromes. Let us suppose that for a $c$-epichristoffel word $w$, there exist palindromic words $u$, $v$ such that $w=uv$. Let $x=\psi _c(w)=\psi _c(uv)=\psi _c(u) \psi _c(v)$ (resp. $x=\overline \psi_c(w)=\overline \psi_c(u)\overline \psi_c(v)$), for $c \in \mathcal{A}$. Then $x=\psi _c(u) cc ^{-1} \psi _c(v)$ (resp. $x=\overline \psi _c(u) c ^{-1} c\overline \psi _c(v)$), where $\psi _c(u) c$, $c ^{-1} \psi _c(v)$ (resp. $\overline \psi _c(u) c ^{-1}$, $c \overline \psi _c(v)$) are palindromic words by Lemma} \ref{lemcon}. \rule{1ex}{1ex} \par\medskip Let now show how some of the properties of Christoffel words can be generalized to epichristoffel words. Recall that for Christoffel {words}, we have: \begin{theo} \label{thdldl}{\rm \cite{dldl2006}} Let $w$ be a non empty finite word. The following conditions are equivalent: \vspace{-0.2cm} \begin{itemize} \item [\rm i)] $w$ is a factor of a Sturmian sequence; \item [\rm ii)] the fractionnary root $z_w$ of $w$ is conjugate to a Christoffel word. \end{itemize} \end{theo} First, note that the equivalence in Theorem \ref{thdldl} cannot be generalized to epichristoffel words. Indeed, let us consider the episturmian sequence $$s=aabaacaabaacaabaabaa \cdot{ caabaacaabaaa}\cdots$$ Then $w=caabaacaabaaa$ is a factor of $s$, but its fractionnary root $z_w=w$ is not $c$-epichristoffel, as we will see later in Example \ref{exa35}. On the other hand, the converse holds for episturmian sequences and epichristoffel words. \begin{theo} Let $w$ be a non empty word such that its fractionnary root is $c$-epichristoffel. Then $w$ is a factor of an episturmian sequence. \end{theo} {\nl\it Proof.\ } Let $w=z^k_w$, with $k\geq 1 \in \mathbb{Q}$, $z_w$ the fractionnary root of $w$. Let us suppose that $z_w$ is $c$-epichristoffel. Thus there exist $x \in \mathcal{A}^*$ and $a \in \mathcal{A}$ such that $\phi^{(0)}\phi^{(1)}\cdots \phi^{(n)}(a)=z_w$, with $\phi^{(i)}\in \{\psi_{x[i]}, \overline \psi_{x[i]}\}$. Then $w$ is a factor of $z_w^{\lceil k \rceil}=(\phi^{(0)}\phi^{(1)}\cdots \phi^{(n)} (a))^{\lceil k \rceil}=\phi^{(0)}\phi^{(1)}\cdots \phi^{(n)}(a^{\lceil k \rceil})$. It is sufficient to take an episturmian sequence having $a^{\lceil k \rceil}$ as a factor and apply the morphism $\phi^{(0)}\phi^{(1)}\cdots \phi^{(n)}$: we obtain that $\phi^{(0)}\phi^{(1)}\cdots \phi^{(n)} (a^{\lceil k \rceil})$ is a factor of an episturmian sequence and so is $w$. \rule{1ex}{1ex} \par\medskip \begin{prop} Let $w \in \mathcal{A}^*$ be a $c$-epichristoffel word. Then, the set of factors of length $\leq |w|$ of its conjugacy class is closed under mirror image. \end{prop} {\nl\it Proof.\ } First note that the set of factors of length $\leq |w|$ of the epichristoffel class of $w$ is the same as the one of $w^2$. Since any $c$-epichristoffel word $w$ is the product of $2$ palindromes (by Proposition \ref{propal}), let $w=p_1p_2$, with $p_1$, $p_2$ palindromes. Then $w^2=p_1p_2p_1p_2$ and it follows that $\widetilde{w}=\widetilde{p_1p_2}=p_2p_1$ is a factor of $w^2$. Thus, the mirror image of any factor of $w$ is also a factor of $w^2$ and consequently, is in the epichristoffel class of $w$. \rule{1ex}{1ex} \par\medskip {\rem The right palindromic closure of a $c$-epichristoffel word is often {a} prefix of $w^2$, but it is not the case in general. It suffices to take the word $w=abcbab$ for which $w^{(+)}=abcbab\cdot cba$. } For Christoffel words, we have: \begin{lem} {\rm \cite{dlm1994}}A Christoffel word can always be written as the product of two Christoffel words. \end{lem} But: \begin{lem} \label{lemneg} An epichristoffel word cannot always be written as the product of two epichristoffel words. \end{lem} {\nl\it Proof.\ } It is sufficient to consider the epichristoffel word $aabacab$. The only decompositions in $c$-epichristoffel factors are $a\cdot abacab$ and $aab\cdot acab$, but $abacab$ and $acab$ are not Lyndon words, assuming $a< b< c$. \rule{1ex}{1ex} \par\medskip \begin{lem} Any $c$-epichristoffel word having length $> 1$ can be non-uniquely written as the product of two $c$-epichristoffel words. \end{lem} {\nl\it Proof.\ } For the non unicity, it is sufficient to consider the example of the word $aabacab$ given in the proof of Lemma \ref{lemneg}. By definition, any $c$-epichristoffel word can be written as $\phi^{(0)}\phi^{(1)} \cdots \phi^{(n-1)}(a)$, with $a \in \mathcal{A}$, $\phi^{(i)}\in \{\psi_{w[i]}, \overline \psi_{w[i]} \}$, $w \in \mathcal{A}^n$ and $w[n-1]\neq a$. Assume $\phi^{(n-1)}=\psi_{w[n-1]}$. To prove the existence of the product, it is then sufficient to consider the words $\phi^{(0)} \phi^{(1)}\cdots \phi^{(n-1)}(w[n-1])$ and $\phi^{(0)}\phi^{(1)}\cdots \phi^{(n-2)}(a)$, since \vspace{-0.2cm} \begin{eqnarray*} \phi^{(0)}\phi^{(1)}\cdots \phi^{(n-1)}(a)&=& \phi^{(0)}\phi^{(1)}\cdots \phi^{(n-2)}(w[n-1]a)\\ &=& \phi^{(0)}\phi^{(1)}\cdots \phi^{(n-2)}(w[n-1]) \cdot \phi^{(0)}\phi^{(1)}\cdots \phi^{(n-2)}(a). \end{eqnarray*} The case $\phi^{(n-1)}=\overline \psi_{w[n-1]}$ is analogue: we would have obtained a conjugate. {\flushright \rule{1ex}{1ex} \par\medskip} \vspace{-0.2cm} \section{Epichristoffel $k$-tuples} Recall from \cite{bl1993} that for a given $(p,q)$, with $p, q \in \mathbb N$, there exists a Christoffel word with occurrence numbers of letters $p$ and $q$ if and only if $p$ and $q$ are relatively primes. Moreover, it is possible to construct the corresponding Christoffel word, using a Cayley graph (see \cite{br2006}). In this section, we give an algorithm which determines if there exists or not an epichristoffel word $w$ over the alphabet $\mathcal{A} =\{a_0,a_1,\ldots,a_{k-1}\}$ such that $p=(p_0,p_1,\dots, p_{k-1})$ with $p_i=|w|_{a_i}$, for $0\leq i \leq k-1$. If so, we also give an algorithm that constructs it. \begin{defn} Let $p=(p_0,p_1,\ldots,p_{k-1})$ be a $k$-tuple of non negative integers. Then the {\it operator} $T: \mathbb N^k \rightarrow \mathbb Z^k$ is defined over the $k$-tuple $p$ as \vspace{-0.2cm} $$T(p)=T(p_0,p_1,\ldots,p_{k-1})=(p_0,p_1,\ldots, p_{i-1},\left(p_i - \sum _{j=0, j \neq i}^{k-1} p_j\right ), p_{i+1}, \ldots,p_{k-1}),$$ \vspace{-0.2cm} where $p_i \geq p_j$, $\forall j\neq i$. \end{defn} \begin{prop} \label{ktuplets} Let $p$ be a $k$-tuple. There exists an epichristoffel word with occurrence numbers of letters $p$ if and only if iterating $T$ over $p$ yields a $k$-tuple $p'$ with $p'_j=0$ for $j\neq m$ and $p'_m=1$, for a unique $m$ such that $0\leq m \leq k-1$. \end{prop} The idea of using the operator $T$ comes from the algorithm computing the greatest common divisor of $3$ integers as described in \cite{cmr1999} and of the tuples described in \cite{jj2000}. {\exa \label{exa35} There is no epichristoffel word with the occurrence numbers of letters $(2,2,9)$. Indeed, $T(2,2,9)=(2,2,5)$, $T^2(2,2,9)=T(2,2,5)=(2,2,1)$, $T^3(2,2,9)=T(2,2,1)=(2,-1,1)$. On the other hand, the $6$-tuple $q=(1,1,2,4,8,16)$ does so: \begin{eqnarray*} T(1,1,2,4,8,16)&=&(1,1,2,4,8,0)\\ T^2(q)&=&T(1,1,2,4,8,0)=(1,1,2,4,0,0)\\ T^3(q)&=&T(1,1,2,4,0,0)=(1,1,2,0,0,0)\\ T^4(q)&=&T(1,1,2,0,0,0)=(1,1,0,0,0,0)\\ T^5(q)&=&T(1,1,0,0,0,0)=(1,0,0,0,0,0). \end{eqnarray*} } Some lemmas are required in order to prove Proposition \ref{ktuplets}. \begin{lem} \label{freq} Let $w=\phi(u)$, with $\phi \in \{\psi_{a_0}, \overline \psi_{a_0}\}$, $\mathcal{A}=\{a_0,a_1,\ldots, a_{k-1}\}$ and $u\in \mathcal{A}^*$. Then \vspace{-0.2cm} \begin{itemize} \item [\rm i)] $\displaystyle |w|_{a_0}=\sum_{i=0}^{k-1} |u|_{a_i}=|u|$; \vspace{-0.2cm} \item [\rm ii)] $\displaystyle |w|_{a_0}=|u|_{a_0}+\sum_{i=1}^{k-1} |w|_{a_i}$. \end{itemize} \end{lem} {\nl\it Proof.\ } The first equality comes from the definition of $\psi_{a_0}$ and $\overline \psi_{a_0}$. For each letter $\alpha \neq a_0$, $\psi_{a_0}(\alpha)=a_0\alpha$, $\overline \psi_{a_0}=\alpha a_0$ and $\overline\psi_{a_0}=\psi_{a_0}(a_0)=a_0$: $\phi$ adds as much $a_0$ as the occurrence numbers of the other letters in the word $u$. The second equality follows from the first one, since $|w|_{a_i}=|u|_{a_i}$ for $i\neq0$. \vspace{-0.8cm} {\flushright \rule{1ex}{1ex} \par\medskip } \begin{lem} \label{letterMax} Let $w\in \mathcal{A}^*$ be a $c$-epichristoffel word. Then, there exist a $c$-epichristoffel word $u \in \mathcal{A}^*$, $|u| >1$ and an episturmian morphism $\phi \in \{\psi_{a_0}, \overline \psi_{a_0}\}$, with $a_0 \in \mathcal{A}$, such that $w=\phi(u)$ if and only if $|w|_{a_0} > |w|_{a_i}$ for all $a_i \in \mathcal{A}$, $i \neq 0$. \end{lem} {\nl\it Proof.\ } \begin{itemize} \item [($\Longrightarrow$)] By contradiction. Let us suppose there exists $u$ with $|u|>1$ such that $w=\phi(u)$ and $|w|_{a_0}$ is not maximum. Then, there exists at least one letter $a_i \in \mathcal{A}$ such that $|w|_{a_i} \geq |w|_{a_0}$. Without loss of generality, let us suppose that $i=1$. By Lemma \ref{freq}, $|w|_{a_0}=\sum_{i=0}^{k-1} |u|_{a_i}=|u|_{a_0}+|w|_{a_1}+\sum_{i =2}^{k-1}|u|_{a_i}$ that implies $|w|_{a_0}-|w|_{a_1}=|u|_{a_0}+\sum_{i=2}^{k-1}|u|_{a_i} \leq 0$, which is possible only if $|u|_{a_i}=0$ for all $i \neq 1$ and then $|w|_{a_1}=|w|_{a_0}$. Hence, we would have that $u={a_1}^n$ and $w=\phi({a_1}^n)$. The only possibility is that $n=1$, since a $c$-epichristoffel word is primitive. Then $|u|=1$: contradiction. Hence, if $w=\phi(u)$, with $|u|>1$, $|w|_{a_0}$ is maximum. \item [($\Longleftarrow$)] Let us now suppose that $|w|_{a_0} > |w|_{a_i}$ for all $a_i \in \mathcal{A}$, $i \neq 0$. Since $w$ is $c$-epichristoffel, there exist an episturmian morphism $\phi \in \{ \psi_{a_i}, \overline \psi_{a_i}\}$ and a $c$-epichristoffel word $u \in \mathcal{A}^*$ such that $\phi(u)=w$. Let us suppose that $i \neq 0$. Using Lemma \ref{freq}, $|w|_{a_i}=|w|_{a_0}+|u|_{a_i}+ \sum_{1\leq j \leq k-1, j\neq i}|w|_{a_j}$. Since $|w|_{a_0}> |w|_{a_i}$, it implies that $|u|_{a_i}+\sum_{1\leq j\leq k-1,j\neq i}|w|_{a_j}<0$, which is impossible. Thus $i=0$. \rule{1ex}{1ex} \par\medskip \end{itemize} An interesting consequence of Lemma \ref{letterMax} is the following. \begin{prop} \label{propUnique} Let $u$ and $v$ be $c$-epichristoffel words. If $|u|_\alpha=|v|_\alpha$ for all $\alpha \in \mathcal{A}$, then $u$ and $v$ are conjugate. In other words, a $k$-tuple of occurrence numbers of letters determines at most one epichristoffel conjugacy class. \end{prop} {\nl\it Proof.\ } {By induction. The result is true when $|u|=|v| \leq 2$. Assume by now that $|u| \geq 3$. By definition of epichristoffel words, there exist letters $a$ and $b$, and epichristoffel words $u', v'$ such that $u=\phi(u')$, $v=\phi'(v'), \phi \in \{\psi_a,\overline \psi_a\}$ and $\phi' \in \{\psi_b,\overline\psi_b\}$. From $|u| \geq 3$ and definitions of morphisms $\psi_a, \overline \psi_a, \psi_b, \overline \psi_b$, we get $|u'| \geq 2, |v'| \geq 2$. From Lemma} \ref{letterMax} {and the fact that $|u|_\alpha=|v|_\alpha$ for all letters $\alpha$, it comes that $a=b$ (and $|u|_a=|v|_a \geq |u|_\alpha=|v|_\alpha$ for all letters $\alpha$). Now from definition of $u'$ and $v'$ and properties of $u$ and $v$, we deduce that $|u'|_\alpha=|v'|_\alpha$ for all letters $\alpha$. By inductive hypothesis, $u'$ and $v'$ are conjugate. Proposition} \ref{propguil} {allows to conclude.} \rule{1ex}{1ex} \par\medskip The algorithm induced by the iteration of Lemma \ref{letterMax} leads to a construction of words which are images of a letter by an episturmian morphism, that is $c$-epichristoffel words. Indeed, iterating $T$ gives a construction of an $c$-epichristoffel word with $p$ describing the occurrence numbers of letters. We take $p$ as the initial $k$-tuple. The iteration over $p$ of the operator $T$ described previously yields a finite sequence of $k$-tuples $p^{(0)}$, $p^{(1)}$, $p^{(2)}, \dots$ We do as in Proposition \ref{ktuplets}, applying the operator $T$ and moreover, we keep an important information that allows us to construct the word: the letter with maximal number of occurrences. Let \vspace{-0.2cm} $$p^{(s)} \xrightarrow [ ]{\text{$i$}} p^{(s+1)}$$ denote the relation $T(p^{(s)})=p^{(s+1)}$, where $p^{(s)}_i$ is the maximal integer of $p^{(s)}$. Then, performing $T$ until $p^{(r)}_i=0$ for all $i$ except for one $i_{r-1}$ for which $p^{(r)}_{i_{r-1}}=1$, we get the sequence of $k$-tuples $$p^{(0)} \xrightarrow [ ]{\text{$i_0$}} p^{(1)} \xrightarrow [ ]{\text{$i_1$}} p^{(2)} \xrightarrow [ ]{\text{$i_2$}} \cdots \xrightarrow [ ]{\text{$i_{r-2}$}} p^{(r-1)} \xrightarrow [ ]{\text{$i_{r-1}$}} p^{(r)}.$$ Then, $$\psi_{a_{i_0}}(\psi_{a_{i_1}}(\dots(\psi_{a_{i_{r-1}}}(\alpha))\dots))$$ is a $c$-epichristoffel word having $p$ as occurrence numbers of letters, with $\alpha$ the letter such that $p^{(r)}_{i_{r-1}}=1$. The epichristoffel word is the Lyndon word of the conjugacy class of the word obtained. Here, Proposition \ref{propUnique} insures that it is sufficient to consider the standard episturmian morphism in order to construct a $c$-epichristoffel word with $p$ describing the occurrences of the letters. \\ \noindent {\it {Proof of Proposition} \ref{ktuplets}.} Follows directly from Lemmas \ref{freq}, \ref{letterMax} and from the ideas described in the previous paragraph. The only difficulty concerns the last iteration, that is when $w=\phi(u)$, with $|w|_{a_0}$ not maximum. As seen in the previous proof, it implies that $u=a_1$ and $w=\phi(a_1) \in \{a_0a_1, a_1a_0\}$, which is clearly a $c$-epichristoffel word. Notice here that $\psi_{a_0}(a_1)=\overline \psi_{a_1}(a_0)$ and $\overline \psi_{a_0}(a_1)= \psi_{a_1}(a_0)$ are conjugate. \rule{1ex}{1ex} \par\medskip {\exa For the triplet $(5,10,16)$ describing the occurrence numbers of respectively the letters $a,b$ and $c$, the sequence obtained is $$(5,10,16) \xrightarrow[ ]{\text{$c$}} (5,10,1) \xrightarrow[ ]{\text{$b$}} (5,4,1) \xrightarrow [ ]{\text{$a$}}(0,4,1)\xrightarrow[ ]{\text{$b$}} (0,3,1)\xrightarrow[ ]{\text{$b$}} (0,2,1)\xrightarrow[ ]{\text{$b$}} (0,1,1) \xrightarrow[]{\text{$b$}} (0,0,1).$$ Performing the algorithm, we find the word \vspace{-0.2cm} \begin{eqnarray*} \psi_{cbabbbb}(c)&=&\psi_{cbabbb}(bc)\\ &=&\psi_{cbabb}(\psi_b(bc))\\ &=&\psi_{cbab}(\psi_b(bbc))\\ &=&\psi_{cba}(\psi_b(bbbc))\\ &=&\psi_{cb}(\psi_a(bbbbc))\\ &=&\psi_{c}(\psi_b(ababababac))\\ &=&\psi_{c}(babbabbabbabbabc)\\ &=& cbcacbcbcacbcbcacbcbcacbcbcacbc.\\ \end{eqnarray*} Since it is obtained by a standard episturmian morphism to a letter, this standard episturmian word is a representant of the epichristoffel conjugacy class. Moreover, its conjugate which is a Lyndon word, and so, an epichristoffel word, is $acbcbcacbcbcacbcbcacbcbcacbc\cdot cbc$ for the order $a<b<c$. } Note that in the previous example, the choice of the last transition is arbitrary: we could have chosen the transition $(0,1,1) \xrightarrow[]{\text{$c$}}(0,1,0)$ instead of $(0,1,1) \xrightarrow[]{\text{$b$}} (0,0,1)$ and we would have obtained a conjugate of $\psi_{cbabbbb}(c)$ which is also $c$-epichristoffel. \section{Criteria to be in an epichristoffel class} Let us recall a characterization of words in the conjugacy class of a Christoffel word. \begin{theo} \label{factSturm} {\rm \cite{dldl2006}} Let $w \in \mathcal{A}^*$ be a primitive word. Every conjugate $w' $ is a factor of a Sturmian sequence, not necessarily the same, if and only if $w$ is conjugate to a Christoffel word. \end{theo} The goal of this section is to prove the following generalization of Theorem \ref{factSturm}. \begin{theo} \label{leth} Let $w$ be a finite primitive word different from a letter. Then there exists an episturmian sequence $z$ such that all the conjugates of $w$ are factors of $z$ if and only if $w$ is a $c$-epichristoffel word. \end{theo} Note that in order to generalize Theorem \ref{factSturm} to a $k$-letter alphabet, $k \geq 3$, an additional condition is necessary: the conjugates must be factor of the {\bf same} episturmian sequence. For example, every conjugates of the word $abc$ are factors of episturmian sequences, but $abc$ is not a $c$-epichristoffel word, since $T(1,1,1)=(1,1,-1)$. Let us recall the following results of Justin and Pirillo that allow us to write any episturmian sequence as the image by an episturmian morphism of an other episturmian sequence. \begin{cor} {\rm \cite{jp2002}}\label{cor37} Let $s \in \mathcal{A}^\omega$ and $\Delta=x[0]x[1]x[2]\cdots$, $x[i] \in \mathcal{A}$. Then $s$ is a standard episturmian sequence with directive sequence $\Delta$ if and only if it exists an infinite sequence of sequences $s^{(0)}=s, s^{(1)}, s^{(2)}, \ldots$ such that for any $i \in \mathbb N$, {$s^{(i-1)}=\psi_{x[i]}(s^{(i)})$}. \end{cor} It can also be generalized to non standard episturmian sequences. In order to do so, let us recall what is a {\it spinned word}. Let $\overline \mathcal{A}= \{\overline a \, | \, a \in \mathcal{A}\}$. A letter $\overline x$ is considered as $x$ with {\it spin} $1$ while $x$ itself is considered as $x$ with spin $0$. Then, an {\it infinite spinned word} $\check s=\check s[0] \check s[1] \check s[2] \cdots $ is an element of $(\mathcal{A} \cup \overline \mathcal{A})^\omega$. \begin{theo} {\rm \cite{jp2002}} \label{dirseq}A sequence $t\in \mathcal{A}^\omega$ is episturmian if and only if there exist a spinned sequence $\check{\Delta}=\check x[0]\check x[1]\check x[2]\cdots$, $\check x[i] \in \{\mathcal{A} \cup \overline \mathcal{A}\}$ and an infinite sequence of recurrent sequences $t^{(0)}=t$, $t^{(1)},t^{(2)}, \ldots$ such that for $i \in \mathbb N$, $t^{(i-1)}=\psi_{x[i]}(t^{(i)})$ if $\check x[i]$ has spin $0$ (resp. $\overline \psi_{x[i]}(t^{(i)})$ if $\check x[i]$ has spin $1$). Moreover $t$ is equivalent to the standard episturmian sequence with directive sequence $\Delta=x[0]x[1]\cdots$. \end{theo} {Theorem} \ref{dirseq} {allows us to write the directive sequence of a non standard episturmian sequence, as we do in the following lemma.} \begin{lem} \label{gp1} Let $\check \Delta(s)=(\check a)^k\check b \check z$ be the directive sequence of an episturmian sequence $s$, with $a \neq b \in \mathcal{A}$ and $z \in \mathcal{A}^\omega$. Then the blocks of $c\neq a$ have length $1$ and the blocks of $a$'s have length $\ell$, $k$ or $(k+1)$, where $\ell \leq k+1$ is the length of the block of $a$'s prefix of the sequence. \end{lem} {\nl\it Proof.\ } Let us consider the equivalent standard episturmian sequence $t$ directed by $\Delta(t)=a^kbz$. {By Corollary} \ref{cor37}, {$t=\psi_{a^kb}(t')$ for a standard episturmian word $t'$. Since $\psi_{a^kb}(a)=a^kba$, $\psi_{a^kb}(b)=a^kb$ and for $c \notin \{a,b\}$, $\psi_{a^kb}(c)=a^kba^kc$, the statement is true for $t$. Since the langage of $s$ and $t$ are equals, it only remains to consider the prefix of $s$ where a block of length $<k$ can appear. Indeed, for the episturmian sequence $s$, since it is directed by $\check \Delta (s)=(\check a)^k\check b \check z$, we easily deduce that $s$ begins by a prefix of $a$'s of length $\ell$ equals to the number of $\check a$ having spin $0$ in the prefix $(\check a)^k$ of its directive sequence, which is less or equal to $k$. \rule{1ex}{1ex} \par\medskip } {\rem An episturmian sequence may not have blocks of $a$'s of length $(k+1)$. It is the case if its directive sequence has the form $\check a^k\check z$, with $|\check z|_{\check a}=0$. } One can be easily convinced of the following statement. \begin{lem} \label{gp2} In an episturmian sequence $w=\psi_\alpha(t)$ or $w=\overline \psi _\alpha (t)$, any letter different from $\alpha$ is preceded and followed by the letter $\alpha$, except for the first letter of the sequence, if it is different from $\alpha$. \end{lem} \begin{lem} \label{lemDeco} Let $z=\psi _{a_0}(t)$ be a standard episturmian sequence and $w=a_0ya_1$ a factor of $z$, with $a_0\neq a_1 \in \mathcal{A}$ and $y \in \mathcal{A}^*$. Then, there exists a factor $u$ of $t$ such that $\psi_{a_0}(u)=w$. \end{lem} {\nl\it Proof.\ } If $z=\psi_{a_0}(t)$, $t=t[0]t[1]t[2]\cdots$ and $\hbox{\rm Card}(\mathcal{A})=k$, then by the definition of $\psi$, $z=\psi_{a_0}(t[0])\psi_{a_0}(t[1])\cdots \in \{a_0,a_0a_1, a_0a_2,\ldots, a_0a_{k-1}\}^\omega$. Since $w$ starts with $a_0$ and ends by $a_1$, then any factor $w$ of $z=\psi_{a_0}(t)$ can be written as $w \in \{a_0,a_0a_1, a_0a_2, \ldots, a_0a_{k-1}\}^*$. Thus we can construct a word $u$ by associating to $a_0a_i$ the letter $a_i$ for $i\neq 0$ and to $a_0$ the letter $a_0$. Thus, $w$ is the image of the word $u$ by the morphism $\psi_{a_0}$. \rule{1ex}{1ex} \par\medskip \begin{prop} \label{prop4} Let $z=\psi _a(t)$, where $t$ and $z$ are standard episturmian sequences. Let $w$ be a factor of $z$ not power of a letter, such that $|w|>1$ and all its conjugates are also factors of $z$. Then, there exists a factor $u$ of $t$ such that $w=\psi _a(u)$ or $w=\overline{\psi}_a(u)$. \end{prop} {\nl\it Proof.\ } Let $\beta, \gamma \in \mathcal{A}$, with $\beta,\gamma \neq a$, $y \in \mathcal{A} ^*$ and $w$ factor of $z$. There are $4$ cases to consider. \begin{itemize} \item [i)] $w=\beta y\gamma$: its conjugate $y\gamma \beta$ is not a factor of $z$, since any occurrence of the letter $\beta$ is preceded by the letter $a$, by Lemma \ref{gp2}. Then $w$ does not satisfied the hypothesis. \item [ii)] $w=ay\beta$: by Lemma \ref{lemDeco}, there exists $u$ factor of $t$ such that $\psi _a(u)=w$. \item [iii)] $w=\beta ya$: symmetric to the case ii). If $w=\beta ya$ is a factor of $z=\psi _a(t)$ and satisfies the hypothesis, then there exists $u$ factor of $t$ such that $\overline{\psi _a}(u)=w$. \item [iv)] $w=aya$: rewrite $w=a ^ my'a ^n$, with $m$, $n$ $\geq 1$ and $m, n$ maximum. The factor $y' $ is {not} empty, since $w$ is supposed not to be a power of a letter. Let us suppose that there exists $\beta \in \mathcal{A}$, $\beta \neq a$ such that $w\beta=a ^my'a ^n\beta$ is a factor of $z$. Since {by Lemma} \ref{gp1} any block of $a$ has length $k$ or $(k+1)$, {for some $k\in \mathbb N\setminus \{0\}$}, we have that $n=k$ or $n=k+1$. On the other hand, by the hypothesis, the conjugate $y'a ^{m+n} $ of $w=a ^m y' a ^n$ is also a factor of $z$. Thus $m+n \leq k+1$. But since $m\neq 0$, the only possibility is that $n=k$ and $m=1$. Consequently $w=ay'a ^k$. Its conjugate $y' a ^{k+1}$ is also a factor of $z$ and since $y'$ does not start by $a$ by the maximality of $m$, it should be preceded by $a$: $ay' a ^{k+1}=ay'a ^ka=wa$ is a factor of $z$. Since $z$ is episturmian, $wa$ factor of $z$ implies that there exist $\ell \in \mathbb N$ and $\beta \neq a \in \mathcal{A}$ such that $wa^\ell\beta$ is so. {By Lemma} \ref{lemDeco}, { there exists a word $u'=ua^{\ell-1}\beta$ such that $\psi_a(u')=wa^\ell\beta$. Since $\psi_a(a^{\ell-1}\beta)=a^\ell\beta$, $w=\psi_a(u)$. } \end{itemize} {\flushright \rule{1ex}{1ex} \par\medskip} We can now prove our main Theorem.\\ \noindent {\it {Proof of Theorem} \ref{leth}}. \begin{itemize} \item [($\Longrightarrow$)] \begin{itemize} \item [i)] Let us suppose that all conjugates of $w$ are factor of a standard episturmian sequence $z=\psi_a(t)$. We proceed by induction on the number of morphisms. Since $z=\psi_a(t)$, by Proposition \ref{prop4}, there exists $u$ such that $w=\psi_a(u)$ or $w=\overline \psi_a(u)$. Let us now prove that all conjugates $u'$ of $u$ are also factors of $t$. Since $u, u'$ are conjugate, using Proposition \ref{propguil}, we have $\psi_a(u')$ is a conjugate of $ \psi_a(u)$. Hence, again by Proposition \ref{prop4}, there exists a factor $u''$ of $t$ with $\psi_a(u')=\psi_a(u'')$ or $ \psi_a(u')=\overline \psi_a(u'')$. The second case is possible only if $u''$ is a power of $a$ and then the first case holds. This first case by injectivity of $\psi$ implies $u'=u''$, that is $u''$ is a factor of $t$. We then find a sequence of episturmian morphisms $\phi _0, \phi _1,.. ,\phi _k \in \{\psi_a,\overline \psi_a \, |\, a \in \mathcal{A}\}^{k+1}$ and a sequence of words $w, w_1, w_2,... ,w_k$ such that $|w| \geq |w_1| \geq |w_2| \geq \ldots \geq |w_k|=1$, $w=\phi _0(\phi _1(...(\phi _k(w_k))...))$ and $w_i=\phi_i(\phi_{i+1}( \ldots (\phi_k(w_k))))$. Thus, $w$ is the image of a letter by an episturmian morphism, implying that $w$ is $c$-epichristoffel. \item [ii)] If $z$ is not standard, by Definition \ref{episst}, we know that there exists an episturmian sequence $z'$ such that $F(z)=F(z')$. Thus, we can then consider the sequence $z'$ and conclude as in i). \end{itemize} \item [($\Longleftarrow$)] Since $w$ is $c$-epichristoffel, we can write $w=f(a)$, where $f \in \mathscr E$ and $a \in \mathcal{A}$. Let $s$ be an episturmian sequence having the factor $aa$ and let consider the episturmian sequence $f(s)$. Thus, it contains the factor $ww$ and we conclude. \end{itemize} \vspace{-1cm}{\flushright \rule{1ex}{1ex} \par\medskip} \section{Concluding remarks} In this paper, we have most of the time consider the $c$-epichristoffel words, also known as the conjugates of the finite standard episturmian words. Some of the properties of standard Sturmian words can be generalized naturally to the $c$-epichristoffel ones. We unfortunately didn't find a characterization of the epichristoffel word of each conjugacy class. Geometrical properties of Christoffel words are well known and very interesting. It would be nice to know if there is a similar geometrical interpretation for the epichristoffel words. In this paper, we only verify if a few properties of the Christoffel words could be generalized or not to the epichristoffel ones. Since the literature of Christoffel words is wide, there are still a lot of open problems about epichristoffel words. For instance: do they satisfy a kind of balanced property? for a fixed $k \geq 3$, does there exist an epichristoffel word over a $k$-letter alphabet of any given length? is it possible to give a closed formula for the number of epichristoffel words of a given length? Episturmian morphisms have been extensively studied for instance in \cite{jj2001,jp2002,gr20032,gr2003,jp2004,jj2005,gr20072}. It might be useful to use their properties to work on the epichristoffel words. Epichristoffel words are still more interesting since they seem to be related to the Fraenkel conjecture. This conjecture states that for a finite $k$-letter alphabet, there exists a unique infinite word, up to letter permutation and conjugation, that is balanced and has pair-wise distinct letter frequencies. This unique word, if it exists, is conjectured to be periodic and can be written as $p^\omega$, with $p$ an epichristoffel word. Then, knowing more about epichristoffel words might help to prove the Fraenkel conjecture. \section*{Acknowledgments} This paper is an extended version of a paper presented in Mons (Belgium) during the 12th Mons Theoretical Computer Science days \cite{gp2008}. The author would also like to thank Christophe Reutenauer for giving her the idea of considering this interesting class of words, for useful discussions and remarks. Many thanks also to the two anonymous referees whose suggestions and constructive remarks helped to improve considerably the quality of the paper. \bibliographystyle{alpha} {\footnotesize
{ "timestamp": "2009-04-24T10:50:25", "yymm": "0805", "arxiv_id": "0805.4174", "language": "en", "url": "https://arxiv.org/abs/0805.4174", "abstract": "Sturmian sequences are well-known as the ones having minimal complexity over a 2-letter alphabet. They are also the balanced sequences over a 2-letter alphabet and the sequences describing discrete lines. They are famous and have been extensively studied since the 18th century. One of the {extensions} of these sequences over a $k$-letter alphabet, with $k\\geq 3$, are the episturmian sequences, which generalizes a construction of Sturmian sequences using the palindromic closure operation. There exists a finite version of the Sturmian sequences called the Christoffel words. They are known since the works of Christoffel and have interested many mathematicians. In this paper, we introduce a generalization of Christoffel words for an alphabet with 3 letters or more, using the episturmian morphisms. We call them the {\\it epichristoffel words}. We define this new class of finite words and show how some of the properties of the Christoffel words can be generalized naturally or not for this class.", "subjects": "Combinatorics (math.CO)", "title": "On a generalization of Christoffel words: epichristoffel words", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.974434788304004, "lm_q2_score": 0.8221891283434877, "lm_q1q2_score": 0.80116968922324 }
https://arxiv.org/abs/2302.02872
Limiting distributions of conjugate algebraic integers
Let $\Sigma \subset \mathbb{C}$ be a compact subset of the complex plane, and $\mu$ be a probability distribution on $\Sigma$. We give necessary and sufficient conditions for $\mu$ to be the weak* limit of a sequence of uniform probability measures on a complete set of conjugate algebraic integers lying eventually in any open set containing $\Sigma$. Given $n\geq 0$, any probability measure $\mu$ satisfying our necessary conditions, and any open set $D$ containing $\Sigma$, we develop and implement a polynomial time algorithm in $n$ that returns an integral monic irreducible polynomial of degree $n$ such that all of its roots are inside $D$ and their root distributions converge weakly to $\mu$ as $n\to \infty$. We also prove our theorem for $\Sigma\subset \mathbb{R}$ and open sets inside $\mathbb{R}$ that recovers Smith's main theorem~\cite{Smith} as special case. Given any finite field $\mathbb{F}_q$ and any integer $n$, our algorithm returns infinitely many abelian varieties over $\mathbb{F}_q$ which are not isogenous to the Jacobian of any curve over $\mathbb{F}_{q^n}$.
\section{Introduction} \subsection{Background} This paper is motivated by the following question. What compact sets $\Sigma$ in the complex plane $\mathbb{C}$ contain infinitely many sets of conjugate algebraic integers, and how are they distributed in $\Sigma$? The ideas of this project originated in the works of many, including Schur~\cite{Schur}, Fekete~\cite{Fekete1923}, Siegel~\cite{MR12092}, Robinson~\cite{MR175881}, Serre~\cite{MR4093205}, Smyth~\cite{MR736460} and Smith~\cite{Smith} on algebraic integers. We introduce the basic concepts and state our theorems which answer some open problems raised in previous works. \newline Let $\Sigma$ be a compact subset of the complex plane. Let \begin{equation}\label{transfn} d_{\Sigma}(n):= \max_{z_1,\dots,z_n\in \Sigma}\prod_{i<j}|z_i-z_j|^{\frac{2}{n(n-1)}}. \end{equation} Fekete proved the following limit exists and called it the transfinite diameter of $\Sigma$: \begin{equation}\label{transf} d_{\Sigma}:=\lim_{n\to \infty} d_{\Sigma}(n). \end{equation} For example, the transfinite diameter of a circle of radius $r$ is $r$. The transfinite diameter of $\Sigma$ equals the \textit{capacity} of $\Sigma$~\cite[Chapter 2]{MR2730573}. Fekete~\cite{Fekete1923}, generalizing some results of Schur~\cite{Schur}, proved that if $d_\Sigma<1$, then there is only a finite number of irreducible monic integral polynomials such that all of their roots lie in $\Sigma$. \\ Note that Fekete's condition $d_{\Sigma}<1$ is optimal. For example, let $\Sigma$ be the unit circle. Then primitive roots of unity give an infinite number of irreducible integral polynomials such that their roots lie all in $\Sigma$. The strict converse is not true. For example, the circle of radius $r>1$ around the origin where $r$ is not an algebraic integer gives a counter-example. However, Fekete and Szeg\"o \cite{MR72941} proved that if $\Sigma$ is symmetric about the real axis and $d_{\Sigma}\geq 1$, then any open set $D$ including $\Sigma$ will contain infinitely many sets of conjugate algebraic integers. Robinson~\cite{Robinson} proved an analogous theorem about real point sets using the properties of Chebyshev's polynomials of $\Sigma$. \\ Recently, Smith~\cite{Smith} studied the weak* limit of a sequence of uniform probability measures on a set of conjugate algebraic integers inside $\Sigma\subset \mathbb{R},$ where $d_{\Sigma}>1$ and $\Sigma$ is a countable union of intervals. Smith used two results form the geometry of numbers (flatness theorem and Minkowski's second theorem) and Robinson's method~\cite{Robinson} to give necessary and sufficient conditions for a probability distribution $\mu$ on the real line to be a weak* limit of such uniform probability measures on totally real conjugate algebraic integers. \\ We generalize Smith's work to the complex plane in the next subsection. Our method is different from Smith's method. In particular, we do not use a flatness theorem or the properties of square-free integral polynomials in our paper. Both of them are crucial ingredients in Smith's work. Instead, we prove a new result in Theorem~\ref{mdim} which is of independent interest. Theorem~\ref{mdim} allows us to improve some results of Smith~\cite[Proposition 3.5]{Smith}. \\ Furthermore, our results do not rely on the existence of Chebychev's polynomials of $\Sigma$; see subsection~\ref{chebsec} and Proposition~\ref{chpolreal}. In fact, our method gives a new proof of Robinson's result and Smith's results without using the properties of Chebychev's polynomials; see Proposition~\ref{chpolreal}. Our method is based on a initial probabilistic sampling of roots with respect to the equilibrium measure and then deforming the roots with a greedy algorithm along a gradient vector. \\ Our algorithm and its implementation has no prior analogue to the best of our knowledge. Our numerical results shows some new features of the algebraic integers that we discuss further in subsection~\ref{complex}. \subsection{Main results}\label{result}\subsubsection{Arithmetic probability measures} Suppose that $\Sigma\subset \mathbb{C}$ is compact and symmetric about the real axis with $d_{\Sigma}\geq 1.$ Let $\mathcal{P}_\Sigma$ be the space of probability measures supported on $\Sigma$ equipped with the weak* topology. Let \[ \Sigma(\rho):=\{z\in \mathbb{C}: |z-\sigma|<\rho \text{ for some }\sigma\in\Sigma \}. \] If $\Sigma\subset \mathbb{R}$, let \[ \Sigma_{\mathbb{R}}(\rho):=\{x\in \mathbb{R}: |x-\sigma|<\rho \text{ for some }\sigma\in\Sigma \}. \] Note that $\Sigma_{\mathbb{R}}(\rho)\subset \mathbb{R}$, and only defined for $\Sigma\subset \mathbb{R}.$ \begin{definition} A measure $\mu \in \mathcal{P}_\Sigma$ is called an \textit{(real) arithmetic probability measure} if $\mu$ is the weak* limit of a sequence of distinct uniform probability measures on a complete set of conjugate (totally real) algebraic integers lying eventually inside ($\Sigma_{\mathbb{R}}(\varepsilon)$ for every $\varepsilon > 0$) $\Sigma(\varepsilon)$ for every $\varepsilon > 0$. The set of all arithmetic probability measures is denoted (by $\mathcal{A}_{\Sigma_{\mathbb{R}}}\subset \mathcal{P}_\Sigma$) by $\mathcal{A}_{\Sigma}\subset \mathcal{P}_\Sigma.$ \end{definition} Next, we define a convex subset of $\mathcal{P}_\Sigma$ that includes $\mathcal{A}_{\Sigma}.$ Let $P(x)$ be a polynomial with complex coefficients of degree $n$. Define its associated root probability measure on the complex plane to be \[ \mu_P:=\frac{1}{n} \sum_{i=1}^{n} \delta_{\alpha_i}, \] where $\alpha_i$ for $1 \leq i\leq n$ are roots of $P$ and $\delta_{\alpha}$ is the delta probability measure at $\alpha.$ Suppose that $\mu\in \mathcal{A}_{\Sigma}.$ Then $\mu_{P_n}\stackrel{\ast}{\rightharpoonup} \mu$ as $n\to\infty$ for some sequence $\{P_n \}$ of distinct irreducible monic polynomials with integral coefficients. Since $P_n$ has real coefficients, its roots are symmetric about the real axis, so $\mu$ should also be symmetric about the real axis. Let $Q$ be a polynomial with integral coefficients. Note that if $P_n \nmid Q,$ \[ \int \log |Q(x)| d\mu_{P_n}(x)= \frac{\log|\text{Res}(Q,P_n)|}{\deg(P_n)} \geq 0, \] where $\text{Res}(Q,P_n) \in \mathbb{Z}$ is the resultant of $P_n$ and $Q.$ By taking $n\to \infty,$ Serre~\cite{MR4093205} proved that \begin{equation}\label{conds} \int \log |Q(x)| d\mu(x) \geq 0 \end{equation} for every non-zero $Q(x)\in \mathbb{Z}[x].$ Let $\mathcal{B}_{\Sigma}\subset \mathcal{P}_{\Sigma}$ be the set of all $\mu\in \mathcal{P}_{\Sigma}$ that are symmetric about the real axis and satisfy~\eqref{conds} for every non-zero $Q(x)\in \mathbb{Z}[x].$ Since $d_{\Sigma}\geq 1,$ it follows that the equilibrium measure of $\Sigma$ belongs to $\mathcal{B}_{\Sigma}$. Note that $\mathcal{B}_{\Sigma}$ is a convex subset of $\mathcal{P}_{\Sigma}$ and $\mathcal{A}_{\Sigma}\subset \mathcal{B}_{\Sigma}.$ Serre~\cite{MR2428512} proved that there exists $\Sigma \subset \mathbb{R}^+$ and $\mu \in \mathcal{B}_{\Sigma}$ such that $\int_{\mathbb{R}} x d\mu<1.8984.$ Smith~\cite{Smith} proved that if $\Sigma\subset \mathbb{R}$ satisfies certain technical conditions (including Serre's construction) then $\mathcal{A}_{\Sigma_{\mathbb{R}}}= \mathcal{B}_{\Sigma}.$ This gaves a definite answer to the Schur--Siegel--Smyth Trace Problem. Finding $\mu \in \mathcal{B}_{\Sigma}$ with the minimal expected value $\int_{\mathbb{R}} x d\mu$ is a linear programming problem and finding its optimal solution is an open problem; see~\cite{Smith}. It is the analog of finding the magic function in the linear programming problem formulated by Cohn-Elkies for the sphere packing problem. \begin{theorem}\label{general} Suppose that $\Sigma\subset \mathbb{C}$ is compact and symmetric about the real axis. If $d_{\Sigma}<1$, then $\mathcal{A}_{\Sigma}= \mathcal{B}_{\Sigma}=\emptyset.$ Otherwise, $d_{\Sigma}\geq 1$ and \(\mathcal{A}_{\Sigma}= \mathcal{B}_{\Sigma}\neq \emptyset.\) Moreover, if $\Sigma\subset \mathbb{R}$ is compact then \(\mathcal{A}_{\Sigma}= \mathcal{B}_{\Sigma}=\mathcal{A}_{\Sigma_{\mathbb{R}}}.\) \end{theorem} \begin{remark} Suppose that $\Sigma \subset \mathbb{R}$ is a finite union of closed intervals with $d_{\Sigma}>1$ and $\mu \in \mathcal{B}_{\Sigma}.$ Smith's main theorem~\cite[Theorem 1.6]{Smith} is equivalent to~\cite[Proposition 2.4]{Smith} the existence of a sequence of conjugate algebraic integers lying inside $\Sigma$ and equidistributing to $\mu.$ We show that $\mathcal{B}_{\Sigma}=\mathcal{A}_{\Sigma_{\mathbb{R}}}$ implies this statement. As we discussed, the circle of radius $r>1$ around the origin where $r$ is not an algebraic integer gives a counter-example~\cite{MR72941}. Knowing this, Smith raised the question of extending his result to $\Sigma\subset \mathbb{C}.$ Let $int_{\Sigma}$ be the interior of $\Sigma.$ If $\mu(int_\Sigma) =1$ and $d_{int_\Sigma}>1$, then Theorem~\ref{general} (with some extra work and using a sequence of open sets inside $int_\Sigma$ and a diagonal argument) implies the existence of a sequence of conjugate algebraic integers lying inside $int_\Sigma$ and equidistributing to $\mu.$ This answers Smith's question when $\mu(int_\Sigma) =1.$ \end{remark} Smith's method does not directly imply Theorem~\ref{general}. For example, his method requires the complement of $\Sigma\subset \mathbb{C}$ to be connected. This condition is trivially satisfied for compact $\Sigma\subset \mathbb{R},$ but fails for some $\Sigma\subset \mathbb{C}$; e.g for the unit circle. We introduce new ideas to prove the above theorem. This is discussed in Section~\ref{method}. \\ Next, we introduce some new notation to state a quantitative version of the Theorem~\ref{general} under some assumptions. Let \begin{equation}\label{logpot} U_{\mu}(z):= \int \log|z-w| d\mu(w), \end{equation} which is called the logarithmic potential of $\mu$ (in other works it is with a negative sign). A measure $\mu$ is a H\"older probability measure if there exists $\delta>0$ and $A>0$ such that \[ \mu([a,b]\times[c,d]) \leq A\max(|b-a|, |c-d|)^{\delta} \] for every $a<b, c<d.$ \\ \begin{comment} Fix a smooth weight function $\omega$ with compact support on the complex plane, where $w(x+iy)\geq 0$ for every $x+iy\in \mathbb{C},$ and \[ \int \omega(x+iy) dxdy=1. \] Let $\omega_\delta(z) = \frac{1}{\delta^2}\omega(\frac{z}{\delta})$ for any $\delta > 0$. The smooth discrepancy between two probability measures $\mu_1$ and $\mu_2$ on the complex plane is defined by \[ \text{disc}_{\delta,\omega}(\mu_1,\mu_2):= \sup_{a<b, c<d} \left|\omega_\delta\ast\mu_1([a,b]\times[c,d])-\omega_\delta\ast\mu_2([a,b]\times[c,d])\right| \] where $[a,b]\times[c,d]:=\{ z\in \mathbb{C}: a \leq \text{Re}(z) \leq b,c \leq \Im(z)\leq d\},$ and $\omega\ast\mu$ is the convolution of $\mu$ and $\omega.$ \\ \end{comment} We cite the following definition from Smith~\cite[Definition 2.7]{Smith}. \begin{definition} Fix a H\"older measure $\mu$ with support contained in the compact subset $\Sigma\subset \mathbb{C}$. Given a complex polynomial $P$ of degree less than or equal to $n$, define the $n$-norm of $P$ with respect to $(\mu,\Sigma)$ by \[ \|P\|_n:=\max_{z\in \mathbb{C}}\left(e^{-nU_{\mu}(z)}|P(z)| \right). \] It follows from~\cite[Theorem 4.1]{logpotentials} that if $\Sigma\subset \mathbb{C}$ is compact and has empty interior with connected complement and for a sequence of monic polynomials $P_n$ of degree $n$ for $n\geq 1$ \[ \limsup_n \|P_n\|_n^{1/n} \leq 1, \] then the roots of $P_n$ are equidistibuted to $\mu.$ \end{definition} \begin{theorem}\label{main1} Suppose $\Sigma\subset \mathbb{C}$ is compact and has empty interior with connected complement. Suppose that $\mu \in \mathcal{B}_{\Sigma}$ is a H\"older measure with exponent $\delta,$ and $D$ is any open set containing $\Sigma$. For every large enough integer $n\geq 1$, there exists an irreducible polynomial $h_n$ of degree n such that all roots of $h_n$ are inside $D$, and \[ \|h_n\|_n\leq e^{n^{1-\delta'}} \] for some $\delta'>0$ which only dependents on the H\"older exponent of $\mu.$ As a result, the roots of $h_n$ are equidistibuted to $\mu.$ Moreover, if $\Sigma\subset \mathbb{R}$ is compact then it is possible to take any open set $D\subset \mathbb{R}$ in the real point set topology containing $\Sigma$ and prove the same result. \end{theorem} \subsubsection{Complexity of approximating arithmetic probability measures}\label{complex} The next goal is to study the complexity of finding a sequence of irreducible monic polynomials $\{P_n\}$ such that $\{ \mu_{P_n} \}$ converges weakly to a given $\mu \in \mathcal{B}_{\Sigma}.$ \begin{theorem}\label{main2} Let $\Sigma$ be as in Theorem \ref{main1}. Suppose that $\mu \in \mathcal{B}_{\Sigma}$ is a H\"older probability measure and $D$ is any open set containing $\Sigma$. There is a polynomial time algorithm in $n$ that returns an integral, irreducible, monic polynomial $g_n$ of degree $n$ such that all roots of $P_n$ are inside $D$, and \[ \|g_n\|_n\leq e^{\frac{Cn\log(\log(n))^3}{\log(n)}} \] where $C > 0$ depends only on $\mu$ and is independent of $n$. As a result, the roots of $g_n$ are equidistibuted to $\mu.$ Moreover, if $\Sigma\subset \mathbb{R}$ is compact then it is possible to take any open set $D\subset \mathbb{R}$ in the real point set topology containing $\Sigma$ and prove the same result. \end{theorem} We implemented a version of this algorithm for arithmetic probability measures $\mu\in \mathcal{B}_{\Sigma}$ constructed by Serre~\cite{MR2428512}. Explicitly, let $\Sigma=[a,b],$ \[ \mu=c\mu_{[a,b]}+(1-c)\nu_{[a,b]}, \] $a=0.1715,$ $ b=5.8255,$ $c=0.5004,$ $\mu_{[a,b]}$ is the equilibrium measure on $[a,b]$, and $\nu_{[a,b]}$ is the pushforward of the equilibrium measure on $[b^{-1},a^{-1}]$ under the map $z\to 1/z.$ Figure~\ref{fig} shows the density of $\mu$, and Figure~\ref{fig2} shows the number of roots of the output polynomial of our algorithm for $n=100$. \begin{figure} \centering \begin{subfigure}[b]{0.4\textwidth} \centering \includegraphics[width=\textwidth]{rootdis.png} \caption{Density function of $\mu$} \label{fig} \end{subfigure} \hfill \begin{subfigure}[b]{0.4\textwidth} \centering \includegraphics[width=\textwidth]{poly_degree_100.png} \caption{Number of roots of $P_{100}(x)$} \label{fig2} \end{subfigure} \caption{This compares the density function of $\mu$ with the histogram of roots of our algorithm's output with input $n=100$.}\label{interval} \end{figure} \subsection{Applications to abelian varieties over finite fields} Recently, Shankar and Tsimerman~\cite{shankar_tsimerman_2018}, \cite{Tsimerman1} conjectured that for every $g\geq 4$, every Abelian variety defined over $\overline{\mathbb{F}_q}$ is isogenous to a Jacobian of a curve defined over $\overline{\mathbb{F}_q}.$ By Honda Tate theory the isogeny class of simple abelian varieties over $\mathbb{F}_q$ and isogenies over $\mathbb{F}_q$ is corresponded to algebraic integers such that all their conjugates have norm $\sqrt{q}.$ We apply our main theorem and Honda Tate theory to prove the following. \begin{theorem} Given any finite field $\mathbb{F}_q$ and any integer $n$, our algorithm returns infinitely many abelian varieties over $\mathbb{F}_q$ which are not isogenous to the Jacobian of any curve over $\mathbb{F}_{q^m}$ for every $1\leq m\leq n$. \end{theorem} Using the algorithm discussed in section \ref{main} of the paper, the authors found an explicit polynomial, with highest order terms being $$x^{290} - 28x^{289} - {484}x^{288} + 20784x^{287} + \cdots,$$ which is irreducible with all roots inside $[-2\sqrt{3},2\sqrt{3}]\subset\mathbb R$ representing an abelian variety that is not isogenous to the Jacobian of any curve over $\mathbb F_3$ or $\mathbb F_9$. \begin{comment} We present our results in this section. Let $P(x)=\prod_{i=1}^{n}(x-\alpha_i)$ be a monic polynomial with real coefficients of degree $n$. We define its associated root probability measure on the complex plane to be \[ \mu_P:=\frac{1}{n} \sum_{i=1}^{n} \delta_{\alpha_i}, \] where $\delta_{\alpha}$ is the delta probability measure at $\alpha.$ \subsubsection{Weak* limit of conjugate algebraic integers} Let $\mu$ be a probability measure with compact support in the complex plane. In this paper, we find necessary and sufficient conditions for $\mu$ to be the weak limit of $\{ \mu_{P_n} \}$ of some sequence $\{P_n \}$ of distinct irreducible monic polynomials with integral coefficients. First, we list some necessary conditions. Suppose $\mu$ is the weak* limit of $\{ \mu_{P_n} \}$ of some sequence $\{P_n \}$ of irreducible monic polynomials with integral coefficients. Since $P_n$ has real coefficients, its roots are invariant under complex conjugation, so $\mu$ should also be invariant under the complex conjugation. The other family of conditions follows from the repulsion of algebraic integers. More precisely, let $Q$ be a polynomial with integral coefficients. We note that if $P_n \nmid Q,$ \[ \int \log |Q(x)| d\mu_{P_n}(x)= \frac{\log|\text{Res}(Q,P_n)|}{\deg(P_n)} \geq 0, \] where $\text{Res}(Q,P_n) \in \mathbb{Z}$ is the resultant of $P_n$ and $Q.$ By taking $n\to \infty,$ it follows that \begin{equation}\label{conds} \int \log |Q(x)| d\mu(x) \geq 0 \end{equation} for every non-zero $Q(x)\in \mathbb{Z}[x].$ Our first result shows that the above inequality implies a stronger version for every integral polynomial in $m\geq 1$ variables. Namely, we show that \eqref{conds} implies \begin{equation}\label{sconds} \int \log |Q_m(x_1,\dots,x_m)| d\mu(x_1)\dots d\mu(x_m) \geq 0 \end{equation} for every non-zero $Q_m(x_1,\dots,x_m)\in \mathbb{Z}[x_1,\dots,x_m].$ Our main theorem states that these necessary conditions are sufficient. \begin{theorem}\label{main1} Suppose that $\mu$ is a \textcolor{blue}{H\"older} probability measure with compact support on the complex plane. The following are equivalent: \begin{enumerate} \item \label{listcond1} $\mu$ is invariant by complex conjugation and~\eqref{conds} holds for every non-zero $Q(x)\in \mathbb{Z}[x].$ \item \label{listcond2} $\mu$ is invariant by complex conjugation and ~\eqref{sconds} holds for every $m$ and every non-zero $Q_m(x_1,\dots, x_m)\in \mathbb{Z}[x_1,\dots,x_m].$ \item \label{listcond3} $\mu$ is the weak* limit of $\{ \mu_{P_n} \}$ for some family $\{P_n \}$ of irreducible monic polynomials with integral coefficients. \end{enumerate} \end{theorem} \subsubsection{Complexity of finding algebraic integers approximating a given distribution} Our next goal is to study the complexity of finding a sequence of irreducible monic polynomials $\{P_n\}$ such that $\{ \mu_{P_n} \}$ converges weakly to a given probability measure $\mu$ satisfying the necessary conditions \eqref{listcond1} in Theorem~\ref{main1}. We introduce some new notations in order to state our next result. The discrepancy of two probability measures $\mu_1$ and $\mu_2$ on the complex plane is defined as \[ \text{Disc} (\mu_1,\mu_2):= \sup_{a<b, c<d} \left|\mu_1([a,b]\times[c,d])-\mu_2([a,b]\times[c,d])\right| \] where $[a,b]\times[c,d]:=\{ z\in \mathbb{C}: a \leq \Re(z) \leq b,c \leq \Im(z)\leq d\}.$ We say $\mu$ is a H\"older probability measure if there exists $\delta>0$ and $C>0$ such that \[ \mu([a,b]\times[c,d]) \leq C\max(|b-a|, |c-d|)^{\delta} \] for every $a<b, c<d.$ \begin{theorem}\label{main2} Suppose that $\mu$ is a H\"older probability measure with compact support satisfying the necessary conditions of Theorem~\ref{main1}. We develop and implement a polynomial time algorithm in $n$ that returns an integral irreducible monic polynomial $P_n$ of degree $n$ such that \[ \text{Disc}(\mu,\mu_{P_n}) \leq \frac{C_1}{\log{n}}, \] where $C_1$ only depends on $\mu.$ \end{theorem} \end{comment} \subsection{Method of proofs}\label{method} In this section, we sketch proofs of our main theorems. Our main tool for studying the distribution of the roots of polynomials is Jensen's formula. Let $P(z)$ be a polynomial with complex coefficients and fix $w\in \mathbb{C}.$ If $z_1, \dots, z_m$ are the roots of $P(z)$ inside the disc $|z-w|<R$ and there is no zero on $|z-w|=R,$ then Jensen's formula states \begin{equation*} \frac{1}{2\pi} \int_{0}^{2\pi} \log |P(Re^{i\theta}+w)| d\theta - \log |P(w)| =\log \frac{R^m}{|z_1-w|\dots |z_m-w|}. \end{equation*} In particular, \begin{equation}\label{jensen} \frac{1}{2\pi} \int_{0}^{2\pi} \log |P(Re^{i\theta}+w)| d\theta - \log |P(w)| \geq 0. \end{equation} Suppose that $P$ is a monic polynomial. We have $|P(w)|=|z_1-w|\cdots |z_{n}-w|$, where $n=\deg(P).$ Hence for a monic polynomial $P$, we have \[ \frac{1}{2\pi} \int_{0}^{2\pi} \frac{\log |P(Re^{i\theta}+w)|}{n} d\theta =\frac{m}{n}\log R+ \frac{1}{n}\sum_{i=m+1}^n\log|z_{i}-w|, \] where $|z_i-w|<R$ for $1\leq i\leq m$ and $|z_i-w|>R$ for $m+1\leq i\leq n.$ We rewrite the above as \begin{equation}\label{Jensen} \frac{1}{2\pi} \int_{0}^{2\pi} \frac{\log |P(Re^{i\theta}+w)|}{n} d\theta = \int \log(z;w,R) d\mu_{P}(z), \end{equation} where \[ \log(z;w,R):=\begin{cases} \log|z-w| &\text{ if } |z-w|>R, \\ \log{R} &\text{otherwise.}\end{cases} \] It follows from \eqref{Jensen} that $\{ \mu_{P_n} \}$ converges weakly to a given probability measure $\mu$ if \begin{equation}\label{conv1} \frac{\log|P_n(z)|}{n}= U_{\mu}(z)+o(1) \end{equation} for every $z\in \mathbb{C}$ such that $|z-z_{i,n}|>e^{-\sqrt{n}}$, where $z_{i,n}$ are complex roots of $P_n.$ We note that $\frac{\log|P_n(z)|}{n}$ is a harmonic function with singularities at roots of $P_n$ and \[\frac{\log|P_n(z)|}{n} = \log|z|+ O\left(\frac{1}{|z|}\right).\] We note that $ U_{\mu}(z)$ has a similar behavior as $|z| \to \infty.$ Since $U_{\mu}(z)$ and $\frac{\log|P_n(z)|}{n}$ are both harmonic functions with similar asymptotic behavior at infinity, it follows from the mean value theorem for harmonic functions that \eqref{conv1} is equivalent to \begin{equation}\label{cvx} \frac{\log|P_n(z)|}{n}\leq U_{\mu}(z)+o(1) \end{equation} for every $z\in \mathbb{C}$ assuming $\mathbb{C}\setminus\Sigma$ is connected and the interior of $\Sigma$ is empty. Let \begin{equation}\label{defKn} K_n:=\left\{ p(x)\in \mathbb{R}[x]: \deg(p)\leq n, \frac{\log |p(z)|}{n} \leq U_{\mu}(z) \text{ for every } z\in \mathbb{C} \right\}. \end{equation} $K_n$ is a symmetric, convex region inside the vector space of polynomials with real coefficients and degree less than or equal to $n$. For proving Theorem~\ref{main1}, we use Minkowski's second theorem to prove the existence of integral polynomials of degree $n$ inside $K_n$ up to a sub-exponential factor in $n$. The key observation is that the necessary condition \eqref{conds} implies $K_n$ is well-rounded which means the successive minima of the lattice of integral polynomials with respect to $K_n$ are close to each other and a lower bound on the volume of $K$ implies the existence of integral polynomials of degree $n$ inside $K_n$ up to a sub-exponential factor in $n$. The following theorem is an important technical result that allows us to use Minkowski's second theorem. \begin{theorem}\label{mdim} Suppose that $\mu$ is a probability measure with compact support on the complex plane. The following are equivalent: \begin{enumerate} \item \label{listcond1} $\mu$ is invariant by complex conjugation and~\eqref{conds} holds for every non-zero $Q(x)\in \mathbb{Z}[x].$ \item \label{listcond2} $\mu$ is invariant by complex conjugation and \begin{equation}\label{sconds} \int \log |Q_m(x_1,\dots,x_m)| d\mu(x_1)\dots d\mu(x_m) \geq 0 \end{equation} for every non-zero $Q_m(x_1,\dots,x_m)\in \mathbb{Z}[x_1,\dots,x_m].$ \end{enumerate} \end{theorem} Recall that \[ \Sigma(\rho):=\{z\in \mathbb{C}: |z-\sigma|<\rho \text{ for some }\sigma\in\Sigma \}. \] Note that for any open set $D$ containing $\Sigma,$ there exist $\rho>0$ such that $\Sigma(\rho)\subset D.$ \\ For $\Sigma\subset \mathbb{C}$, we use Rouch\'e's theorem to show that all roots of our irreducible polynomial are inside $\Sigma(\rho).$ This is similar to the method of Fekete and Szeg\"o~\cite[Theorem I]{MR72941}. \\ For $\Sigma\subset \mathbb{R}$, Robinson used the Chebychev's polynomials and constructed an integral polynomial of degree $n$ with $n$ sign changes on $\Sigma(\rho)\subset \mathbb{R}$ to prove roots are real and are inside $\Sigma(\rho).$ Our results do not rely on the existence of Chebychev's polynomials of $\Sigma$; see subsection~\ref{chebsec} and Proposition~\ref{chpolreal}. In fact, our method gives a new proof of Robinson's result and Smith's results without using the properties of Chebychev's polynomials; see Proposition~\ref{chpolreal}. Our method is based on a initial probabilistic sampling of roots with respect to the equilibrium measure and then deforming the roots with a greedy algorithm along a gradient vector. Proposition~\ref{chpolreal} allows us to improve some results of Smith~\cite[Proposition 4.1]{Smith}. \\ For proving Theorem~\ref{main2}, we apply a lattice algorithm due to Schnorr and find an irreducible, integral polynomial inside $e^{\frac{Cn\log(\log(n))^3}{\log(n)}}K_n$ in polynomial time in $n$ for some constant $C$ dependent on $\mu$. In fact, we take the polynomial to be monic and Eisenstein at the prime 2. Note that this differs from the strongest bound for which we prove existence which is $e^{Cn^{1-\delta_0}}K_n$ for some $\delta_0>0$. Other versions of Schnorr's algorithm can achieve stronger bounds at a super-polynomial run time. \section{Minkowski's theorem applied to the lattice of integral polynomials} \subsection{ Logarithmic potentials} Recall the notation in subsections~\ref{result} and~\ref{method}. In this section, we assume that $\mu$ is supported on a compact set $\Sigma\subset \mathbb{C}$ and is a H\"older probability measure on the complex plane with exponent $\delta>0.$ Hence for every $a<b, c<d$ and some $A>0$ \[ \mu([a,b]\times[c,d]) \leq A\max(|b-a|, |c-d|)^{\delta}. \] Recall the definition of the logarithmic potential of $\mu$ from~\eqref{logpot} $$ U_{\mu}(z)= \int \log|z-w| d\mu(w). $$ We prove a lemma on the asymptotic behaviour of logarithmic potentials. \begin{lemma}\label{lem1} Suppose that $\mu$ is a probability measure with compact support inside $|z|<r$ for $z\in \mathbb{C}.$ If $|w|>2r$, we have \[ U_{\mu}(w)=\log|w|+O\left(\frac{r}{|w|}\right). \] \end{lemma} \begin{proof} We have \begin{align*} \left|U_{\mu}(w)-\log|w|\right|\leq \int \left|\log |w-x| -\log|w| \right|d\mu(x) \\ =\int \left|\log |1-\frac{x}{w}|\right| d\mu(x) \ll \frac{r}{|w|}. \end{align*} \end{proof} \begin{lemma}\label{holderr} Suppose that $\mu$ is a H\"older probability measure on the complex plane with exponent $\delta>0$. Then for any $z_1,z_2\in \mathbb{C}$ with $|z_1-z_2|<1/2,$ we have \[ \left|U_{\mu}(z_1)-U_{\mu}(z_2)\right|\ll |z_1-z_2|^{\delta'}, \] where $\delta'<\min(1/2,\delta/2)$ is any positive exponent. \end{lemma} \begin{proof} Suppose that $z_1,z_2\in \mathbb{C}$ and $|z_1-z_2|<1/2.$ We have \begin{align*} \left|U_{\mu}(z_1)-U_{\mu}(z_2)\right|&=\left| \int \log\frac{|z_1-w|}{|z_2-w|} d\mu(w)\right|. \end{align*} We define the following covering of the complex plane: \begin{align*} U_0:&=\left\{w\in \mathbb{C}: |w-z_2|>2 \sqrt{|z_1-z_2|} \right\}, \\ U_1:&=\left\{w\in \mathbb{C}: |z_1-z_2|<\min(|w-z_2|,|w-z_1|)\leq 2\sqrt{|z_1-z_2|} \right\}, \\ U_k:&= \left\{w\in \mathbb{C}: |z_1-z_2|^k<\min(|w-z_2|,|w-z_1|)\leq |z_1-z_2|^{k-1}\right\}, \end{align*} where $k\geq 2$. Hence, \begin{align*} \left|U_{\mu}(z_1)-U_{\mu}(z_2)\right|\leq \sum_{i\geq 0}\left| \int_{U_i} \log\frac{|z_1-w|}{|z_2-w|} d\mu(w) \right|. \end{align*} For $w\in U_0$, we have $\left|\frac{z_1-z_2}{z_2-w}\right|<1/2$. We write the Taylor expansion of $\log$ and obtain \[ \left| \int_{U_0} \log\frac{|z_1-w|}{|z_2-w|} d\mu(w) \right|=\left| \int_{U_0} \log \left| 1+\frac{z_1-z_2}{z_2-w}\right| d\mu(w) \right| \leq \sum_{m\geq 1}\int_{U_0} \frac{1}{m}\left| \frac{z_1-z_2}{z_2-w}\right|^m d\mu(w)\ll |z_1-z_2|^{1/2}. \] For $w\in U_k$, where $k\geq 1$, we have \[ \left| \int_{U_k} \log\frac{|z_1-w|}{|z_2-w|} d\mu(w) \right| \leq \mu(U_k)k\big|\log|z_1-z_2| \big|\ll |z_1-z_2|^{\frac{k\delta}{2}}k\big|\log|z_1-z_2|\big|, \] where we used $\mu(U_k)\ll |z_1-z_2|^{\frac{k\delta}{2}}$, because $\mu$ is a H\"older probability measure with exponent $\delta>0$. Therefore, \[ \left|U_{\mu}(z_1)-U_{\mu}(z_2)\right|\ll \sqrt{|z_1-z_2|} +\sum_{k\geq 1}|z_1-z_2|^{k\delta/2}k\big|\log|z_1-z_2|\big|\ll |z_1-z_2|^{1/2}+|z_1-z_2|^{\delta/2}\big|\log|z_1-z_2|\big|. \] Therefore, we have \[ \left|U_{\mu}(z_1)-U_{\mu}(z_2)\right|\ll |z_1-z_2|^{\delta'}, \] where $\delta'<\min(1/2,\delta/2)$ is any positive exponent. \end{proof} \subsection{Minkowski's theorem} Recall the definition of $K_n$ from~\eqref{defKn}: \[ K_n=\left\{ p(x)\in \mathbb{R}[x]: \deg(p)\leq n, \text{ and } \frac{\log |p(z)|}{n} \leq U_{\mu}(z) \text{ for every } z\in \mathbb{C} \right\}. \] We identify the space of polynomials with real coefficients of degree less than or equal to $n$ with $\mathbb{R}^{n+1}$ by sending a polynomial to its coefficients. It is easy to see that $K_n$ is a symmetric convex region of $\mathbb{R}^{n+1}$. \begin{lemma}\label{multlem} Suppose that $\mu$ is a probability measure with compact support inside $|z|<r$ for $z\in \mathbb{C}.$ Let $p(x)\in \lambda K_n$ for some $\lambda>0,$ where $\deg(p)<n$. Then \[ xp(x)\in r\lambda K_n. \] \end{lemma} \begin{proof} Let \[ h(z):=U_{\mu}(z)-\frac{\log |zp(z)|}{n}. \] We note that $h(z)$ is a harmonic function on $|z|>r $ and outside roots of $p.$ By Lemma~\ref{lem1}, we have \[ h(z)=\frac{n-\deg(p)-1}{n} \log|z|+\frac{\log|a_p|}{n}+O(1/|z|), \] where $a_p$ is the top coefficient of $p.$ By the above and since $\deg(p)< n$ and $h(z)$ is harmonic, the minimum of $h(z)$ is obtained inside $ |z|<r$. Let $C:=\inf_{z\in \mathbb{C}} h(z)=\inf_{|z|<r} h(z).$ Since $p(x)\in \lambda K_n$, we have \[ |zp(z)| \leq r \lambda e^{n U_{\mu}(z)} \] for any $|z|<r.$ This implies that \[ C\geq \frac{\log r\lambda}{n}. \] Therefore, \[ e^{nU_{\mu}(z)-\log |zp(z)|}= e^{nh(z)}\geq e^{nC}\geq r\lambda \] for any $z\in \mathbb{C}.$ This completes the proof of our lemma. \end{proof} \begin{lemma}\label{derlem} Suppose that $p(x)\in \lambda K_n$ for some $\lambda>0.$ We claim that \[ p'(x)\in An^C\lambda K_n, \] where $p'$ is the derivative of $p$ and $A$ and $C$ are constants that only depend on $\mu.$ \end{lemma} \begin{proof} Let $D(z_0,r)$ be the disk centered at some $z_0\in \mathbb{C}$ and radius $r>0.$ By Bernstein inequality~\cite[Corollary 5.1.6]{MR1367960} for any $r>0$ \[ \sup_{z\in D(z_0,r)} \left|p'(z)\right| \leq \frac{\deg p}{r}\sup _{z\in D(z_0,r)} |p(z)|, \] for some constant $C$ that only depends on $\mu.$ Indeed, we have \[ |p'(z_0)|\leq \frac{n}{r}\sup _{z\in D(z_0,r)} |p(z)| \leq \lambda \frac{n}{r}\sup _{z\in D(z_0,r)} e^{nU_{\mu}(z) }. \] By Lemma~\ref{holderr}, we have for any $z\in D(z_0,r)$ \[ e^{nU_{\mu}(z) } = e^{n(U_{\mu}(z)-U_{\mu}(z_0)) }e^{nU_{\mu}(z_0) } \leq e^{A'nr^{\delta'}}e^{nU_{\mu}(z_0) }. \] for some $A'>0$ and $\delta'>0$ that only depends on $\mu.$ By taking $r=n^{-C'}$ for any $C'>1/\delta'$, it follows that \[ |p'(z_0)|\leq \lambda \frac{n}{r} e^{A'nr^{\delta'}} e^{nU_{\mu}(z_0) } \leq A\lambda n^{C'+1}e^{nU_{\mu}(z_0) }, \] for any $z_0\in \mathbb{C},$ where $A=e^{A'}$. This implies that $ p'(x)\in An^C\lambda K_n,$ where $C=C'+1.$ \end{proof} Let $\Gamma_n$ be the lattice of integral polynomials of degree less than or equal to $n$ in $\mathbb{R}^{n+1}$. The successive minima of $K_n$ on $\Gamma_n$ are defined by setting the $k$-th successive minimum $\lambda_k$ to be the infimum of the numbers $\lambda$ such that $\lambda K_n$ contains $k+1$ linearly-independent vectors of $\Gamma_n$. We have $0 < \lambda_0\leq \lambda_1 \leq \dots \leq \lambda_n <\infty$. Minkowski's second theorem states that \begin{equation}\label{minksec} \frac{2^{n+1}\textup{vol}(K_n)^{-1}}{(n+1)!} \leq \lambda_0\lambda_1\dots \lambda_n\leq 2^{n+1}\textup{vol}(K_n)^{-1}. \end{equation} \begin{proposition}\label{successive} We have \[ \lambda_{m+1}\leq A n^{C}\lambda_{m} \] for any $0\leq m< n$, and some constants $A$ and $C$ that only depend on $\mu.$ \end{proposition} \begin{proof} Suppose that $p_i\in \Gamma_n$ for $0\leq i\leq m$ are the set of linearly independent integral polynomials such that $p_i\in \lambda_i K_n.$ If for some $0\leq i\leq m$, $p_i'$ is linearly independent to $\{p_i:0\leq i\leq m \}$, then by Lemma~\ref{derlem} \[ p'_i(x)\in An^C\lambda_m K_n. \] Hence, \[ \lambda_{m+1}\leq A n^C\lambda_m, \] and this implies our proposition. Otherwise, $V_m:=\text{span}_{\mathbb{R}}\left<p_0,\dots,p_m\right>$ is closed under derivation. It follows that \[ V_m=\text{span}_{\mathbb{R}}\left<1,x,\dots,x^m\right>. \] This implies that there exists $0\leq l\leq m$ such that $\deg(p_j)\leq \deg (p_l)= m$ for every $0\leq j\leq m.$ Let $q(x):=xp_l(x).$ Since $\deg q> m,$ $q$ is linearly independent to $V_m.$ By Lemma~\ref{multlem}, \[ q(x)\in A\lambda_m K_n \] for some constant $A>0$ that only depends on $\mu.$ Hence, \[ \lambda_{m+1}\leq A\lambda_m \] and this completes the proof of our proposition. \end{proof} \subsection{Lower bound on the volume of $K_n$}\label{bound_Kn} In this subsection, we give a lower bound on the Euclidean volume of $K_n$ as a subset of $\mathbb{R}^{n+1}.$ Our method is to find $n+1$ linearly independent polynomials of degree at most $n$ inside $K_n$ and compute the volume of the simplex given by the convex hull of the origin and these $n+1$ points. Since $K_n$ is symmetric, there are $2^{n+1}$ simplexes with the same volume and disjoint interior inside $K_n$ by choosing different signs for the $n+1$ vertices. The total volume of these simplexes give a lower bound for the volume of $K_n.$ \subsubsection{Finding a simplex inside $K_n$} For simplicity in this section, we assume that $n$ is odd and we find a simplex inside $K_n$ and estimate its volume. Recall that $\mu$ is invariant by the complex conjugation. Let $[a,b]\times[-c,c]\subset \mathbb{C}$ be a rectangle containing $\Sigma.$ If $\Sigma\subset \mathbb{R}$, we take $[a,b]\subset \mathbb{R}$ containing $\Sigma$. Fix $0<M\in \mathbb{Z}$ and $0<L \in \mathbb{R}$ where $n^{1/10}<M<n^{1/3}$, and $L<nM^{-2}.$ We partition each side of the rectangle into $2M$ equal length sub-intervals, and obtain a partition of the rectangle as \[ [a,b]\times[-c,c]= \bigcup_{0\leq i,j< 2M}B_{ij}, \] where $B_{ij}:=[a_i,a_{i+1}]\times[c_{j},c_{j+1}]$ and $a_i:=a+\frac{i(b-a)}{2M}$ and $c_j:=-c+\frac{j(2c)}{2M}.$ Since $\mu$ is a H\"older measure with exponent $\delta$, we have \begin{equation}\label{holder} \mu(B_{ij})\ll M^{-\delta} \end{equation} for every $i,j.$ Similarly, for $\Sigma\subset \mathbb{R}$, we write \[ [a,b]=\bigcup_{0\leq i< 2M}B_{i}, \] where $B_{i}:=[a_i,a_{i+1}]$ and $a_i:=a+\frac{i(b-a)}{2M}.$ Let \begin{equation}\label{n_ij} n_{ij}:=\lfloor (n+1)\mu(B_{ij})\rfloor+\varepsilon_{ij} \ll nM^{-\delta}, \end{equation} where $\varepsilon_{ij} \in \mathbb{Z}$ and $|\varepsilon_{ij}| \leq L $ are chosen such that $\sum_{i,j}n_{ij}=n+1$ and $n_{ij}=n_{i(2M-j-1)}.$ Since we assumed $n$ is odd, this is possible. By a covering argument, it is possible to find $z_{ijk}\in B_{ij}$ for $0\leq i,j<2M$ and $1\leq k\leq n_{ij},$ such that $\overline{z_{ijk}}=z_{i(2M-j-1)k}$ and for every $w\in \partial B_{ij}$ and every $1\leq k,k'\leq n_{ij}$ \begin{equation}\label{Bij} \begin{split} |z_{ijk}-w| &\gg M^{-1} \\ |z_{ijk}-z_{ijk'}|&\gg n^{-\frac{1}{2}}M^{-1+\frac{\delta}{2}}\gg n^{-5/6} . \end{split} \end{equation} Similarly, for $\Sigma\subset \mathbb{R}$, we define $n_{i}:=\lfloor (n+1)\mu(B_{i})\rfloor+\varepsilon_{i} \ll nM^{-\delta}$ and $z_{ik}\in B_i$ for $k\leq n_i$ such that $|\varepsilon_{i}|<L,$ $\sum_{i}n_{i}=n+1$ and for every $w\in \partial B_{i}$ and every $1\leq k,k'\leq n_{i}$ \begin{equation}\label{Bireal} \begin{split} |z_{ik}-w| &\gg M^{-1} \\ |z_{ik}-z_{ik'}|&\gg n^{-1} M^{-1+\delta}\gg n^{-4/3} . \end{split} \end{equation} For two compact subsets $B, B'\subset \mathbb{C}$, let \[d(B,B'):=\inf_{\substack{z\in B\\ z'\in B'}} |z-z'|.\]Let \[ I(\mu):=\int \log|z_1-z_2|d\mu(z_1)d\mu(z_2)=\int U_\mu(z)d\mu(z). \] \begin{proposition} We have \begin{equation}\label{diag} \sum_{\substack{ij,i'j' \\ d(B_{ij},B_{i'j'})<M^{-1} }}\sum_{\substack{k,k' \\ i'j'k' \neq ijk}}\frac{\log|z_{ijk}-z_{i'j'k'}|}{(n+1)^2}=O(\log(n)M^{-\delta}), \end{equation} and \begin{equation}\label{maineq} \sum_{ijk}\sum_{ijk\neq i'j'k'} \frac{\log|z_{ijk}-z_{i'j'k'}|}{(n+1)^2}=I(\mu)+O\left(\log(n)M^{-\delta}+ \frac{M^2\log(n)}{n}+ M^{-\delta/2}\log(M)^{1/2}\right). \end{equation} Similarly, for $\Sigma\subset \mathbb{R}$, we have \begin{equation}\label{diagreal} \sum_{\substack{i,i' \\ d(B_{i},B_{i'})<M^{-1} }}\sum_{\substack{k,k' \\ i'k' \neq ik}}\frac{\log|z_{ik}-z_{i'k'}|}{(n+1)^2}=O(\log(n)M^{-\delta}), \end{equation} and \begin{equation}\label{maineqreal} \sum_{ik}\sum_{ik\neq i'k'} \frac{\log|z_{ik}-z_{i'k'}|}{(n+1)^2}=I(\mu)+O\left(\log(n)M^{-\delta}+ \frac{M^2\log(n)}{n}+ M^{-\delta/2}\log(M)^{1/2}\right). \end{equation} \end{proposition} \begin{proof} Note that for a pair $(i,j)$, there are at most $O(1)$ pairs $(i',j')$ where $d(B_{ij},B_{i'j'})<M^{-1}$. By \eqref{n_ij} and \eqref{Bij}, we have \begin{equation*} \sum_{\substack{ij,i'j' \\ d(B_{ij}, B_{i'j'})<M^{-1}}}\sum_{\substack{k,k' \\ i'j'k' \neq ijk}}\frac{\log|z_{ijk}-z_{i'j'k'}|}{(n+1)^2}\ll\sum_{\substack{ij,i'j' \\ d(B_{ij}, B_{i'j'})<M^{-1}}} \frac{n_{ij}n_{i'j'}\log(n)}{(n+1)^2}=O(\log(n)M^{-\delta}). \end{equation*} This completes the proof of the first part of our proposition. We have \begin{equation}\label{summ} I(\mu)= \sum_{ij}\sum_{i'j'}\int_{z\in B_{ij}}\int_{z'\in B_{i'j'}} \log|z-z'|d\mu(z)d\mu(z'). \end{equation} Similarly, we have \[ \sum_{\substack{ij,i'j' \\ d(B_{ij},B_{i'j'})<M^{-1}}}\int_{z\in B_{ij}}\int_{z'\in B_{i'j'}} \log|z-z'|d\mu(z)d\mu(z')=O(\log(n)M^{-\delta}). \] Suppose that $d(B_{ij},B_{i'j'})>M^{-1}.$ We have \begin{multline} \frac{1}{(n+1)^2}\sum_{k}\sum_{ k'} \log|z_{ijk}-z_{i'j'k'}|=\int_{z\in B_{ij}}\int_{z'\in B_{i'j'}} \log|z-z'|d\mu(z)d\mu(z') \\ +O\left(\frac{M_1L}{n}\left(\mu(B_{ij})+\mu(B_{i'j'}) \right)+\mu(B_{ij})\mu(B_{i'j'})M_2\right). \end{multline} where $M_1:=\left|\sup_{\substack{z\in B_{ij}\\ z'\in B_{i'j'}}} \log |z-z'|\right|$ and $M_2:=\sup_{\substack{z_1,z_2\in B_{ij}\\ z_1',z_2'\in B_{i'j'}}}\left|\log|z_1-z_1'|-\log|z_2-z_2'|\right|.$ We have \[ M_1\ll \log(n) \] and \[ M_2\ll \frac{M^{-1}}{d(B_{ij},B_{i'j'})}. \] Therefore, \begin{multline*} \left|\sum_{ijk}\sum_{ijk\neq i'j'k'} \frac{\log|z_{ijk}-z_{i'j'k'}|}{(n+1)^2}-I(\mu) \right| \ll \log(n)M^{-\delta} \\+\sum_{\substack{ij,i'j' \\ d(B_{ij},B_{i'j'})>M^{-1}}} \frac{\log(n)L}{n}\left(\mu(B_{ij})+\mu(B_{i'j'}) \right)+\mu(B_{ij})\mu(B_{i'j'})\frac{M^{-1}}{d(B_{ij},B_{i'j'})}. \end{multline*} Note that \[ \sum_{\substack{ij,i'j' \\ d(B_{ij},B_{i'j'})>M^{-1}}} \frac{\log(n)L}{n}\left(\mu(B_{ij})+\mu(B_{i'j'}) \right) \ll \frac{\log(n)M^2L}{n}. \] By Cauchy-Schwarz inequality, \[ \sum_{ij,i'j'}\mu(B_{ij})\mu(B_{i'j'})\frac{M^{-1}}{d(B_{ij},B_{i'j'})}\leq \left( \sum_{ij,i'j'} \mu(B_{ij})^2\mu(B_{i'j'})\right)^{1/2}\left(\sum_{ij,i'j'}\frac{M^{-2}\mu(B_{i'j'})}{d(B_{ij},B_{i'j'})^2} \right)^{1/2},\] where the sum is over $ij,i'j'$ such that $d(B_{ij},B_{i'j'})\geq M^{-1}.$ We have \[ \sum_{ij,i'j'} \mu(B_{ij})^2\mu(B_{i'j'}) \ll M^{-\delta} \sum_{ij,i'j'} \mu(B_{ij})\mu(B_{i'j'})=M^{-\delta} . \] Moreover, \[ \sum_{ij,i'j'}\frac{M^{-2}\mu(B_{i'j'})}{d(B_{ij},B_{i'j'})^2} \leq \max_{i'j'} \sum_{ij}\frac{M^{-2}}{d(B_{ij},B_{i'j'})^2}\ll \log(M). \] Therefore, \[ \left|\sum_{ijk}\sum_{ijk\neq i'j'k'} \frac{\log|z_{ijk}-z_{i'j'k'}|}{(n+1)^2}-I(\mu) \right| \ll \log(n)M^{-\delta}+\frac{\log(n)M^2L}{n}+M^{-\delta/2}\log(M)^{1/2}. \] \end{proof} Let \begin{equation}\label{measpol} p_{M,n,\mu}(x):=\prod_{i,j,k} (x-z_{ijk}), \end{equation} and for $e,f< 2M$ and $g\leq n_{ef}$ \begin{equation}\label{pmdef} \begin{split} p_{M,n,\mu}^{efg+}(x)&:=\frac{\frac{p_{M,n,\mu}(x)}{(x-z_{efg})}+\frac{p_{M,n,\mu}(x)}{(x-\overline{z_{efg}})}}{2}, \\ p_{M,n,\mu}^{efg-}(x):&=\frac{\frac{p_{M,n,\mu}(x)}{(x-z_{efg})}-\frac{p_{M,n,\mu}(x)}{(x-\overline{z_{efg}})}}{2i\Im(z_{efg})}. \end{split} \end{equation} Similarly, for $\Sigma\subset \mathbb{R}$, let \begin{equation}\label{measpolreal} p_{M,n,\mu}(x):=\prod_{i,k} (x-z_{ik}), \end{equation} and for $e< 2M$ and $g\leq n_{e}$ \begin{equation}\label{pmdefreal} p_{M,n,\mu}^{eg}(x):=\frac{p_{M,n,\mu}(x)}{(x-z_{eg})}. \end{equation} \begin{lemma} We have $p_{M,n,\mu}(x)\in \mathbb{R}[x]$ and \[ p_{M,n,\mu}^{efg+}(x)=\frac{p_{M,n,\mu}(x)}{(x-z_{efg})(x-\overline{z_{efg}})}(x-\Re(z_{efg}))=p_{M,n,\mu}^{e(2M-f-1)g+}(x) \] and \[ p_{M,n,\mu}^{efg-}(x)=\frac{p_{M,n,\mu}(x)}{(x-z_{efg})(x-\overline{z_{efg}})}=-p_{M,n,\mu}^{e(2M-f-1)g-}(x). \] In particular, $p_{M,n,\mu}^{efg\pm}(x)\in \mathbb{R}[x],$ $\deg(p_{M,n,\mu}^{efg+}(x))=n$, and $ \deg p_{M,n,\mu}^{efg-}(x)=n-1$. \end{lemma} \begin{proof} It follows easily from $\overline{z_{ijk}}=z_{i(2M-j-1)k}.$ \end{proof} \begin{proposition}\label{simplex} Suppose that $n \ge 2\max\{|a|,|b|,|c|\}.$ We have \[ p_{M,n,\mu}^{efg\pm}(x) \in \lambda^{efg\pm}K_n \] for some $\lambda^{efg\pm}>0$, where $$\frac{\log(\lambda^{efg\pm})}{n}\leq C\left(\frac{M^2L\log(n)}{n}+ M^{-\delta/2}\log(M)^{1/2} \right)$$ for some constant $C$ that only depends on $\mu.$ In particular, by taking $M=\lfloor n^{1/3}\rfloor$ and $L=n^{1/3-\delta/6},$ we obtain \begin{equation}\label{simplex_potential} |\lambda^{efg\pm}| \leq e^{Cn^{1-\frac{\delta}{6}}\log(n)}. \end{equation} \end{proposition} \begin{proof} We give the proof for $p_{M,n,\mu}^{efg+}(x)$; the other case follows from a similar argument. Note that $\deg(p_{M,n,\mu}^{efg+})=n$ and \[ \frac{\log |p_{M,n,\mu}^{efg+}(z)|}{n} = \int \log|z-x| d\mu_{p_{M,n,\mu}^{efg+}}= U_{\mu_{p_{M,n,\mu}^{efg+}}}(z). \] Hence if $|z|\geq n$ and $n \ge 2\max\{|a|,|b|,|c|\}$, then by Lemma~\ref{lem1}, \[ \frac{\log |p_{M,n,\mu}^{efg+}(z)|}{n} - U_{\mu}(z)=O\left(\frac{1}{n}\right). \] So without loss of generality, we assume that $|z|<n$. For $z\in \mathbb{C},$ we have \begin{multline}\label{sum} \frac{\log |p_{M,n,\mu}^{efg+}(z)|}{n} - U_{\mu}(z)=\sum_{i,j}\left(\sum_{k\leq n_{ij}} \frac{\log |z-z_{ijk}|}{n} -\int_{B_{ij}} \log|z-x|d\mu(x)\right) \\ +\frac{\log \frac{|z-\Re(z_{efg})|}{|z-z_{efg}||z-\overline{z_{efg}}|}}{n}. \end{multline} Let $d(z,B_{ij}):=\inf_{w\in B_{ij}} |z-w|.$ As noted previously, there are at most $O(1)$ pairs $(i,j)$ such that $d(z,B_{ij})<M^{-1}$. First, we give an upper bound on the sum in \eqref{sum} over $i,j$ where $d(z,B_{ij})\leq M^{-1}$. Suppose $z_{hrs}\in B_{ij}$ has minimal distance to $z$ among $z_{ijk}$ where $k\leq n_{ij}$ and $d(z,B_{ij})\leq M^{-1}$. By \eqref{holder}, \eqref{n_ij} and \eqref{Bij}, we have \begin{equation}\label{sum1} \sum_{i,j}\left(\sum_{k\leq n_{ij}} \frac{\log |z-z_{ijk}|}{n} -\int_{B_{ij}} \log|z-x|d\mu(x)\right)=\frac{\log |z-z_{hrs}|}{n}+O(\log(n)M^{-\delta}), \end{equation} where the sum is over $i,j$, where $d(z,B_{ij})\leq M^{-1}.$ This is because by \eqref{Bij}, $|z_{ijk}-z_{ijk'}|\gg n^{-1}.$ Next, we give an upper bound on the above sum over $i,j$, where $d(z,B_{ij})\geq M^{-1}$. Suppose that $d(z,B_{ij})\geq M^{-1},$ we have \[ \left|\sum_{k\leq n_{ij}} \frac{\log |z-z_{ijk}|}{n} -\int_{B_{ij}} \log|z-x|d\mu(x)\right|\leq \frac{M_1L}{n}+\mu(B_{ij})M_2. \] where $M_1:=\left|\sup_{z_1\in B_{ij}} \log |z-z_1|\right|$ and $M_2:=\sup_{z_1,z_2\in B_{ij}}\left|\log|z-z_2|-\log|z-z_1|\right|.$ Because we know $d(z,B_{ij})\geq M^{-1},$ we have \[ M_1\ll \max (|\log(z)|, \log(M))\ll \log(n). \] Suppose that $z_1,z_2\in B_{ij}.$ Let $r_1:=|z-z_1|$ and $r_2:=|z-z_2|$ and assume that $r_1 \leq r_2.$ By the mean value theorem, we have \[ M_2=\frac{\left|r_1-r_2\right|}{r_3} \] for some $r_3,$ where $r_1 \leq r_3 \leq r_2.$ We have $|r_1-r_2|\leq |z_1-z_2|\leq M^{-1}$ and \[ r_3\gg d(z,B_{ij}). \] Hence, \[ M_2\ll \frac{M^{-1}}{d(z,B_{ij})}. \] Therefore, \begin{equation}\label{b1} \sum_{i,j}\left|\sum_{k\leq n_{ij}} \frac{\log |z-z_{ijk}|}{n} -\int_{B_{ij}} \log|z-x|d\mu(x) \right|\ll \frac{M^2L\log(n)}{n}+ \sum_{i,j}\mu(B_{ij})\frac{M^{-1}}{d(z,B_{ij})}, \end{equation} where the sum is over $i,j,$ where $d(z,B_{ij})\geq M^{-1}$. By the Cauchy-Schwarz inequality, \[ \sum_{i,j}\mu(B_{ij})\frac{M^{-1}}{d(z,B_{ij})}\leq \left( \sum_{i,j} \mu(B_{ij})^2\right)^{1/2}\left(\sum_{i,j}\frac{M^{-2}}{d(z,B_{ij})^2} \right)^{1/2}\ll M^{-\delta/2}\log(M)^{1/2},\] where we used $\sum_{i,j}\mu(B_{ij})^2\ll M^{-\delta} \sum_{i,j}\mu(B_{ij})=M^{-\delta}$ and $\sum_{i,j}\frac{M^{-2}}{d(z,B_{ij})^2} \ll \log(M)$ where the sum is over $i,j,$ with $d(z,B_{ij})\geq M^{-1}.$ Therefore by \eqref{b1} and the above inequality, we have \begin{equation}\label{sum2} \sum_{i,j}\left|\sum_{k\leq n_{ij}} \frac{\log |z-z_{ijk}|}{n} -\int_{B_{ij}} \log|z-x|d\mu(x)\right| \ll \frac{M^2L\log(n)}{n}+ M^{-\delta/2}\log(M)^{1/2}, \end{equation} where the sum is over $i,j,$ where $d(z,B_{ij})\geq M^{-1}.$ Finally, by \eqref{sum}, \eqref{sum1} and \eqref{sum2}, we obtain \begin{multline}\label{potapp} \frac{\log |p_{M,n,\mu}^{efg+}(z)|}{n} - U_{\mu}(z)= \frac{\log \frac{|z-z_{hrs}||z-\Re(z_{efg})|}{|z-z_{efg}||z-\overline{z_{efg}}|}}{n} \\ +O\left(\log(n)M^{-\delta}+ \frac{M^2L\log(n)}{n}+ M^{-\delta/2}\log(M)^{1/2}\right). \end{multline} By the definition of $z_{hrs},$ $\frac{|z-z_{hrs}||z-\Re(z_{efg})|}{|z-z_{efg}||z-\overline{z_{efg}}|}\leq 1$ and by letting $M=\lfloor n^{1/3}\rfloor$ and $L=n^{1/3-\delta/6}$, we obtain \begin{equation*} \frac{\log |p_{M,n,\mu}^{efg+}(z)|}{n} \leq U_{\mu}(z)+O(n^{-\frac{\delta}{6}}\log(n)). \end{equation*} \end{proof} \subsubsection{Lower bound on the volume of the simplex} Let \[ A:=\left[p_{M,n,\mu}^{ijk\pm} \right], \] where $0\leq i<2M, 0\leq j< M, k\leq n_{ij}$ be the square matrix of size $(n+1)\times (n+1)$ with coefficients of $p_{M,n,\mu}^{efg\pm}$ as its column vectors. \begin{proposition}\label{detprop} We have \[ |\det(A)|=2^{-\frac{n+1}{2}}\prod_{ijk} \frac{1}{|\Im(z_{ijk})|^{1/2}}\prod_{ijk< i'j'k'} |z_{ijk}-z_{i'j'k'}|, \] where $0\leq i<2M, 0\leq j< 2M, k\leq n_{ij}$ \end{proposition} \begin{proof} Let \[ B:=\left[\frac{p_{M,n,\mu}(x)}{(x-z_{ijk})}\right], \] where $0\leq i<2M, 0\leq j\leq M-1, k\leq n_{ij}$ be the square matrix of size $(n+1)\times (n+1)$ with coefficients of $\frac{p_{M,n,\mu}(x)}{(x-z_{ijk})}$ as its column vectors. By \eqref{pmdef} for fixed indices $efg$, we have \[ \left[p_{M,n,\mu}^{efg+},p_{M,n,\mu}^{efg-} \right]=\left[\frac{p_{M,n,\mu}(x)}{(x-z_{efg})},\frac{p_{M,n,\mu}(x)}{(x-\overline{z_{efg}})}\right] \begin{bmatrix}1/2 & \frac{1}{2i\Im(z_{efg})} \\ 1/2 & -\frac{1}{2i\Im(z_{efg})} \end{bmatrix}. \] Hence, \begin{equation}\label{detA} |\det(A)|=|\det(B)|2^{-\frac{n+1}{2}}\prod_{ijk} \frac{1}{|\Im(z_{ijk})|^{1/2}} \end{equation} where $0\leq i<2M, 0\leq j< 2M, k\leq n_{ij}.$ Let \[ V:=[z_{ijk}^e] \] be the Vandermonde matrix of size $(n+1)\times (n+1),$ where $0\leq e<n+1.$ It is well-known that \[ |\det V|=\prod_{ijk}\prod_{ijk< i'j'k'} |z_{ijk}-z_{i'j'k'}|. \] We have \[ VB=\textup{diag}\left[\prod_{ijk\neq efg} (z_{efg}-z_{ijk})\right]. \] Hence, \[ |\det(B)|=\prod_{ijk}\prod_{ijk< i'j'k'} |z_{ijk}-z_{i'j'k'}|. \] Therefore, by \eqref{detA} \[ |\det(A)|=2^{-\frac{n+1}{2}}\prod_{ijk} \frac{1}{|\Im(z_{ijk})|^{1/2}}\prod_{ijk< i'j'k'} |z_{ijk}-z_{i'j'k'}|. \] \end{proof} \begin{proposition}\label{vol_K_n} We have \[ \textup{vol}(K_n) \gg n^{-cn^{2-\frac{\delta}{6}}}e^{\frac{n^2}{2}I(\mu)} \] where $c$ is a constant dependents only on $\mu$. \end{proposition} \begin{proof} By Proposition~\ref{simplex}, $e^{Cn^{1-\frac{\delta}{6}}\log(n)}K_n$ includes the simplex with vertices $\left\{0, p_{M,n,\mu}^{ijk\pm} \right\}$ with volume $\frac{|\det(A)|}{(n+1)!}$, where $0\leq i<2M, 0\leq j\leq M-1, k\leq n_{ij}.$ Since $K_n$ is symmetric, there are $2^{n+1}$ simplexes with the same volume and disjoint interior inside $e^{Cn^{1-\frac{\delta}{6}}\log(n)}K_n$ by choosing different signs for the $n+1$ vertices $\left\{0,\pm p_{M,n,\mu}^{ijk\pm} \right\}$. Hence, \[ \textup{vol}(K_n) \geq \frac{2^{n+1}}{(n+1)!}e^{-Cn^{2-\frac{\delta}{6}}\log(n)}|\det(A)|. \] By Proposition~\eqref{detprop}, \[ \textup{vol}(K_n) \geq \frac{2^{\frac{n+1}{2}}}{(n+1)!}e^{-Cn^{2-\frac{\delta}{6}}\log(n)}\prod_{ijk} \frac{1}{|\Im(z_{ijk})|^{1/2}}\prod_{ijk< i'j'k'} |z_{ijk}-z_{i'j'k'}|. \] We have that $|\Im(z_{ijk})| \le c$ where $\text{supp}(\mu)\in [a,b]\times[-c,c]$ with $c\ge 1$. Therefore $$\prod_{ijk}\frac{1}{|\Im(z_{ijk})|^{1/2}}\prod_{ijk<i'j'k'}|z_{ijk}-z_{i'j'k'}|\ge c^{-n/2}\prod_{ijk<i'j'k'}|z_{ijk}-z_{i'j'k'}|.$$ From \eqref{maineq} using $M=\lfloor n^{1/3}\rfloor$, we have $\prod_{ijk<i'j'k'}|z_{ijk}-z_{i'j'k'}|= e^{\frac{n^2}{2}I(\mu)+O(n^{2-\frac{\delta}{6}}\log(n))}$ and so for some (potentially negative) constant $C'$, this is at least $e^{\frac{n^2}{2}I(\mu)+C'n^{2-\frac{\delta}{6}}\log(n)}$. This gives $$\textup{vol}(K_n)\gg \frac{2^\frac{n+1}{2}}{(n+1)!c^{n/2}}e^{\frac{n^2}{2}I(\mu)+(C'-C)n^{2-\frac{\delta}{6}}\log(n)}\gg \frac{n^{(C'-C)n^{2-\frac{\delta}{6}}}e^{\frac{n^2}{2}I(\mu)}}{\left(n+1\right)^{n+1}c^n}\gg n^{\tilde C n^{2-\frac{\delta}{6}}}e^{\frac{n^2}{2}I(\mu)}$$ since $\frac{(n+1)^{n+1}}{n^n}\sim (n+1)e$ for some constant $\tilde C$. \end{proof} \begin{proposition}\label{mink} Suppose that $\mu \in \mathcal{B}_{\Sigma}.$ There exists an integral polynomial $p_0$ with $\deg(p_0)\leq n$ such that \[ \frac{\log |p_0(x)|}{n} \leq U_{\mu}(x)-\frac{I(\mu)}{2}+C\log(n)n^{-\delta/6}, \] where $C$ is a constant independent of $n.$ Moreover, \[ I(\mu)\geq 0. \] \end{proposition} \begin{proof} By Proposition ~\eqref{vol_K_n} there are constants $c,c'$ such that for all $n>0$, $\textup{vol}(K_n) \ge c'n^{cn^{2-\frac{\delta}{6}}}e^{\frac{n^2}{2}I(\mu)}$. By Minkowski's second theorem~\eqref{minksec}, there exists $p_0\in\mathbb Z[x]\cap \lambda_0K_n$ of degree at most $n$, where \begin{equation}\label{upperb} \lambda_{0}\leq 2 \left( \textup{vol}(K_n) \right)^{-1/(n+1)} \ll n^{cn^{1-\frac{\delta}{6}}}e^{-\frac{n}{2}I(\mu)}. \end{equation} By definition of $K_n,$ \[ \frac{\log |p_0(x)|}{n} \leq U_{\mu}(x)+ \frac{\log(\lambda_0)}{n} \] By using~\eqref{upperb}, we obtain \[ \frac{\log |p_0(x)|}{n} \leq U_{\mu}(x)-\frac{I(\mu)}{2}+c\log(n)n^{-\frac{\delta}{6}}. \] By taking the expected value of the above inequality, we have \[ 0\leq \int \frac{\log |p_0(x)|}{n} d\mu(x) \leq \frac{I(\mu)}{2}+c\log(n)n^{-\frac{\delta}{6}}, \] where we used the assumption $\mu \in \mathcal{B}_{\Sigma}$ for the first inequality. By letting $n\to \infty$, we have \[ I(\mu)\geq 0. \] This completes the proof of our proposition. \end{proof} \subsection{Proof of Theorem~\ref{mdim} } In this section, we prove Theorem~\ref{mdim}. Suppose that $\mu$ is a probability measure with compact support on the complex plane and that for every non-zero $Q(x)\in \mathbb{Z}[x]$, \[ \int \log |Q(z)| d\mu(z) \geq 0. \] First, we reduce the proof of Theorem~\ref{mdim} to the case of H\"older probability measure by the following proposition. \begin{proposition}\label{holderreduction} Suppose that $\Sigma\subset \mathbb{C}$ is compact and $\mu\in \mathcal{B}_{\Sigma}.$ For any $\rho>0,$ there exists a sequence of H\"older probability measure $\mu_n\in\mathcal{B}_{\Sigma(\rho)}$ such that \( \lim_{n\to \infty} \mu_n=\mu\). \end{proposition} \begin{proof} Let $\lambda_{r}$ be the uniform probability measure on the circle of radius $r$ centered at the origin. Let $\mu*\lambda_r$ be the convolution of $\mu$ with $\lambda_r.$ We show that $U_{\mu*\lambda_r}(z)\geq U_{\mu}(z)$ for every $z\in \mathbb{C}$ and every $r\geq 0.$ In fact, we have \[ U_{\mu*\lambda_r}(z)-U_{\mu}(z)=\int \left( \int_{0}^1\log|z-x+re^{2\pi i \theta}|- \log|z-x| d\theta \right)d\mu(x)\geq 0, \] where we used Jensen's formula~\eqref{jensen} to show the inner integral is positive. Therefore, $\mu*\lambda_r \in \mathcal{B}_{\Sigma(\rho)}$ for any $r\leq \rho.$ Define \[ \mu_n:=\mu*\lambda_{\rho/n}. \] It is clear from the definition that \( \lim_{n\to \infty} \mu_n=\mu\) and $\mu_n$ is a H\"older measure with exponent at least 1. This completes the poof of our proposition. \end{proof} \begin{proof}[Proof of Theorem~\ref{mdim}] Part~\eqref{listcond1} is the special case of part~\eqref{listcond2} for $m=1.$ So, it is enough to show that part~\eqref{listcond1} implies part~\eqref{listcond2}. Suppose that $\mu$ satisfies the conditions of part part~\eqref{listcond1}, which is equivalent to $\mu\in \mathcal{B}_{\Sigma}$ by definition. Next, we reduce the theorem to the case where $\mu$ is H\"older. By Proposition~\ref{holderreduction}, there exists a sequence of H\"older probability measure $\mu_n\in\mathcal{B}_{\Sigma(\rho)}$ such that \( \lim_{n\to \infty} \mu_n=\mu\). Suppose that Theorem~\ref{mdim} holds for H\"older probability measures. Then \[ \int \log |Q(z_1,\dots,z_m)|d\mu_n(z_1)\dots d \mu_n(z_m)\geq0, \] for every $n.$ Note that $\log |Q(z_1,\dots,z_m)|$ is an upper semicontinuous function. Hence, by monotone convergence theorem, we have \[ \int \log |Q(z_1,\dots,z_m)|d\mu(z_1)\dots d \mu(z_m) \geq \limsup_{n\to \infty }\int \log |Q(z_1,\dots,z_m)|d\mu_n(z_1)\dots d \mu_n(z_m)\geq 0. \] So without loss of generality we assume that $\mu\in \mathcal{B}_{\Sigma}$ is H\"older. The proof is by induction on $m,$ and part~\eqref{listcond1} is the base of our induction hypothesis. We assume that part~\eqref{listcond2} holds for $m$ and we show that it holds for $m+1.$ Suppose that $Q(x_1,\dots,x_{m+1})\in \mathbb{Z}[x_1,\dots,x_{m+1}].$ Without loss of generality, we assume that $Q$ is irreducible in $\mathbb{Z}[x_1,\dots,x_{m+1}]$ and does not belong to $\mathbb{Z}[x_{m+1}].$ Let \[ U_{Q}(z_{m+1}):=\int \log |Q(z_1,\dots,z_m,z_{m+1})|d\mu(z_1)\dots d\mu(z_m). \] It is enough to show that \begin{equation}\label{post} \int U_{Q}(z) d\mu(z)\geq 0. \end{equation} We fix $z_i\in \mathbb{C}$ for $1\leq i\leq m$ and consider $Q$ as a polynomial in $x_{m+1}$ variable and assume that it has degree $d$ in $x_{m+1}.$ Then \[ Q(z_1,\dots,z_m,x_{m+1})=\sum_{i=0}^d a_i(z_1\dots,z_m)x_{m+1}^i=a_d(z_1,\dots,z_m)\prod_{i=1}^d (x_{m+1}-\xi_i(z_1,\dots,z_{m})) \] where $\xi_i(z_1,\dots,z_{m}) \in \mathbb{C}$ are complex roots of $Q$. We define the measure $\mu_Q$ on $\mathbb{C}$ as follows \begin{equation}\label{defmuq} \mu_Q:= \int \sum_{i} \delta_{\xi_i(z_1,\dots,z_{m})} d\mu(z_1)\dots d\mu(z_m) \end{equation} where $\delta_{a}$ is the delta probability measure at point $a\in \mathbb{C}.$ We have \[ U_Q(z)=a_Q+ \int \log|z-x|d\mu_{Q}(x) \] where \begin{equation}\label{defaq} a_Q:=\int \log|a_d(z_1,\dots,z_m)|d\mu(z_1)\dots d\mu(z_m). \end{equation} Note that by induction hypothesis, we have \[ a_Q\geq 0. \] By Proposition~\ref{mink}, there exists an integral polynomial $P_0(x)$ with $\deg(P_0)\leq n$ such that \[ \frac{\log |P_0(x)|}{n} \leq \int\log |z-x| d\mu(z)+Cn^{-\delta}. \] We take the expected value of the above inequality with respect to $d\mu_Q$ and obtain \[ a_Q+\int \frac{\log|P_0(x)|}{n} d\mu_Q(x) \leq a_Q+\int \log|z-x| d\mu(z)d\mu_{Q}(x) +\frac{Cd}{n^{\delta}}=\int U_{Q}(z) d\mu(z)+\frac{Cd}{n^{\delta}}. \] It is enough to show that \begin{equation}\label{pn} na_Q+\int \log |P_0(x)| d\mu_Q(x) \geq 0, \end{equation} since \eqref{post} follows by letting $n\to \infty.$ In fact, we prove that for every integral polynomial $P(x),$ \begin{equation}\label{pns} \deg(P)a_Q+\int \log |P(x)| d\mu_Q(x) \geq 0. \end{equation} This implies \eqref{pn}, since $\deg(P_n)\leq n$ and $a_Q\geq0.$ For proving~\eqref{pns}, we consider $P$ as an integral polynomial in $x_{m+1}$. Define \[ R(Q,P)(x_1,\dots,x_m):= Res(Q(x_1,\dots,x_{m+1}),P(x_{m+1}))\in \mathbb{Z}[x_1,\dots,x_m] \] to be the resultant of $Q$ and $P$ as polynomials in $x_{m+1}$ variable. By our assumption, $Q$ is irreducible and does not divide $P(x_{m+1}).$ Hence, $Res(Q(x_1,\dots,x_{m+1}),P(x_{m+1}))\neq 0.$ By our induction hypothesis, \[ \int \log R(Q,P)(z_1,\dots,z_m)d\mu(z_1)\dots d\mu(z_m)\geq 0. \] We note that \[ \log R(Q,P)(z_1,\dots,z_m)=\deg(P)\log|a_d(z_1,\dots,z_d)| + \sum_{i=1}^d \log|P(\xi_i(z_1,\dots,z_{m}))| \] Hence, \[ \int \deg(P)\log|a_d(z_1,\dots,z_d)| + \sum_{i=1}^d \log|P(\xi_i(z_1,\dots,z_{m}))|d\mu(z_1)\dots d\mu(z_m)\geq 0. \] The above implies~\eqref{pns} by definitions of $\mu_Q$ and $a_Q$ in \eqref{defmuq} and \eqref{defaq}. This complete our proof and implies \[ \int \log |Q(z_1,\dots,z_m,z_{m+1})|d\mu(z_1)\dots d \mu(z_m)d\mu(z_{m+1})\geq0. \] \end{proof} \subsection{Lower bound on the product of Minkowski's minima factors} First, we give a lower bound on the product of the Minkowski minima factors by using Theorem~\ref{mdim}. \begin{proposition}\label{lowerd} We have \[ \prod_{i=0}^{m-1} \lambda_i \geq \frac{1}{m!} e^{(-nm+\frac{(m-1)m}{2})I_\mu}. \] \end{proposition} \begin{proof} By definition of $\{\lambda_i: 0\leq i\leq m-1\}$, there exists linearly independent integral polynomials $\{P_0,\dots,P_{m-1}\}$, where \[ |P_i(z)|\leq \lambda_i e^{nU_\mu(z)}. \] for every $z\in \mathbb{C}.$ Let $\{z_1,\dots,z_m \}\subset \mathbb{C}.$ Therefore, \begin{equation}\label{detinq} \det [P_i(z_j)]\leq m!\prod_{i=0}^{m-1} \lambda_i e^{nU_\mu(z_i)}. \end{equation} On the other hand, we have \[ \det [P_i(x_j)]=Q_m(x_1,\dots,x_m)\prod_{i<j} (x_i-x_j) \] for some $Q_m\in \mathbb{Z}[x_1,\dots,x_m].$ By taking the average of the above, we obtain \[ \int \log |\det [P_i(z_j)]| d\mu(z_1)\dots d\mu(z_m)=\frac{(m-1)m}{2}I_{\mu}+\int \log |Q_m(z_1,\dots,z_m)| d\mu(z_1)\dots d\mu(z_m), \] where \[ I(\mu):=\int \log|z_1-z_2|d\mu(z_1)d\mu(z_2). \] By second part of Theorem~\eqref{mdim}, we have \[ \int \log |Q_m(z_1,\dots,z_m)| d\mu(z_1)\dots d\mu(z_m)\geq 0. \] Hence, \[ \int \log |\det [P_i(z_j)]| d\mu(z_1)\dots d\mu(z_m)\geq \frac{(m+1)m}{2}I_{\mu}. \] By~\eqref{detinq}, we obtain \[ mnI_{\mu}+\sum_{i} \log |\lambda_i|+ \log (m!)\geq\int \log |\det [P_i(z_j)]| d\mu(z_1)\dots d\mu(z_m)\geq \frac{(m+1)m}{2}I_{\mu} \] Therefore, \[ \prod_i \lambda_i \geq \frac{1}{m!} e^{(-nm+\frac{(m-1)m}{2})I_\mu}. \] \end{proof} \begin{proposition}\label{upperbound} We have \[ \lambda_n \leq n^{cn^{1-\frac{\delta}{12}}}. \] \end{proposition} \begin{proof} Let $m=n-n^{\delta_1}$ where $0<\delta_1<1$ is specified later. By Proposition~\ref{lowerd}, \[ \prod_{i=0}^m \lambda_i \geq \frac{1}{m!} e^{-nm+\frac{(m-1)m}{2}I_\mu}. \] By Minkowski's second theorem and Proposition~\ref{vol_K_n}, \[ \lambda_0\lambda_1\dots \lambda_n\leq 2^{n+1}\textup{vol}(K_n)^{-1} \ll2^{n+1} n^{cn^{2-\frac{\delta}{6}}}e^{-\frac{n^2}{2}I(\mu)}. \] By combining the above inequalities, we have \[ \prod_{i=m+1}^n \lambda_i \ll n^{cn^{2-\frac{\delta}{6}}} e^{-\frac{(n-m)^2}{2}I(\mu)}. \] By Proposition~\ref{mink}, $I(\mu)\geq 0,$ and we have \[ \prod_{i=m+1}^n \lambda_i \ll n^{cn^{2-\frac{\delta}{6}}}. \] Hence, by taking $\delta_1=1-\delta/12,$ we obtain \[ \lambda_{m+1}\leq n^{cn^{2-\frac{\delta}{6}-\delta_1}} \leq n^{cn^{1-\frac{\delta}{12}}}. \] By Proposition~\ref{successive}, we have \[ \lambda_n \leq \lambda_m (An)^{C(n-m)} \leq n^{cn^{1-\frac{\delta}{12}}}. \] This completes the proof of our proposition. Note that we may change constant $c$ to a larger constant in the above lines of argument. \end{proof} \subsection{Equilibrium measure}\label{chebsec} In this subsection, we apply the results of our previous section to the equilibrium measure and prove Proposition~\ref{chpol}. Our proof is based on Fekete and Szeg\"o~\cite[Theorem D]{MR72941} which is also used by Robinson~\cite{Robinson} and Smith~\cite[Proposition 4.1]{Smith} to prove a similar result. All these results are essential in controlling the roots to remain inside $\Sigma(\rho).$ \\ Suppose that $\Sigma$ is compact and its complement is open and connected. Let $D$ be any open set containing $\Sigma$. Let \[ \Sigma(\rho):=\{z\in \mathbb{C}: |z-\sigma|<\rho \text{ for some }\sigma\in\Sigma \}, \] where $\Sigma(\rho)\subset D.$ Note that for any open set $D$ containing $\Sigma,$ there exist $\rho>0$ such that $\Sigma(\rho)\subset D.$ Let $\mu_{eq}$ be the equilibrium measure of $\Sigma.$ Fix any $\rho>0$ and $z_0\in \Sigma.$ It is well known that \begin{equation}\label{eqbd} U_{\mu_{eq}}(z)\geq \log (d_{\Sigma})+\delta_{\rho,\Sigma} \end{equation} for any $z\notin\Sigma(\rho),$ where $\delta_{\rho,\Sigma}>0 $ is a constant that only depends on $\Sigma$ and $\rho.$ Let $p_{M,n,\mu_{eq}}(x)$ be the polynomial constructed in~\eqref{measpol} associated to the equilibrium measure. By~\eqref{potapp} and~\eqref{eqbd} there exists some $R>d_{\Sigma}$ and $M_0,n_0\in\mathbb{Z}$ such that \[ p_{M_0,n_0,\mu_{eq}}(z)\geq R^{n_0} \] for any $z\notin\Sigma(\rho).$ Fix $z_0\in \Sigma(\rho)$ and $p_{M_0,n_0,\mu_{eq}}(z),$ and define \begin{equation}\label{defTm} T_m(x):=(x-z_0)^rp_{M_0,n_0,\mu_{eq}}(x)^q, \end{equation} where $m=n_0q+r$ is the Euclidean division of $m$ by $n_0$. Let $m\geq h$ be two positive integers. It follows that \begin{equation}\label{cheby} T_m(z)\geq CR^{m-h}T_h(z) \end{equation} for any $z\notin\Sigma(\rho),$ where $C$ only depends on fixed parameters $\rho$ and $n_0.$ \begin{proposition}\label{chpol} Given compact subset $\Sigma\subset \mathbb{C}$ with connected complement, a monic polynomial with real coefficients $p(x)$ and $\rho>0$, there exists fixed integers $l_0$ and $k_0$ independent of $p(x)$ and a monic polynomial $Q_m(x)$ with $\deg(Q_m)=m$ for every $m=kl_0,$ where $k>k_0$ such that \[ |Q_m(z)|\geq \kappa^m \] for every $z\notin \Sigma(\rho)$ where $\kappa>d_{\Sigma}.$ Moreover, the coefficient of $x^i$ in $p(x)Q_m(x)$ is an even integer for every $\deg(p)\leq i\leq \deg(p)+m-1$, and all complex roots of $Q_m(x)$ are inside $\Sigma(\rho).$ \end{proposition} \begin{proof} Recall $T_m(x)$ in~\eqref{defTm} for $m\in \mathbb{Z},$ and define \[ q_{k,l}(x):=\left(T_k(x)+a_1T_{k-1}(x)+\dots+a_kT_0(x) \right)^l \] where $|a_i|\leq \frac{1}{l}$ are chosen recursively for $1\leq i\leq k$ such that $q_{k,l}(x)p(x)$ has even coefficients for $x^{\deg(p)+\deg(q_{k,l})-i},$ where $1\leq i\leq k.$ Note that this is possible since the coefficient of $x^{\deg(p)+\deg(q_{k,l})-j}$ for $j\leq l$ in $q_{k,l}(x)p(x)$ is given by \[ la_j+F_j(a_1,\dots,a_{j-1}) \] where $F_j(a_1,\dots,a_{j-1})$ is a polynomial in terms of $a_i$ for $i<j.$ By~\eqref{cheby}, we have \begin{align*} |q_{k,l}(z)|&=\left|\left(T_k(z)+a_1T_{k-1}(z)+\dots+a_kT_0(z) \right)^l\right| \\ &\geq \left|T_k(z)\left(1-\frac{C}{l}\sum_{j\geq 1}\frac{1}{R^j}\right)\right|^l \geq |T_k(z)|^l\left(1-\frac{C}{l(R-1)}\right)^{l}. \end{align*} We fix an integer $l_0> \frac{C}{R-1}$, and let \[ C':=\left(1-\frac{C}{l_0(R-1)}\right)^{l_0}>0. \] Hence, \begin{equation}\label{fineq} |q_{k,l_0}(z)| \geq C' T_k(z)^{l_0}\geq C''|p_{M_0,n_0,\mu_{eq}}(z)|^{\left\lfloor\frac{k}{n_0}\right\rfloor l_0} \end{equation} where $C''$ is a fixed constant depending on our fixed parameters. Finally, let $m:=l_0k$ and \[ Q_m(x):=q_{k,l_0}(x)-\sum_{i=0}^{m-k}b_iT_i(x), \] where $|b_i|\leq 1$ are chosen recursively for $0\leq i\leq m-k$ such that $Q_m(x)p(x)$ has even coefficients for $x^{j},$ where $\deg(p)\leq j\leq \deg(p)+ m-k.$ By \eqref{fineq}, \eqref{cheby}, and \eqref{defTm}, we have \begin{align*} |Q_m(z)|\geq |q_{k,l_0}(z)|-\sum_{i=0}^{m-k}|T_i(z)|\geq |p_{M_0,n_0,\mu_{eq}}(z)|^\frac{kl_0}{n_0} \left( C''-\frac{C_1}{R^{k-1}(R-1)} \right). \end{align*} Our proposition follows by taking $k>k_0$ where $k_0$ is a fixed constant. \end{proof} Let $\Sigma\subset \mathbb{R}$ be a finite union of closed intervals. For applications to abelian varieties and the trace problem, we use the following proposition. The statement of this proposition is motivated by~\cite[Proposition 4.1]{Smith} which is based on Robinson~\cite{Robinson}. All these results in Robinson's~\cite{Robinson} Smith's\cite{{Smith}}, and our work are essential in controlling the roots to remain inside $\Sigma_{\mathbb{R}}(\rho).$ Smith and Robinson used the properties of the Chebyshev's polynomial to prove their result and their method are very similar to the proof of Proposition~\ref{chpol}. Our proof is conceptually different. \begin{proposition}\label{chpolreal} Suppose that $\Sigma\subset \mathbb{R}$ is a finite union of closed intervals with $d_{\Sigma}\geq 1$. Given a monic polynomial with real coefficients $p(x)$ of degree $n$, $\rho>0$, and every $m\leq n,$ there exists a monic polynomial $Q_m(x)$ with $\deg(Q_m)=m$ such that \begin{enumerate} \item The degree $i$ coefficient of the product $pQ_m$ is an even integer for $n-m\leq i\leq n-1$. \item \label{2deform} Take $X$ to be the set of roots of $Q_m$. Then $X$ is a subset of $\Sigma_{\mathbb{R}}(\rho)\subset \mathbb{R}$, and we have \begin{equation*} \frac{\log |Q_{m}(z)|}{m} =\frac{\min_{\alpha\in X}\log|z-\alpha|}{m}+ U_{\mu_{eq}}(z)+O(m^{-\frac{\delta}{6}}\log(m)) \end{equation*} for all $z\in \mathbb{C}$ and some $\delta>0,$ where $\mu_{eq}$ is the equilibrium measure for $\Sigma_{\mathbb{R}}(\rho/10)\subset \mathbb{R}.$ \item~\label{3deform} Given any root $\alpha$ of $Q_m$, and given any $\alpha'$ that is either a root of $p$ or a boundary point of $\Sigma_{\mathbb{R}}(\rho)$, we have \( n^{-3} \leq |\alpha-\alpha'|. \) \end{enumerate} \end{proposition} \begin{comment} We cite a version of the previous proposition from Smith's work~\cite[Proposition 4.1]{Smith}. We note that Smith's proof is constructive when the Chebyshev's polynomials of $\Sigma$ are explicit. For example when $\Sigma=[-2,2]\subset \mathbb{R}$, the $n$-th degree Chebyshev's polynomial is~\cite{Robinson}: \[ T_n(x)=x^n+\sum_{k=1}^{\lfloor n/2\rfloor}(-1)^k\frac{n}{k}\binom{n-k-1}{k-1}x^{n-2k}. \] For applications to abelian varieties and Siegel's trace problem, we use the following proposition for closed intervals of real line in our proofs and the implementation of our algorithm. \begin{proposition}\cite[Proposition 4.1]{Smith}\label{chpolreal} Choose a compact finite union of intervals $\Sigma$ with $d_\Sigma>1.$ Then there is a positive integer $D_0$ and positive real $C$ so we have the following: \\ Choose positive integers $m$ and $n$ satisfying $m < n < d_{\Sigma}^{m/2}$, with $m$ divisible by $D_0$ and greater than $C$. Choose any monic real polynomial $P$ of degree $n-m$. Then there is a monic real polynomial $Q$ of degree $m$ satisfying the following three conditions: \begin{enumerate} \item The degree $i$ coefficient of the product $PQ$ is an even integer for $n-m\leq i\leq n-1$. \item Take $X$ to be the set of roots of $Q$.Then $X$ is a subset of $\Sigma$,and we have \[ d_{\Sigma}^mn^{-C} \min_{\alpha\in X}|x-\alpha| \leq |Q(x)| \leq d_{\Sigma}^mn^C \] for all $x\in \Sigma$. \item Given any root $\alpha$ of $Q$, and given any $\alpha'$ that is either a root of $P$ or a boundary point of $\Sigma$, we have \[ n^{-C} \leq |\alpha-\alpha′|. \] \end{enumerate} \end{proposition} \end{comment} \begin{proof} Let $P_{M,m,\mu_{eq}}$ be the polynomial constructed in~\eqref{measpolreal} associated to the equilibrium measure for $\Sigma_{\mathbb{R}}(\rho/10)$ with roots $x_1,\dots,x_m,$ such that $M=m^{1/3},$ and by a covering argument we pick $x_1,\dots,x_m$ such that for every $x_i$ we have \( n^{-3} \leq |x_i-\alpha'| \) for any $\alpha'$ that is either a root of $p$, any other $x_j\neq x_i$ or a boundary point of $\Sigma_{\mathbb{R}}(\rho).$ By~\eqref{potapp} and~\eqref{eqbd}, this polynomial satisfies all the conditions above except the first one. Our method is to deform the roots of $P_{M,m,\mu_{eq}}$ with a greedy algorithm so that it satisfies all the properties of Proposition~\ref{chpolreal}. We start with the degree $n-1$ coefficient of $pP_{M,m,\mu_{eq}}.$ We pick $\lceil 10/\rho\rceil$ of them say $x_1,\dots,x_{10/\rho}$ and deform each of them less than $\rho/10$ so that the degree $n-1$ coefficient of $pQ_m$ becomes even and the roots remain inside $\Sigma_{\mathbb{R}}(\rho).$ We show that this is possible. Let $Q_{m,0}(x):=P_{M,m,\mu_{eq}}(x)=x^m+\sum_{i=1}^{m}a_ix^{m-i}.$ The coefficient of $x^{\deg(p)+m-j}$ for $j\leq l$ in $Q_{m,0}(x)p(x)$ is given by \[ a_j+F_j(a_1,\dots,a_{j-1}) \] where $F_j(a_1,\dots,a_{j-1})$ is a polynomial in terms of $a_i$ for $i<j.$ We define $x_i(t):=x_i+t\frac{\rho}{10}$ for $i\leq 10/\rho$ and call the deformed polynomial $Q_{m,0,t}.$ By the intermediate value theorem there exists $|t|\leq 1$ such that the degree $n-1$ coefficient of $pQ_{m,0,t}$ becomes an even number and roots remains inside $\Sigma_{\mathbb{R}}(\rho).$ Suppose that after deformation, condition~\eqref{3deform} is violated for $x_1,\dots,x_k.$ By \eqref{Bireal}, it is possible to pick $k$ other roots of $Q_{m,0}(x)$ say $y_1,\dots, y_k$ such that there are no other roots close to $y_i$ within distance $n^{-4/3}.$ We updated $x_1$ and $y_1$ by $x_1+t$ and $y_1-t.$ By a covering argument, there exists $t\ll n^{-2}$ such that $x_1+t$ is away from other roots with distance at least $n^{-3}.$ Similarly, update $x_i, y_i$ so that the third condition is satisfied. Denote the updated polynomial by $Q_{m,1}.$ Note that $Q_{m,1}$ satisfies the second and third conditions above, and the degree $n-1$ coefficient of $pQ_{m,1}$ is even. \\ We construct $Q_{m,l}$ for $l\geq 1$ recursively as follows. Suppose that $Q_{m,l}$ satisfies the second and third conditions above above, and the degree $i$ coefficient of the product $pQ_{m,l}$ is an even integer for $n-l\leq i\leq n-1$, we want to make the coefficient of $n-l-1$ even as well without changing the higher degree coefficients by deforming the roots of $Q_{m,l}$ as we did in the previous paragraph. \\ Suppose that $l\leq \log\log(n).$ We pick $n^{\varepsilon }$ distinct $l$-tuples $\vec{x}:=[x_1,\dots,x_l]$ among roots inside $\Sigma_{\mathbb{R}}(\rho/10)$ for some $\varepsilon\geq 0$ such that $|x_i-x_j|\gg 1/l$ . Fix $\vec{x}=[x_1,\dots,x_l]$, and let \( s_k:=\sum_{i=1}^lx_i^k, \) and \( e_k:=\sum_{1\leq i_1<\dots<i_k\leq l}x_{i_1}\dots x_{i_k}, \) where $e_0=1$ and $s_0=l.$ Let \(\nabla s_k=k[x_i^{k-1}]\) be the gradient of $s_k.$ Let \[ v_l:=\left[\frac{1}{\prod(x_1-x_i)},\frac{1}{\prod(x_2-x_i)},\dots, \frac{1}{\prod(x_l-x_i)}\right]. \] It follows that $v_l$ is orthogonal to $\nabla s_k$ for $1\leq k\leq l-1$ and \( \left<v_l,\nabla s_l \right>=l. \) We note that \[ e_k=\frac{1}{k} \sum_{j=1}^k (-1)^{j-1}e_{k-j}s_j. \] This implies that $\left<v_l,\nabla e_k \right>=0$ for every $0\leq k\leq l-1$ and \( \left<v_l,\nabla e_l \right>=1. \) We deform $\vec{x}$ with the following ODE $\frac{d\vec{x}(t)}{dt}=v_l(t)$ and the initial condition $\vec{x}(0)=\vec{x}.$ We note that each coordinate of $v_l$ is less than $l^l \leq \frac{n^{\varepsilon}\rho}{10}$. Therefore, by the intermediate value theorem, it is possible to deform $n^{\varepsilon }$ distinct $l$-tuples $[x_1,\dots,x_l]$ along the direction of $v_l$ at most $\rho/10$ amount and make the coefficient of $n-l-1$ also even. Suppose that after deformation, condition~\eqref{3deform} is violated for $x_1,\dots,x_k,$ where $k\leq n^{\varepsilon}$. By \eqref{Bireal}, it is possible to pick $kl\leq n^{\varepsilon}$ other roots of $Q_{m,0}(x)$ say $y_{i,j}$ for $1\leq i\leq k$ and $1\leq j\leq l$ such that there are no other roots close to $y_{i,j}$ within distance $n^{-4/3}$, $|y_{i,k}-y_{i,k'}|\gg 1/l$ for $k\neq k',$ and $|x_i-y_{i,j}|\gg 1/l$. We deform $l+1$ tuple $\vec{x}_i:=\left[x_i,y_{i,1},\dots, y_{i,l}\right]$ with the following ODE $\frac{d\vec{x}_i(t)}{dt}=v_{l+1,i}(t)$ and the initial condition $\vec{x}_i(0)=\vec{x}_i$ where \[ v_{i,l+1}(t):=\left[\frac{1}{\prod_j(x_i-y_{1,j})},\frac{1}{(y_{i,1}-x_i)\prod_j(y_{i,1}-y_{i,j})},\dots, \frac{1}{(y_{i,l}-x_i)\prod_j(y_{i,l}-y_{i,j})}\right]. \] By a covering argument, there exists $t\ll n^{-2+\varepsilon}$ such that $x_1(t)$ is away from other roots with distance at least $n^{-3}$. Denote the updated polynomial by $Q_{m,l+1}.$ Note that $Q_{m,l+1}$ satisfies second and third conditions above, and the degree $i$ coefficient of the product $pQ_{m,l+1}$ is an even integer for $n-l-1\leq i\leq n-1$. \\ Suppose that $n^{\varepsilon} \geq l\geq \log\log(n)$ for some $\varepsilon >0.$ We pick $x_1,\dots,x_l$ according to the equilibrium measure among the roots inside $\Sigma(\rho/10)$. As before, let $\vec{x}=[x_1,\dots,x_l]$ and \[ v_l:=\left[\frac{1}{\prod(x_1-x_i)},\frac{1}{\prod(x_2-x_i)},\dots, \frac{1}{\prod(x_l-x_i)}\right]. \] By~\eqref{potapp} and~\eqref{eqbd}, the size of $v_l$ is exponentially small and we have \[ |v_l|\leq e^{-ld_{\Sigma}}. \] We deform $\vec{x}$ with the following ODE $\frac{d\vec{x}(t)}{dt}=v_l(t)$ and the initial condition $\vec{x}(0)=\vec{x}.$ By the intermediate value theorem there exists $|t|\leq 1$ so that the degree $n-l-1$ coefficient of $pQ_{m,l+1}$ is even. It is clear that the perturbation remains inside $\Sigma(\rho).$ As before, we can insure the third condition is satisfied by picking $y_{i,j}$ according to the equilibrium measure and deform the roots so that they are away from each other with distance at least $n^{-3}.$ \newline Finally, suppose that $m \geq l\geq n^{\varepsilon}.$ We pick $x_1,\dots,x_l$ according to the equilibrium measure among the roots inside $\Sigma_{\mathbb{R}}(\rho/10)$. As before, let $\vec{x}=[x_1,\dots,x_l]$ and \[ v_l:=\left[\frac{1}{\prod(x_1-x_i)},\frac{1}{\prod(x_2-x_i)},\dots, \frac{1}{\prod(x_l-x_i)}\right]. \] By~\eqref{potapp} and~\eqref{eqbd}, the size of $v_l$ is smaller than any negative power of $n$ \[ |v_l|\leq e^{-ld_{\Sigma}}\leq e^{-n^{\varepsilon}}\ll n^{-A} \] for any $A>0.$ We deform $\vec{x}$ with the following ODE $\frac{d\vec{x}(t)}{dt}=v_l(t)$ and the initial condition $\vec{x}(0)=\vec{x}.$ By mean value theorem there exits $|t|\leq 1$ so that the degree $n-l-1$ coefficient of $pQ_{m,1}$ is even. This time the perturbation does not change conditions \eqref{2deform} and \eqref{3deform}, since the perturbation is smaller than $n^{-A}$ for any $A.$ This completes the proof of our proposition. \end{proof} \section{Proofs of Theorem~\ref{main1} and Theorem~\ref{general}} \subsection{Proof of Theorem~\ref{main1}} In this subsection, we assume that $\Sigma\subset \mathbb{C}$ is compact and has empty interior with connected complement. We also assume that $\mu \in \mathcal{B}_{\Sigma}$ is a H\"older measure. \begin{comment} \begin{proposition}\label{discp} Suppose that $g_n(x)$ is a monic polynomial of degree n with complex coefficients and let $\mu $ and $\Sigma$ be as above. Moreover, suppose that \[ \frac{\log |g(z)|}{n} \leq U_{\mu}(z)+O(n^{-a}) \] for some $a>0.$ Then \[ \frac{\log |g(z)|}{n} = U_{\mu}(z)+O(n^{-a}), \] where $\inf_{i}|z-\xi_i|>n^{-A}$ for any fixed $A>0.$ Moreover, \[ \text{disc}_{\delta,\omega}(\mu_{h_n},\mu)=O(n^{-b}) \] for some $b>0$ that only depends on $a$ and the Holder exponent of $\mu.$ \end{proposition} \begin{proof} \end{proof} \end{comment} \begin{proof}[Proof of Theorem~\ref{main1}] Suppose that $\mu \in \mathcal{B}_{\Sigma}$ is a H\"older measure with exponent $\delta,$ and $D$ is any open set containing $\Sigma$. Fix $\rho>0$ such that $\Sigma(\rho)\subset D.$ We apply Proposition~\ref{chpol} to $p_{M,n-m,\mu}(x)$ defined in~\eqref{measpol} and $m:= n^{1-\delta_0}$ for some $\delta_0>0$ that we specify later and obtain $Q_m(x)$ such that \[ p_{M,n-m,\mu}(x)Q_m(x)=x^{n}+\sum_{i=0}^{n-1} a_ix^i \] where $a_i$ are even for $n-m\leq i\leq n-1.$ Let \[ r(x):=\frac{1}{2}+\sum_{i=0}^{n-1-m} \frac{a_i}{4} x^i. \]We write $r(x)$ in terms of linearly independent integral polynomials $\{P_0,\dots,P_{n-1-m}\}$ obtained from Minkowski's successive minima of $K_{n-1-m}$, and obtain \[ r(x)=\sum_{i=0}^{n-1-m} \alpha_i P_i(x). \] Let \[ w(x):=\sum_{i=0}^{n-1-m} \lfloor\alpha_i\rfloor P_i(x), \] and \[ h_n(x):= x^{n}+\sum_{i=n-m}^{n-1} a_ix^i +4w(x)-2. \] By the Eisenstein criteria at the prime 2, $w(x)$ is irreducible. Moreover, \[ h_n(x)=p_{M,n-m,\mu}Q_m(x)+\sum_{i=0}^{n-1-m} \beta_iP_i(x), \] where $|\beta_i|=4|\alpha_i-\lfloor\alpha_i\rfloor|<4.$ Note that \[ |P_i(z)|\leq \lambda_i e^{(n-m)U_\mu(z)} \] for every $z\in \mathbb{C},$ and by Proposition~\ref{upperbound}, \[\lambda_i\leq n^{cn^{1-\frac{\delta}{12}}}.\] Moreover, by \eqref{potapp} and Proposition~\ref{chpol} \[ \left| p_{M,n-m,\mu}(z)Q_m(z) \right|\geq e^{(n-m)U_\mu(z)} \kappa^m n^{-Cn^{1-\frac{\delta}{6}}}. \] for every $z\notin \Sigma(\rho)$, where $M=\lfloor n^{1/3}\rfloor.$ By taking $\frac{\delta}{12}>\delta_0>0$ and $m=n^{1-\delta_0}$, it follows that \[ \left| p_{M,n-m,\mu}Q_m(z) \right|> \left|\sum_{i=0}^{n-m-1} \beta_iP_i(z)\right| \] for large enough $n$ and every $z\notin \Sigma(\rho).$ By Rouch\'e's theorem, $h_n(x)$ and $p_{M,n,\mu}Q_m(z)$ has the same number of roots inside $\Sigma(\rho).$ Since all roots of $p_{M,n-m,\mu}Q_m(x)$ are inside $\Sigma(\rho),$ all roots of $h_n(x)$ are also inside $\Sigma(\rho).$ By Lemma~\ref{lem1}, we have \[ \frac{\log|Q_m(z)|}{m}\leq \max(\log|z|,0)+C_{\Sigma}, \] where $C_{\Sigma}$ is a constant that only depends on $\Sigma.$ Since $m=n^{1-\delta_0},$ we have \[ \frac{\log |h_n(z)|}{n} \leq U_\mu(z)+ O(n^{-\delta_0}). \] This completes the proof of Theorem~\ref{main1}. For $\Sigma\subset \mathbb{R}$, our argument is similar. We apply Proposition~\ref{chpolreal} instead of Proposition~\ref{chpol} and $p_{M,n-m,\mu}(x)Q_m(x)$ which has $n$ distinct roots which are $n^{-3}$ apart. We use the intermediate value theorem and the fact that we have $n$ sign changes on $\Sigma(\rho)$ to prove all roots are real and inside $\Sigma(\rho)$ instead of the Rouch\'e's theorem. \end{proof} \subsection{Proof of Theorem~\ref{general}} In this section, we show that Theorem~\ref{main1} implies Theorem~\ref{general}. To prove Theorem~\eqref{main1}, we assumed that $\mu$ is a H\"older measure, the support of $\mu$ has empty interior, and that its complement is connected. Here we show that these conditions are unnecessary for proving Theorem~\ref{general}. By Proposition~\ref{holderreduction}, for any $\rho>0,$ there exists a sequence of H\"older probability measure $\mu_n\in\mathcal{B}_{\Sigma(\rho)}$ such that \( \lim_{n\to \infty} \mu_n=\mu\). So by a diagonal argument the the proof is reduced to the case where $\mu$ is H\"older. \newline Recall $\Sigma\subset [a,b]\times[-c,c]$ as in section \eqref{bound_Kn} and $[a,b]\times[-c,c]= \bigcup_{0\leq i,j< 2M}B_{ij}$, where we define $B_{ij}=[a_i,a_{i+1}]\times[c_{j},c_{j+1}]$ and $a_i=a+\frac{i(b-a)}{2M}$ and $c_j=-c+\frac{j(2c)}{2M}$. Furthermore, $z_{ijk}\in B_{ij}$ for each $1\le k\le n_{ij}=\lfloor(n+1)\mu(B_{ij})\rfloor+\epsilon_{ij}$ are chosen with distinct imaginary parts. We now define a measure which imitates $\mu$ by being uniformly distributed on a union of intervals around the $z_{ijk}$. Let $\tilde \mu$ be the measure supported on $\bigcup_{ijk}\{z:|\Re(z-z_{ijk})|<\frac{1}{n},\Im(z)=\Im(z_{ijk})\}$ where $\tilde \mu$ restricted to each interval is the one-dimensional Lebesgue measure on that interval. Now let $\tilde\nu$ be the equilibrium measure on the support of $\tilde\mu$ and $\tilde \nu_{\epsilon}$ be the normalized pushforward of $\tilde\nu$ under the scaling map $x\mapsto (1+\epsilon)x$. Finally, define $\mu_{\epsilon,\epsilon'}=(1-\epsilon')\tilde\mu+\epsilon'\tilde\nu_\epsilon$ for $\epsilon,\epsilon'\in(0,1)$. \begin{theorem}\label{no_interior_bound} For all $Q\in\mathbb Z[x]$ and $0<\epsilon,\epsilon'<\frac{1}{2}$, $\int \log|Q(x)|d\mu_{\epsilon,\epsilon'}(x)\ge 0$ for large enough $n$. \end{theorem} \begin{proof}First, we relate $U_{\tilde{\mu}}$ with $U_\mu$. We compute $$U_{\tilde\mu}(x)=\int\log|x-z|d\tilde\mu = \sum_{ijk}\int_{\frac{-1}{2n}}^\frac{1}{2n}\log|x-y -z_{ijk}|dy.$$ Let $z_{hrs}$ be the closest root of $P_{M,n,\mu}$ to $x$. Then $\forall ijk\neq hrs$, $|x-z_{ijk}|\gg n^{-5/6}$ by \eqref{Bij} choosing $M\sim n^{1/3}$. So for $ijk\neq hrs$, we have that $$\int_{\frac{-1}{2n}}^\frac{1}{2n}\log|x-y-z_{ijk}|dy - \frac{1}{n}\log|x-z_{ijk}| = \int_{\frac{-1}{2n}}^\frac{1}{2n}\log\left|1-\frac{y}{x-z_{ijk}}\right| dy\le \int_\frac{-1}{2n}^\frac{1}{2n}\frac{|y|dy}{|x-z_{ijk}|} \ll n^{-7/6}$$ Thus if $|x-z_{hrs}|\ge 2\cdot\text{diam}(\text{supp}(\mu))$, we have by Lemma \ref{lem1} that $$U_{\tilde\mu}(x)=U_{\mu}(x)+O(n^{-1/6}),$$ and otherwise, $$U_{\tilde\mu}(x)=\int_{\frac{-1}{2n}}^\frac{1}{2n}\log|x-y-z_{hrs}|dy + \frac{1}{n}\sum_{ijk\neq hrs}\log|x-z_{ijk}|+O(n^{-1/6})$$ In the latter case, $\int_{\frac{-1}{2n}}^\frac{1}{2n}\log|x-y-z_{hrs}|dy \ll\max(n^{-1},\int_{\frac{-1}{2n}}^\frac{1}{2n}\log|y|dy)\ll n^{-1}\log(n)$. Therefore, we have that $U_{\tilde\mu}(x)=U_{\mu_{p_{M,n,\mu}}}(x)-\frac{1}{n}\log |x-z_{hrs}| + O(n^{-1/6})$. From our proof of Proposition \ref{simplex}, \begin{multline*} \frac{\log |p_{M,n,\mu}^{efg+}(x)|}{n} - U_{\mu}(x)= \frac{\log \frac{|x-z_{hrs}||x-\Re(z_{efg})|}{|x-z_{efg}||x-\overline{z_{efg}}|}}{n} \\ +O\left(\log(n)M^{-\delta}+ \frac{M^2\log(n)}{n}+ M^{-\delta/2}\log(M)^{1/2}\right). \end{multline*} Taking $M\sim n^{1/3}$, we have now shown that $$U_{\tilde\mu}(x) = U_\mu(x) +O\left(n^{-1/6}+n^{\frac{-\delta}{6}}\log(n)^\frac{1}{2}\right)$$ Therefore, in either case, we have $U_{\tilde\mu(x)}=U_\mu(x)+O(n^{-1/6}+n^{-\delta/6}\log(n)^{1/2})$. \newline Now for any $Q(x)=\prod_{i=1}^m(x-\alpha_i)\in\mathbb Z[x]$, we have that $$\int\log|Q(x)|d\mu_{\epsilon, \epsilon'} =(1-\epsilon')\int\log|Q(x)|d\tilde\mu + \epsilon'\int\log|Q(x)|d\tilde\nu_\epsilon$$ Using our calculation above, $$\int\log|Q(x)|d\tilde\mu = \sum_{i=1}^mU_{\tilde \mu}(\alpha_i) = \int \log|Q(x)|d\mu + mO(n^{-1/6}+n^\frac{-\delta}{6}\log(n)^{\frac{1}{2}})$$ Since $I_{\tilde\nu}\ge I_\mu\ge0$, we have that \begin{equation}\label{pushcap}I(\tilde\nu_\epsilon) = \int\int\log|(1+\epsilon)(x-y)|d\tilde\nu d\tilde\nu=I(\tilde\nu)+\log|1+\epsilon|>I(\tilde\nu)+\epsilon-\epsilon^2.\end{equation} We now bound $I(\tilde\nu)$. Since $\tilde \nu$ is the equilibrium measure on the support of $\tilde \mu$, $I(\tilde \mu)\le I(\tilde\nu)$. To compute $I(\tilde\mu)$, we first see that $\int_{\frac{-1}{2n}}^\frac{1}{2n}\int_{\frac{-1}{2n}}^\frac{1}{2n}\log|x-y|dxdy\ll n^{-2}\log(n)$. Therefore $$I(\tilde\mu)=\sum_{ijk\neq i'j'k'}\int_{\frac{-1}{2n}}^\frac{1}{2n}\int_\frac{-1}{2n}^\frac{1}{2n}\log|z_{ijk}-z_{i'j'k'}+x-y|dxdy+O(n^{-1}\log(n)).$$ By the same argument before, this is $\frac{1}{n^2}\sum_{ijk\neq i'j'k'}\log|z_{ijk}-z_{i'j'k'}|+O(n^{-1/6})$. Applying \eqref{maineq} with $M\sim n^{-1/3}$, $I(\tilde\mu) = O(n^{-1/6}+n^{\frac{-\delta}{6}}\log(n))$. Therefore for large enough $n$, $|I(\tilde \mu)|\le \frac{\epsilon}{4}$. Using \eqref{pushcap}, we have $I(\tilde\nu_\epsilon)>\frac{\epsilon}{4}$ since $\epsilon<1/2$. Thus since $$\int\log|Q(x)|d\mu_{\epsilon,\epsilon'}=(1-\epsilon')\int\log|Q(x)|d\mu+m\left(\epsilon'I(\tilde\nu_\epsilon)+(1-\epsilon')O(n^{-1/6}+n^\frac{-\delta}{6}\log(n)^\frac{1}{2})\right)$$ by equation (1.4) of \cite{logpotentials}, this shows that $\int\log|Q(x)|d\mu_{\epsilon,\epsilon'}>0$ for sufficiently large $n$. So for each $0<\epsilon,\epsilon'<\frac{1}{2}$, there is a large enough $n$ for which $\int\log|Q(x)|d\mu_{\epsilon,\epsilon'}>0$ as desired. \end{proof} \begin{comment} \subsection{Discrepancy of $\mu_{P_n}$ and $\mu$} Suppose you have two probability measures $\mu$ and $\nu$. Fix a non-zero smooth function $\omega$ with compact support and $\delta > 0$. Recall that the discrepancy between $\mu$ and $\nu$ is $\text{disc}_{\delta,\omega}(\mu,\nu) = \sup_{a<b,c<d} |\omega_\delta\ast\mu([a,b]\times[c,d])-\omega_\delta\ast\nu([a,b]\times[c,d])|$ where $\omega_\delta(x) = \frac{1}{\delta^2}\omega(\frac{x}{\delta})$. \begin{proposition}\label{discprop} If $||U_\mu-U_\nu||_1<\epsilon$, then $\text{disc}_{\epsilon^{1/8},\omega}(\mu,\nu)< C \sqrt\epsilon$ for some positive constant $C$ depending only on $\omega$. \end{proposition} \begin{proof} Fix $B:=[a,b]\times [c,d]$. Let $\ell_z(x) = \log|z-x|$ so that $U_\mu(z) = \ell_z\ast \mu$ and $U_\nu(z) = \ell_z\ast \nu$. It follows that $U_{\omega\ast\mu}(z) = \omega\ast\ell_z\ast\mu$ and similarly $U_{\omega\ast\nu}(z) = \omega\ast\ell_z\ast\nu$. Since $\omega$ is smooth, $U_{\omega\ast\mu}$ is smooth and thus $\Delta (U_{\omega\ast\mu})(z) = (\Delta\omega)\ast \ell_z \ast \mu$. Since $\omega\ast \mu$ is smooth, it is absolutely continuous with respect to the Lebesgue measure $m$. Therefore by Theorem 1.3 in \cite{logpotentials}, we have that $d(\omega\ast\mu) = -\frac{1}{2\pi}\Delta U_{\omega\ast\mu} dm$. This proves the following inequality: $$|\omega\ast\mu(B)-\omega\ast\nu(B)| = \frac{1}{2\pi}\int_{B}|\Delta\omega \ast (U_\mu-U_\nu)|dm \le \frac{1}{2\pi}m(\text{supp}(\omega))||\Delta \omega||_\infty ||U_\mu-U_\nu||_1.$$ Therefore, $|\omega\ast\mu(B) - \omega\ast\nu(B)| \le \frac{m(\text{supp}(\omega))}{2\pi}||\Delta\omega||_\infty||U_\mu-U_\nu||_1<C\epsilon$ for some constant $C>0$ only depending on $\omega$. Now if we do the same calculation with $\omega_\delta$, we get $$|\omega_\delta\ast\mu(B)-\omega_\delta\ast\nu(B)| = \frac{1}{2\pi\delta^4}\int_B\left|(\Delta \omega)\left(\frac{1}{\delta}z\right)\ast(U_\mu-U_\nu)\right|dm < \frac{C\epsilon}{\delta^4}$$ since $\Delta \omega_\delta(z) = \frac{1}{\delta^4}(\Delta\omega)(\frac{1}{\delta}z)$. Therefore, we have shown that $$\text{disc}_{\epsilon^{1/8},\omega}(\mu,\nu) =\sup_{a<b,c<d}|\omega_{\epsilon^{1/8}}\ast\mu([a,b]\times[c,d])-\omega_{\epsilon^{1/8}}\ast\nu([a,b]\times[c,d])| < C\sqrt\epsilon.$$ This completes the proof. \end{proof} \end{comment} \begin{proof}[Proof of Theorem~\ref{general}] Suppose that $d_{\Sigma}<1.$ Note that \[ \log d_{\Sigma}= I(\mu_{eq})=\sup_{\mu\in \mathcal{P}_{\Sigma}}I(\mu), \] where $\mu_{eq}$ is the equilibrium measure of $\Sigma.$ Hence, \[ I(\mu)<0 \] for every $\mu\in \mathcal{P}_{\Sigma}.$ By Theorem~\ref{mdim} for polynomial $Q(x,y)=x-y$, \[ \mathcal{B}_{\Sigma}=\emptyset. \] Since $\mathcal{A}_{\Sigma}\subset \mathcal{B}_{\Sigma},$ this completes the proof of Theorem~\ref{general} if $d_{\Sigma}<1.$ \\ Otherwise, $d_{\Sigma}\geq 1$ and $\mu_{eq}\in \mathcal{B}_{\Sigma}.$ This implies that $\mathcal{B}_{\Sigma}\neq \emptyset.$ Denote by $\tilde\mu_{\epsilon,\epsilon'}$ the measure described in Theorem \ref{no_interior_bound} for the minimal value of $n$ giving the theorem. This gives us that $\tilde\mu_{\frac{1}{m},\frac{1}{m}}\stackrel{\ast}{\rightharpoonup}\mu$ as $m\to\infty$. Further, for each $m$, there is a sequence $P_{m,n}$ of polynomials with $\mu_{P_{m,n}}\stackrel{\ast}{\rightharpoonup}\tilde\mu_{\frac{1}{m},\frac{1}{m}}$ as $n\to \infty$. Taking a subsequence $P_{m,n_m}$, we have $\mu_{P_{m,n_m}}\stackrel{\ast}{\rightharpoonup}\mu$ as $m\to\infty$. Therefore, we have shown that the assumptions that the support of $\mu$ must have non-empty interior and have its complement be connected are unnecessary, and Theorem~\ref{general} follows from Theorem~\ref{main1}. \end{proof} \section{The Algorithm}\label{main} \subsection{Lattice Algorithms}\label{lattice_alg} The algorithm we develop requires access to short vectors in a lattice. As seen in the work by Micciancio ~\cite{SVP}, finding a vector whose length is within a constant multiple of the shortest vector in a lattice is NP-hard to compute. This is known as the shortest vector problem or SVP. It is still desirable in general to have access to short vectors in a lattice. An algorithm known as the LLL-algorithm, developed by Lenstra, Lenstra, and Lov\'asz, finds a basis of vectors in a lattice in polynomial time ~\cite{LLL} where the shortest vector of the output has length within an exponential factor of the shortest vector of the lattice. In the implementation of this algorithm, we use the built-in method in PARI/GP named \texttt{qflll}. When the \texttt{qflll} method is given a matrix $A$ whose columns generate a lattice, it returns a transformation matrix $T$ so that $AT$ is an ``LLL-reduced" basis of the lattice. While the LLL-algorithm is quite reasonable in practice, the asymptotic bounds are not strong enough to prove a sub-exponential factor. \newline In 1986, Schnorr~\cite{Schnorr} devised a parametrized family of algorithms attacking the SVP problem which output reduced bases somewhere between the stringent requirements of Korkine-Zolotarev reduction and the looser requirements of LLL reduction based on the input parameter. Careful parameter selection achieves slightly sub-exponential constant factors in polynomial time. Schnorr proved that if the algorithm produces a basis $b_1,\dots, b_n$, then $||b_1||$ is within a subexponential factor of a shortest vector in the lattice. We prove that in addition, if $\lambda_1,\dots, \lambda_n$ are so that $\lambda_i$ is the least value for which there are $i$ linearly independent vectors of norm at most $\lambda_i$, then for each $i=1,\dots, n$, $||b_i||$ is within a sub-exponential factor of $\lambda_i$. We now introduce some notation from Schnorr's work and recall some of his results. \newline Let $b_1,\dots, b_n$ be a basis. Then let $b_i^*$ be the projection of $b_i$ onto the orthogonal complement of $\text{span}\{b_1,\dots, b_{i-1}\}$ for $i=1,\dots, n$. We can get $b_1=b_1^*,b_2^*,\dots,$ and $b_n^*$ by Gram-Schmidt process without normalizing. Now define $\mu_{i,j}:=\frac{\langle b_i,b_j^*\rangle}{||b_j^*||^2}$. \begin{definition}We say $b_1,\dots, b_n$ is size-reduced if $\mu_{i,j}\le 1/2$ for all $i>j$.\end{definition} \begin{definition} Let $\Lambda_i$ be the lattice spanned by the projections of $b_i,b_{i+1},\dots, b_n$ onto the orthogonal complement of $\text{span}\{b_1,\dots, b_{i-1}\}$. Then $b_1,\dots, b_n$ is Korkine-Zolotarev reduced if $||b_i^*||$ is minimal among non-zero vectors in $\Lambda_i$. \end{definition} \begin{definition}Let $k\mid n$. We say that $b_1,\dots, b_n$ is semi $k$-reduced if it is size-reduced and it satisfies both \begin{equation} ||b_{ik}^*||^2 \le 2||b_{ik+1}^*||^2, \end{equation} and the components of $b_{ik+j}$ for $j=1,\dots, k$ orthogonal to $b_1,\dots, b_{ik-1}$ are Korkine-Zolotarev reduced for each $i=0,\dots,\frac{n}{k}-1$. Call this property (KZ). \end{definition} Lastly, the constant $\alpha_k$ is defined as $\max\frac{||b_1||^2}{||b_k^*||^2}$ where the max is over all Korkine-Zolotarev reduced bases on lattices of rank $k$. Corollary 2.5 in Schnorr's paper shows that $\alpha_k\le k^{1+\ln k}$. \newline \begin{theorem}\label{factors} Let $b_1,\dots,b_n$ be a semi $k$-reduced basis and $\lambda_j$ be the least real number so there are $j$ linearly independent vectors in the lattice spanned by $b_1,\dots, b_n$ of norm at most $\lambda_j$. Then $||b_j||\le k^{\frac{n}{k}\ln k +O(\frac{n}{k}+\ln k)}\lambda_j$. \end{theorem} The essence is a combination of the proofs of Theorems 2.3 and 3.4 in ~\cite{Schnorr}. \begin{proof} Fix $1\le j\le n$. Suppose that $v_1,\dots, v_j$ are linearly independent vectors in the lattice with $||v_i||=\lambda_i$ for $i=1,\dots, j$. For each $i$, write $v_i = \sum_{s=1}^n c_{s,i}b_s$. Let $s_i$ be the largest index for which $c_{s_i,i}\neq 0$. There is some $i=1,\dots, j$ so that $s_i\ge j$ by linear independence. If $c_{s,i}^*$ is so that $v_i = \sum_{s=1}^nc_{s,i}^*b_s^*$, then $c_{s_i,i}=c_{s_i,i}^*\in \mathbb Z$. This implies that $\lambda_j = ||v_j||\ge ||v_i||\ge||c_{s_i,i}^*b_{s_i}^*||\ge ||b_{s_i}^*||$. Now note that for each $0< s<t\le k$ and $m$ so that $(m+1)k\le n$, then $||b_{mk+s}^*||\le \alpha_k||b_{mk+t}^*||$ as stated in the proof of Theorem 2.3 of ~\cite{Schnorr} and call this inequality $(\ast)$. \newline Using the fact that $b_1,\dots, b_n$ is size reduced, $$||b_j||=\left|\left|b_j^*+\sum_{s=1}^j\mu_{j,s}b_s^*\right|\right|\le \sum_{s=1}^j||b_s^*||.$$ To bound $||b_s^*||$ for $1\le s\le j$, suppose $s=m_1k+t_1$ and $s_i=m_2k+t_2$ for $0\le t_2,t_2<k$. Then \begin{align*} ||b_s^*|| &\le \alpha_k||b_{(m_1+1)k}^*|| &\text{by property (KZ) and }(\ast)\\ &\le 2\alpha_k ||b_{(m_1+1)k+1}^*||&\text{by }(6)\\ &\le (2\alpha_k)^{m_2-m_1}||b_{m_2k+1}||&\text{by induction}\\ &\le (2\alpha_k)^{m_2-m_1}\alpha_k||b_{s_i}^*||&\text{by property (KZ) and }(\ast)\text{ since }j\le s_i\\ &\le (2\alpha_k)^{\frac{n}{k}-\lfloor\frac{s}{k}\rfloor}\alpha_k\lambda_j&\because ||b_{s_i}^*||\le \lambda_j \end{align*} Since $\alpha_k \le k^{1+\ln k}$, we have $||b_s^*||\le k^{(\frac{n}{k}+1)(\ln k+2)-\lfloor\frac{s}{k}\rfloor\ln k}$. Therefore, \begin{align*} ||b_j||&\le k^{\frac{n}{k}\ln k + O(\frac{n}{k}+\ln k)}\sum_{s=0}^\infty k^{-\lfloor\frac{s}{k}\rfloor \ln k}\lambda_j\\ &\le k^{\frac{n}{k}\ln k + O(\frac{n}{k}+\ln k)}k\sum_{s=0}^\infty k^{-s\ln k}\lambda_j&\text{since every group of }k\text{ is equal}\\ &\le k^{\frac{n}{k}\ln k +O(\frac{n}{k}+\ln k)}\left(1-k^{-\ln k}\right)^{-1}\lambda_j\\ &\le k^{\frac{n}{k}\ln k +O(\frac{n}{k}+\ln k)}\lambda_j&\text{since }(1-k^{-\ln k})^{-1}\le 3\text{ for }k\ge 2. \end{align*} This completes the proof. \end{proof} According to Schnorr's paper, this algorithm takes $O(n^4\log B+k^{\frac{k}{2}+o(k)}n^2\log B)$ arithmetical steps on $O(n\log B)$ integers where $B$ is the Euclidean length of the longest input vector. In our case, $B$ is a constant depending on $\mu$. Taking $k=\frac{\log(n)}{\log(\log(n))}$, this is $O(n^4 + n^2\log(n)^{\frac{3\log n}{2\log\log n}})$ which is contained in $O(n^4)$ and thus this choice of $k$ makes the algorithm polynomial time in $n$. In this case, the factor in Theorem \ref{factors} is at most $e^{\frac{2n\log(\log (n))^{3}}{\log(n)}}$ for large enough $n$; in particular, it is sub-exponential. \subsection{Geometry of numbers} There are two different sub-methods of our algorithm regarding geometry. The first finds a short basis of a given lattice. The second finds an integral polynomial near a real polynomial where distance is measured with respect to a given basis. In these two algorithms, $P$ is an integral, square-free, monic polynomial with roots $\alpha_1, \dots, \alpha_n$. \subsubsection{Finding Short Lattice Bases}\label{lattice_bases} To find short lattice bases, we implement a version of Proposition ~\ref{mink}. This theorem shows the existence of a short basis of polynomials in a given convex space. In particular, let $K$ be the set of real polynomials of degree at most $n$ with $\log|P(x)|\le nU_\mu(x)$. There are linearly-independent, integral polynomials $P_0,\dots,P_n$ so that for each $0\le k\le n$, if $P_k\in \lambda K$ where $\lambda\in\mathbb{R}$ is minimal, then there is no integral polynomial in $\rho K$ for any $\rho<\lambda$ linearly-independent from $P_0,\dots, P_{k-1}$. We showed that the largest of these $\lambda$'s is sub-exponential in $n$ in Proposition~\ref{upperbound}. Thus $\{P_0,\dots, P_n\}$ forms a basis of lattice points in a subexponential multiple of $K$. By Theorem \ref{factors}, we can compute explicit integral polynomials which are within a sub-exponential factor of $K$ in polynomial time in $n$. This is formalized in Corollary \ref{find_lattice_basis}. \begin{corollary}\label{find_lattice_basis} If $\mu\in\mathcal B_\Sigma$ is a H\"older probability measure, then we can compute a basis $Q_0,\dots, Q_n$ of integral polynomials of degree at most $n$ in polynomial time where each $Q_i\in e^{Cn\frac{\log(\log(n))^3}{\log(n)}}K_n$ for some fixed constant $C$ depending only on $\Sigma$ and $\mu$. More specifically, this runs in $O(n^4)$ time and assuming $n$ is sufficiently large, we can take $C=3$. \end{corollary} \begin{proof} First compute the basis $p_{M,n,\mu}^{efg\pm}$ of polynomials of degree less than $n$ as defined in \eqref{pmdef} choosing $M = \lfloor n^{1/3}\rfloor$. Call this basis $\mathcal S$. We want to find an integral basis which is short in the basis $\mathcal S$ by Proposition \ref{simplex}. Now write $\mathcal E:=\{1,x,\dots, x^{n-1}\}$ in the basis $\mathcal S$ and apply Schnorr's algorithm to get a semi $k$-reduced basis choosing $k=\frac{\log n}{\log\log n}$. We showed at the end of Section \ref{lattice_alg} that this runs in $O(n^4)$ time and that the output polynomials $Q_0,\dots, Q_n$ are so that $Q_i\in e^{2n\cdot\frac{\log(\log(n))^3}{\log n}}n^{cn^{1-\frac{\delta}{12}}}K_n$ by Proposition \ref{upperbound} for each $i$ and fixed $c$ for large enough $n$. It follows that $Q_i\in e^{3n\cdot\frac{\log(\log(n))^3}{\log n}}K_n$ for large enough $n$. \end{proof} One way of computing the integral polynomials is as follows. Suppose $T$ is a matrix so that $AT$ is a reduced form (as in Schnorr's algorithm) of $A$ where $A$ is the matrix of $\mathcal E$ written in base $\mathcal S$. Then $AT$ has linearly independent columns representing such polynomials in base $\mathcal S$. So $T$ has columns representing those polynomials in base $\mathcal E$ as desired. This gives the desired $n+1$ linearly independent integer polynomials. Further note that that doing the change of basis via matrix inversion is numerically unstable for large matrices. Fortunately, we can compute the change of basis matrix explicitly. As is typically done, the time bounds given in this paper is in number of arithmetic operations. Numerical stability is not studied in this paper. \subsubsection{Finding Close Integer Polynomials}\label{int_poly} The following discussion allows us to algorithmically produce polynomials like those whose existence is proved in Theorem 3.2 of \cite{Smith}. The intention of this method is to find an integral polynomial close to a given real polynomial in polynomial time where distance is measured with respect to a given basis. In our case, we use the basis produced by Corollary \ref{find_lattice_basis}. \begin{comment} Define $P_j(z) = P(z)/(z-\alpha_j)$. This forms a basis $\mathcal{P}$ of the vector space of degree $n-1$ polynomials. Indeed, $P_j(\alpha_k) = \begin{cases} 0,&\text{if }k\neq j\\\prod_{l\neq j}(\alpha_j-\alpha_l),&\text{otherwise}\end{cases}$. As in the proof of Theorem 3.2 in Smith, let $P_j^o$ be $P_j$ if $\alpha_j$ is real and $\begin{cases} \frac{1}{\sqrt{2}}(P_j+\overline{P_j}), &\text{if }\Im(\alpha_j)>0\\ \frac{i}{\sqrt{2}}(P_j-\overline{P_j}), & \text{if }\Im(\alpha_j) < 0 \end{cases}$. This is still a basis and we define $\mathcal O:=\{P_1^o,\dots, P_n^o\}$. For notation, if $Q$ is a polynomial and $\mathcal B$ is a basis of the polynomials of degree at most $\deg(Q)$, then let $[Q]_\mathcal B$ be the coordinate vector of $Q$ in basis $\mathcal B$. For example, if $Q(x)=\sum_{i=0}^{n-1} a_i x^i$, then $[Q]_\mathcal E = \begin{pmatrix}a_0\\ \vdots\\a_{n-1}\end{pmatrix}$. \end{comment} \begin{theorem}\label{close_poly} Given a H\"older probability measure $\mu\in\mathcal B_\Sigma$, there is a polynomial time algorithm which takes in a real polynomial $Q(z)$ of degree less than $n$ and produces an integer polynomial $H$ of degree less than $n$ such that $\frac{\log(Q-H)(x)}{n}\le U_\mu(x) + Cn\cdot \frac{\log(\log(n))^3}{\log(n)}$. The run-time is $O(n^4)$ where $C$ is dependent only on $\mu$ and $\Sigma$ and assuming $n$ is sufficiently large, we can choose $C=4$. \end{theorem} \begin{proof} Let $Q_0,\dots, Q_{n-1}$ be the output of our algorithm described in Corollary \ref{find_lattice_basis}. Call this basis $\mathcal Q$. Suppose the $j$-th coordinate of $[Q]_\mathcal Q$ is $c_j$. Take $\tilde c_j$ to be the nearest integer (rounding up if $c_j\in\frac{1}{2}\mathbb{Z}$), and let $H$ be the polynomial so that $[H]_\mathcal Q=\begin{pmatrix}\tilde c_0\\\vdots\\\tilde c_{n-1}\end{pmatrix}$. Then $H$ is an integer polynomial since $[H]_\mathcal Q$ is an integer linear combination of the integral polynomials in the basis $\mathcal Q$. We know $Q-H = \sum_{j=1}^n(c_j-\tilde c_j)Q_j$. Thus $Q-H\le \frac{n}{2}e^{3n\cdot\frac{\log(\log(n))^3}{\log(n)}}e^{nU_\mu}$ and so $Q-H\le e^{\frac{4n\log(\log(n))^3}{\log(n)}}e^{nU_\mu}$ for large enough $n$ as desired. For an analysis of the run-time, we note that Schnorr's algorithm is the bottleneck. It runs in $O(n^4)$ time by Corollary~\ref{find_lattice_basis}. \end{proof} \subsection{Main Algorithm} Here we prove Theorem \ref{main2}. We have already covered the main concepts. By Corollary \ref{find_lattice_basis}, we can get a basis of integral polynomials in a sub-exponential multiple of $K_n$. If we now apply Theorem \ref{close_poly} to $\frac{1}{4}(p_{M,n,\mu} - x^n)+\frac{1}{2}$ to get $H(x)$, we could output $P_n(x):=x^n+4H(x)-2$. This gives us an Eisenstein polynomial in a sub-exponential multiple of $K_n$. If we simply want a polynomial $P$ with small $n$-norm, this would be sufficient; however, we have to introduce some extra steps to ensure $\text{supp}(\mu_{P_n})\subset D$ for a given open $D\supset \Sigma$. The proof will imitate the proof of Theorem \ref{main1}. \begin{proof}[Proof of Theorem \ref{main2}] Suppose that $\mu\in\mathcal B_\Sigma$ is a H\"older measure with exponent $\delta$, let $D$ be any open set containing $\Sigma$, and choose $\rho>0$ so that $\Sigma(\rho)\subset D$. Apply Proposition \ref{chpol} to $p_{M,n-m,\mu}$ with $m$ chosen later to get the polynomial $Q_m(x)$. Note that the proof of Proposition \ref{chpol} explicitly computes $Q_m(x)$ in polynomial time. As in the proof of Theorem \ref{main1}, let $p_{M,n-m,\mu}(x)Q_m(x)$ be given by $x^n+\sum_{i=0}^{n-1}a_ix^i$. Apply Theorem \ref{close_poly} to $$h(x):=\frac{1}{4}\left(p_{M,n-m,\mu}(x)Q_m(x)-x^n-\sum_{i=n-m}^{n-1}a_ix^i\right)+\frac{1}{2}$$ to get an integer polynomial $H(x)$. Finally, $$p(x):=x^n+\sum_{i=n-m}^{n-1}a_ix^i + 4H(x)-2$$ is an Eisenstein polynomial at $2$. We see that \begin{equation}\label{close_res} |p_{M,n-m,\mu}(x)Q_m(x)-p(x)|= |4\left(h(x) - H(x)\right) - 2| \le e^{Cn\cdot\frac{\log(\log(n))^3}{\log(n)}}e^{(n-m)U_\mu(x)}\end{equation} for some constant $C$ depending only on $\mu$ and $\Sigma$ by Theorem \ref{close_poly}. As in the proof of Theorem \ref{main1}, we have $$|p_{M,n-m,\mu}(x)Q_m(x)|\ge \kappa^mn^{-cn^{1-\frac{\delta}{6}}}e^{(n-m)U_\mu(x)}$$ for every $x\not\in \Sigma(\rho)$ where $M=\lfloor n^{1/3}\rfloor$. For large enough $n$, we can choose $m<n$ so that $m=\lceil\frac{2Cn}{\log(\kappa)}\cdot \frac{\log(\log(n))^3}{\log(n)}\rceil$ where we choose $C=4$ as in Theorem \ref{close_poly}. Then for sufficiently large $n$, $$|p_{M,n-m,\mu}(x)Q_m(x)|\ge|p_{M,n-m,\mu}(x)Q_m(x)-p(x)|$$ for all $x\not\in \Sigma(\rho)$. By Rouch\'e's theorem, since $p_{M,n-m,\mu}(x)Q_m(x)$ has all its roots in $\Sigma(\rho)$, so does $p(x)$. Furthermore by \eqref{close_res}, $$||p||_n\le e^{Cn\cdot\frac{\log(\log(n))^3}{\log(n)}} + ||p_{M,n-m,\mu}(x)Q_m(x)||_n\le e^{\tilde{C}n\cdot\frac{\log(\log(n))^3}{\log(n)}}$$ for some constant $\tilde C$ dependent only on $\Sigma$ and $\mu$. \end{proof} \section{Applications}\label{applications} \subsection{Numerical Data} We implemented the algorithm described above with some simplifications.\footnote{All examples in section \ref{applications} can be found at https://github.com/Bryce-Orloski/Limiting-Distributions-of-Conjugate-Algebraic-Integers-Applications.} Firstly, we used the LLL-algorithm implemented in PARI/GP named \texttt{qflll}. Secondly, instead of choosing $z_{ijk}$ as indicated in the paper, we use reasonable sampling methods as described in each numerical example given. We also do not apply the extra step of forcing the roots in a $\rho$-neighborhood of the measure's support. So these algorithmic outputs do not employ the full power of the algorithm, but as we see, it still outputs strong results quickly. All of the examples given in this subsection were computed using Pennsylvania State University's ROAR servers and gave the output in at most two minutes. \newline The first example we discuss is pictured in Figure \ref{interval}. The measure depicted is one in a family which was constructed by Serre~\cite{MR2428512}. In particular, it is the probability distribution $\mu$ on $\Sigma=[a,b]$ with $\mu=c\mu_{[a,b]}+(1-c)\nu_{[a,b]}$ where $a=0.1715,$ $ b=5.8255,$ $c=0.5004,$ $\mu_{[a,b]}$ is the equilibrium measure on $[a,b]$, and $\nu_{[a,b]}$ is the pushforward of the equilibrium measure on $[b^{-1},a^{-1}]$ under the map $z\to 1/z$. We compute the sample points of $\mu$ using by taking the $k^{th}$ sample point to be the inverse distribution of $\mu$ at $k/n$. The algorithm ran on the ROAR servers for roughly $3$ seconds when asked to compute a degree 100 polynomial (with the sample points pre-computed). The output polynomial has all real roots with their histogram being displayed in Figure \ref{fig2}. The endpoints and interval width in Figure \ref{fig2} are approximate. \newline \begin{figure}[t] \centering \begin{subfigure}[b]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{wolfram_lemniscate.png} \caption{Plot of $|z^2-1|=1$} \label{lemniscate_wolfram} \end{subfigure} \hfill \begin{subfigure}[b]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{lemniscate_200_roots.png} \caption{Plotted roots or degree 200 polynomial} \label{lemniscate_output} \end{subfigure} \hfill \begin{subfigure}[b]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{unif_circle_200.png} \caption{Mapping of output roots via $z\mapsto z^2-1$} \label{circle_unif} \end{subfigure} \caption{These demonstrate the effectiveness of the algorithm on a lemniscate.} \label{lemniscate} \end{figure} The second example is pictured in Figure \ref{lemniscate}. Figure \ref{lemniscate_wolfram} depicts the support of the pull-back of the uniform distribution on the unit circle via $|z^2-1|=1$. We sampled this support via this distribution and ran the algorithm on 200 samples. It finished in under two minutes and a plot of its roots are given in Figure \ref{lemniscate_output}. We want this to sample the pull-back of the uniform distribution on the circle. To see that this is well sampled, Figure \ref{circle_unif} plots the image of these roots under the map $z\mapsto z^2-1$ and we see that it roughly approximates the uniform distribution on the unit circle. \newline Our last example is depicted in Figure \ref{annulus}. Here we sampled the annulus $1\le |z|\le 2$ with 100 complex numbers by sampling 50 using the Monte Carlo method and taking their complex conjugates. This is shown in Figure \ref{annulus_pivots}. Using these samples in our algorithm, it returned a degree 100 monic, integral, irreducible polynomial in under two minutes whose roots are plotted in Figure \ref{annulus_output}. \newline \begin{figure}[t] \centering \begin{subfigure}[b]{0.4\textwidth} \centering \includegraphics[width=\textwidth]{unif_rand_100_annulus.png} \caption{Plot of 50 sampled points from the annulus $1\le |z|\le 2$ and their conjugates} \label{annulus_pivots} \end{subfigure} \hfill \begin{subfigure}[b]{0.4\textwidth} \centering \includegraphics[width=\textwidth]{unif_rand_100_annulus_output.png} \caption{Plotted roots of 100 degree polynomial from algorithm given the samples as input} \label{annulus_output} \end{subfigure} \caption{The left shows 50 points sampled uniformly from $1\le |z|\le 2$ and right plots the roots of the output of the algorithm which tried to approximate the distribution of the left plot.} \label{annulus} \end{figure} Notice how we cannot provably show the desired convergence with the LLL-algorithm, and even with Schnorr's algorithm, the convergence we prove is very slow. However, here we are applying the LLL-algorithm and in each of Figures \ref{interval}, \ref{circle_unif}, and \ref{annulus}, we see that the plots of the roots of the polynomial produced by the algorithm are very close to the plots of the corresponding samples. \subsection{Applications to Abelian varieties} Recently, Shankar and Tsimerman~\cite{shankar_tsimerman_2018}, \cite{Tsimerman1} conjectured that for every $g\geq 4$, every Abelian variety defined over $\overline{\mathbb{F}_q}$ is isogenous to a Jacobian of a curve defined over $\overline{\mathbb{F}_q}.$ Given a finite field $\mathbb{F}_q$ and any arbitrary large extension $\mathbb{F}_{q^n}$, we prove that there are infinitely many abelian varieties over $\mathbb{F}_q$ which are not isogenous to the Jacobian of any curve over $\mathbb{F}_{q^n}$. We use Honda Tate theory and construct an arithmetic measure with some conditions on its moments. First, we introduce our moment conditions which are motivated by the work of Tsfasman and Vladut~\cite{MR1465522}. \\ Let $X$ be a finite curve defined over $\mathbb{F}_q$. Let \( N_r(X):= \# X(\mathbb{F}_{q^r}) \) be the number of $\mathbb{F}_{q^r}$ of $X.$ By Weil's formula, we have \begin{equation}\label{positivity} 0 \leq N_r(X)=q^r+1-2q^{r/2}\sum_{j=1}^g\cos(r2\pi \theta_j), \end{equation} where $\sqrt{q}e^{2\pi i \theta_i}$ are Frobenius eigenvalues. We have \[ N_r(X)=\sum_{d|r}dM_d(X), \] where $M_d(X)$ is the number of points with degree $d.$ By Mobius inversion formula, \[ M_r(X)=\frac{1}{r}\sum_{d|r} \mu(d)N_{\frac{r}{d}}(X). \] In particular, we have \[ \sum_{d|r} \mu(d)N_{\frac{r}{d}}(X) \geq 0 \] for every $r\geq0.$ \\ Let $S_{\sqrt{\alpha}}:=\left\{ \sqrt{\alpha}e^{2\pi i\theta} : \theta \in [0,1) \right\}$ be the circle of radius $\sqrt{\alpha}$ centered at the origin, and defined the following probability measure on $S_{\sqrt{\alpha}}$ \[ d\mu_q:=\left(1+\sum_{k=1}^{\infty}\frac{c}{k^2}\cos(k2\pi \theta)\right) d\theta. \] where $c=\frac{6}{\pi^210}.$ Note that \begin{equation}\label{bound} 0 \leq 0.9 d\theta \leq d\mu_q \leq 1.1 d\theta. \end{equation} \begin{proposition}\label{arithme} Suppose that $\alpha>1.65$ then $U_{\mu_\alpha}(z)\geq 0$ for any $z\in \mathbb{C}.$ As a result, $d\mu_\alpha$ is an arithmetic measure. \end{proposition} \begin{proof} It is enough to show that $U_{\mu_\alpha}(z)\geq 0$ for any $z\in \mathbb{C}.$ Suppose that $|z|\geq 2\sqrt{\alpha}$, then \[ U_{\mu_\alpha}(z)=\int_{S_{\sqrt{\alpha}}} \log|z-\sqrt{\alpha}e^{2\pi i\theta}|d\mu_\alpha \geq \log\sqrt{\alpha}\geq 0. \] Otherwise, suppose that $|z|<2\sqrt{\alpha}.$ By \eqref{bound}, we have \[ \begin{split} U_{\mu_\alpha}(z)&=\int_{S_{\sqrt{\alpha}}} \log|z-\sqrt{\alpha}e^{2\pi i\theta}|d\mu_\alpha \\ &=\int_{ |z-\sqrt{\alpha}e^{2\pi i\theta}|<1} \log|z-\sqrt{\alpha}e^{2\pi i\theta}|d\mu_\alpha+ \int_{ |z-\sqrt{\alpha}e^{2\pi i\theta}|>1} \log|z-\sqrt{\alpha}e^{2\pi i\theta}|d\mu_\alpha \\ &\geq 1.1 \int_{ |z-\sqrt{\alpha}e^{2\pi i\theta}|<1} \log|z-\sqrt{\alpha}e^{2\pi i\theta}| d\theta + 0.9\int_{ |z-\sqrt{\alpha}e^{2\pi i\theta}|>1} \log|z-\sqrt{\alpha}e^{2\pi i\theta}| d\theta \\ &=1.1\int_{S_{\sqrt{\alpha}}} \log|z-\sqrt{\alpha}e^{2\pi i\theta}|d\theta- 0.2 \int_{ |z-\sqrt{\alpha}e^{2\pi i\theta}|>1} \log|z-\sqrt{\alpha}e^{2\pi i\theta}| d\theta \\ &\geq 1.1\log(\sqrt{\alpha})-0.2\log(3\sqrt{\alpha})>0. \end{split} \] This completes the proof of our propositions. \end{proof} \begin{corollary} There exists a sequence $\{A_g\}$ of abelian varities over $\mathbb{F}_q$ such that $\dim(A_g)=g$ and their Frobenious eigenvalues equidistributes with $\mu_q.$ \end{corollary} \begin{proof} Let $\alpha_n:=\sqrt{q}-\frac{1}{10n},$ $\Sigma_n:=[-2\alpha_n,2\alpha_n]$ and \( h_n(z):=z+\frac{\alpha_n}{z}. \) Note that $h_n$ is a conformal map with fixed point at infinity and derivative 1 infinity that sends $\mathbb{C}\backslash S_{\alpha_n}$ to the $\mathbb{C}\backslash \Sigma_n$. Let $h_nd\mu_{\alpha_n}$ be the push-forward of $\mu_{\alpha_n}$ by $h_n.$ By conformal in-variance of the potential function and Proposition~\ref{arithme}, $h_n d\mu_{\alpha_n}$ has a positive potential function and hence $h_nd\mu_{\alpha_n}$ is arithmetic. It follows from Theorem~\ref{general} that there exists a sequence of irreducible polynomial $\{p_m\}$ with real roots contained in $\Sigma_n(\frac{1}{100 n})\subset [2\sqrt{q},-2\sqrt{q}]$ with root distribution converging to $h_nd\mu_{\alpha_n}.$ Let \[ f_m(x):=x^{\deg{p_m}} p_m(x+\frac{q}{x}). \] Note that $f_m(x)$ has all its roots on $S_{\sqrt{q}}.$ Now by a diagonal argument and letting $n\to \infty$ and taking $m$ large enough it follows that there exists a sequecen of $\{A_g\}$ of abelian varities over $\mathbb{F}_q$ such that $\dim(A_g)=g$ and their Frobenious eigenvalues equidistributes with $\mu_q.$ \end{proof} \begin{theorem} Let $\{A_g\}$ be any family of abelian varieties over $\mathbb{F}_q$ such that $\dim(A_g)=g$ and their Frobenius eigenvalues distribution converges to $\mu_q.$ Given any integer $r\geq 0$, there exists $N$ such that if $g\geq N$ then $A_g$ is not isogenous to the Jacobian of any curve over $\mathbb{F}_{q^r}$. \end{theorem} \begin{proof} Suppose the contrary that there exists a sub-sequence $\{ A_{g_i}\}$ of abelian varieties over $\mathbb{F}_q$ such that their Frobenius eigenvalues equidistribute with $\mu_q$ and also they are isogenous to Jacobian of curves $\{ X_{g_i}\}$ over $\mathbb{F}_{q^r}$ for some $r\geq 0.$ By~\eqref{positivity}, it follows that \[ 0 \leq \frac{N_r(X_{g_i})}{g_i}= \frac{q^r+1-2q^{r/2}\sum_{j=1}^{g_i}\cos(r2\pi \theta_j)}{g_i}. \] By taking the limit of the above as $g_i\to \infty$, we obtain \[ 0 \leq -2q^{r/2} \lim_{g_i\to \infty}\frac{\sum_{j=1}^{g_i}\cos(r2\pi \theta_j)}{g_i}=-2q^{r/2} \int \cos(r2\pi \theta) d\mu_q= -q^{r/2}\frac{c}{r^2} <0, \] which is a contradiction. This completes the proof of our theorem. \end{proof} As was mentioned in the introduction, we constructed a polynomial using the algorithm described in section \ref{main} of the paper. Like the other results in section \ref{applications}, this was implemented with the LLL-algorithm and without using the construction that forces the roots to lie inside a desired support. The highest order terms were $$x^{290} - 28x^{289} - {484}x^{288} + 20784x^{287} + \cdots.$$ The largest coefficient of this polynomial has 105 digits. The roots all lie inside $[-2\sqrt{3},2\sqrt{3}]\subset\mathbb R$ and so by Honda-Tate theory, it represents an abelian variety over $\mathbb F_3$. Computing $N_2$ of this variety, we get $-2$. Thus it is not the Jacobian of any curve over $\mathbb F_9$. \begin{comment} For $r=1$, this gives us \[ M_1(X)= q+1- 2\sqrt{q}\sum_{j=1}^g \cos(\theta_j)\geq 0. \] In terms of the limiting measure, we have \[ \frac{q+1}{g} \geq \int x d\mu(x). \] As $g\to \infty,$ we have \[ \int x d\mu(x)\leq 0. \] The above inequality will be violated if the expected value would be positive. For $r=2$, we have \[ M_2(X)= N_2(X)-N_1(X)=\left(q^2+1-2q\sum_{j=1}^g\cos(2\theta_j) \right)- \left(q+1-2\sqrt{q}\sum_{j=1}^g \cos(\theta_j)\right)\geq 0. \] The above gives us another moment inequality:\\ \textcolor{blue}{INCOMPLETE SECTION} \end{comment} \bibliographystyle{alpha} \subsection{Finding Square-free Polynomials}\label{square_free} We see in section \ref{int_poly} that Theorem \ref{smith_3_2} requires a square-free polynomial $P$ to get the basis $\mathcal O$. To do this, we need to find monic, square-free integer polynomials of a given degree. Furthermore, we want this polynomial to have a small $n$-norm. The reason for this is that applying Theorem \ref{smith_3_2} gives a polynomial close to the original polynomial in the $n$-norm metric. This is justified in Corollary 3.3 in Smith. Since we know that the basis we created in Section \ref{lattice_bases} is short and has small $n$-norm, we can take a small linear-combination of these basis vectors to get a square-free polynomial of small norm. In practice, we can simply take a degree $n$-polynomial from this basis since it will usually be square-free. \newline \subsection{Chebyshev Polynomials}\label{cheb} Here we discuss the implementation of Proposition 4.1 in Smith. The idea of this proposition is to construct an oscillatory monic polynomial $Q$ given another real monic polynomial $P$ such that $PQ$ has $\deg(Q)$ of the highest degree coefficients being even (after the highest order term). \textcolor{red}{Explain why the oscillation is important. The oscillation forces $Q$ to have all of its roots in our interval $I$ and also that when we approximate with Theorem \ref{smith_3_2}, the number of roots in $I$ does not change}. \newline The way we achieve this is with Chebyshev polynomials. Given an interval $I$ with capacity $\kappa$, there is a unique polynomial $f$ of degree $n$ which achieves $\max |f(x)|=2\kappa^r$ precisely $n+1$ times where $\kappa$ is the capacity of $I$. We will call this the Chebyshev polynomial on interval $I$. \textcolor{red}{A proof for this can be seen in Lemma 4.3 of Smith}. We algorithmically compute $Q$ as described above in the manner suggested in Smith's proof of Proposition 4.1 in Smith. \subsection{Main Algorithm} First, we find the short square-free, integer polynomial as described in section \ref{square_free}. We then find polynomial $Q$ of degree $m$ (\textcolor{red}{explain where $m$ comes from}) as described in section \ref{cheb} using $P=\tilde{P}_{\mu,n+2}$. Suppose $$PQ=x^{n+m}+a_{n+m-1}x^{n+m-1}+\dots+a_nx^n+b_{n-1}x^{n-1}+\dots+b_0$$ where all of the $a_i$'s are even integers. Then apply the algorithm discussed in section \ref{int_poly} to $\frac{1}{4}\left(\sum_{i=0}^nb_ix^i\right)+\frac{1}{2}$ get an integer polynomial $H$ close to the original. Finally, $$x^{n+m}+a_{n+m-1}x^{n+m-1}+\dots+a_nx^n+4H-2$$ is an Eisenstein polynomial at $2$ close in $(n+m)$-norm to $PQ$. If the norm between $PQ$ and the Eisenstein polynomial is small enough, the roots are forced to remain inside $I$.
{ "timestamp": "2023-02-07T02:31:16", "yymm": "2302", "arxiv_id": "2302.02872", "language": "en", "url": "https://arxiv.org/abs/2302.02872", "abstract": "Let $\\Sigma \\subset \\mathbb{C}$ be a compact subset of the complex plane, and $\\mu$ be a probability distribution on $\\Sigma$. We give necessary and sufficient conditions for $\\mu$ to be the weak* limit of a sequence of uniform probability measures on a complete set of conjugate algebraic integers lying eventually in any open set containing $\\Sigma$. Given $n\\geq 0$, any probability measure $\\mu$ satisfying our necessary conditions, and any open set $D$ containing $\\Sigma$, we develop and implement a polynomial time algorithm in $n$ that returns an integral monic irreducible polynomial of degree $n$ such that all of its roots are inside $D$ and their root distributions converge weakly to $\\mu$ as $n\\to \\infty$. We also prove our theorem for $\\Sigma\\subset \\mathbb{R}$ and open sets inside $\\mathbb{R}$ that recovers Smith's main theorem~\\cite{Smith} as special case. Given any finite field $\\mathbb{F}_q$ and any integer $n$, our algorithm returns infinitely many abelian varieties over $\\mathbb{F}_q$ which are not isogenous to the Jacobian of any curve over $\\mathbb{F}_{q^n}$.", "subjects": "Number Theory (math.NT)", "title": "Limiting distributions of conjugate algebraic integers", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.988491852291787, "lm_q2_score": 0.810478913248044, "lm_q1q2_score": 0.8011518021999936 }
https://arxiv.org/abs/1907.02004
On Hamiltonian cycles in balanced $k$-partite graphs
For all integers $k$ with $k\geq 2$, if $G$ is a balanced $k$-partite graph on $n\geq 3$ vertices with minimum degree at least \[ \left\lceil\frac{n}{2}\right\rceil+\left\lfloor\frac{n+2}{2\lceil\frac{k+1}{2}\rceil}\right\rfloor-\frac{n}{k}=\begin{cases} \lceil\frac{n}{2}\rceil+\lfloor\frac{n+2}{k+1}\rfloor-\frac{n}{k} & : k \text{ odd }\\ \frac{n}{2}+\lfloor\frac{n+2}{k+2}\rfloor-\frac{n}{k} & : k \text{ even } \end{cases}, \] then $G$ has a Hamiltonian cycle unless $k=2$ and 4 divides $n$, or $k=\frac{n}{2}$ and 4 divides $n$. In the case where $k=2$ and 4 divides $n$, or $k=\frac{n}{2}$ and 4 divides $n$, we can characterize the graphs which do not have a Hamiltonian cycle and see that $\left\lceil\frac{n}{2}\right\rceil+\left\lfloor\frac{n+2}{2\lceil\frac{k+1}{2}\rceil}\right\rfloor-\frac{n}{k}+1$ suffices. This result is tight for all $k\geq 2$ and $n\geq 3$ divisible by $k$.
\section{Introduction} The study of Hamiltonian cycles in balanced $k$-partite graphs begins with the following classic results of Dirac, and Moon and Moser. Dirac \cite{D} proved that for all graphs $G$ on $n\geq 3$ vertices, if $\delta(G)\geq \ceiling{\frac{n}{2}}$, then $G$ has a Hamiltonian cycle. Moon and Moser \cite{MM} proved that for all balanced bipartite graphs $G$ on $n\geq 4$ vertices, if $\delta(G)\geq \frac{n+2}{4}$, then $G$ has a Hamiltonian cycle. Over 30 years later Chen, Faudree, Gould, Jacobson, and Lesniak \cite{CFGJL} beautifully tied these results together by proving that for all $k\geq 2$, if $G$ is a balanced $k$-partite graph on $n$ vertices with \begin{equation}\label{cfgjl} \delta(G)> \frac{n}{2}+\frac{n}{2\ceiling{\frac{k+1}{2}}}-\frac{n}{k}, \end{equation} then $G$ has a Hamiltonian cycle. It turns out that while their result is nearly optimal, in most cases the degree condition can be improved by 1. The purpose of this note is simply to provide the precise minimum degree condition in all cases thereby filling the lacuna in the above result (and as we point out in the Appendix, it is unfortunately not as simple as replacing the strict inequality in \eqref{cfgjl} with a weak inequality). \begin{thm}\label{main} Let $k$ be an integer with $k\geq 2$. For all balanced $k$-partite graphs $G$ on $n$ vertices, if \begin{equation}\label{cf} \delta (G) \geq \ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k}=\Dk, \end{equation} then $G$ has a Hamiltonian cycle unless $k=2$ and 4 divides $n$, or $k=\frac{n}{2}$ and 4 divides $n$. \end{thm} Since a graph on $n$ vertices can be viewed as a $k$-partite graph with $k=n$, note that when $k=n$, we have $\ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k}=\ceiling{\frac{n}{2}}$ and thus Theorem \ref{main} reduces to Dirac's theorem. When $k=2$, we have $\ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k}=\floor{\frac{n+2}{4}}$ and thus when 4 does not divide $n$, Theorem \ref{main} reduces to Moon and Moser's theorem; and when 4 does divide $n$, Ferrara, Jacobson, and Powell \cite{FJP} characterized all balanced bipartite graphs $G$ on $n\geq 4$ vertices such that $\delta(G)\geq \frac{n}{4}$, yet $G$ does not have a Hamiltonian cycle. So our proof will only handle the cases when $3\leq k\leq \frac{n}{2}$. We will also prove the following which will handle the case when $k=\frac{n}{2}$ and 4 divides $n$. Together with the results in \cite{FJP}, this gives a complete characterization of balanced $k$-partite graphs $G$ on $n$ vertices which satisfy $\delta (G) \geq \ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k}$, but do not have a Hamiltonian cycle. \begin{prop}\label{prop:n=2k} Let $n\geq 8$ be divisible by $4$, let $k=\frac{n}{2}$, and let $G$ be a balanced $k$-partite graph on $n$ vertices. If $\delta(G)\geq \frac{n}{2}+\floor{\frac{n+2}{k+2}}-\frac{n}{k}=\frac{n}{2}-1$ and $G$ does not have a Hamiltonian cycle, then $G$ belongs to one of the families of examples described in Example \ref{n=2k}. \end{prop} \subsection{Overview} We give the lower bound examples in Section \ref{sec:example}, we collect the main lemmas in Section \ref{sec:example} (while it is a combination of existing results, Lemma \ref{domcycle} may be of independent interest), we deal with the first two exceptions in Section \ref{sec:filter} before starting the main proof in Section \ref{sec:main}. Finally, in the Appendix, we collect some numerical lemmas which are needed because of the floors and ceilings appearing in \eqref{cf}. This project grew out of an earlier work of the first author together with Krueger, Pritikin, and Thompson \cite{DKPT}, where we considered Hamiltonian cycles in unbalanced $k$-partite graphs for $k\geq 3$. The upcoming Example \ref{tight_gen} first appeared in a more general form in \cite{DKPT}. In fact, it was this example which indicated to us that \eqref{cfgjl} is not always tight. By using Theorem \ref{chv} and Lemma \ref{domcycle}, we were able to streamline the original proof of Chen et al.\ with the correct degree condition; however, because of the unexpected (to us) exceptional cases which arose when $k=\frac{n}{2}$, our overall proof didn't end up being any shorter than the original. Again, we emphasize that Chen et al.\ have a beautiful result which places Dirac's theorem and Moon and Moser's theorem on a common spectrum. It is only because of the fundamental nature of these results that we have expended the effort necessary to provide the tight degree condition in all cases. \subsection{Notation} For $S\subseteq V(G)$, we let $N(S) = \bigcup_{v}N(v)$ and $\overline{S}= V(G)\setminus S$. Given disjoint sets $A, B\subseteq V(G)$, we let $\delta(A,B)=\min\{|N(v)\cap B|: v\in A\}|$. Given a cycle $v_1v_2\dots v_kv_1$, $i\in [k]$, and an integer $t$, we assume that the addition in the indices, such as $v_{i+t}$, is taken modulo $k$. \section{Tightness examples}\label{sec:example} \begin{exa}\label{tight_gen} For all $k\geq 2$ and all $n$ divisible by $k$, there exists a family $\mathcal{F}$ of balanced $k$-partite graphs on $n$ vertices such that for all $F\in \mathcal{F}$, $$\delta(F)\geq \ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k}-1,$$ but $F$ does not have a Hamiltonian cycle. \end{exa} \begin{figure}[ht] \centering \begin{tikzpicture}[scale = .25] \draw[pattern=north west lines, pattern color=gray] (-18, 1) rectangle (-16, 2); \draw[pattern=north west lines, pattern color=gray] (-16, 1) rectangle (-12, 1.5); \draw[pattern=north west lines, pattern color=gray] (-18, -6) rectangle (-4, 1); \draw (-21, 1) node[left]{$\floor{\frac{\ceiling{\frac{n+1}{2}}}{\ceiling{\frac{k+1}{2}}}}$} -- (-19, 1); \draw[dashed] (-20, 1) -- (8, 1); \draw (-18, -6) node[below right]{$V_1$} rectangle (-16, 4); \draw (-16, -6) node[below right]{$V_2$} rectangle (-14, 4); \draw (-14, -6) rectangle (-12, 4); \draw (-12, -6) rectangle (-10, 4); \draw (-10, -6) rectangle (-8, 4); \draw (-8, -6) rectangle (-6, 4); \draw (-6, -6) node[below]{$~~~~~~~~~V_{\ceiling{\frac{k+1}{2}}}$} rectangle (-4, 4); \draw (-4, -6) rectangle (-2, 4); \draw (-2, -6) rectangle (0, 4); \draw (0, -6) rectangle (2, 4); \draw (2, -6) rectangle (4, 4); \draw (4, -6) rectangle (6, 4); \draw (6, -6) node[below]{$~~~~V_k$} rectangle (8, 4); \end{tikzpicture} \caption{The family of graphs $\mathcal{F}$. The shaded sets represent $X_1, \dots, X_{\ceiling{\frac{k+1}{2}}}$.} \end{figure} \begin{proof} First note that \begin{equation}\label{equal} \floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}=\floor{\frac{\ceiling{\frac{n+1}{2}}}{\ceiling{\frac{k+1}{2}}}}. \end{equation} Since if $k$ is even, then both sides of the equation equal $\floor{\frac{n+2}{k+2}}$; if $k$ is odd and $n$ is even, then both sides of the equation equal $\floor{\frac{n+2}{k+1}}$; and if $k$ is odd and $n$ is odd then we get that $\floor{\frac{n+1}{k+1}} = \floor{\frac{n+2}{k+1}}$, which is true since $\frac{n+2}{k+1}$ is not an integer. Let $\mathcal{F}$ be the family of graphs which can be obtained from a complete $k$-partite graph with parts $V_1, \dots, V_k$ such that $|V_i|=\frac{n}{k}$ for all $i\in [k]$, by selecting some $X_i\subseteq V_i$ for all $i\in [\ceiling{\frac{k+1}{2}}]$ such that $|X_1|\geq \dots\geq |X_{\ceiling{\frac{k+1}{2}}}|=\floor{\frac{\ceiling{\frac{n+1}{2}}}{\ceiling{\frac{k+1}{2}}}}$ and $|X_1\cup\dots\cup X_{\ceiling{\frac{k+1}{2}}}|=\ceiling{\frac{n+1}{2}}$. Add all edges between parts except for those between a vertex in $X_i$ and $X_j$ for all $i, j \in [\ceiling{\frac{k+1}{2}}]$. Note that every $F\in \mathcal{F}$ has an independent set of size $\ceiling{\frac{n+1}{2}}$ and thus does not contain a Hamiltonian cycle. Finally to see that the degree condition is satisfied, let $i\in [\ceiling{\frac{k+1}{2}}]$ and let $v\in X_{i}$. We have by \eqref{equal} \begin{align*} d(v)&=(1-\frac{1}{k})n-|X_1\cup \dots X_{i-1}\cup X_{i+1}\cup \dots\cup X_{\ceiling{\frac{k+1}{2}}}|\\ &\geq (1-\frac{1}{k})n-\left(\ceiling{\frac{n+1}{2}}-\floor{\frac{\ceiling{\frac{n+1}{2}}}{\ceiling{\frac{k+1}{2}}}}\right) =\ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k}-1. \qedhere \end{align*} \end{proof} \begin{exa}\label{n=2k} Let $n\geq 8$ be divisible by $4$ and let $k=\frac{n}{2}$. \begin{enumerate}[label=\emph{(\roman*)}] \item\label{e1} There exists a family $\mathcal{F}_1$ of balanced $k$-partite graphs on $n$ vertices such that for all $F_1\in \mathcal{F}_1$, $\delta(F_1)\geq \frac{n}{2}+\floor{\frac{n+2}{k+2}}-\frac{n}{k}=\frac{n}{2}-1$, but $\kappa(F_1)\leq 1$ and thus $F_1$ does not have a Hamiltonian cycle. \item\label{e2} There exists a 2-connected balanced 4-partite graph $F_2$ on 8 vertices with $\alpha(F_2)=3$ such that $F_2$ does not have a Hamiltonian cycle. \item\label{e3} There exists a family $\mathcal{F}_3$ of balanced $k$-partite graphs on $n$ vertices such that for all $F_3\in \mathcal{F}_3$, $\delta(F_3)\geq \frac{n}{2}+\floor{\frac{n+2}{k+2}}-\frac{n}{k}=\frac{n}{2}-1$, $\kappa(F_2)\geq 2$, and $\alpha(F_3)=\frac{n}{2}$, but $F_3$ does not have a Hamiltonian cycle. \end{enumerate} \end{exa} \begin{figure}[ht] \centering \includegraphics[scale=1]{F2.pdf} \caption{The graph $F_2$.}\label{fig_F2} \end{figure} \begin{proof} \begin{enumerate} \item Let $V_i=\{x_i, y_i\}$ for all $i\in [k]$. Add all edges inside $\{x_1, \dots, x_k\}$, add all edges from $y_k$ to $\{y_1, \dots, y_{k-1}\}$, and for all $i\in [k-1]$ add at least $k-1$ edges from $y_i$ to $\{y_1, \dots, y_{i-1}, y_{i+1}, \dots, y_{k-1}\}\cup \{x_{k}\}$. Let $\mathcal{F}_1$ be the family of graphs thus obtained. Note that every graph $F_1\in \mathcal{F}_1$ has $\delta(F_1)=k-1=\frac{n}{2}-1$ and $\kappa(F_1)\leq 1$ and thus $F_1$ does not have a Hamiltonian cycle. \item We let $F_2$ be the graph in Figure \ref{fig_F2} which can be seen to be a balanced 4-partite graph (with vertices of the same shape being in the same part of the partition) which is 2-connected and has $\alpha(F_2)=3$. Note that $F_2$ has no Hamiltonian cycle since $G-x_1-x_4$ has three components. \item Let the parts be labeled $X_1, \dots, X_{k/2}$, $Y_1, \dots, Y_{k/2}$ and let $X=\cup_{i=1}^{k/2} X_i$ and $Y=\cup_{i=1}^{k/2}Y_i$. Let $y'\in Y_{k/2}$, let $y''\in Y\setminus \{y'\}$, and let $x'\in X$. Add all edges between $X$ and $Y\setminus \{y', y''\}$, all edges from $y''$ to $X\setminus \{x'\}$, and all edges from $y'$ to $\{x'\}\cup (Y\setminus Y_{k/2})$. Furthermore, we may add any number of other edges between the parts $Y_1, \dots, Y_{k/2}$ and we may add the edge $x'y''$. Let $\mathcal{F}_3$ be the family of graphs thus obtained. Let $F_3\in \mathcal{F}_3$ and let $H$ be the bipartite graph induced by $[X,Y]$. It is easily seen that $\delta(F_3)\geq \frac{n}{2}-1$, $\kappa(F_2)\geq 2$, and $\alpha(F_3)=\frac{n}{2}$. Since $X$ is an independent set, if $F_3$ has a Hamiltonian cycle, it must be in $H$; however, since $y'$ has degree 1 in $H$, there is no Hamiltonian cycle in $H$. \end{enumerate} \end{proof} \section{General lemmas}\label{sec:lemmas} In this section we state three general results which are useful for finding Hamiltonian cycles, beginning with two classics. \begin{thm}[Dirac \cite{D}]\label{dir} Let $n\geq d\geq 3$. If $G$ is 2-connected and $\delta(G)\geq d/2$, then $G$ has a cycle of length at least $d$. \end{thm} \begin{thm}[Chv\'atal \cite{C}]\label{chv} Let $G=(U,V,E)$ be a bipartite graph on $n\geq 4$ vertices with vertex sets $U=\{u_1, \dots, u_{n/2}\}$ and $V=\{v_1, \dots, v_{n/2}\}$. If for all $1\leq k<n/2$, $$d(v_k)\leq k \Rightarrow d(u_{n/2-k})\geq \frac{n}{2}-k+1,$$ then $G$ has a Hamiltonian cycle. \end{thm} The main lemma which we use to begin the proof of Theorem \ref{main} and Proposition \ref{prop:n=2k} is the following combination of the well known result of Nash-Williams \cite{NW} and a (slight weakening of a) result of Bauer, Veldman, Morgana, Schmeichel \cite{BVMS}. We provide a proof for completeness. We say that a cycle $C$ in a graph $G$ is \emph{strongly dominating} if $V(G)\setminus V(C)$ is an independent set and no two vertices of $\bigcup_{u\in V(G)\setminus V(C)}N(u)$ appear consecutively on $C$. \begin{lem}[see {\cite[Lemmas 1,2,3,4]{NW}} and {\cite[Lemma 8]{BVMS}}]\label{domcycle} Let $G$ be a graph on $n$ vertices. If $G$ is 2-connected and $\delta(G)\geq \frac{n+2}{3}$, then every longest cycle of $G$ is strongly dominating. \end{lem} \begin{proof} Let $C=v_1v_2\dots v_kv_1$ be a longest cycle in $G$ and let $P=u_1u_2\dots u_r$ be a longest path in $G-C$. If $r\leq 1$, then we are done; so suppose $r\geq 2$. We have $k+r\leq n$ and note that by Theorem \ref{dir}, we have $k\geq 2\delta(G)\geq \frac{2n+4}{3}$ and thus \begin{equation}\label{nrlower} 3r+4\leq n. \end{equation} Let $X=N(u_1)\cap V(C)$ and $Y=N(u_r)\cap V(C)$ and note that \begin{equation}\label{XYlower} |X|, |Y|\geq \delta(G)-(r-1). \end{equation} The key observation is that by the maximality of $C$, no two vertices in $X\cup Y$ are consecutive along $C$, and furthermore if $v_i\in X\cap Y$, then none of $v_{i-r}, \dots, v_{i-1}, v_{i+1}, \dots, v_{i+r}$ are in $X\cup Y$. First suppose that $X\subseteq Y$ or $Y\subseteq X$; without loss of generality $X\subseteq Y$. In this case we have by \eqref{XYlower}, \begin{equation}\label{Xineq} n-r\geq k\geq (r+1)|X|\geq (r+1)(\delta(G)-(r-1)). \end{equation} First suppose $r=2$, in which case \eqref{Xineq} becomes $n-2\geq 3\left(\frac{n+2}{3}-1\right)=n-1$, a contradiction. Now suppose $r\geq 3$ in which case \eqref{Xineq} becomes $$n\leq \frac{3r^2-5r-5}{r-2}=3r+1-\frac{3}{r-2},$$ contradicting \eqref{nrlower}. Now suppose that $X\setminus Y\neq \emptyset$ and $Y\setminus X\neq \emptyset$. There are vertices $v_i, v_j\in V(C)$ with the following properties: $v_i\in X\setminus Y$ and the next vertex $v_{i'}\in X\cup Y$ which appears after $v_i$ satisfies $v_{i'}\in Y$ (meaning that $i'\geq i+r+1$), and $v_j\in Y\setminus X$ and the next vertex $v_{j'}\in X\cup Y$ from $X\cup Y$ which appears after $v_j$ satisfies $v_{j'}\in X$ (meaning that $j'\geq j+r+1$). Each vertex of $((X\setminus Y)\cup (Y\setminus X))\setminus \{v_i, v_j\}$ is followed by at least one vertex from $V(C)\setminus (X\cup Y)$ and each vertex of $(X\cap Y)\cup \{v_i, v_j\}$ is followed by at least $r$ vertices from $V(C)\setminus (X\cup Y)$. So we have \begin{align*} n-r\geq k&\geq 2(|X\setminus Y|+|Y\setminus X|-2)+(r+1)(|X\cap Y|+2)\\ &=|X|+|Y|+|X\cup Y|+(r-2)|X\cap Y|+2(r-1)\\ &\geq \min\{4\delta(G)-2(r-1), 3\delta(G)-1\}\\ &\geq \min\{4(n+2)/3-2(r-1), n+1\}, \end{align*} where the second to last inequality is seen by using \eqref{XYlower} and splitting into cases whether $|X\cap Y|=0$ or not. However, $4(n+2)/3-2(r-1)\leq n-r$ implies $n\leq 3r-14$, contradicting \eqref{nrlower}. To see that the second part of the definition of strongly dominating is satisfied, suppose that $C=v_1\dots v_kv_1$ is a longest cycle and suppose $V(G)\setminus V(C)=\{u_1, \dots, u_r\}$ is an independent set. If $|V(G)\setminus V(C)|\leq 1$, we are done, so suppose $r\geq 2$. Let $X=N(u_1)\cap V(C)$ and suppose (without loss of generality) for contradiction that $v_1\in X$ and $v_2\in N(u_2)$. By the maximality of $C$, this implies that $v_3\not \in N(u_2)$ and for all $i\geq 3$, if $v_i\in X$, then $v_{i+1}, v_{i+2}\not\in N(u_2)$. Since $k\leq n-2$, this implies that $$\frac{n+2}{3}\leq |N(u_2)|\leq k-(2|N(u_1)|-1)\leq n-1-2\left(\frac{n+2}{3}\right)=\frac{n-7}{3},$$ a contradiction. \end{proof} \section{Filtering out $\mathcal{F}_1$ and $F_2$}\label{sec:filter} We want to say that every balanced $k$-partite graph satisfying \eqref{cf} is both 2-connected and every longest cycle in strongly dominating; however, there are two exceptions and we deal with those exceptions before beginning the main proof in next Section. First, we show that every balanced $k$-partite graph satisfying \eqref{cf} is either $2$-connected or belongs to the family $\mathcal{F}_1$ in Example \ref{n=2k}.\ref{e1}. \begin{lem}\label{F1} Let $k\geq 3$, let $n$ be an integer such that $n\geq 2k$, and let $G$ be a balanced $k$-partite graph on $n$ vertices. If $$\delta(G) \geq \ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k},$$ then $G$ is 2-connected unless $n=2k$ and $G\in \mathcal{F}_1$ (see Example \ref{n=2k}.\ref{e1}). \end{lem} \begin{proof} Let $V_1, \dots, V_k$ be the parts of $G$. Suppose for contradiction that $\kappa(G)\leq 1$. Let $A, B, C$ be a partition of $V(G)$ such that $|C|\leq 1$ and $G-C$ is not connected. First suppose that there exists $i\in [k]$ such that $V_i\subseteq A\cup C$ or $V_i\subseteq B\cup C$. Without loss of generality suppose $V_i\subseteq B\cup C$ and let $u\in A$ and $v\in B\cap V_i$. We have $$2\delta(G)\leq d(u)+d(v)\leq |A|+|C|-1+|B|+|C|-\frac{n}{k}=(1-\frac{1}{k})n+|C|-1\leq (1-\frac{1}{k})n,$$ contradicting Fact \ref{ineq}.\ref{f1} when $k$ is even and $n\geq 3k$, and contradicting Fact \ref{ineq}.\ref{f2} when $k$ is odd and $n\geq 2k$. So unless $k$ is even and $n=2k$, we must have that for all $i\in [k]$, $V_i\cap A\neq \emptyset$ and $V_i\cap B\neq \emptyset$. Either $C=\emptyset$ and we let $u\in A$ and $v\in B$, or $C\neq \emptyset$ and suppose without loss of generality that $V_1\cap C\neq \emptyset$ in which case we let $u\in V_1\cap A$ and $v\in V_1\cap B$. Either way we have $$2\delta(G)\leq d(u)+d(v)\leq n-\frac{n}{k},$$ contradicting Fact \ref{ineq}.\ref{f1} when $k$ is even and $n\geq 3k$, and contradicting Fact \ref{ineq}.\ref{f2} when $k$ is odd and $n\geq 2k$. Finally, suppose $k$ is even and $n=2k$ which implies $\delta(G)\geq \frac{n}{2}-1$. For all $u\in A$ we have $\frac{n}{2}-1\leq d(u)\leq |A|+|C|-1$ which implies $|A|\geq \frac{n}{2}-|C|$ and for all $v\in B$ we have $\frac{n}{2}-1\leq d(v)\leq |B|+|C|-1$ which implies $|B|\geq \frac{n}{2}-|C|$. If $C=\emptyset$, this implies $|A|=\frac{n}{2}$ and $|B|=\frac{n}{2}$. If $C\neq \emptyset$, we have $|A|\geq \frac{n}{2}-1$ and $|B|\geq \frac{n}{2}-1$, so without loss of generality suppose $|A|+|C|=\frac{n}{2}$ and $|B|=\frac{n}{2}$. If there exists $i\in [k]$ such that $V_i\subseteq A\cup C$ or $V_i\subseteq B$; say $V_i\subseteq A\cup C$, then for $u\in V_i\cap A$, we have \[\frac{n}{2}-1\leq d(u)\leq |A|+|C|-|V_i|=\frac{n}{2}-2,\] a contradiction. Thus for all $i\in [k]$, we have $|V_i\cap (A\cup C)|=1$ and $|V_i\cap B|=1$. Let $V_i=\{x_i,y_i\}$ for all $i\in [k]$ and let $X=\{x_1, \dots, x_k\}$ and $Y=\{y_1, \dots, y_k\}$ and suppose $X=A\cup C$ and $Y=B$. So it must be the case that every vertex $u\in A$ is adjacent to precisely the vertices in $X\setminus \{u\}$ which means $G[X]$ is a clique. Also every vertex $v\in Y$ is adjacent to at least $\frac{n}{2}-1$ of the $\frac{n}{2}$ vertices in $(Y\cup C)\setminus \{v\}$. Thus $G\in \mathcal{F}_1$. \end{proof} We now prove a lemma which shows that when $G$ is a 2-connected balanced $k$-partite graph satisfying \eqref{cf}, we either have that every longest cycle in $G$ is strongly dominating or $G$ is isomorphic to the graph $F_2$ in Example \ref{n=2k}.\ref{e2}. \begin{lem}\label{F2} Let $k\geq 3$ and let $n\geq 2k$ and let $G$ be a balanced $k$-partite graph on $n$ vertices. If $G$ is 2-connected and $\delta(G)\geq \ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k}$, then every longest cycle of $G$ is strongly dominating unless $n=8$ and $G\cong F_2$ (see Example \ref{n=2k}.\ref{e2}). \end{lem} \begin{proof} We will show that, unless $n=8$ and $k=4$, we have $\delta(G)\geq \ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k}\geq \frac{n+2}{3}$ and thus we are done by Lemma \ref{domcycle}. First suppose $k$ is odd, in which case $\ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k}=\ceiling{\frac{n}{2}}+\floor{\frac{n+2}{k+1}}-\frac{n}{k}$. First note that when $k=3$ and $n=6$ or $n=9$ we have by direct inspection that $\ceiling{\frac{n}{2}}+\floor{\frac{n+2}{k+1}}-\frac{n}{k}\geq \frac{n+2}{3}$. So in the remaining cases we have \begin{equation}\label{n10} n\geq 10\geq 10-\frac{24k-60}{k^2+k-6}=\frac{2k(5k-7)}{k^2+k-6}. \end{equation} Thus, using Fact \ref{kodd}, we have \begin{align*} \ceiling{\frac{n}{2}}+\floor{\frac{n+2}{k+1}}-\frac{n}{k}-\frac{n+2}{3}&\geq \frac{n}{2}+\frac{n+2-(k-1)}{k+1}-\frac{n}{k}-\frac{n+2}{3}\\ &=\left(\frac{1}{6}-\frac{1}{k(k+1)}\right)n-\frac{2}{3}-\frac{k-3}{k+1}\\ &\stackrel{\eqref{n10}}{\geq} \left(\frac{1}{6}-\frac{1}{k(k+1)}\right)\frac{2k(5k-7)}{k^2+k-6}-\frac{2}{3}-\frac{k-3}{k+1}=0, \end{align*} as desired. Now suppose $k$ is even, in which case $\ceiling{\frac{n}{2}}+\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}-\frac{n}{k}=\frac{n}{2}+\floor{\frac{n+2}{k+2}}-\frac{n}{k}$. Note that aside from the case $n=8$ and $k=4$ we have \begin{equation}\label{n12} n\geq 12\geq 12-\frac{2(k+18)(k-4)}{k^2+2k-12}=\frac{2k(5k-2)}{k^2+2k-12}. \end{equation} Thus \begin{align*} \frac{n}{2}+\floor{\frac{n+2}{k+2}}-\frac{n}{k}-\frac{n+2}{3}&\geq \frac{n}{2}+\frac{n+2-k}{k+2}-\frac{n}{k}-\frac{n+2}{3}\\ &=\left(\frac{1}{6}-\frac{2}{k(k+2)}\right)n-\frac{2}{3}-\frac{k-2}{k+2}\\ &\stackrel{\eqref{n12}}{\geq} \left(\frac{1}{6}-\frac{2}{k(k+2)}\right)\frac{2k(5k-2)}{k^2+2k-12}-\frac{2}{3}-\frac{k-2}{k+2}=0, \end{align*} as desired. Finally suppose $n=8$ and $k=4$ and let $C$ be a longest cycle of $G$. Since $G$ is 2-connected and $\delta(G)\geq \frac{8}{2}+\floor{\frac{10}{6}}-\frac{8}{4}=3$, Theorem \ref{dir} implies that $C$ has length at least $6$. If $C$ had length at least 7, it would be a strongly dominating cycle, so suppose $C$ has length 6. Let $C=x_1x_2x_3x_4x_5x_6$ and let $x'$ and $x''$ be the two vertices in $V(G)\setminus V(C)$. If $x'x''\not\in E(G)$, then by the maximality of $C$ it is easily seen that, without loss of generality, $N(x')=N(x'')=\{x_1,x_3,x_5\}$ and thus $C$ is strongly dominating; so suppose that $x'x'' \in E(G)$. Without loss of generality suppose $x'x_1 \in E(G)$. If either $x''x_2\in E(G)$ or $x''x_6\in E(G)$, then $G$ has a Hamiltonian cycle; and if either $x''x_3\in E(G)$ or $x''x_5\in E(G)$, then $G$ has a cycle longer than $C$, a contradiction. Since $\delta(G)\geq 3$, this forces $x''x_1,x''x_4 \in E(G)$. By the same argument we get $x'x_4 \in E(G)$. If $x_6x_3 \in E(G)$, then $x_6x_3x_2x_1x'x''x_4x_5x_6$ is a Hamiltonian cycle, so $x_6x_3 \not \in E(G)$, and by symmetry $x_2x_5 \not \in E(G)$. If $x_6x_2 \in E(G)$, then $x_6x_2x_3x_4x'x''x_1x_6$ is a cycle longer than $C$, a contradiction. So $x_6x_2 \not \in E(G)$ and by symmetry $x_3x_5 \not \in E(G)$. Since $\delta(G)\geq 3$, this forces $x_6x_4,x_5x_1,x_2x_4,x_3x_1 \in E(G)$. Therefore $G \cong \mathcal{F}_2$. \end{proof} \section{Proof of Theorem \ref{main} and Proposition \ref{prop:n=2k}}\label{sec:main} Let $k\geq 3$ and let $G$ be a balanced $k$-partite graph on $n$ vertices. Let $V_1,V_2, \dots, V_k$ denote the parts and note that $|V_i|=\frac{n}{k}=:m $ for all $i\in [k]$. Since the case $k=n$ is Dirac's theorem, we suppose $k\leq \frac{n}{2}$ and since the case $k=2$ is handled in \cite{MM} and \cite{FJP}, we suppose $k\geq 3$. Furthermore, if $k=\frac{n}{2}$, we suppose that $G\not\in \mathcal{F}_1$ and $G\not\cong F_2$ (see Example \ref{n=2k}). Now let $C$ be a maximum length cycle and suppose for contradiction that $C$ is not Hamiltonian. By Lemma \ref{F1} and Lemma \ref{F2} we may assume that $C$ is strongly dominating. Without loss of generality, let $z\in V_1\setminus V(C)$. Let $S=(V(G)\setminus V(C))\cup \{v_{i+1}: v_i\in N(z)\}$ and $R=(V(G)\setminus V(C))\cup \{v_{i-1}: v_i\in N(z)\}$ and note that \begin{equation} \label{SR} |S|, |R|\geq \delta(G)+1. \end{equation} Since $C$ is strongly dominating, both $S$ and $R$ are independent sets. For each $i \in [k]$, set $$S_i = S \cap V_i \text{ and } R_i = R \cap V_i.$$ Define $\ell=|\{i\in [k]: S_i\neq \emptyset\}|$ and $\ell'=|\{i\in [k]: R_i\neq \emptyset\}|$ and without loss of generality suppose $$\ell\leq \ell'.$$ Furthermore, without loss of generality, we may suppose that \begin{align*} S_i \neq \emptyset \text{ for all } i\in [\ell] \text{ and } S_j=\emptyset \text{ for all } j\in [k]\setminus [\ell]. \end{align*} \begin{cla}\label{elllower} $\ell, \ell'\geq \ceiling{\frac{k}{2}}$ \end{cla} \begin{proof} We claim that $|R|,|S|\geq \delta(G)+1>(\ceiling{\frac{k}{2}}-1)\frac{n}{k}$, which implies the result. Indeed, we have \begin{equation}\label{k/2} \delta(G) + 1 - (\ceiling{\frac{k}{2}} - 1)\frac{n}{k} \geq \ceiling{\frac{n}{2}} + \floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}} + 1 - \frac{n}{k}\ceiling{\frac{k}{2}}. \end{equation} When $k$ is even, \eqref{k/2} reduces to $\floor{\frac{n+2}{k+2}} + 1 > 0$, and when $k$ is odd, by Fact \ref{kodd}, \eqref{k/2} reduces to $\frac{n}{2}+\frac{n}{k+1}-\frac{k-3}{k+1}-\frac{n}{k}\left(\frac{k+1}{2}\right)+1=\frac{n}{k+1}-\frac{n}{2k}+1-\frac{k-3}{k+1}=\frac{(k-1)n}{2k(k+1)}+\frac{4}{k+1}>0$. \end{proof} \begin{cla}\label{Si}~ \begin{enumerate}[label=\emph{(\roman*)}] \item \label{c1} For all $y \in S$, $|\overline {N(y)} \setminus S| \leq n-2\delta(G)-1.$ For all $y \in R$, $|\overline {N(y)} \setminus R| \leq n-2\delta(G)-1.$ \item \label{c2} For all $i\in [k]$, if $S_i\neq \emptyset$, then $|S_i| \geq 2\delta(G)+1-(1-\frac{1}{k})n\geq \frac{1}{2}(\frac{n}{k}-(\floor{\frac{n-1}{2}}-\delta(G))).$ For all $i\in [k]$, if $R_i\neq \emptyset$, then $|R_i| \geq 2\delta(G)+1-(1-\frac{1}{k})n\geq \frac{1}{2}(\frac{n}{k}-(\floor{\frac{n-1}{2}}-\delta(G))).$ \end{enumerate} \end{cla} \begin{proof} \begin{enumerate} \item Since $C$ is a longest cycle of $G$, the vertex subsets $N(y)$, $S$, and $\overline{N(y)} \setminus S$ are pairwise disjoint for all $y \in S$. Thus $ n= |N(y)| + |S| + |\overline{N(y)} \setminus S|\geq 2\delta(G)+1+|\overline{N(y)} \setminus S|$, where the inequality holds by \eqref{SR}. Thus $|\overline {N(y)} \setminus S| \leq n-2\delta(G)-1.$ Similarly, $N(y)$, $R$, and $\overline{N(y)} \setminus R$ are pairwise disjoint for all $y \in R$, so $|\overline{N(y)} \setminus R| \leq n - 2\delta(G) - 1$ \item Let $y\in S_i$. We have that $V_i\setminus S_i\subseteq \overline {N(y)} \setminus S$ so by (i) we have that $|S_i|\geq \frac{n}{k}-|\overline {N(y)} \setminus S|\geq 2\delta(G)+1-(1-\frac{1}{k})n$ as desired. Similarly, if $y \in R_i$, then by (i) we have that $|R_i|\geq \frac{n}{k} - |\overline{N(y)} \setminus R| \geq 2\delta(G) + 1 - (1 - \frac{1}{k})n$. Finally, we have $2\delta(G)+1-(1-\frac{1}{k})n\geq \frac{1}{2}(\frac{n}{k}-(\floor{\frac{n-1}{2}}-\delta(G)))$ by Fact \ref{ineq}.\ref{f3}.\qedhere \end{enumerate} \end{proof} \begin{cla}\label{zVi}~ For all $i\in [k]$, if $S_i\cap R_i\neq \emptyset$, then $|N(z)\cap V_i| \leq \floor{\frac{n-1}{2}}-\delta(G).$ \end{cla} \begin{proof} Let $2\leq i\leq k$ such that $S_i\cap R_i\neq \emptyset$ and let $y \in S_i\cap R_i$. So $y$ is a successor along $C$ of some vertex in $N(z)$, and a predecessor along $C$ of some vertex in $N(z)$ as well. Since $C$ is a longest cycle of $G$, neither $N(z)$ nor $N(y)$ contains two consecutive vertices of $C$, so $N(y)\cap (S\cup R) = \emptyset$. Thus, $$n-1\geq |V(C)| \geq 2 |N(z)\cup N(y)| = 2(d(y) + |N(z) \setminus N(y)|) \geq 2(d(y) +|N(z)\cap V_i|).$$ Rearranging gives the result. \end{proof} \begin{cla}\label{ellupper} $\frac{\ell+\ell'}{2}< \ceiling{\frac{k+1}{2}}$ \end{cla} \begin{proof} Let $i_1\leq i_2\leq \dots\leq i_{\ell'}$ be the indices such that $R_{i_j}\neq \emptyset$ for all $j\in [\ell']$. By Claim \ref{Si}.\ref{c2} and Claim \ref{zVi} and the fact that $z\in S_1\cap R_1$ we see that each of the sets $S_2, \dots, S_\ell$, $R_{i_2}, \dots, R_{i_{\ell'}}$ contributes at least $\frac{1}{2}\left(\frac{n}{k}-\left(\floor{\frac{n-1}{2}}-\delta(G)\right)\right)$ to $|\overline{N(z)} \setminus V_1|$. So we have $$ \delta(G)\leq d(z) \leq (1-\frac{1}{k})n-\frac{1}{2}\left(\frac{n}{k}-\left(\floor{\frac{n-1}{2}}-\delta(G)\right)\right)(\ell+\ell'-2). $$ Solving the above inequality for $\frac{\ell+\ell'}{2}$, we have $$ \frac{\ell+\ell'}{2}\leq \frac{\ceiling{\frac{n+1}{2}}}{\floor{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}}+1}\leq \frac{\frac{n+2}{2}}{\frac{n+2}{2\ceiling{\frac{k+1}{2}}}-\frac{2\ceiling{\frac{k+1}{2}}-1}{2\ceiling{\frac{k+1}{2}}}+1}=\frac{n+2}{n+3}\ceiling{\frac{k+1}{2}}<\ceiling{\frac{k+1}{2}},$$ as desired. \end{proof} Since we are supposing without loss of generality that $\ell\leq \ell'$, we have by Claim \ref{elllower} and Claim \ref{ellupper} that $$\ceiling{\frac{k}{2}}\leq \ell<\ceiling{\frac{k+1}{2}}.$$ Thus if $k$ is odd, we have a contradiction. So for the rest of the proof we will suppose that $k$ is even and consequently by Claim \ref{elllower} and \ref{ellupper}, we have $\ell=\frac{k}{2}$. \subsection{$k$ is even and $\ell=\frac{k}{2}$} Let $$A = \bigcup_{i=1}^{\ell}V_i ~\text{ and }~ B = \bigcup_{i= \ell+1}^kV_i,$$ and let $H$ be the bipartite graph induced by $[A,B]$. Label the vertices of $A$ as $u_1, \dots, u_{n/2}$ such that $d_H(u_1)\leq \dots\leq d_H(u_{n/2})$ and label the vertices of $B$ as $v_1, \dots, v_{n/2}$ such that $d_H(v_1)\leq \dots\leq d_H(v_{n/2})$. Recall that $S\subseteq A$. Since we are in the case where $k$ is even, \eqref{cf} reduces to $$\delta(G)\geq \frac{n}{2}+\floor{\frac{n+2}{k+2}}-\frac{n}{k}.$$ \begin{cla}\label{AB}~ \begin{enumerate}[label=\emph{(\roman*)}] \item\label{SB} $\delta(S, B)\geq \frac{n}{2}+2\floor{\frac{n+2}{k+2}}-\frac{2n}{k}+1$, with equality only if $S_i=V_i$ for some $i\in [\ell]$ \item\label{ASB} $\delta(A\setminus S, B)\geq \floor{\frac{n+2}{k+2}}\geq |A\setminus S|+1$, \item\label{BA} $\delta(B,A)\geq \floor{\frac{n+2}{k+2}}\geq |A\setminus S|+1$. \end{enumerate} \end{cla} \begin{proof} \begin{enumerate} \item This follows from Claim \ref{Si}.\ref{c1} since for all $y\in S$, $$d(y, B)\geq |B|-|\overline{N(y)} \setminus S|\geq |B|-(n-2\delta(G)-1)\geq \frac{n}{2}+2\floor{\frac{n+2}{k+2}}-\frac{2n}{k}+1.$$ If we have equality above, this implies that $\overline{N(y)} \setminus S\subseteq B$, which in particular implies that if $y\in V_i$, then $V_i\setminus S_i=\emptyset$. \item We have \begin{align*} \delta(A\setminus S, B)&\geq \delta(G)-(\frac{n}{2}-\frac{n}{k})\\ &\geq \floor{\frac{n+2}{k+2}} \geq \frac{n}{k}-\floor{\frac{n+2}{k+2}} \geq \frac{n}{2}-\delta(G)\stackrel{\eqref{SR}}{\geq} \frac{n}{2}-|S|+1=|A\setminus S|+1, \end{align*} where the third inequality holds by Fact \ref{ff}.\ref{ff1}. \qedhere \item Since $\delta(B,A)\geq \delta(G)-(\frac{n}{2}-\frac{n}{k})$, the rest of the calculation is the same as in \ref{ASB}. \end{enumerate} \end{proof} Before proceeding with the rest of the proof, we finally filter out $\mathcal{F}_3$. \begin{cla}\label{n/2-1} If $\delta(G)\geq \frac{n}{2}-1$, then either $G$ has a Hamiltonian cycle or $k=\frac{n}{2}$ and $G\in \mathcal{F}_3$. \end{cla} \begin{proof} By \eqref{SR} and the fact that $\ell=\frac{k}{2}$, we have $|S|=\frac{n}{2}$ and thus $A=S$ which means $A$ is an independent set. So by Claim \ref{AB}.\ref{SB}, we have $\delta(A,B)\geq \frac{n}{2}-1$. Furthermore we have by Claim \ref{AB}.\ref{BA} that $\delta(B,A)\geq \floor{\frac{n+2}{k+2}}$. If $\delta(B,A)\geq 2$, then by Theorem \ref{chv}, $G$ has a Hamiltonian cycle. So suppose $\delta(B,A)=1$ which implies $n=2k$. In this case there is a vertex $y' \in B$ such that $y'$ only has one neighbor in $A$, say $x'$. We have $\delta(A,B)\geq \frac{n}{2}-1=|B|-1$ so every vertex in $A\setminus \{x'\}$ is adjacent to every vertex in $B\setminus \{y'\}$. Since $d(y', A)=1$ and $d(y')=\frac{n}{2}-1$, it must be the case that $y'$ is adjacent to everything in $B$ except the other vertex in its own part. Now we have all the edges between $A \setminus \{x'\}$ and $B \setminus \{y'\}$, the edge $x'y'$, all the edges from $y'$ to $B$ excluding the vertex in its own part, and we have all but possibly one edge from $x'$ to $B \setminus \{y'\}$, so $G \in \mathcal{F}_3$. \end{proof} Now for the rest of the proof we may suppose that $n\geq 3k$ (i.e. $m\geq 3$). We now use Theorem \ref{chv} to show that $H$, and therefore $G$, has a Hamiltonian cycle. Suppose there exists $i\in [\frac{n}{2}]$ such that $d_H(v_i)\leq i$. By Claim \ref{AB}.\ref{BA}, we must have \begin{equation}\label{i} i\geq \delta(B, A)\geq \floor{\frac{n+2}{k+2}}. \end{equation} \noindent \tbf{Case 1} ($\frac{n}{2}-i\leq |A\setminus S|$) By Claim \ref{AB}.\ref{ASB} we have $d_H(u_{\frac{n}{2}-i})\geq \delta(A\setminus S, B) \geq |A\setminus S|+1\geq \frac{n}{2}-i+1$. \noindent \tbf{Case 2} ($\frac{n}{2}-i\geq |A\setminus S|+1$) \tbf{Case 2.1} ($k\geq 6$) Note that since $k\geq 6$, when $3k\leq n\leq 5k$, we have $\delta(G)\geq \frac{n}{2}-1$ and thus we are done by Claim \ref{n/2-1}. So for remainder of this case, suppose $n\geq 6k$ (i.e. $m\geq 6$). By Claim \ref{AB}.\ref{SB} $$d_H(u_{\frac{n}{2}-i})\geq \delta(S, B)\geq \frac{n}{2}+2\floor{\frac{n+2}{k+2}}-\frac{2n}{k}+1\geq \frac{n}{2}-\floor{\frac{n+2}{k+2}}+1\stackrel{\eqref{i}}{\geq} \frac{n}{2}-i+1,$$ where the third inequality holds by Fact \ref{ff}.\ref{ff2} since $m\geq 6$ and $k\geq 6$. Thus the conditions of Theorem \ref{chv} are satisfied and therefore $H$ has a Hamiltonian cycle. \tbf{Case 2.2} ($k=4$) By Claim \ref{AB}.\ref{SB}, we either don't have equality and thus $$d_H(u_{\frac{n}{2}-i})\geq \delta(S, B)\geq \frac{n}{2}+2\floor{\frac{n+2}{k+2}}-\frac{2n}{k}+2\geq \frac{n}{2}-\floor{\frac{n+2}{k+2}}+1\stackrel{\eqref{i}}{\geq} \frac{n}{2}-i+1,$$ where the third inequality holds by Fact \ref{ff}.\ref{ff2} since $m\geq 3$ and $k\geq 4$, and thus we are done as in the previous case; or $\delta(S, B)= \frac{n}{2}+2\floor{\frac{n+2}{k+2}}-\frac{2n}{k}+1$ and without loss of generality, $S_1=V_1$. When $3k\leq n\leq 4k$, we have $\delta(G)\geq \frac{n}{2}-1$ and thus we are done by Claim \ref{n/2-1}. So for remainder of this case, suppose $n\geq 5k$ (i.e. $m\geq 5$). Also note that since $k=4$, \eqref{cf} reduces to $$\delta(G) \geq \frac{n}{4}+\floor{\frac{n+2}{6}}.$$ \begin{cla}\label{42}~ \begin{enumerate}[label=\emph{(\roman*)}] \item \label{42i} $|S_2|\geq \floor{\frac{n+2}{6}}+1$. \item \label{42ii} $\delta(S_1, B)\geq 2\floor{\frac{n+2}{6}}+1$. \item \label{42iii} $\delta(S_2, B)\geq \delta(G)\geq \frac{n}{4}+\floor{\frac{n+2}{6}}$ \end{enumerate} \end{cla} \begin{proof} \begin{enumerate} \item Since $|S|\geq \delta(G)+1$, we have $$|S_2|=|S|-|S_1|\geq |S|-\frac{n}{4}\geq \delta(G)+1-\frac{n}{4}=\floor{\frac{n+2}{6}}+1.$$ \item Each vertex in $S_1$ has at most $|V_{2}\setminus S_{2}|$ neighbors in $V_{2}$ and thus by \ref{42i}, at least $\frac{n}{4}+\floor{\frac{n+2}{6}}-(\frac{n}{4}-|S_{2}|)\geq \floor{\frac{n+2}{6}}+\floor{\frac{n+2}{6}}+1=2\floor{\frac{n+2}{6}}+1$ neighbors in $B$. \item Since $S_1=V_1$, the vertices in $S_2$ have no neighbors in $A$ and thus all of their neighbors are in $B$. \qedhere \end{enumerate} \end{proof} We are in the case where $\frac{n}{2}-i\geq |A\setminus S|+1$, so if \begin{equation}\label{i2} \frac{n}{2}-i\leq \frac{n}{2}-\floor{\frac{n+2}{6}}-1, \end{equation} then by Claim \ref{42}.\ref{42ii} we have $$d_H(u_{\frac{n}{2}-i})\geq \delta(S_1, B)\geq 2\floor{\frac{n+2}{6}}+1\geq \frac{n}{2}-\floor{\frac{n+2}{6}}\stackrel{\eqref{i2}}{\geq} \frac{n}{2}-i+1,$$ where the third inequality holds by Fact \ref{ff}.\ref{ff2} (in particular $3\floor{\frac{n+2}{6}}+1\geq 3(\frac{n+2-4}{6})+1=\frac{n}{2}$). Otherwise together with \eqref{i}, we have $\frac{n}{2}-i=\frac{n}{2}-\floor{\frac{n+2}{6}},$ so by Claim \ref{42}.\ref{42i},\ref{42iii} we have $$d_H(u_{\frac{n}{2}-i})\geq \delta(S_2, B)\geq\frac{n}{4}+\floor{\frac{n+2}{6}}\geq \frac{n}{2}-\floor{\frac{n+2}{6}}+1= \frac{n}{2}-i+1,$$ where the third inequality holds by Fact \ref{ff}.\ref{ff1} since $m\geq 5$ and $k=4$. This completes the proof of Theorem \ref{main} and Proposition \ref{prop:n=2k}. \qed
{ "timestamp": "2020-05-28T02:07:59", "yymm": "1907", "arxiv_id": "1907.02004", "language": "en", "url": "https://arxiv.org/abs/1907.02004", "abstract": "For all integers $k$ with $k\\geq 2$, if $G$ is a balanced $k$-partite graph on $n\\geq 3$ vertices with minimum degree at least \\[ \\left\\lceil\\frac{n}{2}\\right\\rceil+\\left\\lfloor\\frac{n+2}{2\\lceil\\frac{k+1}{2}\\rceil}\\right\\rfloor-\\frac{n}{k}=\\begin{cases} \\lceil\\frac{n}{2}\\rceil+\\lfloor\\frac{n+2}{k+1}\\rfloor-\\frac{n}{k} & : k \\text{ odd }\\\\ \\frac{n}{2}+\\lfloor\\frac{n+2}{k+2}\\rfloor-\\frac{n}{k} & : k \\text{ even } \\end{cases}, \\] then $G$ has a Hamiltonian cycle unless $k=2$ and 4 divides $n$, or $k=\\frac{n}{2}$ and 4 divides $n$. In the case where $k=2$ and 4 divides $n$, or $k=\\frac{n}{2}$ and 4 divides $n$, we can characterize the graphs which do not have a Hamiltonian cycle and see that $\\left\\lceil\\frac{n}{2}\\right\\rceil+\\left\\lfloor\\frac{n+2}{2\\lceil\\frac{k+1}{2}\\rceil}\\right\\rfloor-\\frac{n}{k}+1$ suffices. This result is tight for all $k\\geq 2$ and $n\\geq 3$ divisible by $k$.", "subjects": "Combinatorics (math.CO)", "title": "On Hamiltonian cycles in balanced $k$-partite graphs", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9884918522917869, "lm_q2_score": 0.810478913248044, "lm_q1q2_score": 0.8011518021999935 }
https://arxiv.org/abs/2010.03347
Weakly reinforced Pólya urns on countable networks
We study the long-time asymptotics of a network of weakly reinforced Pólya urns. In this system, which extends the WARM introduced by R. van der Hofstad et. al. (2016) to countable networks, the nodes fire at times given by a Poisson point process. When a node fires, one of the incident edges is selected with a probability proportional to its weight raised to a power $\alpha < 1$, and then this weight is increased by $1$.We show that for $\alpha < 1/2$ on a network of bounded degrees, every edge is reinforced a positive proportion of time, and that the limiting proportion can be interpreted as an equilibrium in a countable network. Moreover, in the special case of regular graphs, this homogenization remains valid beyond the threshold $\alpha = 1/2$.
\section{Proof of \cref{mainThm} (1)}\label{sec:bounded} In this section, we establish part (1) of \cref{mainThm}. That is, we show homogenization for $\alpha < 1/2$ and graphs of bounded degree. First, in \cref{sec:existence_equi}, we prove existence of a non-vanishing equilibrium. Then, in \cref{sec:bounded_limit}, we show that $X_t$ converges to such an equilibrium, which is in fact uniquely determined. \subsection{Existence of equilibrium}\label{sec:existence_equi} Henceforth, we call an equilibrium $\mu \in \mb R_+^E$ \emph{non-vanishing} if $\mu(e) > 0$ for every $e \in E$. Before discussing existence of non-vanishing equilibria on general graphs, we present the $\Delta$-regular case as a particularly easy example. \begin{example}[$\Delta$-regular graphs]\label{lemma:equi_zd} Let $G$ be a $\Delta$-regular graph. Then, $\mu \equiv 2/\Delta$ defines a non-vanishing equilibrium. Indeed, checking \cref{eqn:equi} leads to $$ \sum_{v\in e} \frac{\mu^\alpha}{\sum_{e'\in E_v}\mu^\alpha}\\ = \sum_{v\in e} \frac1{\sum_{e'\in E_v}1}\\ = \frac2\Delta = \mu. $$ \end{example} The general case of bounded-degree graphs does not admit such an easy analysis as the equilibrium could in principle assume an infinite number of different values. Nevertheless, we deduce existence from a compactness argument. \begin{proposition}[Existence of non-vanishing equilibria]\label{lemma:existence_for_all} Every graph with degrees uniformly bounded by $\Delta$ exhibits at least one non-vanishing equilibrium $\mu$ {{with $\mu(e) \ge 2/\Delta^{1/(1-\alpha)}$ for all $e\in E$}}. \end{proposition} \begin{proof} % % By \cref{eqn:equi}, the equilibria correspond to fixed points of the continuous operator \begin{align} T\colon \mb R_+^E&\longrightarrow \mb R_+^E\\ \mu(\cdot)&\longmapsto \sum_{v\in \cdot} \frac{{\mu(\cdot)}^\alpha} {\sum_{e'\in E_v}{\mu(e')}^\alpha}. \end{align} % % As a product of locally convex and Hausdorff spaces, $\mb R_+^E$ is again locally convex and Hausdorff. Now, define the closed set {$C = {[2\Delta^{-1/(1 - \alpha)}, 2]}^E$}, which as a product of convex and compact sets, is again convex and compact. We claim that $T(C) \subseteq C.$ Once this claim is established, Schauder's fixed point theorem yields a non-vanishing equilibrium in $C$. To prove $T(C) \subseteq C$, note that for every $\mu\in C$ and $e = \{v_1, v_2\} \in E$ we have \begin{align} T(\mu)(e) = \frac{\mu(e)^\alpha} {\sum_{e'\in E_{v_1}}{\mu(e')}^\alpha} + \frac{\mu(e)^\alpha} {\sum_{e'\in E_{v_2}}{\mu(e')}^\alpha} \le 2\;, \end{align} and \begin{align} T(\mu)(e) = \sum_{i \le 2} \frac{\mu(e)^\alpha} {\sum_{e'\in E_{v_i}}{\mu(e')}^\alpha} \ge \sum_{i \le 2} \frac{\mu(e)^\alpha} {2^\alpha\ms{deg}(v_i)} \ge \frac{2\mu(e)^\alpha}{2^\alpha {\Delta}} \ge \frac2{{{\Delta}}^{1/(1 - \alpha)}}, \end{align} as claimed. \end{proof} \subsection{Proof of Theorem \ref{mainThm} (1)}\label{sec:bounded_limit} The proof of part (1) of Theorem \ref{mainThm} is based on three pivotal ingredients. First, as each edge is incident to two vertices firing at rate 1, its weight grows at rate at most 2. In other words, for every $e \in E$, almost surely, $$X_+(e) : = \limsup_{t \to \infty}X_t(e) \le 2.$$ \begin{lemma}[Rate upper bound]\label{lemma:loc_firstboundonC} Let $e \in E$ be arbitrary. Then, $\mb P(X_+(e) \le 2) = 1$. \end{lemma} Second, we derive a positive lower bound for the growth rate \begin{align} X_-(e) :=\liminf_{t\rightarrow\infty}X_t(e). \end{align} This task is slightly more subtle than the upper bound, since arbitrarily small rates can occur with positive probability. To approach this challenge, we fix throughout this section a non-vanishing equilibrium $\mu$ as in \cref{lemma:existence_for_all} and write $$E_e := \{e' \in E:\, e' \cap e \ne \emptyset\}$$ for the family of edges adjacent to a given edge $e \in E$. Then, an edge $e \in E$ is \emph{$\delta$-stable} if \begin{align} \min_{e' \in E_e} \frac{X_-(e')}{\mu(e')} \ge \delta. \end{align} \begin{lemma}[Rate lower bound] \label{noPercLem} There exists $\delta_0 > 0$ such that with probability $1$, all connected components of $\delta_0$-unstable edges are finite. \end{lemma} Finally, we rely on a bootstrapping procedure to push the weights iteratively closer to 1. To that end, define \begin{align} \mc C(e) : = [X_-(e), X_+(e)] \end{align} as the smallest interval containing the accumulation points of $X_t(e)$. \begin{lemma}[Bootstrap on bounded degree graphs] \label{bootStrapLem} Let $\alpha < 1/2$ and $\varrho>1$. If $e\in E$ is such that $\mc C(e') \subseteq [\varrho^{-1}\mu(e'), \varrho\mu(e')]$ holds for all $e' \in E_e$, then $\mc C(e) \subseteq [\varrho^{-2\alpha}\mu(e), \varrho^{2\alpha}\mu(e)]$. \end{lemma} Before establishing Lemmas~\ref{lemma:loc_firstboundonC}--\ref{bootStrapLem}, we elucidate how they can be combined to yield the proof of part (1) of \cref{mainThm}. \begin{proof}[Proof of \cref{mainThm}(1)] Let $\delta_0 > 0$ be as in \cref{noPercLem}. { Assume without loss of generality that $\delta_0 \le \Delta^{-1/(1-\alpha)}$ so that $2\le \mu(e)\Delta^{1/(1-\alpha)}\le \mu(e)/\delta_0$ for any $e\in E$ where the first inequality follows from $\mu(e)\ge 2/\Delta^{1/(1-\alpha)}$ given by \cref{lemma:existence_for_all}}. We first claim that $X_-(e) \ge \delta_0\mu(e)$ almost surely for every $e \in E$. { Indeed, assume the contrary and let $S$ be a connected component of $\delta_0$-unstable edges. By \cref{noPercLem}, $S$ is finite and we choose $e^*\in S$ (randomly) such that $\varrho:=\frac{\mu(e^*)}{X_-(e^*)}$ is maximal.} Then, $\varrho^{-1} \ge \delta_0$ almost surely since otherwise \cref{lemma:loc_firstboundonC} and the bootstrap property of \cref{bootStrapLem}{, combined with $2\le \mu(e)/\delta_0\le \varrho\mu(e)$,} would give that $\mc C(e^*) \subseteq [\varrho^{-2\alpha}\mu(e^*), 2]$, thereby contradicting that \[\frac{X_-(e^*) }{ \mu(e^*)} = \varrho^{-1} < \varrho^{-2\alpha}.\] {{Hence, the infimum \[\varrho_* := \inf\{\varrho > 1\colon\mc C(e) \subseteq [\varrho^{-1}\mu(e), \varrho\mu(e)] \text{ for every $e \in E$}\}\] is at most $\delta_0^{-1}$. }}Now, $\varrho_* = 1$ almost surely since otherwise the bootstrap property yields that $\mc C(e) \subseteq [\varrho_*^{-2\alpha}\mu(e), \varrho_*^{2\alpha}\mu(e)]$ for every $e \in E$, thereby contradicting the choice of $\varrho_*$. \end{proof} It remains to establish the auxiliary results. We start by proving Lemma \ref{lemma:loc_firstboundonC}. \begin{proof}[Proof of Lemma \ref{lemma:loc_firstboundonC}] One extremal case is that each clock-ring event for $P_v$ increments the value for an edge $e = \{v, w\}\in E$. The process \begin{align} Y_t := \frac{P^{}_v([0, t]\times[0, 1]) + P_w([0, t]\times[0, 1])}t \end{align} counts the normalized occurrences for this upper bound and has an expected value of \begin{align} \mb E[Y_t] = \frac{\mb E[P^{}_v([0, t]\times [0, 1])] + \mb E[P^{}_w([0, t]\times [0, 1])]}t = \frac{2t}t = 2\;. \end{align} Then, the strong law of large numbers for homogeneous Poisson point processes, gives that almost surely $ \lim_{t \rightarrow \infty}Y_t = 2$. Now, $ X_t(e)\le Y_t + \frac1t $ for all $t \ge 0$, implies that almost surely, $ \limsup_{t \rightarrow \infty} X_t(e)\le \lim_{t \rightarrow \infty} Y_t = 2. $ \end{proof} Next, we verify the bootstrap property. \begin{proof}[Proof of Lemma \ref{bootStrapLem}] Let $\varrho > 1$ and $e = \{v_1, v_2\} \in E$ be such that $\mc C(e') \subseteq [\varrho^{-1}\mu(e'), \varrho\mu(e')]$ for all $e' \in E_e$. Moreover, for $\varepsilon > 0$ set $\varrho_\varepsilon = (1 + \varepsilon) \varrho$. Then, there exists a random time $T < \infty $ such that $X_t(e') \in [\varrho_\varepsilon^{-1}\mu(e'), \varrho_\varepsilon\mu(e')]$ for all $t \ge T$ and $e' \in E_e$. In particular, for every $i \in \{1, 2\}$, { \begin{align} \ms{pol}_{v_i,e}(X_t) = \frac{{N_t(e)}^\alpha} {\sum\limits_{\mathclap{e'\in E_{v_i}}}{N_t(e')}^\alpha} \ge \frac{{(\varrho_\varepsilon^{-1}\mu(e))}^\alpha} {\sum\limits_{\mathclap{e'\in E_{v_i}}}{{(\varrho_\varepsilon\mu(e'))}^\alpha}} = \varrho_\varepsilon^{-2\alpha} \frac{{\mu(e)}^\alpha} {\sum\limits_{\mathclap{e'\in E_{v_i}}}{{\mu(e')}^\alpha}}. \end{align}} Therefore, using that $\mu$ is an equilibrium, $\{N_t(e)\}_{t \ge T}$ is dominated from below by a Poisson process with intensity $$\varrho_\varepsilon^{-2\alpha}\frac{\mu(e)^\alpha} { \sum\limits_{{\mathclap{\substack{e'\in E_{v_1}}}}} { {\mu(e')^\alpha}{}}} + \varrho_\varepsilon^{-2\alpha}\frac{\mu(e)^\alpha} { \sum\limits_{{\mathclap{\substack{e'\in E_{v_2}}}}} { {\mu(e')^\alpha}{}}} = \varrho_\varepsilon^{-2\alpha} \mu(e). $$ Hence, almost surely, $X_-(e) \ge \varrho_\varepsilon^{-2\alpha} \mu(e)$. Similar arguments yield that $X_+(e) \le \varrho_\varepsilon^{2\alpha} \mu(e)$. \end{proof} The proof of Lemma \ref{noPercLem} is the most challenging part of the auxiliary results. The main ingredient is \cref{lemma:loc_secondboundonC} below. It states that $X_-(e)$ is bounded away from 0 with a high probability, even when conditioning on $\mc F_{e^c}$, where for $e \in E$ we let \[\mc F_{e^c} := \sigma(\{P_v\}_{v \not \in \cup_{e' \in E_e}e'})\] denote the $\sigma$-algebra generated by all Poisson processes at nodes that are at distance at least~2 away from the edge $e$. \begin{lemma}[Compact containment of $\mc C(e)$] \label{lemma:loc_secondboundonC} For every $\varepsilon > 0$ there exists $\delta > 0$ such that almost surely $$\inf_{e \in E}\mb P( X_-(e) \ge \delta\,|\, \mc F_{e^c}) \ge 1 - \varepsilon.$$ In particular, $\mb P( X_-(e) > 0\,|\, \mc F_{e^c}) = 1$ for every $e \in E$. \end{lemma} Hence, we conclude from dependent percolation theory in the form of \cite[Theorem 0.0]{domProd} that the edges violating the lower bound are restricted to finite, well-separated islands. To make the presentation self-contained, we give an elementary direct proof. \begin{proof}[Proof of Lemma \ref{noPercLem}] Let $v \in V$ be arbitrary. We resort to a first-moment argument and show that for sufficiently small $\delta$ the expected number of length-$n$ self-avoiding paths of $\delta$-unstable edges tends to 0 as $n \to \infty$. Since the number of length-$n$ self-avoiding paths in $G$ starting from $v$ is at most $\Delta^n$, it suffices to show that the probability for any fixed self-avoiding path to consist of $\delta$-unstable edges only is of the order at most $(2\Delta)^{-n}$. Note that the number of edges that are at graph distance at most 3 of a given edge is bounded above by $2\Delta^3$. Hence, any self-avoiding path of length $n$ contains at least $n / (2\Delta^3)$ edges that are of pairwise distance at least 4, so that by Lemma \ref{lemma:loc_secondboundonC}, the probability that they are all $\delta$-unstable is at most $\varepsilon(\delta)^{n / (2\Delta^3)}$. In particular, choosing $\delta > 0$ such that $\varepsilon(\delta)^{1 / (2\Delta^3)} \le 1 / (2\Delta)$ concludes the proof. \end{proof} It remains to prove Lemma \ref{lemma:loc_secondboundonC}. To that end, we invoke a conditioning argument in the spirit of \cite[Lemma 5.2]{benaim} to provide a bootstrap result propagating large edge weights at a current time point to a considerable duration into the future. For $k, \ell \ge 1$ put \begin{align} a_{k, \ell} := \Delta^{-\frac1{1 - \alpha}} 2^{k(\ell - 1) - k \sum_{2 \le i \le \ell}\alpha^i}. \end{align} \begin{lemma}[Bootstrapped lower bound] \label{lemma:loc_HilfsLemma} There exists a constant $c = c(\alpha, \Delta) > 0$ such that for all $k$ large enough, all $\ell\ge 1$ and all $e \in E$, almost surely, \begin{align}\label{eqn:condProb} &\mb P(N_{2^{k(\ell + 1)}}(e) \le a_{k, \ell + 1} \text{ and }N_{2^{k\ell}}(e) \ge a_{k, \ell} \,|\, \mc F_{e^c} ) \le e^{ -c2^{k\ell(1 - \alpha)} }. \end{align} \end{lemma} \begin{proof} Let $\varepsilon > 0$ be arbitrary. Then, the Poisson concentration inequality \cite[Lemma 1.2]{penrose} implies that the event \begin{align} A_e:=\{N_{2^{k(\ell + 1)}}({e'}) \le 2^{1 + \varepsilon + k(\ell + 1)}\ \text{for all}\ e'\in E_{e}\}\label{eqn:upper_bound_weights} \end{align} has a probability tending to $1$ with an error decaying exponentially in $2^{k\ell}$. Further, under the event $\{N_{2^{k\ell}}(e)\ge a_{k, \ell}\} \cap A_e$, \cref{eqn:polya_weight} has a lower bound for times $t \in T_\ell : = [2^{k\ell}, 2^{k(\ell + 1)}]$ given by \begin{align} {\mathsf{pol}_{v, e}(N_t)} = \frac {N_t(e)^\alpha}{\sum_{e'\in E_v}N_t({e'})^\alpha} \ge \frac {a_{k, \ell}^\alpha}{a_{k, \ell}^\alpha + \Delta 2^{(1 + \varepsilon + k(\ell + 1))\alpha}}. \end{align} Since $\Delta a_{k, \ell + 1} = a_{k, \ell}^\alpha 2^{k (\ell + \alpha) (1 - \alpha)}$, putting $\alpha_\varepsilon = \alpha (1 + \varepsilon)$, we deduce that \begin{align} {\mathsf{pol}_{v, e}(N_t)}&\ge \frac1{1 + 2^{\alpha_\varepsilon} 2^{k (\ell + 1) \alpha} a_{k, \ell + 1}^{-1} 2^{k (\ell + \alpha)(1 - \alpha)} } = \frac1{1 + 2^{\alpha_\varepsilon + k(\ell + 1 - (1 - \alpha)^2)} a_{k, \ell + 1}^{-1} }. \end{align} Hence, \begin{align} {\mathsf{pol}_{v, e}(N_t)} \ge \frac {2^{-\alpha_\varepsilon + k((1 - \alpha)^2 - \ell - 1)} a_{k, \ell + 1}} {1 + 2^{-k\alpha}} = :{G_{k, \ell}}\;. \end{align} In particular, for $t\in T_\ell$, the Poisson processes $P_v$ and $P_w$ having points in $[2^{k\ell}, t]\times U_e$ where $U_e\subseteq[0, 1]$ and $|U_e|\ge G_{k, \ell}$ implies weight increases of the edge $e$ (by one or more) in the time interval $[2^{k\ell}, t]$. Therefore, using \cref{eqn:upper_bound_weights}, we find a constant $c$ such that \begin{align} &\mb P\left(N_{2^{k(\ell + 1)}}(e) \le a_{k, \ell + 1}\text{ and } N_{2^{k\ell}}(e)\ge a_{k, \ell}\right)\\ &\quad\le\mb P\left(N_{2^{k(\ell + 1)}}(e)-N_{2^{k\ell}}(e)\le a_{k, \ell + 1}\text{ and } N_{2^{k\ell}}(e)\ge a_{k, \ell}\right)\\ &\quad\le \mb P\left(P_v(T_\ell\times U_e) + P_w(T_\ell\times U_e)\le a_{k, \ell + 1} \text{ and } N_{2^{k\ell}}(e)\ge a_{k, \ell}\right) + {O}\big(e^{-c 2^{k\ell}}\big)\\ &\quad \le \mb P \left( \mathsf{Poi}\left( 2|T_\ell| G_{k, \ell}\right) \le a_{k, \ell + 1}\label{eqn:poisson_param} \right) + {O} \big( e^{-c 2^{k\ell}} \big). \end{align} Since \begin{align} 2|T_\ell|G_{k, \ell} = a_{k, \ell + 1}2^{1 - \alpha_\varepsilon + k(1 - \alpha)^2} \frac{1 - 2^{-k}}{1 + 2^{-k\alpha}} \end{align} there exists $\tilde c > 1$ such that $ 2|T_\ell|G_{k, \ell} > \tilde c\cdot a_{k, \ell + 1} $ holds for all $k$ large enough. Thus, by \cite[Lemma 1.2]{penrose}, \begin{align} \mb P\left(N_{2^{k(\ell + 1)}}(e)\le a_{k, \ell + 1}\text{ and } N_{2^{k\ell}}(e)\ge a_{k, \ell}\right)&\le \exp\left((1 - \tilde c + \log(\tilde c))a_{k, \ell + 1}\right)\;. \end{align} Now, we conclude the proof by noting that $a_{k, \ell + 1} \ge \Delta^{-\frac1{1 - \alpha}} 2^{k\ell(1 - \alpha)}$ and that the coefficient of $a_{k, \ell + 1}$ is negative. \end{proof} \cref{lemma:loc_HilfsLemma} propagates the lower bounds on the edge weights through time. As $t \rightarrow \infty$, these lower bounds allow to exclude vanishing edge weights. \begin{proof}[Proof of Lemma \ref{lemma:loc_secondboundonC}] Defining $F = \cap_{\ell \ge 1}\{N_{2^{k\ell}}(e)\ge a_{k, \ell}\}$ and $\delta_0 = \liminf_{\ell \to \infty}a_{k, \ell}2^{-k\ell} > 0,$ we note that $X_-(e) \ge \delta_0$ holds under the event $F$. Hence, it remains to establish a lower bound on $\mb P(F|\mc F_{e^c})$. Since $a_{k, 1} \le 1$, the bound $N_{2^k}(e) \ge a_{k, 1}$ holds for any $k \ge 1$. Hence, if the event $F$ does not occur, then there exists $\ell_0 \ge 1$ such that $N_{2^{k(\ell_0 + 1)}}(e) \le a_{k, \ell_0 + 1}$ and $N_{2^{k\ell_0}}\ge a_{k, \ell_0}$. In particular, by Lemma \ref{lemma:loc_HilfsLemma}, almost surely, $$1 - \mb P(F\,|\, \mc F_{e^c}) \le \sum_{\ell \ge 1}\mb P(N_{2^{k(\ell + 1)}}(e) \le a_{k,\ell + 1} \text{ and }N_{2^{k\ell}}\ge a_{k, \ell}\,|\,\mc F_{e^c}) \le \sum_{\ell \ge 1}e^{ -c2^{k\ell (1 - \alpha)} },$$ which becomes smaller than $\varepsilon$ for sufficiently large $k > 0$. \end{proof} \section{Introduction}\label{sec:intro} P\'olya's urn process is the paradigm model for a random process incorporating reinforcement effects. However, when thinking in the direction of applications, a single urn often does not represent complex interactions accurately. For instance, in the field of social sciences, the formation of friendship networks could be related to reinforcement effects in social interactions \cite{rolles}. In economy, in a network of companies competing on a variety of products, the global reputation could influence the market shares of the products differently \cite{benaim, lucas}. Finally, in the domain of neuroscience, it is plausible that synapses that were successful in the past should be selected again in the future and reinforced with a higher probability \cite{warm1}. In the present paper, we focus on the network formation model proposed in \cite{warm1} and study its long-time behavior in the regime of weak reinforcement. In that model, starting from a base network, nodes are picked sequentially at random. Once a node is selected, we choose one of the incident edges with a probability proportional to its weight to some power $\alpha > 0$ and increase that weight by 1. The analysis of \cite{warm1} concerns the asymptotic proportion of times that edges are reinforced in the regime of strong reinforcement, where $\alpha > 1$. In this regime, the limiting proportion is random and coincides with some stable equilibrium in an associated dynamical system, which is typically concentrated on a small subset of the edges if $\alpha$ is large. In contrast, we consider the regime of weak reinforcement, where $\alpha < 1$ and find that in many settings the reinforcement proportions converge to a uniquely determined limit equilibrium, which is supported on all edges of the base network. Figure \ref{frontFig} illustrates the network at an early and at a late time point. \begin{figure}[!htpb] \input{early}\hspace{0.8cm} \input{late} \caption{Weakly reinforced P\'olya model after time $2$ (left) and time $100$ (right). Dashed edges have not been reinforced until the considered time. The widths of solid edges are proportional to time-normalized weights.} \label{frontFig} \end{figure} Hence, together with the analysis in \cite{warm1, warm2}, our main result is a first step towards a network-based analog of Rubin's dichotomy for classical P\'olya urns:\, while for strong reinforcement some edges are only reinforced finitely often, in the weakly reinforced regime all edges are reinforced a positive proportion of time. Although this description outlines the broader picture, more research is needed to carve out the precise conditions for this dichotomy. Indeed, \cite{rubin} describes an example of a meticulously designed network together with firing rates where even for $\alpha > 1$ there is percolation of edges that are reinforced a positive proportion of time. Similarly to the setting of \cite{main_paper}, one major difficulty in the analysis stems from the choice of the base graph. In contrast to \cite{benaim, warm1}, we do not restrict our attention to finite base graphs, but work on a countable network with uniformly bounded degrees. In particular, the highly developed machinery of stochastic approximation algorithms invoked in \cite{benaim, warm1} is not available as the state space of the associated dynamical system would be infinite-dimensional. In \cite{main_paper}, the very strong reinforcement leads to an effective decomposition of the countable network into finite islands separated by edges that are never reinforced. This trick recovers the finite-dimensional setting. However, the strategy cannot work in the regime of weak reinforcement, where we expect all edges to be reinforced a positive proportion of time. Hence, in the present paper, we follow an entirely different plan to prove our main result. Namely, convergence of the normalized edge weights to an equilibrium for (1) all graphs of uniformly bounded degree if $\alpha < 1/2$, (2) regular graphs if $\alpha < 1/2+\varepsilon$ for some $\varepsilon>0$, and (3) $\mb Z$ if $\alpha < 1$. This plan consists of three critical steps. First, we invoke a compactness argument to obtain an equilibrium on the entire countable network. Second, we establish that all edges are reinforced a positive proportion of time. This step rests on a percolation argument, where we decompose the network again into finite islands separated by edges that are now reinforced a positive proportion of time. Finally, to obtain convergence to the equilibrium, we rely on a homogenizing bootstrap argument. It formalizes the intuition that for weak reinforcement, deviations from the equilibrium decrease over time. The rest of the manuscript is organized as follows. \cref{sec:results} introduces the model and states the main contribution of the paper, a homogenization result in the sub-linear regime for three different combinations of graphs and exponents. In \cref{sec:bounded}, we consider graphs of bounded degree and $\alpha < 1/2$. Finally, \cref{seg:regular} deals with regular graphs of degree $\Delta$ and covers $\alpha < 1/2+\varepsilon$ for $\Delta > 2$ and $\alpha < 1$ for the graph $\mb Z$. \subsection*{Acknowledgement.}The authors thank both anonymous referees for the careful reading of the manuscript and the constructive feedback. Their comments and suggestions substantially helped to improve the presentation of the material. The authors thank M.~Holmes and V.~Kleptsyn for illuminating discussions and ideas for future work. \section{Regular graphs}\label{seg:regular} Throughout this section, we assume $G$ to be $\Delta$-regular. The key to the proof of parts (2) and (3) of \cref{mainThm} is the following bootstrap property. \begin{lemma}[Bootstrap on regular graphs]\label{lemma:multi_tightenBounds} Let $a < 2/\Delta$, $b > 2/\Delta$ and $e \in E$ be such that $\mc C(e) \subseteq [a,b]$ holds almost surely for all $e' \in E_e$. Furthermore, define the function $$f(r, s) :=\frac{2r^\alpha}{r^\alpha + (\Delta - 1)s^\alpha}.$$ Then, $\mc C(e)\subseteq[f(a, b), f(b, a)]$ holds almost surely. In particular, for $G = \mb Z$ there exist $a'\in [a, 1)$ and $b' \in (1,b]$ such that $a'/b' \ge {(a / b)}^\alpha$ and $\mc C(e)\subseteq[a', b']$ holds almost surely. \end{lemma} Before proving Lemma \ref{lemma:multi_tightenBounds} we explain how it implies part (3) of \cref{mainThm}. \begin{proof}[Proof of part (3) of \cref{mainThm}] % % First, $X_-(e)$ is strictly positive by \cref{noPercLem} using a similar argument to the proof of part (1) of \cref{mainThm}. Hence, \cref{lemma:multi_tightenBounds} gives that $X_-(e) \ge X_-(e)^\alpha 2^{-\alpha}$, i.e., $X_-(e) \ge 2^{-\alpha / ( 1 - \alpha)}$. % % In other words, almost surely $\mc C(e) \subseteq [a_1, b_1]$ where $a_1 := 2^{-\alpha / ( 1 - \alpha)}$ and $b_1 := 2$. Applying Lemma \ref{lemma:multi_tightenBounds} iteratively yields sequences ${\{a_i\}}_{i\ge 1}$ and ${\{b_i\}}_{i\ge 1}$ such that \begin{enumerate} \item ${\{a_i\}}_{i\ge 1}$ is increasing and bounded above by $1$, \item ${\{b_i\}}_{i\ge 1}$ is decreasing and bounded below by $1$, \item $b_{i + 1}/a_{i + 1}\le (b_i/a_i)^\alpha < b_i/a_i$, and \item $\mc C(e)\subseteq \bigcap_{i \ge 1}[a_i, b_i]$ holds almost surely for all $ e\in E$. \end{enumerate} Since the first three items imply that $a_i$ and $b_i$ converge to 1, we arrive at \begin{align} \mb P(\mc C(e) = 1) = \mb P(\cap_{i\ge 0}\{ \mc C(e)\subseteq[a_i,b_i]\}) = \lim_{i \to \infty}\mb P( \mc C(e)\subseteq[a_i,b_i]) = 1,\label{eqn:regular_almost_sure_limit} \end{align} as asserted. \end{proof} Next, we prove Lemma~\ref{lemma:multi_tightenBounds}. \begin{proof}[Proof of Lemma \ref{lemma:multi_tightenBounds}] Let \begin{align} G_{a,b} := \cap_{e'\in E_e}\{\mc C(e')\subseteq[a,b]\} \end{align} denote the event that $\mc C(e')\subseteq [a, b]$ holds for all $e'\in E_e$. Under this event, $X_t^{\{v,w\}}$ gains mass at a rate of at least \begin{align} a'' = f(at, bt) = f(a, b), \end{align} for $t$ large enough. More precisely, under $G_{a,b}$ one can find a sequence ${\{\varepsilon_t\}}_{t\ge 0}$ with $\varepsilon_t\searrow 0$ such that almost surely \begin{align} N_t^{\{v,w\}}\ge P_v([0,t]\times U_{v,t}) + P_w([0,t]\times U_{w,t}) + 1 \end{align} for all $t>0$ where $|U_{v,t}| = |U_{w,t}| = a''/2- \varepsilon_t$. Analogous arguments to the one in the proof of \cref{lemma:loc_firstboundonC} give that almost surely \begin{align} \liminf_{t\to\infty} X_t^{\{v,w\}} \ge a', \end{align} where {$a':=a\vee a''$}. Similar arguments give the upper bound $b' := b\wedge b''$ where $b'' := f(b,a)$. { In the special case $\Delta=2$ (i.e.\ $G=\mb Z$) we find $a'/b'\ge {(a/b)}^{\alpha}$ as desired.} \end{proof} Now, let $G$ be $\Delta$-regular for some $\Delta \ge 2$. In this case we do not have $a'/b'\ge {(a/b)}^{\alpha}$ to prove \cref{mainThm}, but we still have the sequences ${\{a_i\}}_{i\ge 1}$ and ${\{b_i\}}_{i\ge 1}$ with $b_{i+1}/a_{i+1}\le b_i/a_i$. The idea is to show that for $\alpha$ sufficiently close to $1/2$ the inequality is strict. \begin{lemma}\label{lemma:improve_at_least_one} For $\alpha=1/2+\varepsilon$ with $\varepsilon$ small enough and $0<a<2/\Delta<b<2$ we have $f(a,b)>a$ or $f(b,a)<b$. \end{lemma} \begin{proof} Assume both statements are wrong, i.e. $f(a,b)\le a$ and $f(b,a)\ge b$, then this gives \begin{align} \rho \le\rho^{2\alpha} \frac{1+\frac{\rho^{-\alpha}}{\Delta-1}}{1+\frac{\rho^{\alpha}}{\Delta-1}} \label{eqn:regular_to_contradict} \end{align} for $\rho:=b/a$. Insert $\alpha=1/2+\varepsilon$ and take log on both sides to get \begin{align} 0\le 2\varepsilon\log\rho + \log\Big(1+\frac{\rho^{-1/2-\varepsilon}}{\Delta-1}\Big) - \log\Big(1+\frac{\rho^{1/2+\varepsilon}}{\Delta-1}\Big)\;. \end{align} Note that equality holds for $\rho=1$ so if the right-hand side is decreasing in $\rho$ then this contradicts the above inequality for $\varrho>1$. To that end, note that the derivative \begin{align} \frac1\rho\Big(2\varepsilon -(1/2+\varepsilon) \Big(\frac{1}{1+(\Delta-1)\rho^{-1/2-\varepsilon}} + \frac1{1+(\Delta-1)\rho^{1/2+\varepsilon}} \Big)\Big) \end{align} of the right-hand side is bounded above for all $\rho\ge 1$ by \begin{align} \frac1\rho\Big(2\varepsilon -\frac{(1/2+\varepsilon)}{\Delta}\Big) \end{align} which is negative for $\varepsilon$ small enough and hence we get a contradiction to \ref{eqn:regular_to_contradict}. \end{proof} Not improving both bounds at each application of the bootstrap property means that the proof for a uniform lower bound on $\mathcal C(e)$ in part (1) of \cref{mainThm} does not work since we do not immediately get a contradiction. The following Lemma helps remedy this. \begin{lemma}\label{lemma:lower_threshold_for_improvement} For $a<2\Delta^{-1/(1-\alpha)}$ we have $f(a,2)>a$. \end{lemma} \begin{proof} Straightforward calculation as \begin{align} f(a,2)=\frac{2a^{\alpha}}{a^{\alpha}+(\Delta-1)2^{\alpha}} > \frac{2a^{\alpha}}{2^{\alpha}\Delta} =a \frac{2^{1-\alpha}}{a^{1-\alpha}\Delta} >a.&\qedhere \end{align} \end{proof} \begin{proof}[Proof of part (2) of \cref{mainThm}] Take the $\varepsilon$ from \cref{lemma:improve_at_least_one} and consider $\alpha=1/2+\varepsilon$. $X_-(e)$ is bounded above by \cref{lemma:loc_firstboundonC} and strictly positive by \cref{noPercLem} and \cref{lemma:lower_threshold_for_improvement} using analogous arguments as in the proof of part (1) of this theorem. So we find $a \le 2/\Delta$ and $b \ge 2/\Delta$ such that $\mathcal{C}(e)\subseteq[a,b]$ for all $e\in E$. We choose $a$ maximal and $b$ minimal with that property. To derive a contradiction, assume that $a < b$. Then, by \cref{lemma:multi_tightenBounds} and \cref{lemma:improve_at_least_one} we can tighten the bounds as \begin{align} \mathcal{C}(e)\subseteq [a\vee f(a,b), b\wedge f(b,a)] \subsetneq [a, b]\;\label{eqn:new_bounds_on_C}. \end{align} This contradicts the maximality/minimality of $a$ and $b$, and since $f(a,b)\le 2/\Delta \le f(b,a)$ therefore gives $a = b = 2/\Delta$. \end{proof} \section{Model definition and main result}\label{sec:results} Let $G = (V, E)$ be a countable graph with uniformly bounded degrees. That is, the vertex set $V$ is countable and $$\Delta := \sup_{v \in V} \ms{deg}(v) < \infty.$$ If $\Delta=\ms{deg}(v)$ for every $v\in V$ then the graph is regular with degree $\Delta$, which we write as $\Delta$-regular. We investigate a system of random variables \begin{align} N_t := {\{N_t(e)\}}_{e\in E} \end{align} of interacting P\'{o}lya-type urns on the edge set $E$ at continuous time $t \ge 0$ on a probability space $(\Omega, \mb P)$. The dynamics of $N_t$ are a continuous-time analog of the process considered in~\cite{warm1}. Loosely speaking, every vertex has a Poisson clock and whenever that clock rings at a node $v \in V$, the dynamics choose and increment the weights on one of the incident edges $E_v = \{e\in E:\,v\in e\}$ by 1. The choice of the edge happens using the power-weighted P\'{o}lya-scheme \begin{align}\label{eqn:polya_weight} \ms{pol}_{v,e}(N_t): = \frac{{N_t(e)}^\alpha} {\sum_{e'\in E_v}N_t(e')^\alpha}\;, \qquad \alpha\ge 0\;. \end{align} Henceforth, we tacitly assume that $\alpha < 1$. More precisely, the weights $N_t(e)$ are initialized at $1$ and the dynamics of $N_t$ are governed by an iid family of Poisson processes ${\{P_v\}}_{v\in V}$ on $[0, \infty)\times[0,1]$ with intensity $1$ counting clock-ring events at vertex $v\in V$ in a time window $[0,\infty)$ with marks in the range $[0,1]$. If the process $P_v$ contains an atom of the form $(t, u) \in [0, \infty) \times [0, 1]$, then increment the mass of an edge $e_i \in E_v = \{e_1, \dots e_{\ms{deg}(v)}\}$ by 1 if $u\in U_{v,e_i}$, where ${\{U_{v, e_i}\}}_{i \le \ms{deg}(v)}$ is a partition of $[0,1]$ given by \begin{align} U_{v, e_i} & = \Big(\sum_{j \le i - 1}\ms{pol}_{v, e_{j}}(N_t),\sum_{j \le i } \ms{pol}_{v, e_{j}}(N_t) \Big]\;. \end{align} For the existence of the process $\{N_t\}_{t \ge 0}$ on bounded-degree graphs, see \cite{main_paper}. Finally, a non-negative vector $\mu\in\mb R_+^E$ defines an \emph{equilibrium on $G$} if \begin{align}\label{eqn:equi} \mu(e) := \sum_{v\in e}\frac{{\mu(e)}^\alpha} {\sum_{e'\in E_v}{\mu(e')}^\alpha} \end{align} holds for all $e\in E$ with $\mu(e) > 0$. This is a straightforward extension of the notion of finite equilibria from~\cite{warm1} to countable networks. Now, let \begin{align} X_t:=\frac{N_t}t \end{align} denote the time-normalized system of weights at time $t > 0$. We say that $X_t$ exhibits \emph{homogenization} if the equilibrium measure $\mu$ exists, is unique and $X_t \to \mu$ almost surely. \begin{conjecture}[Homogenization] Let $G$ be a countable graph with uniformly bounded degrees. Then, $X_t$ exhibits homogenization. \end{conjecture} In the present work, we verify this conjecture for three combinations of $G$ and $\alpha$. \begin{theorem}[Homogenization]\label{mainThm} $X_t$ exhibits homogenization in the following cases. \begin{enumerate} \item The degree of $G$ is uniformly bounded and $\alpha < 1/2$. \item $G$ is $\Delta$-regular and $\alpha < 1/2+\varepsilon$ for some $\varepsilon>0$. \item $G = \mb Z$ and $\alpha < 1$. \end{enumerate} \end{theorem}
{ "timestamp": "2021-06-28T02:19:29", "yymm": "2010", "arxiv_id": "2010.03347", "language": "en", "url": "https://arxiv.org/abs/2010.03347", "abstract": "We study the long-time asymptotics of a network of weakly reinforced Pólya urns. In this system, which extends the WARM introduced by R. van der Hofstad et. al. (2016) to countable networks, the nodes fire at times given by a Poisson point process. When a node fires, one of the incident edges is selected with a probability proportional to its weight raised to a power $\\alpha < 1$, and then this weight is increased by $1$.We show that for $\\alpha < 1/2$ on a network of bounded degrees, every edge is reinforced a positive proportion of time, and that the limiting proportion can be interpreted as an equilibrium in a countable network. Moreover, in the special case of regular graphs, this homogenization remains valid beyond the threshold $\\alpha = 1/2$.", "subjects": "Probability (math.PR)", "title": "Weakly reinforced Pólya urns on countable networks", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9884918522917869, "lm_q2_score": 0.8104789086703225, "lm_q1q2_score": 0.801151797674953 }
https://arxiv.org/abs/1710.03249
Optimal Graphs for Independence and $k$-Independence Polynomials
The independence polynomial $I(G,x)$ of a finite graph $G$ is the generating function for the sequence of the number of independent sets of each cardinality. We investigate whether, given a fixed number of vertices and edges, there exists optimally-least (optimally-greatest) graphs, that are least (respectively, greatest) for all non-negative $x$. Moreover, we broaden our scope to $k$-independence polynomials, which are generating functions for the $k$-clique-free subsets of vertices. For $k \geq 3$, the results can be quite different from the $k = 2$ (i.e. independence) case.
\section{Introduction} Given a property of subsets of the vertex or edge set -- such as independent, complete, dominating for vertices and matching for edges -- one is often interested in maximizing or minimizing the size of the set in a given graph $G$. However, one can get a much more nuanced study of the subsets by studying the number of such sets of each cardinality in $G$, and in this guise, one often encapsulates the sequence by forming a generating function. Independence, clique, dominating and matching polynomials have all arisen and been studied in this setting. In all cases, the generating polynomial $f(G,x)$ is naturally a function on the domain $[0,\infty)$. If the number of vertices $n$ and edges $m$ are fixed, one can ask whether there exists an extremal graph. Let $\mathcal{S}_{n,m}$ denote the set of simple graphs of order $n$ ($n$ vertices) and size $m$ ($m$ edges). Let $H\in \mathcal{S}_{n,m}$. $H$ is {\em optimally-greatest} (or {\em optimally-least}) if $f(H,x) \geq f(G,x)$ ($f(H,x) \leq f(G,x)$, respectively) for all graphs $G\in \mathcal{S}_{n,m}$ and {\em all} $x \geq 0$ (for any particular value of $x \geq 0$, of course, there is such a graph $H$, as the number of graphs of order $n$ and size $m$ is finite, but we are interested in {\em uniformly} optimal graphs). Such questions (related to simple substitutions of generating polynomials) have attracted considerable attention in the areas of network reliability \cite{ath,suffel,boesch91,browncox,gross,myrvold} and chromatic polynomials \cite{sakaloglu,chromatic}. Here we consider optimality for independence polynomials \[ I(G,x) = \sum_{j} i_{j}x^{j},\] and a broad generalization, for fixed $k \geq 2$, to $k$-independence polynomials \[ I_k(G,x) = \sum_{j} i_{k,j}x^{j},\] where $i_{k,j}$ is the number of subsets of the vertex set of size $j$ that induce a $k$-clique-free subgraph (that is, the induced subgraph contains no complete subgraph of order $k$). Clearly independence polynomials are precisely the $2$-independence polynomials, while the $3$-independence polynomials are generating functions for the numbers of triangle-free induced subgraphs. In this paper, we look at the optimality of independence polynomials and more generally $k$-independence polynomials. In the case of the former, optimally-greatest graphs always exist for independence polynomials, but we do not know whether optimally-least graphs necessarily exist for independence polynomials as well (although we will prove for some $n$ and $m$ they do). In contrast, we shall show that for $k\geq 3$, optimality behaves quite differently, in that for some $n$ and $m$, optimally-least and optimally-greatest graphs for $k$-independence polynomials do not exist. \section{Optimally for Independence Polynomials} \subsection{Optimally-Greatest Graphs for Independence Polynomials} Our first result shows that for independence polynomials, optimally-greatest graphs indeed do always exist. \begin{theorem} For all $n \geq 1$ and all $m \in \{0,\ldots,{n \choose 2}\}$, an optimally-greatest graph always exists. \end{theorem} \begin{proof} A key observation is that a sufficient condition for $H$ to be optimally-greatest (or optimally-least) for independence polynomials is that $i(H,x) = \sum i_{j}(H) x^{j}$ is {\em coefficient-wise greatest}, that is, for all other graphs $G$ of the same order and size, the generating polynomials $f(G,x) = \sum i_{j}(G) x^{j}$ satisfies $i_{j}(H) \geq i_{j}(G)$ for all $j$ (respectively, $i_{j}(H) \leq i_{j}(G)$). (The coefficient-wise condition is not, in general, necessary for optimality as, for example, $5x^{2}+x+5 \geq x^{2}+4x+1$ for $x \geq 0$, but clearly $5x^{2}+x+5$ is not coefficient-wise greater than or equal to $x^{2}+4x+1$.) Consider the following graph construction. For a given $n$ and $m$, take a fixed linear order $\preceq$ of the vertices, $v_{n} \preceq v_{n-1} \preceq \cdots \preceq v_{1}$, and select the $m$ largest edges in lexicographic order. We will denote this graph as $G_{n,m,\preceq}$ (it is dependent on the linear order, but of course all such graphs are isomorphic). It was shown in \cite{cutler} that $G_{n,m,\preceq}$ is the graph of order $n$ and size $m$ with the most number of independent sets of size $j$, for {\em all} $j \geq 0$. It follows that the independence polynomial for $G_{n,m,\preceq}$ is coefficient-wise optimally-greatest for independence polynomials, for all graphs of order $n$ and size $m$. \null\hfill$\Box$\par\medskip \end{proof} Moreover, we can compute the independence polynomial of this optimally-greatest graph. To do so, we need the following well-known recursion to compute the independence polynomial of a graph. Let $G$ be a graph with $v\in V(G)$. Then \begin{eqnarray} I(G,x)=xI(G-N[v],x)+I(G-v,x) \label{ipoly} \end{eqnarray} where $N[v]$ is the closed neighbourhood of $v$ and $G-v$ is $G$ with $v$ removed. For graph $G_{n,m,\preceq}$ with $m < {{n} \choose {2}}$, write $m=(n-\ell)+(n-\ell+1)+\ldots (n-1)+j$ with $j \in \{0,\ldots,n-\ell-2\}$. Since the edges are added in lexicographic order, there will be $k$ vertices that have degree $n-1$ and one vertex of degree $j+\ell$. The remaining $n-\ell-1$ vertices form an independent set. Let $v$ be the vertex of degree $j+\ell$. Using vertex, $v$ and recursion (\ref{ipoly}), we obtain our result, \begin{eqnarray*} I(G_{n,m,\preceq},x)& = &I(G_{n,m,\preceq}-v,x)+xI(G_{n,m,\preceq}-N[v],x)\\ & = & I(K_{\ell} \cup \overline{K_{n-\ell-1}},x) + I(K_{\ell} \cup \overline{K_{n-\ell-j-1}},x)\\ & = & (1+\ell x)(1+x)^{n-\ell-1} + x(1+\ell x)(1+x)^{n-\ell-j-1}\\ & = & (1+\ell x)(1+x)^{n-\ell-j-1} \left( (1+x)^{j} + x\right). \end{eqnarray*} Of course, for $m = {n \choose 2}$, $G_{n,m,\preceq} = K_{n}$ and so $I(G_{n,m,\preceq},x) = 1+nx$. \vspace{0.25in} It is instructive that this result can also be proved purely algebraically, and we devote the remainder of this section to do so. (We refer the reader \cite{brownbook} for an introduction to complexes and their connection to commutative algebra.) The {\em independence complex} of graph $G$ of order $n$ and size $m$, $\Delta_{2}(G)$, has as its faces the independent sets of $G$ (these are obviously closed under containment, and hence a complex). The {\em $f$-vector} of $\Delta_{2}(G)$ is $(1,n,{{n} \choose {2}} - m,f_{3},\ldots,f_{\beta})$, where $f_{i}$ is the number of faces of cardinality $i$ in the complex (and $\beta$ is the independence number of $G$, which is the same as the dimension of the complex). We will show that we can maximize the independence polynomial on $[0,\infty)$ by maximizing ({\em simultaneously}, for some graph $G$) all of the $f_{i}$'s, via an excursion into commutative algebra. We begin with some definitions. Fix a field ${\mathbf k}$. Let $A$ be a {\em ${\mathbf k}$-graded algebra}, that is, $A$ is a commutative ring containing ${\mathbf k}$ as a subring, that can be written as a vector space direct sum $\displaystyle{A = \bigoplus_{d\geq 0} A_{d}}$, over ${\mathbf k}$, with the property that $A_{i}A_{j} \subseteq A_{i+j}$ for all $i$ and $j$ (we call elements in some $A_{i}$ {\em homogeneous}, and $A_{i}$ is called the {\em $d$-th graded component} of $A$). The graded ${\mathbf k}$-algebra $A$ is {\em standard} if it is generated (as a ring) by a finite set of elements in $A_{1}$. Our prototypical example of a standard ${\mathbf k}$-graded algebra is the polynomial ring ${\mathbf k}[x_{1},x_{2},\ldots,x_{n}]$ in variables $x_{1},x_{2},\ldots,x_{n}$. Note that for any standard graded ${\mathbf k}$-algebra $\displaystyle{A = \bigoplus_{d\geq 0} A_{d}}$ that is a quotient of a polynomial ring by a homogenous ideal, a ${\mathbf k}$-basis for $A_{d}$ is simply the monomials in $A_{d}$. The {\em Stanley-Reisner complex} of an ideal $I$ of a standard graded ${\mathbf k}$-algebra $A$ (generated by $x_{1},x_{2},\ldots,x_{n}$ in $A_{1}$) is the (simplicial) complex whose faces are the square-free monomials in $x_{1},x_{2},\ldots,x_{n}$ not in $I$ (the properties of being an ideal ensures that this set is closed under containment). Let $I$ be an ideal of ${\mathbf k}$-algebra $A$; $I$ is {\em homogeneous} if it is generated by homogeneous elements of $A$. We write $\displaystyle{I = \bigoplus_{d\geq 0} I_{d}}$, where $I_{d} = A_{d} \cap I$ is the {\em $d$-th graded component of $I$} (it is a ${\mathbf k}$-subspace of $I$). For a homogeneous ideal $I$ of $A$, a square-free monomial $M$ of degree $d$ in $x_{1},\ldots,x_{n}$ belongs to exactly one of $I_{d}$ and the Stanley-Reisner complex of $I$ (where we identify a face of the complex with the product of its elements). As the total number of monomial of $Q$ of degree $d$ is fixed, we see that maximizing the number of faces of size $d$ in the Stanley-Reisner complex of $I$ corresponds to minimizing the number of monomials of degree $d$ in $I$. The {\em Hilbert function} of the homogeneous ideal $I$ is the function $H_{I}:{\mathbb N} \rightarrow {\mathbb N}$, where $H_{I}(d) = \mbox{dim}_{\mathbf k}I_d$. We call $I$ {\em Gotzmann} (see \cite{hoefelpaper}) if for all other homogeneous ideals $J$ of $A$ and all $d \geq 0$, if $H_{I}(d)=H_{J}(d)$ then $H_{I}(d+1)\leq H_{J}(d+1)$. For an ideal of $Q$ which is Gotzmann, its Hilbert function is smallest for each value of $d \in {\mathbb N}$. We will now focus in on a standard graded ${\mathbf k}$-algebra related to independence in graphs. Fix $n \geq 1$. The {\em Kruskal-Katona ring}, $Q = {\mathbf k}[x_{1},\ldots,x_{n}]/\langle x_{1}^{2},\ldots,x_{n}^{2}\rangle$, is generated by the square-free monomials; it is clearly a standard graded $k$-algebra. Let $G$ be a graph on vertices $\{x_1,x_2,\ldots x_n\}$. The {\em edge ideal} $I_{G}$ is the ideal of $Q$ generated by $\{x_ix_j \mid x_ix_j \in E(G) \}$. If a (square-free) monomial of $Q$ is not in $I_G$ then that set of vertices cannot contain an edge in $G$, so it is an independent set (and vice versa). This means that the Stanely-Reisner complex of our edge ideal $I_{G}$ in the Kruskal-Katona ring is precisely the independence complex $\Delta_{2}(G)$ of our graph $G$. If we can show that the edge ideal $I_{G}$ is Gotzmann in $Q$, then this means that for each $d \geq 0$, $I_{G}$ contains the fewest monomials of degree $d$ for all $d$ compared to {\em any} other such edge ideal. Hence by a previous observation, the $f$-vector of the independence complex of $G$ will have the largest entries component-wise compared to any other graph of order $n$ and size $m$. Thus our graph will have an independence polynomial that is optimally-greatest. Let $I$ be an ideal in a standard graded ${\mathbf k}$-algebra $A$, generated by $x_{1},\ldots,x_{n} \in A_{1}$. Then $I$ is a { \em lexicographic ideal} if for any monomials $u$ and $v$ in $x_{1},\ldots,x_{n}$, whenever $v \in I$ and $u$ is lexicographically bigger than $v$, we have $u \in I$ as well. It is known that lexicographic ideals are Gotzmann in the Kruskal-Katona rings \cite{macaulaylex}. As the edges of our family of graphs $G_{n,m,\preceq}$ are added in lexicographic order, it follows that the edge ideal of $I_G$ in $Q$ is lexicographic, and hence Gotzmann. Therefore the $f$-vector is maximized, given $f_0, f_1, f_2$, that is, given, $n$ and $m$, and so $G_{n,m,\preceq}$ is optimally-greatest. \vspace{0.25in} \subsection{Optimally-Least Graphs for Independence Polynomials} We now turn our attention to the existence of optimally-least graphs for the independence polynomial, and we find here that we can only prove the existence of optimally-least graphs for certain values of $m = m(n)$ (and we do not know if there are values of $n$ and $m$ for which optimally-least graphs do not exist). We begin with dense graphs. It is obvious that for a graph $G$ of order $n$ and size $m$ (which has ${n \choose 2}-m$ many independent sets of cardinality $2$) that the independence polynomial of such a graph $G$ is, for $x \geq 0$, at least \[ 1+nx+\left( {n \choose 2}-m \right)x^2,\] and by a previous observation, if there is a graph with this independence polynomial, then it is the optimally-least. Turan's famous theorem states (see, for example \cite{thebook}) that the maximum number of edges of a graph with no triangles is $\lceil \frac{n}{2} \rceil \lfloor \frac{n}{2} \rfloor = \lfloor \frac{n^{2}}{4} \rfloor$, with equality iff $G$ is a complete bipartite graph with sides of equal or nearly equal cardinality. It follows, by taking complements, that provided \[ m \geq {{n} \choose {2}} - \Bigl \lfloor \frac{n^{2}}{4} \Bigr \rfloor \] then the graph formed by adding any $m - ({{n} \choose {2}} - \lfloor \frac{n^{2}}{4} \rfloor)$ edges to the complement of $K_{ \lceil \frac{n}{2} \rceil, \lfloor \frac{n}{2} \rfloor}$, namely $\overline{K_{ \lceil \frac{n}{2} \rceil, \lfloor \frac{n}{2} \rfloor}} ~=~ K_{\lceil \frac{n}{2} \rceil} \cup K_{\lfloor \frac{n}{2} \rfloor}$ is the optimally-least graph. This gives the following result: \begin{theorem} For a given $n\geq 2$ and $m\geq {{n} \choose {2}} - \lfloor \frac{n^{2}}{4} \rfloor$, the graph with the optimally-least independence polynomial is formed from $K_{\lceil \frac{n}{2} \rceil} \cup K_{\lfloor \frac{n}{2} \rfloor}$ by adding in any $m - ({{n} \choose {2}} - \lfloor \frac{n^{2}}{4} \rfloor)$ edges. The independence polynomial of such a graph is \[ 1+nx+\left( {n \choose 2}-m \right)x^2.\] \null\hfill$\Box$\par\medskip \end{theorem} We can extend this result by utilizing a result of Lov\'{a}sz and Simonovits \cite{lovsim}, which, answering a conjecture of Erd\"{o}s, showed that for $1 \leq k < n/2$, the ($K_{4}$-free) graph of order $n$ and size $\lfloor n^{2}/4 \rfloor$ with the fewest number of triangles is the graph formed from $K_{ \lceil \frac{n}{2} \rceil, \lfloor \frac{n}{2} \rfloor}$ by adding in edges to a largest cell so that no triangles are formed within the cell. (In such a case, the number of triangles formed in the graph is $k \lfloor n/2 \rfloor$.) As well, a theorem of Fisher and Solow \cite{fishersolow} states that for $a \geq b \geq 1$, the least number of triangles in a $K_{4}$-free graph of order $n=2a+b$ and size $m=2ab+a^{2}$ occurs in $K_{a,a,b}$. By taking complements, we derive: \begin{theorem} For a given $n\geq 2$ and $m = {{n} \choose {2}} - \lceil \frac{n}{2} \rceil \lfloor \frac{n}{2} \rfloor - k$, where $1 \leq k \leq \lfloor n/2 \rfloor$, the graph with the optimally-least independence polynomial is formed from $K_{\lceil \frac{n}{2} \rceil} \cup K_{\lfloor \frac{n}{2} \rfloor}$ by deleting any $k$ edges in $K_{\lceil \frac{n}{2} \rceil}$ so that no independent set of size $3$ is formed in that part. The independence polynomial of such a graph is \[ 1+nx+\left( {n \choose 2}-m \right)x^2+ k \lfloor n/2 \rfloor x^{3}.\] As well, for any for $a \geq b \geq 1$, the graph with the optimally-least independence polynomial with order $n=2a+b$ and size $m = a(a-1)+b(b-1)/2$ is $2K_{a} \cup K_{b}$, which has independence polynomial \[ 1+nx+\left( {n \choose 2}-m \right)x^2+\left( 2{a \choose 3} + {b \choose 3}\right)x^3.\] \null\hfill$\Box$\par\medskip \end{theorem} To find sparse families of optimally-least graphs we will look at a graph operation which can be done to decrease the value of the independence polynomial on $[0,\infty)$. Let $H_1$ be a graph which consists of an induced subgraph $G_1$, containing an edge $e = vw$, and two other vertices $y$ and $z$ that are isolated, and let $H_2$ be the graph formed from $H_1$ by removing edge $e$ and adding in an edge between $y$ and $z$ (see Figure~\ref{polyinc} -- we set $G_{2} = G_{1} - e$). Clearly $G_1$ and $G_2$ have the same number of vertices and edges. We will show that this removal of an edge to form a $K_{2}$ can never increase the independence polynomial (on $[0,\infty)$). \begin{figure}[ht] \centering \includegraphics[width=4in]{polyinc.eps} \caption{Graphs for Lemma~\ref{lelmlem}. $H_1$ is the graph $G_1 \cup \{y,z\}$ (on the left) and $H_2$ is the graph $G_2 \cup \{y,z\}$ (on the right).} \label{polyinc} \end{figure} \begin{lemma} \label{lelmlem} For the graphs in Figure~\ref{polyinc}, we have $I(H_2,x)\leq I(H_1,x)$ on $[0,\infty)$. \end{lemma} \begin{proof} Note that by Equation (\ref{ipoly}), our deletion-contraction formula for independence polynomials, we have \begin{eqnarray*} I(H_1,x)& = & (1+2x+x^2)I(G_1,x)\\ & = & (1+2x+x^2)(I(G_1-v,x)+xI(G_1-N[v],x))\\ &=& (1+2x+x^2)(I(G_2-v,x)+xI(G_2-N[v]-w,x)) \end{eqnarray*} and \begin{eqnarray*} I(H_2,x)& = &(1+2x)(I(G_2-v,x)+xI(G_2-N[v],x)). \end{eqnarray*} We also find that \begin{eqnarray}\label{inequal} I(G_2-N[v],x)\leq (1+x)I(G_2-N[v]-w,x) \end{eqnarray} since $G_2-N[v]$ is a subgraph of $(G_2-N[v]-w) \cup K_{1}$. Consider $F(x)=I(H_1,x)-I(H_2,x)$. Using our expression for $i(H_{1},x)$ and $i(H_{2},x)$ and the inequality (\ref{inequal}), we can see that \begin{eqnarray*} F(x) & = & (1+2x)I(G_2-v,x)+x^2I(G_2-v,x)+\\ && x(1+2x+x^2)\ipoly{G_2-N[v]-w}{x}-(1+2x)\ipoly{G_2-v}{x}\\ &&-x(1+2x)\ipoly{G_2-N[v]}{x}\\ &=& x^2\ipoly{G_2-v}{x}+x(1+x)\ipoly{G_2-N[v]-w}{x}\\ &&+x^2(1+x)\ipoly{G_2-N[v]-w}{x}-x(1+2x)\ipoly{G_2-[v]}{x}\\ &\geq & x^2\ipoly{G_2-v}{x}+x\ipoly{G_2-N[v]}{x}\\ &&+x^2\ipoly{G_2-N[v]}{x}-x\ipoly{G_2-[v]}{x}-2x^2\ipoly{G_2-N[v]}{x}\\ &=& x^2\ipoly{G_2-v}{x}-x^2\ipoly{G_2-N[v]}{x}. \end{eqnarray*} Since $G_2-N[v]$ is a subgraph of $G_2-v$, we have $\ipoly{G_2-v}{x}\geq \ipoly{G_2-N[v]}{x}$. It follows that $F(x)\geq 0$, so $\ipoly{H_1}{x}\geq \ipoly{H_2}{x}$. \null\hfill$\Box$\par\medskip \end{proof} It follows that if $m\leq n/2$, by pulling out $K_2$'s, we derive: \begin{theorem} \label{lessthanhalf} For a given $n\geq 2$ and $m \leq \frac{n}{2} $, the optimally-least graph for the independence polynomial for $x \geq 0$ is $mK_2 \cup (n-2m)K_1 $. \null\hfill$\Box$\par\medskip \end{theorem} \section{Optimality for $k$-Independence Polynomials} We now look at the optimality of $k$-independence polynomials and find the situation is much different for $k \geq 3$ than for $k = 2$. We will show that in contrast, to $k=2$, for all $k \geq 3$, there {\em does not} always exist optimally-greatest nor optimally-least graphs for the $k$-independence polynomial. Before we do so, we make the following observation, which shows that the nonexistence of optimal graphs can sometimes be derived by considering only certain coefficients of the polynomials. \begin{observation} \label{max} Suppose that $G$ and $H$ be graphs on $n$ vertices and $m$ edges and let $k\geq 2$, with \[ {\rm I}_{k}(G,x)=\sum_{j=0}^ni_j(G)x^{j} \] and \[ {\rm I}_{k}(H,x)=\sum_{j=0}^ni_j(H)x^{j} \] Then \begin{itemize} \item if $i_j(G)=i_j(H)$ for $j<\ell$ but $i_{\ell}(G)>i_{\ell}(H)$, then {\rm I}$_{k}(G,x)>${\rm I}$_{k}(H,x)$ for $x$ arbitrarily small and \item if $i_j(G)=i_j(H)$ for $t>j$ but $i_{t}(G)>i_{t}(H)$, then {\rm I}$_{k}(G,x)>${\rm I}$_{k}(H,x)$ for $x$ arbitrarily large. \end{itemize} \end{observation} Note that $i_{j} = {n \choose j}$ for $j < k$. For a graph $G$, let $r_{G} = r^{k}_{G}$ denote the largest value of $j$ such that there exists an induced subgraph of $G$ of order $j$ that does not contains a $k$-clique, that is, $r_{G}$ is the largest value of $j$ such that $i_j(G) > 0$. Thus to show that for $k$-independence polynomials ($k \geq 3$) there does not always exist optimally-greatest graphs, we will show that for some $n$ and $m = m(n)$, there is a unique graph $G \in \mathcal{S}_{n,m}$ with the largest value of $r_{G}$ in $\mathcal{S}_{n,m}$, thus optimally-greatest for sufficiently large values of $x$, but there is another graph $H \in \mathcal{S}_{n,m}$ with more $k$-independent sets than $G$, and hence optimally-greatest the $k$-independence polynomial for arbitrarily small values of $x\geq 0$. \begin{theorem} For $k \geq 3$ and $l \geq 2$, and any $n > (k-1)l(l-1)$, there does not exist an optimally-greatest graph for the $k$-independence polynomial of order $n$ and size $m = {n \choose 2}-(k-1){l \choose 2}$. \end{theorem} \begin{proof} We recall an old well known result by Turan (see \cite{bollobas2}, for example) that states that the unique graph $T_{n,k}$ of order $n$ with the maximum number $m_{n,k}$ of edges in a graph of order $n$ without a $k$-clique is the complete $(k-1)$-partite graph with cells of order $\lfloor n/(k-1) \rfloor$ or $\lceil n/(k-1) \rceil$. For fixed $n > (k-1)l(l-1)$, we set $m = {n \choose 2}-(k-1){l \choose 2}$ and consider the class $\mathcal{S}_{n,m}$. We define the graph $G = G(n,m)$ as the join $T_{(k-1)l,m} + K_{n-(k-1)l}$ of the Turan graph $T_{(k-1)l,m}$ with $K_{n-(k-1)l}$, that is, $G$ is formed from the disjoint union of $T_{(k-1)l,k}$ with $K_{n-(k-1)l}$ by adding in all edges between them (equivalently, the complement of $G$, $\overline{G}$, consists of $k-1$ cliques of order $l$, together with $n-(k-1)l$ isolated vertices). A quick calculation shows that $G$ has $m$ edges, and hence belongs to $\mathcal{S}_{n,m}$. We claim first that among all graphs $F^{\prime}$ in $\mathcal{S}_{n,m}$, $G$ has the largest value of $r_{F^{\prime}}$. Note that $r_{G} = (k-1)l$ as the Turan graph $T_{(k-1)l,k}$ has no $k$-clique. Now if there is a graph $F$ in $\mathcal{S}_{n,m}$ with $r_{F} > r_{G}$, then $F$ has an induced subgraph $S$ on say $s > (k-1)l$ vertices with no $k$-clique, and hence $S$ has at most as many edges as $T_{s,m}$. However, then $\overline{F}$ has at least as many edges as $\overline{T_{s,m}}$, which is strictly more than the number of edges in $\overline{T_{(k-1)l,m}}$ (to see this, think of the complements of Turan graphs being the disjoint unions of cliques, and observe that for $t > (k-1)l$, one can form $\overline{T_{t,m}}$ from $\overline{T_{(k-1)l,m}}$ by successively adding vertices to the cliques to keep them as nearly equal as possible). Thus $\overline{F}$ would contain more edges than $\overline{G}$, the disjoint union of $\overline{T_{(k-1)l,m}}$ and isolated vertices, and hence $F$ would fewer edges than $G$, a contradiction as both $F$ and $G$ have the same number of vertices and edges. We conclude no such $F$ exists, so $G$ has the maximal $r_F$ value. Moreover, by the argument given, if $G^{\prime}$ were a graph in $\mathcal{S}_{n,m}$ with $r_{G^{\prime}} = r_{G}$, then if $S$ is any induced subgraph of $G^{\prime}$ of size $r_{G}=(k-1)l$ that has no $k$-clique, $\overline{S}$ must have precisely as many edges as $\overline{T_{(k-1)l,k}}$, which is the number of edges of $\overline{G}$. We conclude that, from Turan's Theorem, that $G^\prime$ is isomorphic to $G$, which is therefore the unique graph in $\mathcal{S}_{n,m}$ with the largest $r_F$-value, and hence the unique graph optimally-greatest for the $k$-independence polynomial for $x$ sufficiently large. So if there is an optimally-greatest graph the $k$-independence polynomial for $\mathcal{S}_{n,m}$, it must be $G$. We note that if a graph of order $n$ and size $m$ has a minimum number of $k$-cliques, then it has a maximum number of $k$-independent sets of order $k$. We will show now that there is another graph $H \in \mathcal{S}_{n,m}$ with fewer $k$-cliques than $G$, and hence more $k$-independent sets of size $k$, and so $G$ cannot be optimally-greatest the $k$-independence polynomial as $I_{k}(H,x) > I_{k}(G,x)$ for $x> 0 $ sufficiently small. Let $H$ be the graph of order $n$ such that $\overline{H}$ is the disjoint union of $(k-1){l \choose 2}$ $K_{2}$'s and isolated vertices (as $n > (k-1)l(l-1) = 2(k-1){l \choose 2}$, we can find such a graph of order $n$). We can think of $\overline{H}$ as splitting the edges of the $l$-cliques of $\overline{G}$ into edges. Now it is easy to see that for graphs $G_{1}$ and $G_{2}$ having $k_{1,i}$ and $k_{2,i}$ cliques of order $i$, respectively, then for any positive integer $t$, the number of $k$-cliques of order $t$ in $G_{1}+G_{2}$ is \[ \sum_{i+j = t} k_{1,i}k_{2,j}.\] It follows that it suffices to show that the following graph $G_{1} = \overline{K_{l}+l(l-2)K_{1}}$ has, for all $i$, at least as many $i$-cliques as the graph $G_{2} = \overline{{l \choose 2}K_{2}}$, and that for $i \geq 3$ the former has strictly more $i$-cliques than the latter. For $i = 1 $ and $i = 2$ they have the same number of $i$-cliques (as they have the same number of vertices and edges). For $i \geq 2$, the number of $i$-cliques in $G_{1}$ is \[ {l(l-2) \choose i} + l {l(l-2) \choose {i-1}}\] while $G_{2}$ has \[ {{l \choose 2} \choose i}2^{i}\] many $i$-cliques. We set \[ f_{l,i} = \frac{{l(l-2) \choose i} + l {l(l-2) \choose {i-1}}}{{{l \choose 2} \choose i}2^{i}}.\] Clearly $f_{l,2} = 1$ as both $G_{1}$ and $G_{2}$ has the same number of edges. We'll show that the sequence is strictly increasing, and hence for $i \geq 3$, always greater than $1$, so that $G_{1}$ has strictly more $i$-cliques than $G_{2}$, concluding the proof. Now via some straightforward but tedious calculations, we find that for $i \geq 2$, \[ \frac{f_{l,i+1}}{f_{l,i}} = \frac{(l(l-2)-i+l(i+1))(l(l-2)-i+1)}{(l(l-1)-2i)(l(l-2)-i+1+li)}.\] It follows that \[ \frac{f_{l,i+1}}{f_{l,i}} > 1 \] iff \begin{eqnarray*} (l(l-2)-i+l(i+1))(l(l-2)-i+1) & > & (l(l-1)-2i)(l(l-2)-i+1+li), \end{eqnarray*} which holds iff \[ i(l-1)(i-1) > 0.\] The latter is true as $i, l \geq 2$. Thus we conclude that $f_{l,i} > 1$ for all $i \geq 2$. Thus $H$ has fewer $k$-cliques than $G$, hence more $k$-independent sets of order $k$ and so an optimally-greatest graph the $k$-independence polynomial does not exist. \null\hfill$\Box$\par\medskip \end{proof} \vspace{0.25in} We turn finally to the issue of optimally-least $k$-independence polynomials. From Observation~\ref{max}, we have that if an optimally-least graph the $k$-independence polynomial exists, then it must have the least number of $k$-independent sets of order $k$ (or equivalently, for the optimally-least graph the $k$-independence polynomial has the maximum number of $k$-cliques), since a graph with the latter will be optimally-least the $k$-independence polynomial for sufficiently small values of $x\geq 0$. In \cite{bollobas2} it was shown that for a graph on $n$ vertices and $m={d \choose 2}+r$ edges ($0 \leq r < d$) and for $k \geq 3$, the maximum number of cliques of size $k$ is ${d \choose k}+{r \choose k-1}$, and a graph which achieves such bounds consists of a $K_d$, a vertex $x$ with $N(x)\subseteq V(K_d)$, and $n-d-1$ isolated vertices (see Figure~\ref{optkleastnear0}). (This graph is not the unique extremal graph for values of $r<k-1$, since the addition of fewer than $k-1$ edges will not produce another $k$-clique, but for values of $r\geq k-1$, this graph is the unique extremal graph, since to obtain the maximum number of $k$-cliques, the edges will have to be added to the same vertex.) Such graphs are candidates for optimally-least graphs the $k$-independence polynomial, but we will show that for $k\geq 3$, that optimally-least graphs the $k$-independence polynomial do not always exist. \begin{figure}[ht] \centering \includegraphics[width=1.8in]{optkleastnear0.eps} \caption{A graph that is optimally-least for the $k$-independence polynomial near $0$.} \label{optkleastnear0} \end{figure} \begin{theorem} For $k\geq 3$, $n \geq {k\choose 2} + k + 1$ vertices and $m={{k\choose 2}+2\choose 2}-1$ optimally-least graphs the $k$-independence polynomial do not exist. \end{theorem} \begin{proof} Let $k\geq 3$. Let $G=\left( K_{{k\choose 2}+2}-e \right) \cup (n-{k \choose 2}-2)K_1$, for some edge $e$ of $K_{{k\choose 2}+2}$, and let $H=K_{{k\choose 2}+1}\cup K_k\cup (n-{k \choose 2}-1-k)K_1$. We know that $G$ is optimally-least for the $k$-independence polynomial, when $x$ is sufficiently close to $0$, and $H$ is not. We will now show that $H$ is optimally-least for the $k$-independence polynomial for arbitrarily large values of $x$. In $G$, the largest size of a vertex set that does not contain a $K_k$ is $n-{k \choose 2}+k-2$, taking any of the $n-{k\choose 2}-2$ isolated vertices, end points of edge $e$ and $k-2$ vertices from $K_{{k\choose 2}+2}$. In $H$, the largest size of a vertex set that does not contain a $K_k$ is size $n-{k \choose 2}+k-3$, taking any of the isolated $n-{k \choose 2}-1-k$ vertices and $k-1$ vertices from each complete graph. Thus by Observation~\ref{max}, $H$ is optimally-least the $k$-independence polynomial for arbitrarily large values of $x$, and $G$ is not, so no optimally-least graphs exist for the $k$-independence polynomial for these $n$ and $m$. \null\hfill$\Box$\par\medskip \end{proof} \section{Conclusion} While we have seen that optimally-greatest graphs exist for independence polynomials, we have only be able to prove the existence of optimally-least polynomials for a restricted collection of $n$ and $m$. Our belief is that such graphs always exist, but there does not seem to be any reasonable family to put forward as extremal. In terms of optimality for $k$-independence polynomials, we have seen that for all $k \geq 3$, there are infinitely many values of $n$ and $m$ such that optimally-greatest graphs do not exists, and similarly for optimally-least graphs. A full characterization of when optimal graphs for the $k$-independence polynomial exist (for $k\geq 3$, and even for optimally-least for $k=2$) remains open. \vspace{0.25in} \noindent {\bf Acknowledgements} \vspace{0.1in} \noindent J.I. Brown acknowledges support from NSERC (grant application RGPIN 170450-2013). D. Cox acknowledges research support from NSERC (grant application RGPIN 2017-04401) and Mount Saint Vincent University.
{ "timestamp": "2017-10-11T02:00:41", "yymm": "1710", "arxiv_id": "1710.03249", "language": "en", "url": "https://arxiv.org/abs/1710.03249", "abstract": "The independence polynomial $I(G,x)$ of a finite graph $G$ is the generating function for the sequence of the number of independent sets of each cardinality. We investigate whether, given a fixed number of vertices and edges, there exists optimally-least (optimally-greatest) graphs, that are least (respectively, greatest) for all non-negative $x$. Moreover, we broaden our scope to $k$-independence polynomials, which are generating functions for the $k$-clique-free subsets of vertices. For $k \\geq 3$, the results can be quite different from the $k = 2$ (i.e. independence) case.", "subjects": "Combinatorics (math.CO)", "title": "Optimal Graphs for Independence and $k$-Independence Polynomials", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9884918536478773, "lm_q2_score": 0.8104788995148792, "lm_q1q2_score": 0.8011517897239546 }
https://arxiv.org/abs/1811.06493
Intermediate dimensions
We introduce a continuum of dimensions which are `intermediate' between the familiar Hausdorff and box dimensions. This is done by restricting the families of allowable covers in the definition of Hausdorff dimension by insisting that $|U| \leq |V|^\theta$ for all sets $U, V$ used in a particular cover, where $\theta \in [0,1]$ is a parameter. Thus, when $\theta=1$ only covers using sets of the same size are allowable, and we recover the box dimensions, and when $\theta=0$ there are no restrictions, and we recover Hausdorff dimension.We investigate many properties of the intermediate dimension (as a function of $\theta$), including proving that it is continuous on $(0,1]$ but not necessarily continuous at $0$, as well as establishing appropriate analogues of the mass distribution principle, Frostman's lemma, and the dimension formulae for products. We also compute, or estimate, the intermediate dimensions of some familiar sets, including sequences formed by negative powers of integers, and Bedford-McMullen carpets.
\section{Intermediate dimensions: definitions and background} \setcounter{equation}{0} \setcounter{theo}{0} We work with subsets of $\mathbb{R}^n$ throughout, although much of what we establish also holds in more general metric spaces. We denote the {\em diameter} of a set $F$ by $|F|$, and when we refer to a {\em cover} $\{U_i\}$ of a set $F$ we mean that $F\subseteq \bigcup_i U_i$ where $\{U_i\}$ is a finite or countable collection of sets. Recall that Hausdorff dimension $\dim_{\rm H}$ may be defined without introducing Hausdorff measures, but using Hausdorff content. For $F\subseteq \mathbb{R}^n$, \begin{align*}\label{hdim} \dim_{\rm H} F = \inf \big\{& s\geq 0 : \mbox{ \rm for all $\epsilon >0$ there exists a cover $ \{U_i\} $ of $F$ such that $\sum |U_i|^s \leq \epsilon$} \big\}, \end{align*} see \cite[Section 3.2]{Fa}. (Lower) box dimension $\lbd$ may be expressed in a similar manner, by forcing the covering sets to be of the same diameter. For bounded $F\subseteq \mathbb{R}^n$, \begin{align*} \lbd F = \inf \big\{ s\geq 0 &: \mbox{ \rm for all $\epsilon >0$ there exists a cover $ \{U_i\} $ of $F$} \\ & \mbox{ \rm such that $|U_i| = |U_j|$ for all $i,j$ and $\sum |U_i|^s \leq \epsilon$} \big\}. \end{align*} see \cite[Chapter 2]{Fa}. Expressed in this way, Hausdorff and box dimensions may be regarded as extreme cases of the same definition, one with no restriction on the size of covering sets, and the other requiring them all to have equal diameters. With this in mind, one might regard them as the extremes of a continuum of dimensions with increasing restrictions on the relative sizes of covering sets. This is the main idea of this paper, which we formalise by considering restricted coverings where the diameters of the smallest and largest covering sets lie in a geometric range $\delta^{1/\theta} \leq |U_i| \leq \delta$ for some $0\leq \theta \leq 1$. \begin{defi}\label{adef} Let $F\subseteq \mathbb{R}^n$ be bounded. For $0\leq \theta \leq 1$ we define the {\em lower $\theta$-intermediate dimension} of $F$ by \begin{align*} \da F = \inf \big\{& s\geq 0 : \mbox{ \rm for all $\epsilon >0$ and all $\delta_0>0$, there exists $0<\delta\leq \delta_0$} \\ & \mbox{ \rm and a cover $ \{U_i\} $ of $F$ such that $\delta^{1/\theta} \leq |U_i| \leq \delta$ and $\sum |U_i|^s \leq \epsilon$} \big\}. \end{align*} Similarly, we define the {\em upper $\theta$-intermediate dimension} of $F$ by \begin{align*} \uda F = \inf \big\{& s\geq 0 : \mbox{ \rm for all $\epsilon >0$ there exists $\delta_0>0$ such that for all $0<\delta\leq \delta_0$,} \\ & \mbox{ \rm there is a cover $ \{U_i\} $ of $F$ such that $\delta^{1/\theta} \leq |U_i| \leq \delta$ and $\sum |U_i|^s \leq \epsilon$} \big\}. \end{align*} \end{defi} With these definitions, $$\underline{\mbox{\rm dim}}_{0} F = \overline{\mbox{\rm dim}}_{0} F = \hdd F, \quad \underline{\mbox{\rm dim}}_{1} F = \lbd F \quad {\mbox{ and }}\quad \overline{\mbox{\rm dim}}_{1} F = \ubd F, $$ where $\ubd$ is the upper box dimension. Moreover, it follows immediately that, for a bounded set $F$ and $\theta \in [0,1]$, \[ \hdd F \leq \da F \leq \uda F \leq \ubd F \quad {\mbox{ and }}\quad \da F \leq \lbd F. \] It is also immediate that $\da F$ and $\uda F$ are increasing in $\theta$, though as we shall see they need not be strictly increasing. Furthermore, $\uda$ is finitely stable, that is $\uda (F_1\cup F_2) = \max\{\uda F_1, \uda F_2\}$, and, for $\theta\in (0,1]$, both $\da F$ and $\uda F$ are unchanged on replacing $F$ by its closure. In many situations, even if $\hdd F < \ubd F$, we still have $\lbd F = \ubd F$ and $\da F = \uda F$ for all $\theta \in [0,1]$. In this case we refer to the box dimension $\bdd F = \lbd F = \ubd F$ and the {\em $\theta$-intermediate dimension} $\mbox{\rm dim}_{\theta} F = \da F = \uda F$. This paper is devoted to understanding $\theta$-intermediate dimensions. The hope is that $\mbox{\rm dim}_{\theta} F$ will interpolate between the Hausdorff and box dimensions in a meaningful way, a rich and robust theory will be discovered, and interesting further questions unearthed. We first derive useful properties of intermediate dimensions, including that $\underline{\mbox{\rm dim}}_{0} F$ and $\overline{\mbox{\rm dim}}_{0} F $ are continuous on $(0,1]$ but not necessarily at $0$, as well as proving versions of the mass distribution principle, Frostman's lemma and product formulae. We then examine a range of examples illustrating different types of behaviour including sequences formed by negative powers of integers, and self-affine Bedford-McMullen carpets. Intermediate dimensions provide an insight into the distribution of the diameters of covering sets needed when estimating the Hausdorff dimensions of sets whose Hausdorff and box dimensions differ. They also have concrete applications to well-studied problems. For example, since the intermediate dimensions are preserved under bi-Lipschitz mappings, they provide another invariant for Lipschitz classification of sets. A very specific variant was used in \cite{KP} to estimate the singular sets of partial differential equations. A related approach to `dimension interpolation' was recently considered in \cite{spec1} where a new dimension function was introduced to interpolate between the box dimension and the Assouad dimension. In this case the dimension function was called the \emph{Assouad spectrum}, denoted by $\mbox{\rm dim}_\textup{A}^{\theta} F$ $(\theta \in (0,1))$. \section{Properties of intermediate dimensions}\label{sec:props} \setcounter{equation}{0} \setcounter{theo}{0} \subsection{Continuity} The first natural question is whether, for a fixed bounded set $F$, $\da F$ and $\uda F$ vary continuously for $\theta\in [0,1]$. We show this is the case, except possibly at $\theta = 0$. We provide simple examples exhibiting discontinuity at $\theta = 0$ in Section \ref{examplessimple}. However, for many natural sets $F$ we find that the intermediate dimensions \emph{are} continuous at $0$ (and thus on $[0,1]$), for example for self-affine carpets, see Section \ref{carpetssec}. \begin{prop}\label{cty} Let $F$ be a non-empty bounded subset of $\mathbb{R}^n$ and let $0\leq \theta<\phi \leq 1$. Then \begin{equation}\label{ineqs} \da F \ \leq \ \db F \ \leq \ \da F +\Big(1-\frac{\theta}{\phi}\Big) (n- \da F). \end{equation} and \begin{equation}\label{ineqs2} \uda F \ \leq \ \udb F \ \leq \ \uda F +\Big(1-\frac{\theta}{\phi}\Big) (n- \uda F). \end{equation} In particular, $\theta \mapsto\da F$ and $\theta \mapsto\uda F$ are continuous for $\theta \in (0,1]$. \end{prop} \begin{proof} We will only prove \eqref{ineqs} since \eqref{ineqs2} is similar. The left-hand inequality of \eqref{cty} is just the monotonicity of $\da F$. The right-hand inequality is trivially satisfied when $\da F=n$, so we assume that $0\leq \da F <n$. Suppose that $0\leq \theta<\phi \leq 1$ and that $0\leq \da F <s <n$. Then, given $\epsilon >0$, we may find arbitrarily small $\delta>0$ and countable or finite covers $\{U_i\}_{i\in I}$ of $F$ such that \begin{equation}\label{epsum} \sum_{i\in I} |U_i|^s<\epsilon \quad \mbox{ and } \quad \delta \leq |U_i| \leq \delta^\theta \quad \mbox{ for all } i\in I. \end{equation} Let $$I_0 =\{i\in I: \delta \leq |U_i| \leq \delta^\phi\} \quad \mbox{ and } \quad I_1 =\{i\in I: \delta^\phi < |U_i| \leq \delta^\theta\}.$$ For each $ i\in I_1$ we may split $U_i$ into subsets of small coordinate cubes to get sets $\{U_{i,j}\}_{j\in J_i}$ such that $U_i \subseteq\bigcup_{j\in J_i} U_{i,j}$, with $|U_{i,j}| \leq \delta^\phi$ and ${\rm card} J_i \ \leq \ c_n|U_i|^n \delta^{-\phi n} \ \leq \ c_n\delta^{n(\theta-\phi)}$, where $c_n = 4^n n^{n/2}$. Let $s<t\leq n$. Then $ \{U_i\}_{i\in I_0} \cup \{U_{i,j}\}_{i\in I_1, j \in J_i}$ is a cover of $F$ such that $\delta \leq |U_i|, |U_{i,j}| \leq \delta^\phi$. Taking sums with respect to this cover: \begin{eqnarray*} \sum_{i\in I_0}|U_i|^t + \sum_{i\in I_1}\sum_{j\in J_{i}}|U_{i,j}|^t &\leq& \sum_{i\in I_0}|U_i|^t + \sum_{i\in I_1} \delta^{\phi t} c_n |U_i|^n \delta^{-\phi n} \\ &\leq& \sum_{i\in I_0}|U_i|^t + c_n \sum_{i\in I_1} |U_i|^s |U_i|^{n-s}\delta^{\phi( t-n)} \\ &\leq& \sum_{i\in I_0}|U_i|^s + c_n \sum_{i\in I_1} |U_i|^s \delta^{\theta(n-s)}\delta^{\phi( t-n)} \\ &\leq& \sum_{i\in I_0}|U_i|^s + c_n \delta^{\phi[t-(n\phi+\theta(s-n))/\phi]}\sum_{i\in I_1} |U_i|^s \\ &\leq& (1+c_n) \sum_{i\in I}|U_i|^s\ < \ (1+c_n)\epsilon \end{eqnarray*} if $t\geq (n\phi+\theta(s-n))/\phi$, from \eqref{epsum}. This holds for some cover for arbitrarily small $\epsilon$ and all $s>\da F$, giving $\db F \leq n + \theta (\da F-n)/\phi$, which rearranges to give \eqref{ineqs}. Finally, note that \eqref{ineqs2} follows by exactly the same argument noting that the assumption $ \uda F <s $ gives rise to $\delta_0>0$ such that for all $\delta \in (0,\delta_0)$ we can find covers $\{U_i\}_{i\in I}$ of $F$ satisfying \eqref{epsum}. \end{proof} \subsection{A mass distribution principle for $\da$ and $\uda$} The \emph{mass distribution principle} is a powerful tool in fractal geometry and provides a useful mechanism for estimating the Hausdorff dimension from below by considering measures supported on the set, see \cite[page 67]{Fa}. We present natural analogues for $\da$ and $\uda$. \begin{prop}\label{mdp} Let $F$ be a Borel subset of $\mathbb{R}^n$ and let $0\leq \theta \leq 1$ and $s\geq 0$. Suppose that there are numbers $a, c, \delta_0 >0$ such that for all $0< \delta\leq \delta_0$ we can find a Borel measure $\mu_\delta$ supported by $F$ with $\mu_\delta (F) \geq a $, and with \begin{equation}\label{mdiscond} \mu_\delta (U) \leq c|U|^s \quad \mbox{ for all Borel sets $U \subseteq \mathbb{R}^n $ with } \delta \leq |U|\leq \delta^\theta. \end{equation} Then $\da F \geq s$. Moreover, if measures $\mu_\delta$ with the above properties can be found only for a sequence of $\delta \to 0$, then the conclusion is weakened to $\uda F \geq s$. \end{prop} \begin{proof} Let $\{U_i\}$ be a cover of $F$ such that $\delta \leq |U_i|\leq \delta^\theta$ for all $i$. Then $$a\ \leq \ \mu_\delta(F) \ \leq\ \mu_\delta\Big(\bigcup_i U_i\Big)\ \leq \ \sum_i \mu_\delta(U_i) \ \leq \ c\sum_i |U_i|^s,$$ so that $\sum_i |U_i|^s \geq a/c>0$ for every admissible cover and therefore $\da F \geq s$. The weaker conclusion regarding the upper intermediate dimension is obtained similarly. \end{proof} Note the main difference between Proposition \ref{mdp} and the usual mass distribution principle is that a family of measures $\{\mu_\delta\}$ is used instead of a single measure. Since each measure $\mu_\delta$ is only required to describe a range of scales, in practice one can often use finite sums of point masses. Whilst the measures $\mu_\delta$ may vary, it is essential that they all assign mass at least $a>0$ to $F$. \subsection{A Frostman type lemma for $\da$} \emph{Frostman's lemma} is another powerful tool in fractal geometry, which asserts the existence of measures of the type considered by the mass distribution principle, see \cite[page 77]{Fa} or \cite[page 112]{mattila}. The following analogue of Frostman's lemma holds for intermediate dimensions and is a useful dual to Proposition \ref{mdp}. \begin{prop}\label{frostman} Let $F$ be a compact subset of $\mathbb{R}^n$, let $0< \theta \leq 1$, and suppose $0< s< \da F$. There exists a constant $c >0$ such that for all $\delta \in (0,1)$ we can find a Borel probability measure $\mu_\delta$ supported on $F$ such that for all $x \in \mathbb{R}^n$ and $\delta^{1/\theta} \leq r \leq \delta$, \[ \mu_\delta (B(x,r)) \leq c r^s . \] Moreover, $\mu_\delta$ can be taken to be a finite collection of atoms. \end{prop} \begin{proof} This proof follows the proof of the classical version of Frostman's lemma given in \cite[pages 112-114]{mattila}. For $m \geq 0$ let $\mathcal{D}_m$ denote the familiar partition of $[0,1]^n$ consisting of $2^{nm}$ pairwise disjoint half-open dyadic cubes of sidelength $2^{-m}$, that is cubes of the form $[a_1, a_1+2^{-m}) \times \cdots \times [a_n, a_n+2^{-m})$. By translating and rescaling we may assume without loss of generality that $F \subseteq [0,1]^n$ and that $F$ is not contained in any $Q \in \mathcal{D}_1$. It follows from the definition of $\da F$ that there exists $\varepsilon>0$ such that for all $\delta \in (0,1)$ and for all covers $\{U_i\}_i$ of $F$ satisfying $\delta^{1/\theta} \leq |U_i| \leq \delta$, \begin{equation} \label{goodcover1} \sum_i |U_i|^s > \varepsilon. \end{equation} Given $\delta \in (0,1)$, let $m \geq 0$ be the unique integer satisfying $2^{-m-1}< \delta^{1/\theta} \leq 2^{-m}$ and let $\mu_m$ be a measure defined on $F$ as follows: for each $Q \in \mathcal{D}_m$ such that $Q \cap F \neq \emptyset$, then choose an arbitrary point $ x_Q \in Q \cap F$ and let \[ \mu_m = \sum_{ Q \in \mathcal{D}_m : Q \cap F \neq \emptyset} 2^{-ms} \delta_{x_Q} \] where $\delta_{x_Q}$ is a point mass at $x_Q$. Modify $\mu_m$ to form a measure $\mu_{m-1}$, supported on the same finite set, defined by \[ \mu_{m-1}\vert_Q = \min\{1, 2^{-(m-1)s} \mu_m(Q)^{-1}\} \mu_m\vert_Q \] for all $Q \in \mathcal{D}_{m-1}$, where $\nu \vert_E$ denotes the restriction of $\nu$ to $E$. The purpose of this modification is to reduce the mass of cubes which carry too much measure. This is done since we are ultimately trying to construct a measure which we can estimate uniformly from above. Continuing inductively, $\mu_{m-k-1}$ is obtained from $\mu_{m-k}$ by \[ \mu_{m-k-1}\vert_Q = \min\{1, 2^{-(m-k-1)s} \mu_{m-k} (Q)^{-1}\} \mu_{m-k} \vert_Q \] for all $Q \in \mathcal{D}_{m-k-1}$. We terminate this process when we define $\mu_{m-l}$ where $l$ is the largest integer satisfying $2^{-(m-l)}n^{1/2} \leq \delta$. (We may assume that $l \geq 0$ by choosing $\delta$ sufficiently small to begin with.) In particular, cubes $Q \in \mathcal{D}_{m-l}$ satisfy $|Q| = 2^{-(m-l)}n^{1/2} \leq \delta$. By construction we have \begin{equation} \label{goodbound1} \mu_{m-l}(Q) \leq 2^{-(m-k)s} = |Q|^s n^{-s/2} \end{equation} for all $k =0, \dots, l$ and $Q \in \mathcal{D}_{m-k}$. Moreover, for all $x \in F$, there is at least one $k \in \{0, \dots, l\}$ and $Q \in \mathcal{D}_{m-k}$ with $x \in Q$ such that the inequality in \eqref{goodbound1} is an equality. This is because all cubes at level $m$ satisfy the equality for $\mu_m$ and if a cube $Q$ satisfies the equality for $\mu_{m-k}$, then either $Q$ or its parent cube satisfies the equality for $\mu_{m-k-1}$. For each $x \in F$, choosing the largest such $Q$ yields a finite collection of cubes $Q_1, \dots, Q_t$ which cover $F$ and satisfy $\delta^{1/\theta} \leq |Q_i| \leq \delta$ for $i = 1, \dots, t$. Therefore, using \eqref{goodcover1}, \[ \mu_{m-l}(F) = \sum_{i=1}^t \mu_{m-l}(Q_i) = \sum_{i=1}^t |Q_i|^s n^{-s/2} > \varepsilon n^{-s/2}. \] Let $\mu_\delta = \mu_{m-l}(F)^{-1} \mu_{m-l}$, which is clearly a probability measure supported on a finite collection of points. Moreover, for all $x \in \mathbb{R}^n$ and $\delta^{1/\theta} \leq r \leq \delta$, $B(x,r)$ is certainly contained in at most $c_n$ cubes in $\mathcal{D}_{m-k}$ where $k$ is chosen to be the largest integer satisfying $0 \leq k \leq l$ and $2^{-(m-k+1)} < r$, and $c_n$ is a constant depending only on $n$. Therefore, using \eqref{goodbound1}, \[ \mu_\delta (B(x,r)) \leq c_n \mu_{m-l}(F)^{-1} 2^{-(m-k) s} \leq c_n \varepsilon^{-1} n^{s/2} 2^s r^s \] which completes the proof, setting $c= c_n \varepsilon^{-1} n^{s/2} 2^s $. \end{proof} \subsection{General bounds} Here we consider general bounds which rely on the Assouad dimension and which have interesting consequences for continuity. Namely, they provide numerous examples where the intermediate dimensions are \emph{dis}continuous at $\theta=0$ and also provide another proof that the intermediate dimensions are continuous at $\theta = 1$. The \emph{Assouad dimension} of $F \subseteq \mathbb{R}^n$ is defined by \begin{eqnarray*} \dim_\textup{A} F &=& \inf \bigg\{ s \geq 0 \ : \ \text{there exists $C>0$ such that for all $x \in F$, } \\ &\,& \qquad \qquad \text{and for all $0<r<R$, we have} \ N_r(F \cap B(x,R)) \leq C \left(\frac{R}{r} \right)^s \bigg\} \end{eqnarray*} where $N_r(A)$ denotes the smallest number of sets of diameter at most $r$ required to cover a set $A$. In general $\underline{\dim}_\textup{B} F \leq \overline{\dim}_\textup{B} F \leq \dim_\textup{A} F \leq n$, but equality of these three dimensions occurs in many cases, even if the Hausdorff dimension and box dimension are distinct, for example if the box dimension is equal to the ambient spatial dimension. See \cite{Fra, robinson} for more background on the Assouad dimension. The following proposition gives lower bounds for the intermediate dimensions in terms of Assouad dimensions. \begin{prop} \label{assouad} Given any non-empty bounded $F \subseteq \mathbb{R}^n$ and $\theta \in (0,1)$, \[ \da F \geq \dim_\textup{A} F - \frac{\dim_\textup{A} F - \underline{\dim}_\textup{B} F}{\theta}, \] and \[ \uda F \geq \dim_\textup{A} F - \frac{\dim_\textup{A} F - \overline{\dim}_\textup{B} F}{\theta}. \] In particular, if $\underline{\dim}_\textup{B} F = \dim_\textup{A} F$, then $\da F =\overline{\dim}_\theta F = \underline{\dim}_\textup{B} F = \dim_\textup{A} F$ for all $\theta \in (0,1]$. \end{prop} \begin{proof} We will prove the lower bound for $\da F$, the proof for $\uda F$ is similar. Fix $\theta \in (0,1)$ and assume that $\underline{\dim}_\textup{B} F >0$, otherwise the result is trivial. Let \[ 0<b< \underline{\dim}_\textup{B} F \leq \dim_\textup{A} F < d < \infty \] and $\delta \in (0,1)$ be given. By the definition of lower box dimension, there exists a uniform constant $C_0$, depending only on $F$ and $b$, such that there is a $\delta$-separated set of points in $F$ of cardinality at least $C_0 \delta^{-b}$. Let $\mu_\delta$ be a uniformly distributed probability measure on these points, i.e. a sum of $C_0\delta^{-b}$ point masses each with mass $C_0^{-1} \delta^{b}$. We use our mass distribution principle with this measure to prove the proposition. Let $U \subseteq \mathbb{R}^n$ be a Borel set with $|U| = \delta^\gamma$ for some $\gamma \in [\theta, 1]$. By the definition of Assouad dimension there exists a uniform constant $C_1$, depending only on $F$ and $d$, such that $U$ intersects at most $C_1 ( \delta^\gamma/\delta)^d$ points in the support of $\mu_\delta$. Therefore \[ \mu_\delta (U) \leq C_1\delta^{(\gamma-1)d} C_0^{-1} \delta^{b} = C_1C_0^{-1} \ |U|^{(\gamma d - d+ b)/ \gamma} \leq C_1C_0^{-1} \ |U|^{(\theta d - d+ b)/ \theta} \] which, using Proposition \ref{mdp}, implies that \[ \da F \geq (\theta d - d+ b)/ \theta = d - \frac{d-b}{\theta}. \] Letting $d \to \dim_\textup{A} F$ and $b \to \underline{\dim}_\textup{B} F$ yields the desired result. \end{proof} This proposition implies that for bounded sets with $\dim_\textup{H} F < \underline{\dim}_\textup{B} F = \dim_\textup{A} F$, the intermediate dimensions $\da F$ and $\uda F$ are necessarily discontinuous at $\theta = 0$. In fact the intermediate dimensions are constant on $(0,1]$ in this case. On the other hand, this gives an alternative demonstration that $\da F$ and $\uda F$ are \emph{always} continuous at $\theta=1$. Moreover, the proposition provides a quantitative lower bound near $\theta=1$. In Section \ref{examplessimple} we will use Proposition \ref{assouad} to construct examples exhibiting a range of behaviours. \subsection{Product formulae} A well-studied problem in dimension theory is how dimensions of product sets behave. The following product formulae for intermediate dimensions may be of interest in their own right, but in Section \ref{examplessimple} they will be used to construct examples. \begin{prop} \label{products} Let $E \subseteq \mathbb{R}^n$ and $F \subseteq \mathbb{R}^m$ be bounded and $\theta \in [0,1]$. Then \[ \da E + \da F\ \leq\ \da (E \times F)\ \leq\ \overline{\dim}_\theta ( E \times F)\ \leq\ \overline{\dim}_\theta E + \overline{\dim}_\textup{B} F. \] \end{prop} \begin{proof} Fix $\theta \in (0,1)$ throughout, noting that the cases when $\theta=0,1$ are well-known, see \cite[Chapter 7]{Fa}. We begin by demonstrating the left-hand inequality. We may assume that $\da E , \da F >0$ as otherwise the conclusion follows by monotonicity. Moreover, since $E,F$ are bounded we may assume they are compact since all the dimensions considered are unchanged under taking closure. Let $0< s< \da E$ and $0<t< \da F$. It follows from Proposition \ref{frostman} that there exist constants $C_s,C_t > 0$ such that for all $\delta \in (0,1)$ there exist Borel probability measures $\mu_\delta$ supported on $E$ and $\nu_\delta$ supported on $F$ such that for all $x \in \mathbb{R}^n$ and $\delta^{1/\theta} \leq r \leq \delta$, \[ \mu_\delta(B(x,r)) \leq C_s r^s \qquad \text{and} \qquad \nu_\delta(B(x,r)) \leq C_t r^t. \] Consider the product measure $\mu_\delta \times \nu_\delta$ which is supported on $E \times F$. For $z \in \mathbb{R}^n \times \mathbb{R}^m$ and $\delta^{1/\theta} \leq r \leq \delta$, \[ (\mu_\delta \times \nu_\delta) (B(z,r)) \leq C_s C_t r^{s+t} \] and Proposition \ref{mdp} yields $\da (E \times F) \geq s+t$; letting $ s\to \da E$ and $t\to \da F$ gives the desired inequality. The middle inequality is trivial and so it remains to prove the right-hand inequality. Let $s > \overline{\dim}_\theta E$ and $d > \overline{\dim}_\textup{B} F$. From the definition of $\overline{\dim}_\textup{B} F$ there exists a constant $\delta_1 \in (0,1)$ such that for all $0<r<\delta_1$ there is a cover of $F$ by at most $r^{-d}$ sets of diameter $r$. Let $\varepsilon>0$. By the definition of $\overline{\dim}_\theta E$ there exists $\delta_0\in (0,\delta_1)$ such that for all $0<\delta < \delta_0$ there is a cover of $E$ by sets $\{U_i\}_{i}$ with $\delta^{1/\theta} \leq |U_i| \leq \delta$ for all $i$ and \[ \sum_{i} |U_i|^s \leq \varepsilon. \] Given such a cover of $E$, for each $i$ let $\{U_{i,j}\}_{j}$ be a cover of $F$ by at most $|U_i|^{-d}$ sets with diameters $|U_{i,j}|=|U_{i}|$ for all $j$. Then $$ E\times F\ \subseteq\ \bigcup_i \bigcup_j \big(U_i \times U_{i,j}\big),$$ with $$ \sum_i\sum_j |U_i \times U_{i,j}|^{s+d} \ \leq\ \sum_i|U_i|^{-d}\big( \sqrt{2}|U_i|\big)^{s+d} \ =\ 2^{(s+d)/2}\sum_i|U_i|^{s} \ \leq\ 2^{(s+d)/2}\varepsilon.$$ Since $\delta^{1/\theta} \leq |U_i \times U_{i,j}| \leq \sqrt{2}\delta$ for all $i,j$, each set $U_i \times U_{i,j}$ may be covered by at most $c_{n+m}$ sets $\{V_{i,j,k}\}_k$ with diameters $\delta^{1/\theta}\leq |V_{i,j,k}| \leq \min\{|U_i \times U_{i,j}|,\delta\}\leq \delta$, where $c_{q}$ is the least number such that every set in $\mathbb{R}^q$ of diameter $\sqrt{2}$ can be covered by at most $c_{q}$ sets of diameter 1. Hence $$\sum_i\sum_j \sum_k|V_{i,j,k}|^{s+d}\ \leq \ c_{n+m}2^{(s+d)/2}\varepsilon.$$ As $\varepsilon$ may be taken arbitrarily small, $\overline{\dim}_\theta (E \times F) \leq s+d$; letting $ s\to \uda E$ and $d\to \overline{\dim}_\textup{B} F$ completes the proof. \end{proof} \section{Examples}\label{sec:egs} \setcounter{equation}{0} \setcounter{theo}{0} In this section we construct several simple examples where the intermediate dimensions exhibit a range of phenomena. All of our examples are compact subsets of $\mathbb{R}$ or $\mathbb{R}^2$ and in all examples the upper and the lower intermediate dimensions coincide for all $\theta \in [0,1]$. \subsection{Convergent sequences}\label{secseq} Let $p>0$ and \begin{equation*} F_p = \bigg\{0, \frac{1}{1^p}, \frac{1}{2^p},\frac{1}{3^p},\ldots \bigg\}. \end{equation*} Since $F_p$ is countable, $\hdd F_p=0$. It is well-known that $\bdd F_p=1/(p+1)$, see \cite[Chapter 2]{Fa}. We obtain the intermediate dimensions of $F_p$. \begin{prop}\label{conseq} For $p>0$ and $0\leq \theta \leq 1$, $$\da F_p = \overline{\dim}_\theta F_p = \frac{\theta}{p+\theta}.$$ \end{prop} \begin{proof} We first bound $\overline{\dim}_\theta F_p$ above. Let $0<\delta<1$ and let $M =\lceil \delta^{-(s +\theta(1-s))/(p+1)}\rceil$. Write $B(x,r)$ for the closed interval (ball) of centre $x$ and length $2r$. Take a covering ${\mathcal U}$ of $F_p$ consisting of $M$ intervals $B(k^{-p}, \delta/2)$ of length $\delta$ for $1\leq k\leq M$ and $\lceil M^{-p} /\delta^\theta\rceil \leq M^{-p}/\delta^\theta +1$ intervals of length $\delta^\theta$ that cover $[0, M^{-p}].$ Then \begin{eqnarray*} \sum_{U\in {\mathcal U}} |U|^s &\leq & M\delta ^s + \delta^{\theta s}\Big(\frac{1}{M^p\delta^\theta}+ 1\Big)\\ &=& M\delta ^s + \frac{\delta^{\theta (s-1)}}{M^p}+ \delta^{\theta s}\\ & \leq & ( \delta^{-(s +\theta(1-s))/(p+1)} +1) \delta^s + \delta^{\theta (s-1)}\delta^{(s +\theta(1-s))p/(p+1)} +\delta^{\theta s}\\ & =& 2\delta^{(\theta (s-1)+sp)/(p+1)} + \delta^s+ \delta^{\theta s} \ \to \ 0 \end{eqnarray*} as $\delta \to 0$ if $s(\theta +p) > \theta$. Thus $\overline{\dim}_\theta F_p \leq\theta/(p+\theta)$. \medskip For the lower bound we put suitable measures on $F_p$ and apply Proposition \ref{mdp}. Fix $s = \theta/(p+\theta)$. Let $0<\delta < 1$ and again let $M =\lceil \delta^{-(s +\theta(1-s))/(p+1)}\rceil$. Define $\mu_\delta$ as the sum of point masses on the points $1/k^p \ (1\leq k<\infty)$ with \begin{equation}\label{mass} \mu_\delta \Big(\left\{\frac{1}{k^p}\right\}\Big)\ = \ \left\{ \begin{array}{cl} \delta^s & \mbox{ if } 1\leq k \leq M \\ 0 & \mbox{ if } M+1\leq k <\infty \end{array} \right. . \end{equation} Then \begin{eqnarray*} \mu_\delta(F_p) &=& M\delta^s \\ & \geq&\delta^{-(s +\theta(1-s))/(p+1)}\delta^s \\ & =& \delta^{(ps +\theta(s-1))/(p+1)} = 1 \end{eqnarray*} by the choice of $s$. To see that \eqref{mdiscond} is satisfied, note that if $2\leq k\leq M$ then, by a mean value theorem estimate, $$\frac{1}{(k-1)^p} - \frac{1}{k^p}\ \geq \ \frac{p}{k^{p+1}}\ \geq \ \frac{p}{M^{p+1}};$$ thus the gap between any two points of $F_p$ carrying mass is at least $p/M^{p+1}$. Let $U$ be such that $\delta \leq |U|\leq \delta^\theta$. Then $U$ intersects at most $1+ |U|/(p/M^{p+1}) = 1+|U|M^{p+1}/p $ of the points of $F_p$ which have mass $\delta^s$. Hence \begin{eqnarray*} \mu_\delta (U)& \leq & \delta^s + \frac{1}{p}|U|\delta^s \delta^{-(s +\theta(1-s))} \\ &=& \delta^s + \frac{1}{p}|U| \delta^{(\theta(s-1))}\\ &\leq& |U|^s + \frac{1}{p} |U||U|^{s-1} \\ &=& \left(1+ \frac{1}{p}\right)|U|^s. \end{eqnarray*} From Proposition \ref{mdp}, $\da F_p \geq s = \theta/(p+\theta)$. \end{proof} \subsection{Simple examples exhibiting different phenomena} \label{examplessimple} We use the convergent sequences $F_p$ from the previous section, the product formulae from Proposition \ref{products}, and that the \emph{upper} intermediate dimensions are finitely stable to construct examples displaying a range of different features which are illustrated in Figure 1. The natural question which began this investigation is `does $\da$ vary continuously between the Hausdorff and lower box dimension?'. This indeed happens for the convergent sequences considered in the previous section, but turns out to be false in general. The first example in this direction is provided by another convergent sequence. \medskip \begin{figure}[t] \centering \includegraphics[width=0.65\textwidth]{idfig4} \caption{Plots of $\da F$ for the four examples in Section \ref{examplessimple}} \end{figure} \emph{Example 1: Discontinuous at 0, otherwise constant.} Let \[ F_{\log} = \{ 0, 1/\log 2, 1/\log 3, 1/\log 4, \dots \}. \] This sequence converges slower than any of the polynomial sequences $F_p$ and it is well-known and easy to prove that $\bdd F_{\log} = \dim_\textup{A} F_{\log}= 1$. It follows from Proposition \ref{assouad} that \[ \da F_{\log} = \overline{\dim}_\theta F_{\log} = 1,\qquad \theta \in (0,1]. \] Since $\dim_0 F_{\log}=\hdd F_{\log}=0$ there is a discontinuity at $\theta = 0$. \medskip \emph{Example 2: Continuous at 0, part constant, part strictly increasing. } In the opposite direction, it is possible that $\da F = \hdd F < \lbd F$ for some $\theta >0$. Indeed, let $F =F_1 \cup E$ where $F_1 = \{0, 1/1, 1/2, 1/3, \dots\}$ as before, and let $E \subset \mathbb{R}$ be any compact set with $\hdd E = \ubd E = 1/3$ (for example an appropriately chosen self-similar set). Then it is straightforward to deduce that \[ \da F = \overline{\dim}_\theta F = \max\left\{ \frac{ \theta}{1+ \theta}, \, 1/3\right\},\qquad \theta \in [0,1]. \] It is also possible for $\da F$ to approach a value strictly in between $\hdd F$ and $\lbd F$ as $\theta \to 0$. This is the subject of the next two examples. \medskip \emph{Example 3: Discontinuous at 0, part constant, part strictly increasing.} For an example that is constant on an interval adjacent to 0, let $F =E \cup F_{1}$ where this time $E\subset \mathbb{R}$ is any closed countable set with $\lbd E = \dim_\textup{A} E = 1/4$. It is immediate that \[ \da F = \overline{\dim}_\theta F = \max\left\{ \frac{ \theta}{1+ \theta}, \, 1/4\right\},\qquad \theta \in (0,1], \] with $\hdd F=0$ and $\bdd F=1/2$. \medskip \emph{Example 4: Discontinuous at 0, strictly increasing.} Finally, for an example where $\da F$ is smooth, strictly increasing but not continuous at $\theta=0$, let \[ F = F_{1} \times F_{\log}\subset \mathbb{R}^2. \] Here $\hdd F=0$ and $\bdd F=3/2$ and Proposition \ref{products} gives \[ \da F = \overline{\dim}_\theta F = \frac{ \theta}{1 + \theta}+ 1,\qquad \theta \in (0,1], \] noting that $\da F_{\log} = \overline{\dim}_\theta F_{\log} = \lbd F_{\log} = \ubd F_{\log} = \dim_\textup{A} F_{\log}= 1$ for $\theta \in (0,1]$. \section{Bedford-McMullen carpets} \label{carpetssec} A well-known class of fractals where the Hausdorff and box dimensions differ are the self-affine carpets; this is a consequence of the alignment of the component rectangles in the iterated construction. The first studies of planar self-affine carpets were by Bedford \cite{bedford} and McMullen \cite{mcmullen} independently, see also \cite{peres} and these Bedford-McMullen carpets have been widely studied and generalised, see for example \cite{Fa1} and references therein. Indeed, the question of the distribution of scales of covering sets for Hausdorff and box dimensions of Bedford-McMullen carpets was one of our motivations for studying intermediate dimensions. Finding an exact formula for the intermediate dimensions of Bedford-McMullen carpets seems a difficult problem, so here we obtain some lower and upper bounds, which in particular establish continuity at $\theta=0$ and that the intermediate dimensions take a strict minimum at $\theta=0$. We introduce the notation we need, generally following that of McMullen \cite{mcmullen}. Choose integers $n > m \geq 2$. Let ${I}=\left\{0,\ldots, m-1 \right\}$ and ${J}=\left\{0,\ldots, n-1 \right\}$. Choose a fixed digit set $D \subseteq I \times J$ with at least two elements. For $\left( p,q \right)\in D$ we define the affine contraction $S_{\left(p,q \right)}\colon [0,1]^2 \rightarrow [0,1]^2$ by \[ S_{ \left( p,q \right)}\left(x,y\right) = \left( \frac{x+p}{m},\frac{y+q}{n} \right) . \] There exists a unique non-empty compact set $F \subseteq [0,1]^2$ satisfying \[ F=\bigcup_{(p,q) \in D}S_{ \left( p,q \right)}(F); \] that is $F$ is the attractor of the iterated function system $\ \left\{ S_{ \left(p,q \right)} \right\}_{ \left( p,q \right)\in D}$. We call such a set $F$ a {\em Bedford-McMullen self-affine carpet}. It is sometimes convenient to denote pairs in $D$ by $\ell = (p_\ell,q_\ell )$. We model our carpet $F$ via the symbolic space $D^{\mathbb{N}}$, which consists of all infinite words over $D$ and is equipped with the product topology. We write $\mathbf{i} \equiv (i_1,i_2,\ldots)$ for elements of $D^{\mathbb{N}}$ and $(i_1,\ldots,i_k)$ for words of length $k$ in $D^k$, where $i_j \in D$. Then the canonical projection $\tau :D^{\mathbb{N}} \rightarrow [0,1]^2$ is defined by \[ \{\tau(\mathbf{i})\}\equiv\{\tau(i_1,i_2,\ldots)\}=\bigcap_{k \in \mathbb{N}} S_{i_1} \circ \cdots \circ S_{i_k}([0,1]^2). \] This allows us to switch between symbolic and geometric notation since \[ \tau(D^{\mathbb{N}})=F. \] Bedford \cite{bedford} and McMullen \cite{mcmullen} showed that \be\label{dimb} \bdd F = \frac{\log m_0}{\log m} + \frac{\log |D| - \log m_0}{\log n} \ee where $m_0$ is the number of $p$ such that there is a $q$ with $(p,q) \in D$, that is the number of columns of the array containing at least one rectangle. They also showed that \be\label{dimh} \hdd F = \log \big(\sum_{p=1}^m n_p^{\log_n m} \big) \Big/ \log m, \ee where $n_p \, (1\leq p\leq m)$ is the number of $q$ such that $(p,q)\in D$, that is the number of selected rectangles in the $p$th column of the array. For each $\ell \in D$ we let $a_\ell$ be the number of rectangles of $D$ in the same column as $\ell$. Then, writing $d = \hdd F$, \eqref{dimh} may be written as \begin{equation} m^d \ = \ \sum_{p=1}^m {n}_p^{\log_n m}\ =\ \sum_{\ell\in D} a_{\ell}^{(\log_n m -1)}, \end{equation} where equality of the sums follows from the definitions of $n_p$ and $a_\ell$. We assume that the non-zero $n_p$ are not all equal, otherwise $\hdd F = \bdd F$; in particular this implies that $a := \max_{\ell \in D}a_\ell \geq 2$. We denote the $k$th-level iterated rectangles by $$ R_k (i_1,\ldots,i_k)\ =\ S_{i_1}\circ \cdots \circ S_{i_k}([0,1]^2).$$ We also write $R_k(\mathbf{i}) \equiv R_k(i_1,i_2, \ldots)$ for this rectangle when we wish to indicate the $k$th-level iterated rectangle containing the point $\tau(\mathbf{i}) = \tau(i_1,i_2, \ldots)$. We will associate a probability vector $\ \left\{ b_{\ell} \right\}_{\ell\in D}$ with $D$ and let $\widetilde{\mu}=\prod_{k\in\mathbb{ N}}\left( \sum_{ i_k\in D}b_{ i_k}\delta_{ i_k}\right)$ be the natural Borel product probability measure on $D^{\mathbb{N}}$, where $\delta_{ \ell}$ is the Dirac measure on $D$ concentrated at $ \ell$. Then the measure \[ \mu=\widetilde{\mu}\circ \tau^{-1} \] is a self-affine measure supported on $F$. Following McMullen \cite{mcmullen} we set \be\label{pi} b_\ell = a_\ell^{(\log_n m -1)}/m^d\qquad (\ell \in D), \ee noting that $\sum_{\ell\in D} b_\ell =1$, to get a measure $\mu$ on $F$; thus the measures of the iterated rectangles are \be\label{rect} \mu\big( R_k (i_1,\ldots,i_k)\big)\ =\ b_{i_1}\cdots b_{i_k}\ =\ m^{-kd} ( a_{i_1} \cdots a_{i_k})^{(\log_n m -1)}. \ee Approximate squares are well-known tools in the study of self-affine carpets. Given $k\in \mathbb{N}$ let $l(k) = \lfloor k \log_n m\rfloor $ so that \be\label{lkbounds} k\log_n m \ \leq\ l(k) \ \leq k\log_n m \ + \ 1 \ee For such $k$ and $\mathbf{i} = (i_1,i_2, \ldots) \in D^{\mathbb{N}}$ we define the {\em approximate square} containing $\tau(\mathbf{i})$ as the union of $m^{-k}\times n^{-k}$ rectangles: \begin{align*} Q_k(\mathbf{i}) = Q_k(i_1,i_2, \ldots) = \bigcup\Big\{ &R_k(i_1',\ldots,i_k'): p_{i_{j}'}=p_{i_{j}} \text{ for } j= 1, \ldots, k \\ & \text{ and } q_{i_{j}'}=q_{i_{j}} \text{ for } j= 1, \ldots, l(k) \ \Big\}, \end{align*} recalling that $\ell = (p_\ell,q_\ell)$. This approximate square has sides $m^{-k}\times n^{-l(k)}$ where $n^{-1}m^{-k} \leq n^{-l(k)}\leq m^{-k}$. Note that, by virtue of self-affinity and since the sequence $(p_{i_1},\cdots p_{i_k})$ is the same for all level-$k$ rectangles $R_k (i_1,\ldots,i_k)$ in the same approximate square, the $a_{i_{l(k)+1}}\cdots a_{i_k}$ level-$k$ rectangles that comprise the approximate square $Q_k(\mathbf{i})$ all have equal $\mu$-measure. Thus, writing $L= \log_n\! m$, \begin{eqnarray} \mu(Q_k(\mathbf{i})) &=& m^{-kd} a_{i_1}^{(L -1)}\cdots a_{i_k}^{(L -1)} \times a_{i_{l(k)+1}}\cdots a_{i_k}\label{squaremes1}\\ &=& m^{-kd} a_{i_1}^{L}\cdots a_{i_{k}}^{L} \times a_{i_{1}}^{-1}\cdots a_{i_{l(k)}}^{-1}.\label{squaremes} \end{eqnarray} We now obtain an upper bound for $\overline{\dim}_\theta F$ which implies continuity at $\theta = 0$ and so on $[0,1]$. Recall that $a = \max_{\ell \in D}a_\ell \geq 2$. \begin{prop}\label{propup} Let $F$ be the Bedford-McMullen carpet as above. Then for $0<\theta <\frac{1}{4}(\log_n\! m)^2$, \be\label{upbound} \overline{\dim}_\theta F \ \leq \ \hdd F + \bigg(\frac{2\log(\log_m n)\log a}{\log n}\bigg) \frac{1}{-\log \theta}. \ee In particular, $ \underline{\dim}_\theta F$ and $ \overline{\dim}_\theta F$ are continuous at $\theta = 0$ and so are continuous on $[0,1]$. \end{prop} \begin{proof} For $\mathbf{i} = (i_1,i_2,\ldots)$, rewriting \eqref{squaremes} gives \be \mu\big(Q_k(\mathbf{i})\big) = m^{-kd} \bigg(\frac{ (a_{i_1}\cdots a_{i_k})^{1/k}}{(a_{i_1}\cdots a_{i_{l(k)}})^{1/l(k)}}\bigg)^{Lk} \big(a_{i_1}\cdots a_{i_{l(k)}}\big)^{(kL/l(k))-1}. \label{rewritemes} \ee We consider the two bracketed terms on the right-hand side of \eqref{rewritemes} in turn. We show that the first term cannot be too small for too many consecutive $k$ and that the second term is bounded below by $1$. For $k>1$ with $l(k) =\lfloor k \log_n m\rfloor$ as usual, define \be\label{fkstar} f_k(\mathbf{i})\ \equiv\ f_k(i_1,i_2,\ldots) \ =\ \bigg(\frac{ (a_{i_1}\cdots a_{i_k})^{1/k}}{(a_{i_1}\cdots a_{i_{l(k)}})^{1/l(k)}}\bigg) \ee We claim that for all $K\geq L/(1-L)$ and all $\mathbf{i} = (i_1,i_2,\ldots) \in D^{\mathbb{N}}$, there exists $k$ with $K\leq k \leq K/ \theta$ such that \be\label{claim} f_k(\mathbf{i})\ \geq \ a^{\log L/\log(L/2\theta)}. \ee Suppose this is false for some $(i_1,i_2,\ldots)$ and $K\geq 1/L(L-1)$, so for all $K\leq k \leq K/ \theta$, $ f_k(\mathbf{i}) < \lambda:= a^{\log L/\log(L/2\theta)}$, that is \be\label{basin} (a_{i_1} a_{i_2}\cdots a_{i_k})^{1/k}\ <\ \lambda (a_{i_1} a_{i_2}\cdots a_{i_l})^{1/l(k)}. \ee Define a sequence of integers $k_{r}\, (r=0,1,2,\ldots)$ inductively by $k_{0} = K$ and for $r \geq 1$ taking $k_{r}$ to be the least integer such that $ \lfloor k_{r} \log_n m \rfloor = k_{r-1}$. Then $k_{r} \leq k_{r-1}/L +1$, and a simple induction shows that $$k_{r}\ \leq \ L^{-r}\big(K + L/(1-L)\big)-L/(1-L) \ \leq\ L^{-r}\big(K + L/(1-L)\big)\qquad (r\geq 0).$$ Fix $N$ to be the greatest integer such that $k_N \leq K/\theta$. Then $$ L^{-(N+1)} \big(K+L\big/(1 -L)\big)\ \geq\ k_{N+1}\ >\ K/\theta$$ so rearranging, provided $K\geq L/(1-L)$, $$L^N\ <\ \frac{\theta\big(K + L\big/(1-L)\big)}{LK}\ \leq \ \frac{2\theta}{L}, $$ that is \be\label{ineq1} N > \frac{\log(2\theta/L)}{\log L}. \ee From \eqref{basin} $$ (a_{i_1} a_{i_2}\cdots a_{i_{k_r}})^{1/{k_r}}\ <\ \lambda (a_{i_1} a_{i_2}\cdots a_{i_{k_{r-1}}})^{1/k_{r-1}} \qquad (1\leq r\leq N) $$ so iterating $$ 1\ \leq\ (a_{i_1} a_{i_2}\cdots a_{i_{k_N}})^{1/k_N}\ <\ \lambda^{N} (a_{i_1} a_{i_2}\cdots a_{i_{K}})^{1/K}\ \leq \ \lambda^{N}a. $$ Combining with \eqref{ineq1} we obtain $$ \lambda \ >\ a^{-1/N} \geq a^{\log L/\log(L/2\theta)} $$ which contradicts the definition of $\lambda$. Thus the claim \eqref{claim} is established. For the second bracket on the right-hand side of \eqref{rewritemes} note that $$ 0\ \leq\ (kL/l(k))-1\ =\ \frac{ k\log_n m -\lfloor k\log_n m \rfloor}{l(k)}\ \leq \frac{1}{l(k)}$$ so that \be\label{brac2} 1\ \leq \ \big(a_{i_1}\cdots a_{i_{l(k)}}\big)^{(kL/l(k))-1}\ \leq a. \ee Putting together \eqref{rewritemes}, \eqref{claim} and \eqref{brac2}, we conclude that for all $K\geq L/(1-L)$ and all $\mathbf{i} \in D^{\mathbb{N}}$ there exists $ K\leq k \leq K/ \theta $ such that $$\mu\big(Q_k(\mathbf{i})\big) \ \geq\ m^{-dk} a^{kL\log L/\log(L/2\theta)} \ \geq\ m^{-dk} a^{2kL\log L/\log(1/\theta)} \ =\ m^{-k(d+\epsilon(\theta))},$$ as $\theta \leq L^2/4$, where $$\epsilon(\theta)\ =\ \frac{ -(\log a)\, 2L\log L}{\log m \log(1/\theta)} \ =\ \frac{2 \log(\log_m n)\log a}{\log n} \frac{1}{-\log \theta}.$$ Geometrically this means that for $K\geq L/(1-L)$ every point $z \in F$ belongs to at least one approximate square, $Q_{k(z)}$ say, with $K \leq {k(z)} \leq K/\theta$ and with $\mu(Q_{k(z)}) \geq m^{-k(d+\epsilon(\theta))}$. Since the approximate squares form a nested hierarchy we may choose a subset $\{Q_{k(z_n)}\}_{n=1}^N \subset \{ Q_{k(z)}: z\in F\}$ that is disjoint (except possibly at boundaries of approximate squares) and which cover $F$. Thus $$1\ =\ \mu(F)\ = \ \sum_{n=1}^N \mu (Q_{k(z_n)})\ \geq\ \sum_{n=1}^N m^{-k(z_n)(d+\epsilon(\theta))} \ \geq\ \sum_{n=1}^N (2^{-1/2}|Q_{k(z_n)}|)^{(d+\epsilon(\theta))} $$ where $| Q_k|$ denotes the diameter of the approximate square $Q_k$, noting that $| Q_k|\leq 2^{1/2} m^{-k}$. It follows that $ \overline{\dim}_\theta F \leq d+\epsilon(\theta)$ as claimed. \end{proof} The following lemma brings together some basic estimates that we will need to obtain a lower bound for the intermediate dimensions of $F$. \begin{lem}\label{mcmass} Let $\epsilon >0$. There exists $K_0\in \mathbb{N}$ and a set $E\subset F$ with $\mu(E) \geq \frac{1}{2}$ such that for all $\mathbf{i}$ with $\tau(\mathbf{i})\in E$ and $k\geq K_0$, \begin{equation}\label{uppermass} \mu(Q_k(\mathbf{i}))\leq m^{-k(d-\epsilon)} \end{equation} and \begin{equation}\label{lowermass} \mu(R_k(\mathbf{i}))\geq \exp(-k(H(\mu)+\epsilon)), \end{equation} where $d=\dim_\textup{H} F$ and $H(\mu) \in (0,\log |D|)$ is the entropy of the measure $\mu$. \end{lem} \begin{proof} McMullen \cite[Lemmas 3,4(a)]{mcmullen} shows that for $\widetilde{\mu}$-almost all $\mathbf{i} \in D^{\mathbb{N}}$ $$\lim_{k\to\infty} \mu(Q_k(\mathbf{i}))^{1/k} \to m^{-d}.$$ Thus by Egorov's theorem we may find a set $\widetilde{E}_1\subset D^{\mathbb N}$ with $\widetilde{\mu}(\widetilde{E}_1) \geq \frac{3}{4}$, and $K_1\in \mathbb{N}$ such that \eqref{uppermass} holds for all $\mathbf{i} \in \widetilde{E}_1$ and $k\geq K_1$. Furthermore, it is immediate from the Shannon-MacMillan-Breimann Theorem and \eqref{rect} that for $\widetilde{\mu}$-almost all $\mathbf{i} \in D^{\mathbb{N}}$, $$\lim_{k\to\infty} \mu(R_k(\mathbf{i})))^{1/k} \to \exp(-H(\mu)),$$ and again by Egorov's theorem there is a set $\widetilde{E}_2\subset D^{\mathbb N}$ with $\widetilde{\mu}(\widetilde{E}_2) \geq \frac{3}{4}$, and $K_2\in \mathbb{N}$ such that \eqref{lowermass} holds for all $\mathbf{i} \in \widetilde{E}_2$ and $k\geq K_2$. The conclusion of the lemma follows taking $E =\tau(\widetilde{E}_1\cap \widetilde{E}_2)$ and $K_0 = \max\{K_1,K_2\}$. \end{proof} We now obtain a lower bound for $\underline{\dim}_{\theta}F$, showing in particular that $\underline{\dim}_{\theta}F> \hdd F$ for all $\theta>0$. \begin{prop}\label{proplb} Let $F$ be the Bedford-McMullen carpet as above. Then for $0\leq \theta \leq 1$, \begin{equation}\label{lowerbnd} \underline{\dim}_{\theta}F\geq \hdd F+\theta\frac{\log |D|-H(\mu)}{\log m}. \end{equation} \end{prop} \begin{proof} Fix $\theta\in(0,1)$, let $E\subset F$ and $K_0$ be given by Lemma \ref{mcmass}, and let $K\geq K_0$. We define a measure $\nu_K$ which assigns equal mass to all level-$K$ rectangles, and then subdivides this mass among sub-rectangles using the same weights as for the measure $\mu$, given by \eqref{pi}. This gives a measure to which we can apply the mass distribution principle, Lemma \ref{mdp}. Thus for $k\geq K$, writing $b_\ell = a_\ell^{(\log_n m -1)}/m^d$ as in \eqref{pi}, \begin{equation}\label{defnuk} \nu_K\big( R_k (i_1,\ldots,i_k)\big)\ :=\ |D|^{-K}b_{i_{K+1}}\cdots b_{i_k}\ =\ |D|^{-K} m^{-(k-K)d} ( a_{i_{K+1}} \cdots a_{i_k})^{L-1}. \end{equation} We now consider an approximate square $Q_k(\mathbf{i})$ containing the point $\mathbf{i}$. This approximate square is a union of rectangles $R_k(\mathbf j)$ which, as explained in our comment before \eqref{squaremes1}, each have the same $\mu$ measure equal to $\mu(R_k(\mathbf i))$. The same argument gives that for any $R_k(\mathbf j)\subset Q_k(\mathbf i)$ \[ \nu_K(R_k(\mathbf j))=\nu_K(R_k(\mathbf i))=\mu(R_k(\mathbf i)) \dfrac{|D|^{-K}}{\mu(R_K(\mathbf i))} \] where the final equality holds since the formula for $\nu_{K}$ differs from that of $\mu$ only in the mass it assigns according to the first $K$ letters. Putting this together allows one to express the $\nu_{K}$-measure of an approximate square of side length $m^{-k}$ in relation to the $\mu$-measure of such a square. For $\tau(\mathbf{i})\in E$ and $k\geq K$, the approximate square $Q_k(\mathbf{i})$ that contains the point $\mathbf{i}$ has $\nu_K$-measure \begin{eqnarray} \nu_{K}(Q_k(\mathbf{i}))&=&\dfrac{|D|^{-K}}{\mu(R_K(\mathbf{i}))}\mu(Q_k(\mathbf{i}))\label{nukmu}\\ &\leq& \dfrac{|D|^{-K} m^{-k(d-\epsilon)}}{\exp(-K(H(\mu)+\epsilon))},\nonumber \end{eqnarray} using Lemma \ref{mcmass}. (Alternatively \eqref{nukmu} may be verified directly using \eqref{rect}, \eqref{squaremes} and \eqref{defnuk}.) Then \begin{eqnarray} \nu_{K}(Q_k(\mathbf{i}))& \leq & m^{-k(d-\epsilon) -K \log_m \big(|D|\exp(-H(\mu)-\epsilon)\big)}\nonumber\\ &\leq& m^{-k\big(d-\epsilon +\frac{K}{k}(\log |D|-H(\mu)-\epsilon)/\log m\big)}\label{inequal}. \end{eqnarray} We need bounds that are valid for all $k\in[K,K/\theta]$ corresponding to approximate squares of sides between (approximately) $m^{-K}$ and $m^{-K/\theta}$. The exponent in \eqref{inequal} is maximised when $k = K/\theta$, so that \begin{equation}\label{munu}\nu_{K}(Q_k(\mathbf{i}))\ \leq \ m^{-k\big(d-\epsilon+\theta(\log |D|-H(\mu)-\epsilon)/\log m\big)}\end{equation} for all $\mathbf{i}$ with $\tau(\mathbf{i})\in E$ and integers $k\in[K,K/\theta]$, where $\mu (E) \geq \frac{1}{2}$. To use our mass distribution principle we need equation \eqref{munu} to hold on a set of $\mathbf{i}$ of large $\nu_K$ mass, whereas currently we have that it holds on a set $E$ of large $\mu$ mass. Let $$E'=\tau\{\mathbf{i} : \mbox{inequality \eqref{munu} is satisfied for all } k\in[K,K/\theta] \}.$$ Firstly we observe that $Q_k(\mathbf{i})$ depends only on $(i_1,\ldots i_k)$, and we are dealing with $k\leq K/\theta$, so the question of whether $\tau(\mathbf{i})\in E'$ is independent of $(i_{\lfloor K/\theta\rfloor+1}, i_{\lfloor K/\theta\rfloor+2},\ldots)$. Secondly, since $Q_k(\mathbf{i})$ is a union of rectangles $R_k(\mathbf{i})$, and $\nu_K\big( R_k (i_1,i_2,\ldots))$ is independent of $(i_1,\ldots ,i_K)$, the question of whether $\tau(\mathbf{i})\in E'$ is independent of $(i_1,\ldots ,i_K)$. Thus we can write \[ E'=\bigcup_{i_{K+1}\ldots i_{\lfloor K/\theta\rfloor}\in I''} \bigg(\bigcup_{i_1\ldots i_K\in D^K} R_{\lfloor K/\theta\rfloor}(i_1,\ldots,i_{\lfloor K/\theta\rfloor})\bigg) \] for some set $I''\subset D^{\lfloor K/\theta\rfloor-K}$. But using \eqref{defnuk} gives \begin{eqnarray*} \nu_K \bigg(\bigcup_{i_1\ldots i_K\in D^K} R_{\lfloor K/\theta\rfloor}(i_1,\ldots, i_{\lfloor K/\theta\rfloor})\bigg)&=& \sum_{i_1\ldots i_K\in D^K}\frac{1}{|D|^K}b_{i_{K+1}}\cdots b_{i_{\lfloor K/\theta\rfloor}}\\ &=& b_{i_{K+1}}\cdots b_{i_{\lfloor K/\theta\rfloor}} \end{eqnarray*} and \eqref{rect} gives \begin{eqnarray*} \mu \bigg(\bigcup_{i_1\ldots i_K\in D^K} R_{\lfloor K/\theta\rfloor}(i_1,\ldots, i_{\lfloor K/\theta\rfloor})\bigg)&=& \sum_{i_1\ldots i_K\in D^K}b_{i_1}\cdots b_{i_K}b_{i_{K+1}}\cdots b_{i_{\lfloor K/\theta\rfloor}}\\ &=& \bigg(\sum_{i_1\ldots i_K\in D^K}b_{i_1}\cdots b_{i_K}\bigg)b_{i_{K+1}}\cdots b_{i_{\lfloor K/\theta\rfloor}}\\ &=& b_{i_{K+1}}\cdots b_{i_{\lfloor K/\theta\rfloor}}, \end{eqnarray*} as $\sum_{i\in D} b_i =1$. Since these quantities are equal we conclude that \[ \nu_K(E')\ =\ \mu(E')\ \geq\ \mu(E)\ \geq\ {\textstyle \frac{1}{2}} \] as required. Since \eqref{munu} holds for all $\mathbf{i} \in E'$, a straightforward variant on our mass distribution principle, where we use approximate squares instead of balls, gives \[ \underline{\dim}_{\theta}F\ \geq\ \hdd F-\epsilon +\theta\frac{\log |D|-H(\mu)-\epsilon}{\log m}. \] Since $\epsilon$ can be chosen arbitrarily small, \eqref{lowerbnd} follows. \end{proof} Note that, in \eqref{lowerbnd}, \[ H(\mu)\ =\ -m^{-d}\sum_{\ell\in D }(a_\ell^{L-1}\big((L-1)\log a_\ell - d\log m\big)\ \leq\ \log | D| \] with equality if and only if $\mu$ gives equal mass to all cylinders of the same length, which happens if and only if each column in our construction contains the same number of rectangles. This happens exactly when the box and Hausdorff dimension coincide. Thus our lower bounds give that $\underline{\dim}_{\theta}F> \hdd F$ whenever $\theta>0$, provided that the Hausdorff and box dimensions of $F$ are different. Since the measures that we have constructed to give lower bounds on $\underline{\dim}_{\theta}F$ are rather crude, it is unlikely that our lower bound for $\underline{\dim}_{\theta}F$ converges to ${\dim}_\textup{B} F$ as $\theta\to 1$. However, a lower bound which \emph{does} approach ${\dim}_\textup{B} F$ as $\theta\to 1$ is given by Proposition \ref{assouad}, noting that $\dim_\textup{A} F > \underline{\dim}_\textup{B} F = {\dim}_\textup{B} F$ provided ${\dim}_\textup{B} F > {\dim}_\textup{H} F$, see \cite{mackay,Fra}. Many questions on the intermediate dimensions of Bedford-McMullen carpets remain, most notably finding the exact forms of $\underline{\dim}_{\theta}F$ and $\overline{\dim}_{\theta}F$. In that direction we would at least conjecture that these intermediate dimensions are equal and strictly monotonic. One might hope to get better estimates using alternative definitions of $\mu$ in Proposition \ref{propup} and $\nu_K$ in Proposition \ref{proplb}, but McMullen's measure and our modifications seemed to work best when optimising mass distribution type estimates across $F$ and over the required range of scales. \section*{Acknowledgements} JMF was financially supported by a \emph{Leverhulme Trust Research Fellowship} (RF-2016-500) and KJF and JMF were financially supported in part by an \emph{EPSRC Standard Grant} (EP/R015104/1). We thank James Robinson for pointing out the reference \cite{KP}. \\ \bibliographystyle{plain}
{ "timestamp": "2019-11-07T02:15:12", "yymm": "1811", "arxiv_id": "1811.06493", "language": "en", "url": "https://arxiv.org/abs/1811.06493", "abstract": "We introduce a continuum of dimensions which are `intermediate' between the familiar Hausdorff and box dimensions. This is done by restricting the families of allowable covers in the definition of Hausdorff dimension by insisting that $|U| \\leq |V|^\\theta$ for all sets $U, V$ used in a particular cover, where $\\theta \\in [0,1]$ is a parameter. Thus, when $\\theta=1$ only covers using sets of the same size are allowable, and we recover the box dimensions, and when $\\theta=0$ there are no restrictions, and we recover Hausdorff dimension.We investigate many properties of the intermediate dimension (as a function of $\\theta$), including proving that it is continuous on $(0,1]$ but not necessarily continuous at $0$, as well as establishing appropriate analogues of the mass distribution principle, Frostman's lemma, and the dimension formulae for products. We also compute, or estimate, the intermediate dimensions of some familiar sets, including sequences formed by negative powers of integers, and Bedford-McMullen carpets.", "subjects": "Metric Geometry (math.MG); Classical Analysis and ODEs (math.CA); Dynamical Systems (math.DS)", "title": "Intermediate dimensions", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9854964186047326, "lm_q2_score": 0.8128673246376008, "lm_q1q2_score": 0.8010778372311661 }
https://arxiv.org/abs/math/9806076
On the volume of the polytope of doubly stochastic matrices
We study the calculation of the volume of the polytope B_n of n by n doubly stochastic matrices; that is, the set of real non-negative matrices with all row and column sums equal to one.We describe two methods. The first involves a decomposition of the polytope into simplices. The second involves the enumeration of ``magic squares'', i.e., n by n non-negative integer matrices whose rows and columns all sum to the same integer.We have used the first method to confirm the previously known values through n=7. This method can also be used to compute the volumes of faces of B_n. For example, we have observed that the volume of a particular face of B_n appears to be a product of Catalan numbers.We have used the second method to find the volume for n=8, which we believe was not previously known.
\section{Introduction} \label{sec:intro} We study the calculation of the volume of the polytope $B_n$ of $n \times n$ doubly stochastic matrices; that is, the set of real nonnegative matrices with all row and column sums equal to one. This polytope is sometimes known as the Birkhoff polytope or the assignment polytope. We will describe and evaluate two methods for computing the volume of $B_n$. In the first method we decompose $B_n$ into a disjoint union of simplices all of the same volume and count the simplices. The fact that this can be done appears in \cite{St1}. This method applies to any face of $B_n$ as well. In the second method we count the number of $n \times n$ nonnegative integer matrices with all row and column sums equal to $t$ (sometimes called magic squares) for suitable values of $t$. These numbers allow us to compute the Ehrhart polynomial of $B_n$, which (essentially) has the volume of $B_n$ as its leading coefficient. It appears that this has been the most common method of computing the volume of $B_n$. Sturmfels reports in \cite{Stu} on other work in which the volume of $B_n$ has been computed for $n$ up to 7. We have also used this method to compute the volume when $n=8$. This study is largely expository since the two methods are not new. However, the details about how we carry out these methods may be of interest. We are not aware of any reports of others who have carried out the simplicial decomposition method. As a byproduct of our program for carrying out the simplicial decomposition method, we are easily able to compute the volume of any face (of any dimension) of $B_n$ provided that $n$ is not too large. This allowed us to discover that a certain special face of $B_n$ has a volume which appears to be given by a simple product formula. This formula is given in Conjecture \ref{conj:1}. Our study resulted from a question \cite{Mi} of Victor Miller, who asked how one could generate a doubly stochastic matrix uniformly at random. It is not hard to see that it would be easy to generate a random doubly stochastic matrix if one could easily calculate the volume of any face of $B_n$. However the method described here for calculating face volumes is practical only for small $n$. In what follows we will make use of some well known properties of the face structure of $B_n$: the vertices of $B_n$ are precisely the $n!$ $n \times n$ permutation matrices; on the other hand, for each pair $(i,j)$ with $1 \le i,j \le n$, the doubly stochastic matrices with $(i,j)$ entry equal to 0 form a {\it facet\/}\ (maximal proper face) of $B_n$ and all facets arise in this way. See \cite{BS} for further properties and references. In general, it is convenient to identify the faces of $B_n$ with certain $n \times n$ matrices of 0's and 1's, as follows. First we identify a 0-1 matrix with the set of entries in the matrix that are 1's. Thus, for two 0-1 matrices $A$ and $B$ of the same size, we can define their union $A \cup B$ as the 0-1 matrix whose set of 1's is the union of the sets of 1's of $A$ and $B$. {\it e.g.,\/}\ \[ \left(\begin{array}{ccc} 1 & 0 & 0 \\ 0 & 1 & 0 \\ 0 & 0 & 1 \end{array}\right) \; \cup \; \left(\begin{array}{ccc} 0 & 1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 1 \end{array}\right) \; = \; \left(\begin{array}{ccc} 1 & 1 & 0 \\ 1 & 1 & 0 \\ 0 & 0 & 1 \end{array}\right). \] Similarly we can speak of one 0-1 matrix containing another and so forth. Now to each face $F$ of $B_n$, we associate the matrix $M$ which is the union of the vertices (permutation matrices) in $F$. The facets of $B_n$ containing $F$ are precisely those associated with the zero entries of $M$. Since any face is the intersection of the facets containing it, any permutation matrix contained in $M$ must be a vertex of $F$. Thus the vertices of $F$ are precisely the permutation matrices contained in $M$, so we can recover $F$ from $M$. In this way we identify the faces of $B_n$ with the set of 0-1 matrices which are unions of permutation matrices. Note that not every 0-1 matrix corresponds to a face of $B_n$. For example \[ \left(\begin{array}{cc} 0 & 1 \\ 1 & 1 \end{array}\right) \] is not a union of permutation matrices, hence not a face of $B_2$. \section{Volume} \label{sec:2} It is easy to see that the dimension of $B_n$ is $(n-1)^2$. Strictly speaking, the volume we wish to compute is the $(n-1)^2$-volume of $B_n$ regarded as a subset of $n^2$-dimensional Euclidean space. Thus, for example, the polytope $B_2$ consists of the line segment joining the matrices \[ \left(\begin{array}{cc} 1 & 0 \\ 0 & 1 \end{array}\right) \quad\hbox{and} \quad \left(\begin{array}{cc} 0 & 1 \\ 1 & 0 \end{array}\right) \] and hence its volume is 2. An $n \times n$ doubly stochastic matrix is determined by its upper left $(n-1)\times (n-1)$ submatrix. The set of $(n-1)\times (n-1)$ matrices obtained this way is the set $A_n$ of all nonnegative $(n-1)\times (n-1)$ matrices with row and column sums $\le 1$ such that the sum of all the entries is at least $n-2$. This is affinely isomorphic to $B_n$. In the Appendix we show that the ratio of the volume of $B_n$ to the volume of $A_n$, regarded as a subset of Euclidean $(n-1)^2$ space, is $n^{n-1}$. In some ways the volume of $A_n$ is easier to understand since its dimension is equal to the dimension of its ambient space. James Maiorana \cite{Ma} (and probably others) noted a Monte Carlo method for approximating the volume of $A_n$. Consider the set $C_n$ of $(n-1)\times (n-1)$ nonnegative matrices with row sums (but not necessarily column sums) $\le 1$. This is the Cartesian product of $n-1$ unit simplices in Euclidean $(n-1)$-space so its volume is $ \frac{1}{(n-1)!^{n-1}}$. It is easy to choose points in $C_n$ uniformly at random. The probability $\alpha_n$ that such a point is in $A_n$ is the ratio of the volume of $A_n$ to that of $C_n$. Thus we can run Monte-Carlo trials to estimate $\alpha_n$ and hence the volume of $A_n$. For large $n$, this Monte Carlo method is impractical since $\alpha_n$ is too small. However, it is useful for checking computations for small $n$. A lower bound for $\alpha_n$ is given by Bona in \cite{Bo}. There is a more natural unit for the volume of $B_n$ and its faces. This is based on the fact that the vertices of $B_n$ are integer matrices. Suppose that $F$ is a $d$-dimensional face of $B_n$. Since its vertices have integer coordinates, the integer points in the affine span of $F$ comprise a $d$-dimensional affine lattice $L$. Given such a lattice there is a minimum volume of any $d$-simplex with vertices in $L$. Lattice points $w_0,\dots,w_d$ are the vertices of one of these minimum volume simplices if and only if every point of $L$ is uniquely expressible in the form $ \sum_{i=0}^d k_i w_i$, where the $k_i$'s are integers whose sum is 1. The {\it relative volume} of a face $F$ is the volume of $F$ expressed in units equal to the volume of a minimal simplex in $L$. The relative volume of a face is the same whether regarded as a face of $B_n$ or as a face of $A_n$, since the mapping from $B_n$ to $A_n$ (by taking the upper left $(n-1) \times (n-1)$ minor) preserves integrality of points. Here are the currently known relative volumes of $B_n$. \[ \begin{array}{cr} n & \mbox{Relative Volume of $B_n$} \\ 1 & 1\\ 2 & 1\\ 3 & 3\\ 4 & 352\\ 5 & 4718075\\ 6 & 14666561365176\\ 7 & 17832560768358341943028\\ 8 & 12816077964079346687829905128694016 \end{array} \] To convert relative volumes to true volumes, we need to know the volume of a minimal simplex of $A_n$. But the affine span of $A_n$ is all of $(n-1)^2$-dimensional space. Hence the volume of a minimal simplex in $A_n$ is $\frac{1}{((n-1)^2)!}$, and the volume of a minimal simplex in $B_n$ is $\frac{n^{n-1}}{((n-1)^2)!}$. \section{Triangulations} \label{sec:3} We call the first method for computing the volume of $B_n$ the {\it triangulation method.\/}\ The method applies to the calculation of the volume of any polytope. The essence is that we decompose the polytope into simplices and sum the volumes of the simplices. For $B_n$ we have used a standard method of decomposing a polytope $P$ into simplices. See for example \cite{St1}. To decompose $P$ into simplices, we choose an arbitrary vertex $v$ and form the collection of facets of $P$ opposite $v$ (facets of $P$ not containing $v$.) We then recursively triangulate each facet. The triangulation of $P$ is then formed by adding our chosen vertex to each simplex in the triangulation of each of the facets. The standard triangulations of $B_n$ and its faces have an unusual property, given in \cite{St1}, for which we provide a self-contained proof below. \begin{proposition} In any standard triangulation of a face $F$ of $B_n$, every simplex has minimal volume in the affine lattice determined by $F$. \end{proposition} {\it Proof:\/}\ Let $F$ be a $d$-dimensional face of $B_n$, $v_0$ any vertex in $F$, and $G$ a facet of $F$ opposite $v_0$. Suppose that a simplex in a standard triangulation of $G$ has vertices $v_1,v_2,\dots,v_d$. We need to prove that the set of integer points of the affine space determined by $F$ is the same as the set of points $\sum_{i=0}^{d} k_i v_i $, where the $k_i$ are integers whose sum is 1. Of course all the integer combinations are in the affine span. The question is whether there are any other points. Any integral point of the affine span can be uniquely expressed in the form $\sum_{i=0}^{d} r_i v_i $, where the $r_i$'s are real numbers with sum 1. Since $v_0$ is not in the face $G$, there is a facet of $B_n$ containing $G$ but not $v_0$. Thus $v_0$ must have at least one entry equal to $1$ in the same position where all $v_i$, $i\ge 1$, have zeroes. Thus, in the hypothetical combination above, $r_0$ must be an integer. If we add $r_0(v_1-v_0)$ to the combination above, we obtain another integral point in the affine span of $G$. It follows, using induction, that $r_1+r_0$, and $r_2,\dots,r_d$ are integers and therefore all the $r$'s are integers, as desired.\\ Note that, as a corollary, in any standard triangulation of a face of the $B_n$, the number of simplices in the triangulation is equal to the relative volume. We also obtain an important computational principle. Given a face $F$ of $B_n$ and a vertex $v$ of $F$, the relative volume of $F$ is the sum of the relative volumes of facets of $F$ opposite $v$. \section{The Triangulation Method for $B_n$} \label{sec:4} We now describe the triangulation method for computing the volume of $B_n$. This is simply an elaboration of the principle that the relative volume of a face is the sum of the relative volumes of the faces opposite any vertex. We apply this principle recursively. To get started we use the fact that the relative volume of any zero-dimensional face of $B_n$ is 1. In the most naive plan we calculate the relative volumes of all faces. We first produce a list of all faces of each dimension. For dimension 0, we know all the relative volumes are 1. Then, for each face $F$ of dimension $d$ we select a vertex and find the opposite facets (of dimension $d-1$). Assuming recursively that their relative volumes have already been computed, we now find the relative volume of $F$ by summing the relative volumes of the facets. There are two serious drawbacks to the naive plan. Perhaps the most pressing problem is that we need to compute the volumes of an extremely large number of faces, since quite a few of the $2^{n^2}$ possible 0-1 matrices are actually faces of $B_n$. Here we have recourse to a single important trick. If we permute the rows and columns of the matrix representing a face, we obtain the matrix of another face with the same volume. Also if we transpose a matrix representing a face, we obtain another face of the same volume. We regard matrices which can be obtained from each other by these operations as equivalent. We can cut down on the cost of our algorithm if we compute the volume only for a single ``canonical'' face in each equivalence class. The next most difficult problem is to produce the lists of faces. The most practical method that we found for producing faces is to start with the single $(n-1)^2$-dimensional face, $B_n$ itself, and successively produce faces of lower dimension by intersecting with a facet of $B_n$. While producing the faces we save the subface information so that we can look up the volumes when we are done. Unfortunately we need to construct a very large partially ordered set of faces before we can calculate any volumes since the only volumes we know are those of the zero-dimensional faces. While the cost in memory is not so bad for $n$ less than 8, when we reach $n=8$, we seem to need about 200 gigabytes of intermediate storage. If the memory were available, the computation of the volumes would be relatively easy. In fact we are able to carry out a substantial fraction of the work before running out of memory. There are two phases to our algorithm. In the first phase we construct a collection of faces together with information about which ones are facets of which others. In particular, we successively compute, for $d=(n-1)^2,(n-1)^2-1,\dots,0$, a collection ${\cal F}_d$ of $d$-dimensional faces of $B_n$. We begin by setting ${\cal F}_{(n-1)^2} = \{ B_n \}$, {\it i.e.,\/}\ consisting of just the all 1's matrix representing $B_n$ itself. Given ${\cal F}_d$ we produce ${\cal F}_{d-1}$ as follows. Start with ${\cal F}_{d-1}=\emptyset$. For each face $F\in {\cal F}_d$ we select a vertex $v\in F$. We then find the facets of $F$ opposite $v$, canonicalize these faces, and add them to ${\cal F}_{d-1}$. Having done this for all $F\in {\cal F}_d$, we sort ${\cal F}_{d-1}$ and remove the duplicates. Then, for each face $f\in{\cal F}_{d-1}$, we save a list of pointers to the faces in ${\cal F}_d$ from which $f$ arose. (Equivalent faces can appear several times as opposite faces of the same face. When this happens, we include the pointer in the list of pointers multiple times.) This completes the first phase. In the second phase we start with ${\cal F}_0$ and work up to higher dimensions, calculating the relative volume of every saved face until we obtain the volume of $B_n$ itself. This is quite fast, requiring just one addition for each saved pointer. Note that once the pointers are constructed we do not need the faces themselves, unless we want to know which face has each of the intermediate volumes we are computing. For larger values of $n$, the accumulators used for calculating the volumes will overflow. But we can get around this problem by using multiple precision arithmetic or by performing the volume calculation several times modulo various primes and combining the results with the Chinese Remainder Theorem. The main computational work of our algorithm takes place in three steps. \begin{enumerate} \item for each face $F\in {\cal F}_d$, find a vertex $v\in F$. \item determine the facets of $F$ opposite $v$. \item put these opposite facets into canonical form. \end{enumerate} We now describe how each of these steps is done. One important decision is the data structure for storing faces. We identify each face with the 0-1 matrix which is the union of its vertices (regarded as permutation matrices). For $n\le 8$ it is convenient to represent each face as $n^2$ bits of a single word, where the words of a (64-bit) computer are regarded as 64-long arrays of bits. In Step 1 we are given a face $f$ represented by a 0-1 matrix and we are looking for a permutation matrix $\pi$ contained in $f$. This could be done with the assignment algorithm or one of the methods for finding maximum matchings, but for the small values of $n$ that we were using, it was quicker to use a backtracking search method, as follows. The matrix $f$ has at least one 1 in its first row. We guess one of these as the location of the 1 in the first row of $\pi$. We then guess the location of the 1 in the second row of $\pi$, bearing in mind that it cannot be in the same column as the 1 in the first row. We continue this way searching for the location of the 1 in subsequent rows. We backtrack if we reach a row in which there are no feasible choices. Now we consider Step 2. Given a face $f$ and a vertex $\pi$ we need to find the facets of $f$ opposite $\pi$. For a moment let us ignore $\pi$ and consider the general problem of constructing the facets of $f$. The main principle is that each facet of $f$ can be obtained by intersecting $f$ with a facet of $B_n$ that does not contain $f$. Consider the facet corresponding to the pair $(i,j)$. The facet does not contain $f$ if $f_{ij}=1$. To intersect $f$ with this facet we start by replacing $f_{ij}$ with 0, obtaining a 0-1 matrix $g$. The face which is the intersection of $f$ with the facet $(i,j)$ is then the largest face $h$ contained in $g$. The matrix $h$, which is the union of the permutation matrices contained in $g$, can be strictly contained in $g$. Given one of the 1's in $g$, to test whether it is in $h$, we search for a permutation matrix in $g$ which uses the 1 in question. This can be done with our backtracking search algorithm. The 1 in question is in $h$ precisely when this search succeeds. When the position of a 1 in $g$ is zero in $h$ we say it is {\it forced} to zero. For example if $n=3$ and $f$ is the 3-dimensional face with matrix \[ \left(\begin{array}{ccc} 0 & 1 & 1 \\ 1 & 1 & 1 \\ 1 & 1 & 1 \end{array}\right) \] then the intersection of $f$ with the facet corresponding to the middle entry of the top row is one-dimensional, with matrix \[ \left(\begin{array}{ccc} 0 & 0 & 1 \\ 1 & 1 & 0 \\ 1 & 1 & 0 \end{array}\right). \] In this example two zeroes in the last column are forced. This example also shows that although every facet of $f$ is the intersection of $f$ with a facet of $B_n$, the converse is not true and the dimension of the intersection can be too small to be a facet of $f$. There are some additional simplifications when we search for the facets of $f$ opposite a given vertex $\pi$ of $f$. If $g$ is a facet of $f$ not containing $\pi$, then $g$ must contain a 0 in place of one of the 1's of $\pi$. Thus there are at most $n$ facets of $f$ opposite $\pi$. Observe that if $g$ is a facet of $f$ not containing $\pi$, then $g \cup \pi$ is a union of permutation matrices and therefore a face of $B_n$ containing $g$ and $\pi$. Thus $g \cup \pi=f$. This implies an important and helpful principle. When we introduce a 0 at a 1 of $\pi$ and this results in a facet of $f$, then the only other positions that might be forced to zero are those of the other 1's of $\pi$. Thus we can loop through the $n$ 1's of $\pi$ one at a time and, for each of these, introduce a 0 and determine what other 1's of $\pi$ are forced to be 0 and produce accordingly a matrix, which we call a candidate. We obtain a set of $n$ candidates among which all the facets opposite $\pi$ must occur. (This list can have duplicates which we remove.) Of these candidates the facets are those which are maximal under inclusion. Indeed, it is clear that every candidate contains a face that has the same intersection with $\pi$. But this face is contained in a facet which has an intersection with $\pi$ that is at least as large. Thus every candidate is contained in at least one facet, and the facets are precisely the maximal candidates. Finally we describe Step 3, which we call ``canonicalization''. The most straightforward way to choose a canonical form for a face $f$ is to apply every element of our group of symmetries to $f$ and choose the image of $f$ with the least value (where the bit pattern $f$ is regarded as an integer.) But this is prohibitively slow. Instead we make use of certain special functions, which we call scores, which assign integers to every row and column of a 0-1 matrix. The scores have the special property that when rows are permuted, the row scores are permuted the same way leaving the column scores unchanged, whereas, when columns are permuted, the column scores are permuted the same way leaving the row scores unchanged. An example of an allowable score is to assign to each row its row sum and to each column its column sum. Given such scores we say that a matrix is in standard form if it satisfies the following three properties: \begin{enumerate} \item the column scores are weakly increasing. \item the row scores are weakly increasing. \item in the case of tied row scores the rows are ordered lexicographically as bit strings. \end{enumerate} For a given 0-1 matrix, once its row and column scores have been computed it is easy to put a matrix and its transpose into standard form by forcing each of the three conditions above in the listed order. For each face constructed, we put both the face and its transpose into standard form and finally choose the smaller of these two, regarded as integers, as the ``canonical'' form that is saved. Note that we are abusing terminology a little here since although the method always replaces a face by an equivalent face, it is conceivable that equivalent faces will canonicalize to distinct faces. When this happens, we still obtain correct volumes, but we end up doing work which could be avoided if the equivalence were recognized. However, if this event is rare, we obtain almost all the savings of true canonicalization as described above (but without the excessive cost). It turns out that just using row and column sums as the score functions fails to recognize a substantial number of equivalences. What we need are scores that tend to assign different values to different rows and columns. Slightly more complicated scores do better. Given a column score, we can produce a more complicated row score by assigning to each row the sum (or any symmetric function) of the values of the column scores of those columns for which 1's occur in the given row. Similarly a row score can be used to produce a more complicated column score. We can also add two row scores to obtain another row score or two column scores to obtain another column score. By combining steps like this we produced scores that were better at distinguishing rows (and columns) without being much more expensive to compute. This concludes our description of the triangulation method. As mentioned earlier it is reasonably practical for $n < 8$. The times required on a 500mhz DEC alpha were as follows: \[ \begin{array}{cr} & \mbox{Time in seconds} \\ n<6 & \mbox{less than 0.1}\\ n=6 & 0.63\\ n=7 & 250.1 \end{array} \] Although the volumes of $B_n$ do not seem to follow a recognizable pattern, it seemed conceivable that there would be faces of $B_n$ for which the relative volumes had interesting properties. One fairly natural class is the set of matrices for which the set of of zeroes of the matrix form a Young tableau in a corner of the matrix. Since our triangulation method applies to any face of $B_n$, we were able to check some natural classes of faces. It turned out that for the simplest non-trivial Young tableau faces the volumes apparently obey a simple rule, although we have not been able to supply a proof. More precisely, suppose that $n\ge 2$ and that $F_n$ is the $n\times n$ matrix whose $(i,j)$ entry is 1 when $j \le i+1$ and 0 otherwise. Then $F_n$ is a union of permutation matrices corresponding to a face of $B_n$ of dimension $\binom{n}{2}$ with $2^{n-1}$ vertices and we have the following\\ \begin{conjecture}\label{conj:1} The relative volume of $F_n$ is the product \[ \prod_{i=0}^{n-2} \frac{1}{i+1}\binom{2i}{i} \] of the first $n-1$ Catalan numbers.\\ \end{conjecture} We have verified this for $n \leq 12$.\\ Finally, we give some miscellaneous observations which may be useful but do not actually enter our algorithm. \begin{itemize} \item In our method, we never needed to calculate the dimension of a face since the way they were produced guaranteed their dimension. However one may wonder how one can efficiently calculate the dimension of a face. One of the most efficient methods makes use of the fact, discussed in \cite{BS}, that the dimension is equal to $e+k-2n$, where $e$ is the number of 1's in the matrix of $F$ and $k$ the number of components in the graph corresponding to $F$. ({\it i.e.,\/}\ the bipartite graph on $2n$ letters in which $i$ is joined to $j$ when the $(i,j)$ entry of the matrix of $F$ is 1.) \item The relative volume of any $d$-face $F$ can be computed in several different ways since it is the sum of the relative volumes of the facets opposite any vertex of $F$. This yields linear relations on the volumes of $(d-1)$-faces of $B_n$. It seems conceivable that these linear relations could be strong enough to yield useful information about the volumes. However from our limited investigation this does not appear to save anything in our computations. \item Since our standard triangulations all involve minimum volume simplices, one might wonder whether all minimum volume simplices with vertices from the vertex set of $B_n$ belong to one of these triangulations. For $n=4$, we found that there are 658584 minimum volume simplices whose vertices are vertices of $B_4$. Of these, only 641112 belong to some standard triangulation. \end{itemize} \section{The Magic Squares Method} \label{sec:5} In the next two sections we describe the magic squares method for calculating the volume of $B_n$. We have no reason to believe that our implementation is substantially different from those used by others. (See \cite{DG}, \cite{Mo},\cite{SS}, and \cite{Stu}.) The only apparent novelty is that we have carried out the computation when $n=8$. We briefly explain here the connection between magic squares and the volume of $B_n$. It is known that for a $d$-dimensional polytope $P$ with integer vertices, for any nonnegative integer $t$, the number $e(P,t)$ of lattice points contained in $t\cdot P$ is a polynomial of degree $d$ in $t$. This polynomial is called the {\it Ehrhart polynomial\/}\ of $P$. Its leading coefficient is the volume of $P$ in units equal to the volume of the fundamental domain of the affine lattice spanned by $P$. Thus if we know the values of $e(P,t)$ for values of $t$ from 0 to $d$, we can find the Ehrhart polynomial by interpolation and in that way determine the volume of $P$. For $P=B_n$, this method is particularly attractive since the polynomial is known to have certain symmetries, which make it necessary to calculate the values of $e(B_n,t)$ for $t$ only up to and including $\binom{n-1}{2}$ rather than $(n-1)^2$. Note that $e(B_n,t)$ is exactly the number of $n \times n$ matrices with nonnegative integer entries and all row and column sums equal to $t$, i.e., the number of $n \times n$ magic squares with sum $t$. In the next section we will describe how to count magic squares relatively efficiently. To see that we need only find $e(B_n,t)$ for values of $t$ up to and including $\binom{n-1}{2}$ we refer to the following identities: \begin{enumerate} \item $e(B_n,t)=0$ for $-n+1 \le t \le -1$. \item $e(B_n,-n-t)=(-1)^{n-1} e(B_n,t)$ for all $t$. \end{enumerate} These identities (conjectured in \cite{ADG}) are easy consequences of Ehrhart's Law of Reciprocity, which states that, for a $d$-dimensional polytope $P$ with integer vertices, and $t>0$, \[ e^*(P,t)=(-1)^d e(P,-t) \] where $e^*(P,t)$ denotes the number of integer points in the interior of $P$. See \cite[Chapter 9]{H}, and \cite{E} for proof and references. \noindent {\it Proof of 1\/}:\\ $e^*(B_n,t)$ is the number of $n\times n$ matrices with positive integer entries and all row and column sums equal to $t$. Since all the entries are $\ge 1$, each row and column sum must be $\ge n$, so $e^*(B_n,t)=0$ for $1 \le t \le n-1$. By Ehrhart's Law of Reciprocity this implies $e(B_n,t)=0$ for $-n+1 \le t \le -1$. \noindent {\it Proof of 2\/}:\\ There is a one-to-one correspondence between $n\times n$ matrices with nonnegative integer entries and row and column sums $t$ and $n\times n$ matrices with positive integer entries and row and column sums $n+t$. (Simply add 1 to each entry in matrices of the first type.) Thus $e(B_n,t)=e^*(B_n,n+t)$. Applying Ehrhart's Law of Reciprocity, the right-hand-side equals $(-1)^{(n-1)^2}e(B_n,-n-t)$, which simplifies to $(-1)^{n-1}e(B_n,-n-t)$. \medskip The effect of the first identity is that we know $n-1$ zeroes of $e(B_n,t)$. We also have $e(B_n,0)=1$. For each $t>0$, if we calculate the value of $e(B_n,t)$, by the second identity we obtain also the value of $e(B_n,-n-t)$. Thus if we calculate $e(B_n,t)$ for $t$ up to $\binom{n}{2}$, we have a total of $n-1+1+2\binom{n}{2}=(n-1)^2+1$ values of the $e(B_n,t)$ so we have enough data to find the polynomial $e(B_n,t)$ by interpolation. \section{Counting Magic Squares} \label{sec:6} We now describe the method we used for counting the number of $n \times n$ magic squares of row and column sum $t$ for $t\le\binom{n-1}{2}$. This seems no different from the methods used by others \cite{DG} to carry out the smaller cases. Given an $m$-tuple $r=(r_1,\dots,r_m)$ and an $n$-tuple $c=(c_1,\dots,c_n)$ of nonnegative integers, we denote by $N(r,c)$ the number of nonnegative integer matrices with row sums $r_1,\dots,r_m$ and column sums $c_1,\dots,c_n$. There are a few computational principles. The first is that $N(r,c)=0$ unless $|r|=\sum_ir_i=\sum_jc_j=|c|$. Next note that $N(r,c)$ is invariant under permutation of either the $r$'s or the $c$'s. Finally the principle that leads to substantial computational savings is that, for any integer $k$ (usually near $m/2$) \[ N(r,c)=\sum_x N((r_1,\dots,r_k),x)N((r_{k+1},\dots,r_n),c-x) \] where the sum is over all nonnegative $n$-tuples $x$ such that $|x|=r_1+\cdots+r_k$ and $x_i \le c_i$, $i=1,\dots,n$. This formula results from classifying the matrices counted by $N(r,c)$ according to the column sums of the submatrix formed from the first $k$ rows. For fixed column sums $x_1,\dots,x_n$, the column sums of the submatrix formed by the remaining rows must be $c_i-x_i$. The total number of matrices in the class corresponding to $x$ is the number of ways of choosing the top submatrix multiplied by the number of ways of choosing the bottom. The counting of magic squares amounts to the calculation of $N(r,c)$ with the ``constant'' $n$-tuples $r=c=(t,\dots,t)$. For this special case there are a few simplifications. We discuss the case when $n$ is even. The same ideas apply with slight modification when $n$ is odd. Suppose that $n=2m$ and we wish to calculate $e(B_n,t)$. From our general principle we have \[ e(B_n,t)=\sum_y N(R,y)N(R,T-y) \] where $R$ is the $m$-tuple of all $t$'s, $T$ is the $n$-tuple of all $t$'s, and $y$ runs over all nonnegative $n$-tuples satisfying $|y|=mt$, and $y_i \le t$ for all $i$. For a $k$-tuple $y=(y_1,\dots,y_k)$, let us denote by $M(y)$ the number of distinct $k$-tuples which arise by permuting the $y_i$'s. So, if $z_1,\dots,z_l$ are distinct, and $y$ is a $k$-tuple consisting of $k_1$ $z_1$'s, $k_2$ $z_2$'s, etc., then $M(y)=k!/(k_1!\dots k_l!)$. In terms of this notation a more computationally efficient version of the preceding equation is \begin{equation}\label{eq:a} e(B_n,t)=\sum_y M(y)N(R,y)N(R,T-y) \end{equation} where now we further restrict $y$ to weakly increasing $n$-tuples. We can apply this principle again to the calculation of $N(R,y)$ and $N(R,T-y)$ that appear in the last formula. We find that \begin{equation} \label{eq:b} N(R,y)=\sum_x M(x)N(x,(y_1,\dots,y_m))N(R-x,(y_{m+1},\dots,y_n)) \end{equation} where now $x$ runs over all weakly increasing nonnegative $m$-tuples with $|x|=y_1+\cdots+y_m$ and $x_i\le t$ for all $i$. We can save an additional factor of 2 by noting that the quantities $N(R,y)$ are the same as $N(R,T-y)$ except in a different order. Thus if we save the former in a suitable array, we can look up the latter ones in the array rather than computing them. Notice that the ingredients for calculating the sums $N(R,y)$ and $N(R,T-y)$ are the quantities $N(x,y)$ where $x$ and $y$ vary over weakly increasing nonnegative $m$-tuples with $x_i,y_i \le \binom{n-1}{2}$. Thus it is sensible to precompute these quantities and save the results before forming the sums for $N(R,y)$ or the sum for $e(B_n,t)$. For example, for $n=8$ we need to precompute the quantities $N(x,y)$ where $x$ and $y$ have length 4. Again it is easier to calculate \[ N(x,y)=\sum N((x_1,x_2),z) N((x_2,x_3),y-z) \] where the sum is over all 4-long vectors $z$ with $|z|=x_1+x_2$ and $z_i \le y_i$ for all $i$. However we do not have available the additional simplification to a sum over increasing sequences $z$. Thus on the right side we require the values $N(x,y)$, for pairs $(x,y)$ where $x$ has length 2 and $y$ has length 4, not necessarily weakly increasing, where the components of $x$ and $y$ vary up to 21. It would be possible to precompute all the needed values and save these as well for later use. This might be advantageous since these results are used several times each. However, for simplicity, we use a subroutine to compute these, in effect repeating the calculation of any $N(x,y)$ whenever needed. This subroutine in turn calls a subroutine for counting $2 \times 2$ matrices with prescribed row and column sums which calculates $N((x_1,x_2),(y_1,y_2))=\min(x_1,x_2,y_1,y_2)+1$ whenever $|x|=|y|$. The precalculation for $n=8$ requires about 20 minutes on a 500Mhz DEC alpha. The remaining calculation also takes about 20 minutes. The first part can be calculated in single precision. In the remaining parts we need some sort of multiple precision method. We perform the calculation modulo several primes and combine the results with the Chinese Remainder Theorem. A similar program for $n=7$ requires 38 seconds. Here are the Ehrhart polynomials $e(B_n,t)$, for $n=1,\dots,8$. For each $n$, the coefficient of the last binomial coefficient in the expression for $e(B_n,t)$ is the relative volume of $B_n$. We express the Ehrhart polynomial of $B_n$ as an integer combination of binomial coefficients $C(t+n-1+k,n-1+2k)=\binom{t+n-1+k}{n-1+2k}$, $k=0,\dots,\binom{n-1}{2}$, because they are a basis for the polynomials satisfying $p(-n-t)=(-1)^{n-1}p(t)$. \begin{eqnarray*} e(B_1,t) &=& C(t,0)\\ e(B_2,t) &=& C({t+1},1)\\ e(B_3,t) &=& C({t+2},2)+3C({t+3},4)\\ e(B_4,t) &=& C({t+3},3)+20C({t+4},5)+152C({t+5},7)+352C({t+6},9)\\ e(B_5,t) &=& C({t+4},4)+115C({t+5},6)+5390C({t+6},8)+\\ && 101275C({t+7},10)+858650C({t+8},12)+\\ && 3309025C({t+9},14)+4718075C({t+10},16)\\ e(B_6,t) &=& C({t+5},5)+714C({t+6},7)+196677C({t+7},9)+\\ && 18941310C({t+8},11)+809451144C({t+9},13)+\\ && 17914693608C({t+10},15)+223688514048C({t+11},17)+\\ && 1633645276848C({t+12},19)+6907466271384C({t+13},21)+\\ && 15642484909560C({t+14},23)+14666561365176C({t+15},25)\\ \end{eqnarray*} \begin{eqnarray*} e(B_7,t) &=& C({t+6},6)+ 5033C({t+7},8)+\\ && 9090305C({t+8},10)+ 4562637436C({t+9},12)+\\ &&876755512997C({t+10},14)+ 80592643025748C({t+11},16)+\\ && 4085047594855542C({t+12},18)+\\ && 125166504299043921C({t+13},20)+\\ && 2460507569635629206C({t+14},22)+\\ && 32199612314177385616C({t+15},24)+\\ && 285953447105799237366C({t+16},26)+\\ && 1727929241168643056768C({t+17},28)+\\ && 6989369809320320632154C({t+18},30)+\\ && 18096158896344747268932C({t+19},32)+\\ && 27093648035077238674360C({t+20},34)+\\ && 17832560768358341943028C({t+21},36)\\ e(B_8,t) &=& C(t+7,7)+\\ && 40312C(t+8,9)+\\ && 544604804C(t+9,11)+\\ && 1572522771472C(t+10,13)+\\ &&1433860489078360C(t+11,15)+\\ && 546197610013169408C(t+12,17)+\\ && 104573799019751624800C(t+13,19)+\\ && 11404657872578818785152C(t+14,21)+\\ && 773100275338739807806336C(t+15,23)+\\ && 34668602440014649185072000C(t+16,25)+\\ && 1075823106306592550013512704C(t+17,27)+\\ && 23865735845675030268755397632C(t+18,29)+\\ && 387264682746696963082402212768C(t+19,31)+\\ && 4666750907574155613393947915904C(t+20,33)+\\ && 42107239094874587731729608526080C(t+21,35)+\\ && 284859465667770778104594682157824C(t+22,37)+\\ && 1435919936068954265096148477657088C(t+23,39)+\\ && 5307981556350553774098942855517184C(t+24,41)+\\ && 13958946247270195588626193027208192C(t+25,43)+\\ && 24706461764218063045041689495950080C(t+26,45)+\\ && 26368507913706408235698183181290240C(t+27,47)+\\ && 12816077964079346687829905128694016C(t+28,49)\\ \end{eqnarray*} \section{Comparison} \label{sec:7} We now compare the two methods described above. The main advantage of the first method seems to be that it applies just as well to any face of $B_n$ as it does to $B_n$ itself. To apply the algorithm to a face $F$ of $B_n$, we simply start at the top level with the 0-1 matrix associated to $F$ and produce lists of canonical subfaces as before. In the second method it is not obvious how well one could do in computing the volume of an arbitrary face $F$ of $B_n$. This would amount to counting the number of magic squares with prescribed zeros and row and column sums $t$ for possibly as many as $\dim(F)$ values of $t$. We would not have $e^*(F,t)=0$ for $1 \le t \le n-1$, nor would we have $e(F,t)=e^*(F,n+t)$, because of the prescribed zeros in $F$. For certain $F$ ({\it e.g.,\/}\ those with the same number of prescribed zeros in every row) we would have an analogous identity, and some automatic roots, but in general we cannot guarantee any cutdown in the number of values of $e(F,t)$ needed to determine the polynomial. Furthermore in the actual counting of magic squares with certain prescribed zeros, we would not be able to exploit the symmetries used in our algorithm above. The second method however has the advantage that, for computing volumes of $B_n$ itself, it is much more feasible in terms of memory. The second method also computes the Ehrhart polynomial. It seems possible that the first method could be modified to compute Ehrhart polynomials of the faces as well as just their volumes. We would need to keep track of the numbers of simplices of each dimension in a standard triangulation instead of just the simplices of the largest dimension. \newpage \section*{Appendix: Ratio of Volumes of \protect\boldmath$B_n$ and \protect\boldmath$A_n$} \label{sec:appendix} Consider the linear mapping $\cal L$ from $(n-1) \times (n-1)$ matrices to $n \times n$ matrices which sends matrix $E_{i,j}$ which is all zero except for a 1 at $(i,j)$ to the matrix $F_{i,j}$ which is all zero except for 1's at $(i,j)$ and $(n,n)$ and $-1$'s at $(i,n)$ and $(n,j)$. If we follow $\cal L$ by the addition of the $n \times n$ matrix that has the block form \[ \left(\begin{array}{ccc} 0 & J_{n-1,1} \\ J_{1,n-1} & 2-n \end{array}\right) \] where $J_{k,l}$ is the all $k \times l$ matrix of all 1's, then we obtain the affine mapping which sends $A_n$ to $B_n$. Thus if we denote the ratio we seek by $R$, we find that $R^2$ is the determinant of the $(n-1)^2 \times (n-1)^2$ matrix of dot products of $F_{i,j} \cdot F_{k,l}=2^{\delta_{ik}}2^{\delta_{jl}}$. But in general if $ x_{ik}$ and $y_{jl}$ are two $m \times m$ matrices and $z$ is the $m^2 \times m^2$ tensor product matrix indexed by pairs $ij$ and $kl$ given by $ z_{ij,kl}=x_{ik}y_{jl} $ then $\det z = (\det x \det y)^m $. Our case is the special case that $x=y=J_{n-1,n-1}+I_{n-1}$. Since the characteristic polynomial of $-J_m$ is $\lambda^{m-1}(\lambda+m)$, the determinant of $J_{n-1}+I_{n-1}$ is $n$. It follows that $R=n^{n-1}$.
{ "timestamp": "1998-06-13T04:05:13", "yymm": "9806", "arxiv_id": "math/9806076", "language": "en", "url": "https://arxiv.org/abs/math/9806076", "abstract": "We study the calculation of the volume of the polytope B_n of n by n doubly stochastic matrices; that is, the set of real non-negative matrices with all row and column sums equal to one.We describe two methods. The first involves a decomposition of the polytope into simplices. The second involves the enumeration of ``magic squares'', i.e., n by n non-negative integer matrices whose rows and columns all sum to the same integer.We have used the first method to confirm the previously known values through n=7. This method can also be used to compute the volumes of faces of B_n. For example, we have observed that the volume of a particular face of B_n appears to be a product of Catalan numbers.We have used the second method to find the volume for n=8, which we believe was not previously known.", "subjects": "Combinatorics (math.CO)", "title": "On the volume of the polytope of doubly stochastic matrices", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9854964215865318, "lm_q2_score": 0.8128673201042492, "lm_q1q2_score": 0.8010778351873715 }
https://arxiv.org/abs/2109.01697
The double-bubble problem on the square lattice
We investigate minimal-perimeter configurations of two finite sets of points on the square lattice. This corresponds to a lattice version of the classical double-bubble problem. We give a detailed description of the fine geometry of minimisers and, in some parameter regime, we compute the optimal perimeter as a function of the size of the point sets. Moreover, we provide a sharp bound on the difference between two minimisers, which are generally not unique, and use it to rigorously identify their Wulff shape, as the size of the point sets scales up.
\section{Introduction} The classical double-bubble problem is concerned with the shape of two sets of given volume under minimisation of their surface area. In the Euclidean space, minimisers are enclosed by three spherical caps, intersecting at an angle of $2\pi/3$. The proof of this fact in ${\mathbb R}^2$ dates back to \cite{Foisy}, and has then been extended to ${\mathbb R}^3$ \cite{Hutchings} and ${\mathbb R}^n$ for $n \geq 4$ \cite{Reichardt}. See also \cite{Cicalese2} for a quantitative stability analysis in two dimensions. A number of variants of the problem has also been tackled, including double bubbles in spherical and hyperbolic spaces \color{black} \cite{Corneli,Corneli2,Cotton,Masters}, \color{black} hyperbolic surfaces \cite{Boyer}, cones \cite{Lopez,Morgan}, the $3$-torus \cite{Carrion,Corneli0}, \color{black} the Gau\ss\ space \cite{Corneli2,Milman}, \color{black} and in the anisotropic Grushin plane \cite{Franceschi}. The aim of this paper is to tackle a lattice version of the double-bubble problem. We restrict our attention to the square lattice ${\mathbb Z}^2$ and define the {\it lattice length} of the interface separating two disjoint sets $C,\,D \subset {\mathbb Z}^2$ as $Q(C,D) = \#\{(c,d)\in C\times D \colon \, |c-d|=1\}$, \color{black} where $|\cdot|$ is the Euclidean norm. \color{black} The {\it lattice double-bubble problem} consists in finding two distinct lattice subsets $A$ and $B$ of fixed sizes $N_A,N_B \in \mathbb{N}$ solving \begin{equation} \label{eq:dbp} \min\{P(A,B) \colon \ A,\, B \subset {\mathbb Z}^2, \ A \cap B = \emptyset, \ \#A = N_A, \ \#B = N_B\}, \end{equation} where the {\it lattice perimeter} $P(A,B)$ is defined by \begin{align}\label{eq:dbp3} P(A,B) &= Q(A,A^c) + Q(B,B^c) - 2\beta Q(A,B)\notag\\ &= Q(A,A^c\setminus B) + Q(B,B^c\setminus A) + (2-2\beta) Q(A,B). \end{align} The latter definition features the parameter $\beta\in (0,1)$. Note that the classical double-bubble case corresponds to the choice $\beta=1/2$. In the following, we allow for the more general $\beta\in (0,1)$, for this will be relevant in connection with applications, see Section \ref{sec:equiv}. In particular, $\beta$ models the interaction between the two sets. The reader is referred to \cite{Futer} where cost-minimizing networks featuring different interaction costs are considered. Analogously to the Euclidean case, we prove that minimisers $(A,B)$ of \eqref{eq:dbp} are connected ($A$, $B$, and $A\cup B$ are connected in the usual lattice sense, see below). Call {\it isoperimetric} those subsets of the lattice which minimize $C \mapsto Q(C,C^c)$ under given cardinality. Without claiming completeness, the reader is referred to the monograph \cite{Harper} and to \cite{Bezrukov0,Biskup,Bobkov,Bollobas,Wang} for a minimal collection of results on discrete isoperimetric inequalities, to \cite{Cicalese,Mainini,Mainini2} for sharp fluctuation estimates, and to \cite{Barrett} for some numerical approximation. A second analogy with the Euclidean setting is that optimal pairs $(A,B)$ {\it do not} consist of the mere union of two isoperimetric sets $A$ and $B$, for the onset of an interface between $A$ and $B$ influences their shape. \begin{figure}[h] \includegraphics[scale=0.4]{aspectratio.png} \caption{A minimiser for $\beta=1/2$} \label{fig:aspectratio} \end{figure} Differently from the Euclidean case, existence of minimisers for \eqref{eq:dbp} is here obvious, for the minimisation problem is finite. Moreover, the geometry of the intersection of interfaces is much simplified, as effect of the discrete geometry of the underlying lattice. In particular, all interfaces meet at multiples of $\pi/2$ angles. At finite sizes $N_A, \, N_B $, boundary effects are relevant and a whole menagerie of minimisers of \eqref{eq:dbp} may arise, depending on the specific values of $N_A, \, N_B $, and $\beta$. Indeed, although uniqueness holds in some special cases, it cannot be expected in general. We are however able to prove an a priori estimate on the symmetric distance of two minimisers, which differ at most by \color{black} $N^{1/2}_A=N^{1/2}_B$ \color{black} points. As size scales up, whereas properly rescaled isoperimetric sets approach the square, $A$ and $B$ converge to suitable rectangles. In the limit $N_A = N_B \to \infty$ (and for $\beta=1/2$), we prove that minimisers of \eqref{eq:dbp} converge to the {\it Wulff shape} configuration of Figure \ref{fig:aspectratio}. That is, uniqueness is restored in the Wulff shape limit. In fact, \color{black} in the crystalline-perimeter case, the \color{black} double-bubble problem for $\beta=1/2$ has been already tackled in \cite{Morgan98}, see also the recent \cite{Duncan0} for an elementary proof of the existence of minimisers. The case $\beta\not = 1/2$ is addressed in \cite{Wecht} instead. In particular, the different possible geometries of the Wulff shape, corresponding to different volume fractions of the two phases, have been identified. Let us now present our main results. We start by associating to each $\mathcal{V} \subset {\mathbb Z}^2$ the corresponding {\it unit-disk graph}, namely the undirected simple graph $G=(\mathcal{V},\mathcal{E})$, where vertices are identified with the points in $\mathcal{V}$, and the set $\mathcal{E}\subset \mathcal{V} \times \mathcal{V}$ of edges contains one edge for each pair of points in $\mathcal{V}$ at distance $1$. We say that a subset $ \mathcal{V} \subset {\mathbb Z}^2$ is {\it connected} if the corresponding unit-disk graph is connected. Moreover, we indicate by $R_z : = {\mathbb Z}\times \{z\}$ and $C_z=\{z\}\times {\mathbb Z}$ rows and columns, for all $z \in {\mathbb Z}$. Our main findings read as follows. \begin{theorem}\label{thm:main} Let $(A,B)$ solve the double-bubble problem \eqref{eq:dbp}. Then, \begin{enumerate} \item[\rm i] {\rm (Connectedness)} The sets $A$, $B$, and $A \cup B$ are connected. Moreover, the sets $A\cap R_z$, $B\cap R_z$, $(A \cup B)\cap R_z$, $A\cap C_z$, $B\cap C_z$, and $(A \cup B)\cap C_z$ are connected (possibly being empty) for all $z\in {\mathbb Z}$;\\ \item[\rm ii] {\rm (Separation)} If $\max\{x \colon\, (x,z)\in A\} \leq \min\{x \colon\, (x,z)\in B\} - 1 $ {\rm for some} $z \in {\mathbb Z} $, then the same holds with equality {\rm for all} $z \in {\mathbb Z} $ (whenever not empty). An analogous statement is valid for columns, possibly after exchanging the role of $A$ and $B$;\\ \item[\rm iii] {\rm (Interface)} Let $I\subset {\mathbb R}^2$ be the set of midpoints of segments connecting points in $A$ with points in $B$ at distance $1$. Then, for all $x\in I$ there exists $y\in I\setminus\{x\}$ with $|x-y| \in \lbrace 1/ \sqrt{2}, 1\rbrace $ and $I$ can be \color{black} included in the image of a piecewise-affine curve \color{black} $\iota \colon [0,1] \to {\mathbb R}^2 $ with monotone components. \end{enumerate} If $N_A=N_B=\color{black} N \color{black}$ and $\beta\leq 1/2$, we additionally have that \begin{enumerate} \item[\rm iv] {\rm (Minimal perimeter)} $P(A,B)=\min\{P_*,P^*\}$, where \begin{align*} P_*:= 4 \left\lceil \frac{\color{black} N \color{black}}{ \left\lfloor \sqrt{\frac{2\color{black} N \color{black}}{2-\beta}}\right\rfloor }\right\rceil +2 \left\lfloor \sqrt{\frac{2 \color{black} N \color{black}}{2-\beta}}\right\rfloor (2-\beta),\\ P^*:= 4\left\lceil \frac{\color{black} N \color{black}}{ \left\lceil \sqrt{\frac{2 \color{black} N \color{black} }{2-\beta}}\right\rceil }\right\rceil+2 \left\lceil \sqrt{\frac{2 \color{black} N \color{black} }{2-\beta}}\right\rceil (2-\beta); \end{align*} \item[\rm v] {\rm (Explicit solution)} There exist $h,\, \ell \in {\mathbb N}$ and $0 \le r < h$ with $\color{black} N \color{black} = h\ell +r$, $|h - \sqrt{2 \color{black} N \color{black}/(2-\beta)}| \le 1$, and $|\ell - \sqrt{\color{black} N \color{black} (2-\beta)/2}| \le 2 $ such that, letting \begin{align*} A'&:=\{(x,y)\in {\mathbb Z}^2 \, \colon \, x \in[-\ell + 1 ,0], \, y \in [1,h] \ \text{or} \ x=-\ell,\, y \in [1,r]\},\\ B'&:=\{(x,y)\in {\mathbb Z}^2 \, \colon \, x \in[1,\ell], \, y \in [1,h] \ \text{or} \ x=\ell+1,\, y \in [1,r]\}, \end{align*} the pair $(A',B')$ solves the double-bubble problem \eqref{eq:dbp};\\ \item[\rm vi] {\rm (Fluctuations)} There exists a constant $C_\beta$ only depending on $\beta$ \color{black} and an isometry $T$ of ${\mathbb Z}^2$ \color{black} such that \begin{equation}\label{eq:fluct}\# (A \triangle T(A')) + \# (B \triangle T( B')) \leq C_\beta N^{1/2}, \end{equation} (see beginning of Section \ref{sec:law} for the definition of \color{black} isometry). \color{black} \end{enumerate} \end{theorem} Theorem \ref{thm:main} is proved in subsequent steps along the paper, by carefully characterising the geometry of optimal pairs $(A,B)$. In fact, our analysis reveals additional geometrical details, so that the statements in the coming sections are often more precise and more general in terms of conditions on the parameters $N_A$, $N_B$, and $\beta$ with respect to Theorem~\ref{thm:main}. We prefer to postpone these details in order not to overburden the introduction. The connectedness of optimal pairs $(A,B)$ is discussed in Section \ref{sec:algo} and Theorem \ref{thm:main}.i is proved in Theorem \ref{thm:connected} and Corollary \ref{cor:rows}. The separation property of Theorem \ref{thm:main}.ii follows from Proposition~\ref{prop:algorithmdecreasesenergy} and Corollaries~\ref{cor:nomissingrows}-\ref{cor:allononeside}. The geometry of the interface between $A$ and $B$, namely Theorem~\ref{thm:main}.iii, is described by Corollary \ref{cor:interface}. In Section \ref{sec:info} we present a collection of examples, illustrating the variety of optimal geometries. In particular, we show that optimal pairs may be not unique and, in some specific parameter range, present quite distinguished shapes. We then classify different admissible pairs in Section \ref{sec:classi} by introducing five distinct classes of configurations. The first of these classes, called Class $\mathcal I$ and corresponding to Figure \ref{fig:aspectratio}, is indeed the reference one and is studied in detail in Section \ref{sec:reg1}. In Proposition \ref{prop:classIregularisationstep2} we prove the existence of optimal pairs in Class $\mathcal I$, among which \color{black} there is \color{black} the explicit one of Theorem \ref{thm:main}.v. The minimal perimeter in Theorem~\ref{thm:main}.iv is then computed by referring to this specific class in Theorem~\ref{thm:classIexact}. The remaining classes are studied in Section \ref{sec:reg2}. We show that some of the classes cannot be optimal in the case $N_A = N_B$, and that the other ones can be modified to a configuration in Class $\mathcal{I}$ by an explicit regularisation procedure. We also observe that for arbitrarily large $N$ solutions may appear which are not in Class $\mathcal{I}$, see Proposition~\ref{prop:largeminimisersiv}. Although optimal pairs $(A,B)$ are not unique, by carefully inspecting our constructions, we are able to prove that, in some specific parameter regime, two optimal pairs differ by at most $C_\beta N^{1/2}$ points, up to isometries, and that this bound is actually sharp. This is studied in Section~\ref{sec:law}, see Theorem~\ref{thm:nonehalf} \color{black} which \color{black} proves Theorem \ref{thm:main}.vi. \color{black} This scaling in fluctuations is specifically related to the presence of an interface between the two sets $A$ and $B$. In fact, in case of a single set $A$, optimal configurations show fluctuations of order $N^{3/4}$, \color{black} see Subsection \ref{sec:opt} for details. \color{black} Although \color{black} the setting of our paper is discrete, our results deliver some understanding of the continuous case, as well. This results by considering the so-called {\it thermodynamic limit} as $\color{black} N \color{black} \to \infty$. For all $ V =\{x_1, \dots, x_{\color{black} N \color{black}}\} \subset {\mathbb Z}^2$, let $ \mu_V = (\sum_{i=1}^{\color{black} N \color{black}} \delta_{x_i/\sqrt{\color{black} N \color{black}}})/\color{black} N \color{black}$ be the corresponding empirical measure on the plane and denote by $\mathcal L$ the two-dimensional Lebesgue measure. We indicate by \begin{align} \mathcal A:=\left(-\sqrt{\frac{2-\beta}{2}},0\right)\times \left(0, \sqrt{\frac{2}{2-\beta}}\right) \ \ \text{and} \ \ \mathcal B:= \left(0, \sqrt{\frac{2-\beta}{2}} \right)\times \left(0, \sqrt{\frac{2}{2-\beta}}\right)\label{eq:wulff} \end{align} the continuous Wulff shapes, see Figure \ref{fig:aspectratio}. Note that $\mathcal L(\mathcal A)=\mathcal L(\mathcal B) = 1$. By combining the explicit construction of Theorem \ref{thm:main}.v and the fluctuation estimate \eqref{eq:fluct} we have the following. \begin{corollary}[Wulff shapes]\label{cor: wulff} Let $\beta\leq 1/2$ and $(A_{\color{black} N \color{black}},B_{\color{black} N \color{black}})$ be solutions of \eqref{eq:dbp} with $N_{A_{\color{black} N \color{black}}} = N_{B_{\color{black} N \color{black}}} = \color{black} N \color{black}$, for all $\color{black} N \color{black}\in \mathbb N$. Then, there exist isometries $T_{\color{black} N \color{black}}$ of ${\mathbb Z}^2$ such that \begin{align} \label{eq:meas} \mu_{T_{\color{black} N \color{black}} A_{\color{black} N \color{black}}} \stackrel{\ast}{\rightharpoonup} {\mathcal L} \mres {\mathcal A} \ \ \text{and} \ \ \mu_{T_{\color{black} N \color{black}} B_{\color{black} N \color{black}}} \stackrel{\ast}{\rightharpoonup} {\mathcal L} \mres {\mathcal B}, \end{align} \color{black} as $N \to \infty$, \color{black} where the symbol $\stackrel{\ast}{\rightharpoonup}$ indicates the weak-$\ast$ convergence of measures. \end{corollary} Note that, by taking $\beta=1$ in \eqref{eq:wulff} (not covered by the corollary, though) we have that $\mathcal A \cup \mathcal B$ form a single square with side $\sqrt{2}$ whereas for $\beta=0$ the Wulff shapes $\mathcal A$ and $\mathcal B$ are two squares of side $1$. Our results also allow to solve the double-bubble problem in the continuous setting of ${\mathbb R}^2$ with respect to a crystalline perimeter notion. More precisely, for every set $D \subset {\mathbb R}^2$ of {\it finite perimeter} we denote by $\partial^* D$ its {\it reduced boundary} \cite{Ambrosio-Fusco-Pallara,Maggi}, and define the {\it crystalline perimeter} and the {\it crystalline length} as $${\rm Per} (D) = \int_{\partial^* D} \| \nu \|_1 \, {\rm d} \mathcal H^1, \quad {\rm L} (\gamma) = \int_{\gamma} \| \nu \|_1 \, {\rm d} \mathcal H^1,$$ where $\nu$ is the outward pointing \color{black} unit \color{black} normal to $\partial^*D$, $\|\nu\|_1 = |\nu_x|+|\nu_y|$, $ \mathcal H^1$ is the one-dimensional Hausdorff measure, and $\gamma \subset \partial^*D$ is measurable. The continuous analogue of \eqref{eq:dbp} is the {\it crystalline double-bubble problem} \begin{align} & \min\Big\{ {\rm Per} (A) + {\rm Per} (B) - 2\beta \, {\rm L} (\partial^*A \cap \partial^*B)\, \colon \ \\ & \qquad \qquad A,\, B \subset {\mathbb R}^2 \ \text{of finite perimeter}, \quad A \cap B = \emptyset, \ \mathcal L(A)=\mathcal L(B)=1\Big\}. \label{eq:dbp2} \end{align} By combining Theorem \ref{thm:main}.v and \ref{thm:main}.vi we obtain the following. \begin{corollary}[Crystalline double bubble]\label{cor: cryst-db} For all $\beta \leq 1/2$, the pair $(\mathcal A,\mathcal B)$ is a solution of \eqref{eq:dbp2}. The minimal energy is given by $4 \sqrt{4-2\beta}$. \end{corollary} For the reference choice $\beta=1/2$, the solution of the crystalline double-bubble problem~\eqref{eq:dbp2} is depicted in Figure \ref{fig:aspectratio}, see also \cite{Duncan,Morgan98}. Corollaries \ref{cor: wulff} and \ref{cor: cryst-db} are proved in Section \ref{sec:wulff}. In the recent \cite{Duncan}, the difference in energy between any properly rescaled optimal discrete configuration and the Wulff shape is estimated. In case $N_A=N_B$ and $\beta\leq 1/2$ such an estimate can be recovered from the exact expressions in Theorem \ref{thm:main}.iv and of Corollary \ref{cor: cryst-db}. Note however that the analysis in \cite{Duncan} covers the case $N_A\not =N_B$ as well, although for $\beta=1/2$ only. \section{ Equivalent formulations of the double-bubble problem}\label{sec:equiv} \subsection{Optimal particle configurations}\label{sec:opt} The double-bubble problem \eqref{eq:dbp} can be equivalently recasted in terms of ground states of \color{black} configurations of particles of two different types. \color{black} Let $A =\{ x_1,\dots, x_{N_A} \}$ and $B=\{ x_{N_A+1}, \dots, x_{N_A+N_B}\}$ indicate the mutually distinct positions of particles of two different particle species and assume that $A,\,B \subset {\mathbb Z}^2$, which in turn restricts the model to the description of zero-temperature situations. To the particle configuration $(A,B)$ we associate the {\it configurational energy} \begin{align}\label{eq: basic eneg} E(A,B) = \frac12 \sum_{i, j=1}^{N_A+ N_B} V_{ \rm sticky}(x_i, x_j), \end{align} where \begin{equation*} V_{ \rm sticky}(x_i, x_j) = \threepartdef{-1}{|x_i - x_j|= 1 \mbox{ and } x_i,x_j \in A \mbox{ or } x_i,x_j \in B,}{-\beta}{|x_i - x_j| = 1 \mbox{ and } x_i \in A, x_j \in B \mbox{ or } x_i \in B, x_j \in A,}{0}{|x_i - x_j| \neq 1.} \end{equation*} The interaction density $V_{\rm sticky}(x_i, x_j) $ corresponds to the so-called {\it sticky} or {\it Heitmann-Radin-type} potential \cite{Heitmann} and models the binding energy of the two particles $x_i$ and $ x_j $. In particular, only first-neighbor interactions contribute to the energy, and {\it intraspecific} (namely, of type $A-A$ or $B-B$) and {\it interspecific} (type $A-B$) interactions are quantified differently, with interspecific interactions being weaker as $\beta <1$. The relation between the minimisation of $E$ and the double-bubble problem \eqref{eq:dbp} is revealed by the equality \begin{equation} E(A,B)+2N_A + 2N_B = \frac12P(A,B).\label{eq:eq} \end{equation} This follows by analysing the contribution to $E$ and $P$ of each point. In fact, one could decompose $$E(A,B) = \sum_{i=1}^{N_A+N_B} e(x_i), \quad P(A,B) =\sum_{i=1}^{N_A+N_B} p(x_i), $$ where the single-point contribution to energy and perimeter is quantified via \begin{align*} e(x) = -\frac12 \# \{\text{same-species neighbors of $x$}\} - \frac{\beta}{2} \# \{\text{other-species neighbors of $x$}\}\\ p(x) = 4 - \# \{\text{same-species neighbors of $x$}\} - \beta \# \{\text{other-species neighbors of $x$}\}. \end{align*} The latter entail \eqref{eq:eq}, which in turn ensures that ground states of $E$ and minimisers of $P$ coincide, for all given sizes $N_A$ and $N_B$ of the sets $A$ and $B$. The geometry of ground states of $E$ results from the competition between intraspecific and interspecific interaction. In the extremal case $\beta=1$, intra- and interspecific interaction are indistinguishable, and one can consider the whole system $(A,B)$ as a single species. The minimisation of $E$ is then the classical {\it edge-isoperimetric problem} \cite{Bezrukov,Harper}, namely the minimisation of $C \mapsto Q(C,C^c)$ under prescribed size $\#C$. Ground states are isoperimetric sets, the ground-state energy is known, the possible distance between two ground states scales as $N^{3/4}$ where $N = \# C$, and one could even directly prove crystallization, \color{black} i.e., the periodicity of ground states, \color{black} under some stronger assumptions on the interaction potentials~\cite{Mainini}. In the other extremal case $\beta=0$, no interspecific interaction is accounted for, and both phases $A$ and $B$ are independent isoperimetric sets. In particular, if $N_A$ and $N_B$ are perfect squares (or for $N_A, \, N_B \to \infty$ and up to rescaling), the phases $A$ and $B$ are squares. In the intermediate case $\beta\in(0,1)$, which is hence the interesting one, intraspecific and interspecific interaction compete and neither $A$ or $B$ nor $A \cup B$ end up being isoperimetric sets. The presence of interspecific interactions adds some level of rigidity. This is revealed by the fact, which we prove, that the distance between different ground states scales like $N^{1/2}$, in contrast with the purely edge-isoperimetric case, where fluctuations are of order $N^{3/4}$ \cite{Mainini}, see also \cite{Cicalese,Davoli,Mainini2,Schmidt}. Although we do not directly deal with crystallization here, for the points $A$ and $B$ are {\it assumed} to be subset of the lattice $\mathbb Z^2$, let us mention that a few rigorous crystallization results in multispecies systems are available. At first, existence of quasiperiodic ground states in a specific multicomponent two-dimensional system has been shown by Radin \cite{Radin86}. One dimensional crystallization of {\it alternating} configurations of two-species has been investigated by B\'etermin, Kn\"upfer, and Nolte \cite{Betermin}, see also \cite{periodic} for some related crystallization and noncyrstallization results. In the two-dimensional, sticky interaction case, two crystallization results in hexagonal and square geometries are given in \cite{kreutz, kreutz2}. Here, however, interspecific interactions favor \color{black} the onset of alternating phases. \color{black} \subsection{Finite Ising model} The double-bubble problem \eqref{eq:dbp} can also be equivalently seen as the ground-state problem for a {\it finite} Ising model with ferromagnetic interactions. In particular, given $C=A \cup B\subset {\mathbb Z}^2$ one describes the state of the system by $u \colon C \to \pm 1$, distinguishing the $+1$ and the $-1$ phase. The Ising-type energy of the system is then given by $$\color{black} F \color{black} (C,u) = -\frac{1-\beta}{ 4}\sum_{\substack{x,y \in C \\ |x-y|=1}} u(x)\, u(y) - \frac{1+\beta}{ 4} \sum_{\substack{x,y \in C \\ |x-y|=1}} \color{black} |u(x)\, u(y))|. \color{black} $$ The first term above is the classical ferromagnetic interaction contribution, while the second sum gives the total number of interactions, irrespective of the phase. This second term is required since in our model same-phase and different-phase interactions are both assumed to give negative contributions to the energy. Under the above provisions, minimisers of the problem $$\min\left\{\color{black} F \color{black} (C,u) \, \colon \, C \subset {\mathbb Z}^2, \ u \colon C \to \pm 1, \ \# \{x \in C \, \colon \, u(x)=1\} = N_A, \ \# \{x \in C \, \colon \, u(x)=-1\} = N_B \right\} $$ corresponds to solutions $(A,B)$ of the double-bubble problem \eqref{eq:dbp}, under the equivalence $A\equiv \{x \in C \, \colon \, u(x)=1\}$ and $B\equiv \{x \in C \, \colon \, u(x)=-1\}$. In fact, each pair of first neighbors contributes $-1$ to $\color{black} F \color{black}$ if it belongs to the same phase and $-\beta$ if it belongs to different phases, namely, $$\color{black} F \color{black} (C,u) = E(A,B).$$ The literature on the Ising model is vast and the reader is referred to \cite{Cerf,McCoy} for a comprehensive collection of results. Ising models are usually investigated from the point of view of their thermodynamic limit $\#C \to \infty$ and at \color{black} positive \color{black} temperature. In particular, models are usually formulated on the whole lattice or on a large box with constant boundary states. Correspondingly, the analysis of Wulff shapes is concerned with the study of a droplet of one phase in a sea of the other one \cite{Cerf2}. Our setting is much different, for our system is finite and boundary effects matter. To the best of our knowledge, we contribute here the first characterisation of ferromagnetic Ising ground states, where the location $C$ of the system is also unknown and results from minimisation. Alternatively to the finite two-state setting above, one could equivalently formulate the minimisation problem in the whole ${\mathbb Z}^2$ by allowing a third state, to be interpreted as interaction-neutral. In particular, we could equivalently consider the minimisation problem $$\min\left\{\color{black} F \color{black} ({\mathbb Z}^2,v) \, \colon \, v \colon {\mathbb Z}^2\to\{-1,0,1\}, \ \# \{x \in {\mathbb Z}^2 \, \colon \, v(x)=1\} = N_A, \ \# \{x \in {\mathbb Z}^2 \, \colon \, v(x)=-1\} = N_B \right\}. $$ The equivalence is of course given by setting $u=v$ on $C:=\{x\in {\mathbb Z}^2\, \colon \, v(x)\not = 0 \}$. \subsection{Finite Heisenberg model} The three-state formulation of the previous subsection can be easily reconciled within the frame of the classical Heisenberg model \cite{Tasaki}. In particular, we shall define the vector-valued state function $s\colon M \to \{s_{-1}, s_0, s_1\}$ where the box $M$ is given as $M:=[0,m]^2 \cap {\mathbb Z}^2$ for $m$ large. We choose the three possible spins as $$ s_0 =(-1,0), \ \ s_1 =\left(\beta, \sqrt{1-\beta^2}\right), \ \ s_{-1}=\left(\beta, -\sqrt{1-\beta^2}\right).$$ The energy of the system is defined as $$H(s) = -\sum_{\substack{x,y \in M \\ |x-y|=1}} s(x)\cdot s(y).$$ For all $s\colon M \to \{s_{-1}, s_0, s_1\}$, let $A:=\{x \in M \, \colon \, s(x)=s_1\} $ and $B:=\{x \in M \, \colon \, s(x)=s_{-1}\}$. We are interested in the minimisation problem \begin{align*}&\min\left\{H(s) \, \colon \, s\colon M \to \{s_{-1}, s_0, s_1\}, \ \#A=N_A , \ \#B=N_B \right\}. \end{align*} By letting $m$ be very large compared with $N_A$ and $N_B$, we can with no loss of generality assume that $s=s_0$ close to the boundary $\partial M$. Let us now show that the latter minimisation problem is indeed equivalent to the double-bubble problem \eqref{eq:dbp}. To this aim, we start by noting that the total number of first-neighbor interactions in $M$ is $2m^2+2m$. First-neighbor interactions between identical states contribute $-1$ to the energy, $s_0 - s_{1}$ and $s_0 - s_{-1}$ interactions contribute $-s_1\cdot s_0 = -s_{-1}\cdot s_0 = \beta$, and $s_1- s_{-1}$ interactions contribute $-s_1\cdot s_{-1} = 1-2{\beta^2} $. We hence have that \begin{align*} H(s) + (2m^2 +2m) &= (\beta+1) \left( Q (A, A^c\setminus B) + Q (B, B^c\setminus A) \right) + (2-2\beta^2) Q(A,B) \\ & = \left(\beta+1 \right) \left( Q (A, A^c\setminus B) + Q (B, B^c\setminus A) +(2-2\beta)Q(A,B) \right) \nonumber\\ &= \left(\beta+1\right) P(A,B), \end{align*} so that minimising $H$ is actually equivalent to solving \eqref{eq:dbp}. \subsection{Minimum \color{black} balanced-separator \color{black} problem} One can rephrase the double-bubble problem \eqref{eq:dbp} as a minimum \color{black} balanced-separator \color{black} problem on an unknown graph as well. Indeed, as interspecific contributions are energetically less favored with respect to intraspecific ones, given the common occupancy ${\mathcal V} =A \cup B$ of the two phases, one is asked to part ${\mathcal V} $ into two regions $A$ and $B$ with given size in such a way that the interface between $A$ and $B$ is minimal. This corresponds to a minimum \color{black} balanced-separator \color{black} problem on the {\it unit-disk graph} corresponding to ${\mathcal V}$, i.e., finding a disjunct partition ${\mathcal V} = A \cup B$ solving $$\min \{ Q(A,B) \, : \, \#A = N_A, \ \#B = N_B\}.$$ This is indeed a classical problem, with relevant applications in operations research and computer science \cite{Nagamochi}. Here, we generalize the above minimum \color{black} balanced-separator \color{black} problem by letting the underlying graph also vary and by simultaneously optimising its perimeter. In particular, we consider $$\min \{P(A,B) \, \colon \, V=A \cup B, \ A \cap B =\emptyset, \ \#A = N_A, \ \#B = N_B\},$$ where $({\mathcal V}, {\mathcal E})$ is again the unit graph related to $A \cup B\subset {\mathbb Z}^2$. Also in this setting, the competition between minimisation of the interface and of the perimeter is evident. Recall $P(A,B) = Q(A,A^c \setminus B) + Q(B,B^c \setminus A) + (2-2\beta) Q(A,B)$. On the one hand, a graph with few edges between $A$ and $B$ would give a short cut $Q(A,B)$, while necessarily having large $Q(A,A^c \setminus B) + Q(B,B^c \setminus A)$. On the other hand, a graph with small $Q(A,A^c \setminus B) + Q(B,B^c \setminus A)$ has $A \cup B$ close to be a square, and for $N_A = N_B$ all possible cuts partitioning it in two are approximately as long as its side. \section{Notation} Let us collect here some notation, to be used throughout the paper. For each pair of disjoint sets $A,B \subset \mathbb{Z}^2$ we call the elements of $A$ and $B$ the {\it $A$-points} and {\it $B$-points}, respectively. We let $N_A = \# A$ and $N_B = \# B$. For any point $p \in A \cup B$, we denote its first and second coordinate by $p = (p_x, p_y)$. We say that two points are {\it connected by an edge} if their distance is equal to one. (Equivalently, we sometimes use the words {\it bond} or {\it connection} in place of {\it edge}.) We say that a set $S \subset A \cup B$ is {\it connected} if it is connected as a graph with edges described above, or equivalently if the corresponding unit-disk graph is connected. For the sake of definiteness, from here on, our notation is adapted to the setting of Subsection~\ref{sec:opt}. In particular, we say that a configuration is {\it minimal} (or {\it optimal}) if it minimises the energy $E$ given in \eqref{eq: basic eneg} in the class of configurations with the same number of $A$- and $B$-points. Recall once more that minimisers of $E$ and solutions of the double-bubble problem \eqref{eq:dbp} coincide. Since the number of points is finite, any configuration lies in a bounded square. Suppose that a configuration $(A,B)$ has $N_{\rm row}$ rows (i.e., there are $N_{\rm row}$ rows in $\mathbb{Z}^2$ with at least one point from $A \cup B$). For $k=1,...,N_{\rm row}$, denote by $R_k$ the $k$-th row (counting from the top). In a similar fashion, $N_{\rm col}$ denotes the number of columns, and $C_k$ indicates the $k$-th column (counting from the left). To simplify the notation, given a finite set $X \subset \mathbb{Z}^2$, we denote $X^{\rm row}_k = X \cap R_k$ and $X_k^{\rm col} = X \cap C_k$. We will typically apply this to the sets $A$, $B$, their union or some of their subsets. Moreover, denote by $n_k^{\rm row}$ the number of $A$-points in the row $R_k$ and by $m^{\rm row}_k$ the number of $B$-points in the row $R_k$. In a similar fashion, $n_k^{\rm col}$ and $m_k^{\rm col}$ denote the number of $A$- and $B$-points in column $C_k$, respectively. In the following, we will frequently modify configurations. Not to overburden the notation, when we use the notation $n^{\rm row}_k$ and $m_k^{\rm row}$ (and similarly for columns) we always refer to the configuration in the same sentence, unless otherwise specified. For two points $p,q \in A \cup B$, we say that $p$ {\it lies to the left} (respectively {\it right}) of $q$ if $p_y = q_y$ and $p_x < q_x$ (respectively $p_x > q_x$). In other words, they are in the same row, and the first coordinate of $p$ is smaller (respectively larger) than the first coordinate of $q$. We say that $p$ lies directly to the left (respectively right) of $q$ if additionally $p$ and $q$ are connected by an edge. Similarly, we say that $p$ {\it lies above} (respectively {\it below}) $q$ if $p_x = q_x$ and $p_y > q_y$ (respectively $p_y < q_y$). Again, we say that $p$ lies {\it directly} above (respectively below) $q$ if additionally these two points are connected by an edge. We will also say that the set $A^{\rm row}_k$ {\it lies to the left} (respectively {\it right}) of $B^{\rm row}_k$ if for every $p \in A^{\rm row}_k$ and $q \in B^{\rm row}_k$ the point $p$ lies to the left (respectively right) of $q$. (Note that by definition $A^{\rm row}_k$ and $B^{\rm row}_k$ are in the same row.) We also say that $A^{\rm row}_k$ lies {\it directly} to the left of $B^{\rm row}_k$ if additionally there is a connection between one of the points in $A^{\rm row}_k$ and one of the points in $B^{\rm row}_k$. An analogous notion is used for columns. Furthermore, we say that a number of points from different rows are {\it aligned} if their first coordinates are equal. We also say that two sets are {\it aligned to the right} (or {\it left}) if their rightmost (leftmost) points are aligned. The same notion is also used for columns. Finally, given a finite set $X \subset \mathbb{Z}^2$, we denote by $X + (a,b)$ the set consisting of all points of $X$ shifted by the vector $(a,b) \in \mathbb{Z}^2$. \section{Connectedness, separation, and interface}\label{sec:algo} In this section, we introduce a procedure in order to modify an arbitrary configuration $(A,B)$ into another configuration $(\hat{A}, \hat{B})$ with specific additional properties, without increasing the energy. In particular, this will prove that for a minimal configuration the sets $A$, $B$, and $A \cup B$ are connected. \subsection{Description of the procedure} The goal of this subsection is to present a procedure allowing to modify a configuration, making it more regular in the following sense: not only the sets $A$ and $B$ are connected, but also for any $k = 1,...,N_{\rm row}$ and any $l = 1,...,N_{\rm col}$ the sets $A^{\rm row}_k$, $B^{\rm row}_k$, $(A \cup B)_k^{\rm row}$, $A^{\rm col}_l$, $B^{\rm col}_l$, and $(A \cup B)_l^{\rm col}$ are connected. We start with the following preliminary result. \begin{proposition} Let $(A,B)$ be a configuration in the sense described above. If there are any empty rows (or columns) between any two rows (or columns) in $(A,B)$, then there exists a configuration $(\hat{A},\hat{B})$ with strictly smaller energy. \end{proposition} \begin{proof} Without restriction we present the argument for rows. Suppose that between rows $R_k$ and $R_{k+1}$ for some $k \in \{ 1,...,N_{\rm row} - 1 \}$ there are $l$ empty rows. Then, we can reduce the energy in the following way: denote by $(A',B')$ the configuration consisting of the top $k$ rows and by $(A'',B'')$ the configuration consisting of the bottom $N_{\rm row}-k$ rows. Then, we remove the empty rows, i.e., replace $(A'',B'')$ with $(A'',B'') + (0,l)$. Clearly, this does not increase the energy of the configuration $(A,B)$. If after this shift there is at least one connection between $A^{\rm row}_k \cup B^{\rm row}_k$ and $A^{\rm row}_{k+1} \cup B^{\rm row}_{k+1}$, the energy even decreases by at least $\beta$. Otherwise, if after this shift there are no connections between $A^{\rm row}_k \cup B^{\rm row}_k$ and $A^{\rm row}_{k+1} \cup B^{\rm row}_{k+1}$, we shift the configuration $(A',B')$ horizontally to make at least one connection. Again, the energy is decreased by at least $\beta$. \end{proof} Hence, in studying minimal configurations, we may assume that there are no empty rows and columns. Now, we are ready to describe a modification procedure making the configuration more regular. Notice that we may write the energy in the following way: \begin{equation*} E(A,B) = \sum_{k=1}^{N_{\rm row}} E^{\rm row}_k(A,B) + \sum_{k=1}^{N_{\rm row} -1} E_k^{\rm \color{black} inter \color{black}}(A,B). \end{equation*} Here, $E^{\rm row}_k(A,B)$ is the part of the energy given by interactions in the row $R_k$, namely \begin{equation}\label{eq: row energy} E^{\rm row}_k(A,B) = \frac{1}{2} \sum_{ x_i,x_j \in A^{\rm row}_k \cup B^{\rm row}_k} V_{ \rm sticky}(x_i, x_j), \end{equation} and $E_k^{\rm \color{black} inter \color{black}}(A,B)$ is the part of the energy given by interactions between rows $R_k$ and $R_{k+1}$, namely \begin{equation*} E_k^{\rm \color{black} inter \color{black}}(A,B) = \sum_{ x_i \in A^{\rm row}_k \cup B^{\rm row}_k, \, x_j \in A^{\rm row}_{k+1} \cup B^{\rm row}_{k+1}} V_{ \rm sticky }(x_i, x_j). \end{equation*} Now, let us see that we may bound $E^{\rm row}_k$ and $E_k^{\rm \color{black} inter \color{black}}$ by expressions depending on $n^{\rm row}_k$ and $m^{\rm row}_k$. First, we estimate $E^{\rm row}_k$. \begin{lemma}\label{lem:horizontal} We have \begin{equation*} E^{\rm row}_k(A,B) \geq \begin{cases} - (n^{\rm row}_k + m^{\rm row}_k) +2- \beta & \text{if $n^{\rm row}_k > 0$, $m^{\rm row}_k > 0$,} \\ - (n^{\rm row}_k + m^{\rm row}_k) +1 & \text{else.} \end{cases} \end{equation*} Moreover, this inequality is an equality if and only if the sets $A^{\rm row}_k$, $B^{\rm row}_k$, and $A^{\rm row}_k \cup B^{\rm row}_k$ are connected. \end{lemma} \begin{proof} We consider two cases. In the first case, we suppose that $m^{\rm row}_k = 0$ (a similar argument works if $n^{\rm row}_k = 0$): then, the desired inequality takes the form $E^{\rm row}_k(A,B) \geq -n^{\rm row}_k + 1$. Since $A^{\rm row}_k$ is a subset of a single row, $n^{\rm row}_k - 1$ is the maximum number of connections between points in $A^{\rm row}_k$ and it is achieved only if $A^{\rm row}_k$ is connected. In the second case, we have $n^{\rm row}_k > 0$ and $m^{\rm row}_k > 0$. Since $A^{\rm row}_k \cup B^{\rm row}_k$ is a subset of a single row, the maximum number of connections (regardless of their type) is $n^{\rm row}_k + m^{\rm row}_k - 1$. It is achieved only if $(A \cup B)_k^{\rm row}$ is connected. Among these, at most $n^{\rm row}_k - 1$ are connections between points in $A^{\rm row}_k$ and at most $m^{\rm row}_k - 1$ are connections between points in $B^{\rm row}_k$. These numbers are achieved if and only if $A^{\rm row}_k$ and $B^{\rm row}_k$ are connected. Each of these connections contributes $-1$ to the energy and there can be at most $n^{\rm row}_k + m^{\rm row}_k - 2$ of them. The remaining connections are between $A^{\rm row}_k$ and $B^{\rm row}_k$ contributing $-\beta$ to the energy. The fact that $\beta < 1$ yields the statement. \end{proof} Now, we make a similar computation for $E_k^{\rm \color{black} inter \color{black}}$. \begin{lemma}\label{lem:vertical} We have \begin{equation*} E_k^{\rm \color{black} inter \color{black}}(A,B) \geq -(1-\beta)\big(\min\lbrace n^{\rm row}_k, n^{\rm row}_{k+1}\rbrace +\min\lbrace m^{\rm row}_k, m^{\rm row}_{k+1}\rbrace \big) - \beta \min\lbrace n^{\rm row}_k + m^{\rm row}_k, n^{\rm row}_{k+1} + m^{\rm row}_{k+1}\rbrace. \end{equation*} Moreover, equality is achieved if and only if the following conditions hold: \\ (1) There are $\min\lbrace n^{\rm row}_k,n^{\rm row}_{k+1}\rbrace$ points in $A^{\rm row}_k$ directly above points in $A^{\rm row}_{k+1}$; \\ (2) There are $\min\lbrace m^{\rm row}_k,m^{\rm row}_{k+1}\rbrace$ points in $B^{\rm row}_k$ directly above points in $B^{\rm row}_{k+1}$; \\ (3) Supposing that $n^{\rm row}_k + m^{\rm row}_k \geq n^{\rm row}_{k+1} + m^{\rm row}_{k+1}$, there is a point in $A^{\rm row}_k \cup B^{\rm row}_k$ directly above every point in $A^{\rm row}_{k+1} \cup B^{\rm row}_{k+1}$. Otherwise, if $n^{\rm row}_k + m^{\rm row}_k < n^{\rm row}_{k+1} + m^{\rm row}_{k+1}$, there is a point in $A^{\rm row}_{k+1} \cup B^{\rm row}_{k+1}$ directly below every point in $A^{\rm row}_k \cup B^{\rm row}_k$. \end{lemma} \begin{proof} First, as there are $n^{\rm row}_k + m^{\rm row}_k$ points in $A^{\rm row}_k \cup B^{\rm row}_k$ and $n^{\rm row}_{k+1} + m^{\rm row}_{k+1}$ points in $A^{\rm row}_{k+1} \cup B^{\rm row}_{k+1}$, there are at most $\min\lbrace n^{\rm row}_k + m^{\rm row}_k, n^{\rm row}_{k+1} + m^{\rm row}_{k+1}\rbrace$ connections between points in $A^{\rm row}_k \cup B^{\rm row}_k$ and $A^{\rm row}_{k+1} \cup B^{\rm row}_{k+1}$, regardless of their type. Among these, we denote the number of connections between points in $A^{\rm row}_k$ and $A^{\rm row}_{k+1}$ by $\tilde{n}_k$ and the number of connections between points in $B^{\rm row}_k$ and $B^{\rm row}_{k+1}$ by $\tilde{m}_k$. We have $\tilde{n}_k \le \min\lbrace n^{\rm row}_k, n^{\rm row}_{k+1}\rbrace$ and $\tilde{m}_k \le \min\lbrace m^{\rm row}_k, m^{\rm row}_{k+1}\rbrace$ with equality if this many points in $A^{\rm row}_{k+1}$ are placed directly under points in $A^{\rm row}_k$ (and similarly for $B^{\rm row}_k$ and $B^{\rm row}_{k+1}$). Each of these connections contributes $-1$ to the energy, i.e., a total contribution of $-\tilde{n}_k - \tilde{m}_k$. Then, there are at most $\min\lbrace n^{\rm row}_k + m^{\rm row}_k, n^{\rm row}_{k+1} + m^{\rm row}_{k+1}\rbrace - (\tilde{n}_k + \tilde{m}_k)$ possible connections which need to be either connections between points in $A^{\rm row}_k$ and $B^{\rm row}_{k+1}$ or between points in $B^{\rm row}_k$ and $A^{\rm row}_{k+1}$. Either way, each of these connections contributes $-\beta$ to the energy. In conclusion, we obtain the desired inequality, with equality only if $\tilde{n}_k = \min\lbrace n^{\rm row}_k, n^{\rm row}_{k+1}\rbrace$, $\tilde{m}_k = \min\lbrace m^{\rm row}_k, m^{\rm row}_{k+1}\rbrace$, and if there are $\min\lbrace n^{\rm row}_k + m^{\rm row}_k, n^{\rm row}_{k+1} + m^{\rm row}_{k+1}\rbrace$ connections between $A^{\rm row}_k \cup B^{\rm row}_k$ and $A^{\rm row}_{k+1} \cup B^{\rm row}_{k+1}$. \end{proof} In light of these estimates, we describe a simple modification procedure making any configuration more regular. For any configuration $(A,B)$, we construct a configuration $(\hat{A},\hat{B})$ having the same number of $A$- and $B$-points in each row as $(A,B)$ such that the energy is lower or equal and $(\hat{A},\hat{B})$ has some additional structure properties. \textit{Step 0:} We start with the first row from the top. We let $\hat{A}_1$ be a connected set in a single row consisting of $n^{\rm row}_1$ atoms and let $\hat{B}_1$ be the connected set in the same row with $m^{\rm row}_1$ points right of $\hat{A}_1$, in such a way that there is a connection between $\hat{A}_1$ and $\hat{B}_1$. By Lemma~\ref{lem:horizontal}, we have $ E^{\rm row}_1 (\hat{A},\hat{B}) \leq E^{\rm row}_1 (A,B)$. \textit{Step $k$} (for $k = 1,..., N_{\rm row} -1$): We suppose that the sets in the previous steps have been constructed in such a way that $\hat{A}_{k}$, $\hat{B}_k$, and $\hat{A}_k \cup \hat{B}_k$ are connected, and $\hat{A}_k$ lies on the left of $\hat{B}_k$. We will now define $\hat{A}_{k+1}$ and $\hat{B}_{k+1}$. To this end, we distinguish four cases. \textit{Case 1:} $n^{\rm row}_k \leq n^{\rm row}_{k+1}$ and $m^{\rm row}_k \leq m^{\rm row}_{k+1}$. We place $n^{\rm row}_k$ points of $\hat{A}_{k+1}$ directly below $\hat{A}_k$. Then, we put the remaining $n^{\rm row}_{k+1} - n^{\rm row}_k$ points to the left of the previously placed points, so that $\hat{A}_{k+1}$ is connected. Similarly, we place $m^{\rm row}_k$ points from $\hat{B}_{k+1}$ directly below $\hat{B}_k$ and the remaining $m^{\rm row}_{k+1} - m^{\rm row}_k$ points to the right of the previously placed points, so that $\hat{B}_{k+1}$ is connected. By Lemma \ref{lem:horizontal}, we have $E^{\rm row}_{k+1}(\hat{A},\hat{B}) \leq E^{\rm row}_{k+1}(A,B)$, and by Lemma \ref{lem:vertical}, we have $E_k^{\rm \color{black} inter \color{black}}(\hat{A},\hat{B}) \leq E_k^{\rm \color{black} inter \color{black}}(A,B)$. \textit{Case 2:} $n^{\rm row}_k > n^{\rm row}_{k+1}$ and $m^{\rm row}_k > m^{\rm row}_{k+1}$. We place all the points of $\hat{A}_{k+1}$ directly below $\hat{A}_k$, starting from the right. Then, we place all the points of $\hat{B}_{k+1}$ directly below $\hat{B}_k$, starting from the left. In this way, the sets $\hat{A}_{k+1}$, $\hat{B}_{k+1}$ and $\hat{A}_{k+1} \cup \hat{B}_{k+1}$ are connected. Again, by Lemma \ref{lem:horizontal} we have $E^{\rm row}_{k+1}(\hat{A},\hat{B}) \leq E^{\rm row}_{k+1}(A,B)$ and by Lemma \ref{lem:vertical} we have $E_k^{\rm \color{black} inter \color{black}}(\hat{A},\hat{B}) \leq E_k^{\rm \color{black} inter \color{black}}(A,B)$. \textit{Case 3:} $n^{\rm row}_k \le n^{\rm row}_{k+1}$ and $m^{\rm row}_k > m^{\rm row}_{k+1}$. First, we put $n^{\rm row}_k$ points of $\hat{A}_{k+1}$ directly below $\hat{A}_k$. Then, we consider two possibilities: - If $n^{\rm row}_k + m^{\rm row}_k \geq n^{\rm row}_{k+1} + m^{\rm row}_{k+1}$, we place the remaining $n^{\rm row}_{k+1} - n^{\rm row}_k$ points of $\hat{A}_{k+1}$ under $\hat{B}_k$, starting from the left so that $\hat{A}_{k+1}$ is connected. Then, we place the $m^{\rm row}_{k+1}$ points of $\hat{B}_{k+1}$ to the right of the previously placed points, so that $\hat{B}_{k+1}$ and $\hat{A}_{k+1} \cup \hat{B}_{k+1}$ are connected. - If $n^{\rm row}_k + m^{\rm row}_k < n^{\rm row}_{k+1} + m^{\rm row}_{k+1}$, we place the $m^{\rm row}_{k+1}$ points of $\hat{B}_{k+1}$ below points in $\hat{B}_k$, starting from the right, so that $\hat{B}_{k+1}$ is connected. Then, we place $m^{\rm row}_k - m^{\rm row}_{k+1}$ points of $\hat{A}_{k+1}$ between the two sets of previously placed points. Finally, we place the remaining points of $\hat{A}_{k+1}$ to the left of all points placed so far, so that $\hat{A}_{k+1} \cup \hat{B}_{k+1}$ is connected. In both cases, by Lemma~\ref{lem:horizontal} we have $E^{\rm row}_{k+1}(\hat{A},\hat{B}) \leq E^{\rm row}_{k+1}(A,B)$ and by Lemma~\ref{lem:vertical} we get $E_k^{\rm \color{black} inter \color{black}}(\hat{A},\hat{B}) \leq E_k^{\rm \color{black} inter \color{black}}(A,B)$. \textit{Case 4:} $n^{\rm row}_k > n^{\rm row}_{k+1}$ and $m^{\rm row}_k \le m^{\rm row}_{k+1}$. We proceed as in Case 3 with the roles of $A$ and $B$ interchanged, with 'left' and 'right' also interchanged. Again, by Lemma \ref{lem:horizontal} we have $E^{\rm row}_{k+1}(\hat{A},\hat{B}) \leq E^{\rm row}_{k+1}(A,B)$ and by Lemma \ref{lem:vertical} we have $E_k^{\rm \color{black} inter \color{black}}(\hat{A},\hat{B}) \leq E_k^{\rm \color{black} inter \color{black}}(A,B)$. \begin{proposition}\label{prop:algorithmdecreasesenergy} The procedure described above modifies a configuration $(A,B)$ into a configuration $(\hat{A},\hat{B})$ with $E(\hat{A},\hat{B}) \leq E(A,B)$. Moreover, if one of the sets $A^{\rm row}_k$, $B^{\rm row}_k$, or $(A \cup B)_k^{\rm row}$, for $k=1,\ldots,N_{\rm row}$, is not connected, or one of the properties {\rm (1)--(3)} in Lemma \ref{lem:vertical} is violated, then $E(\hat{A},\hat{B}) < E(A,B)$. \end{proposition} \begin{proof} The construction ensures that the configuration $(\hat{A},\hat{B})$ has the same number of rows as $(A,B)$. Hence, we compute \begin{align*} E(\hat{A},\hat{B}) & = \sum_{k=1}^{N_{\rm row} } E^{\rm row}_k(\hat{A},\hat{B}) + \sum_{k=1}^{N_{\rm row}-1} E_k^{\rm \color{black} inter \color{black}}(\hat{A},\hat{B}) \leq \sum_{k=1}^{N_{\rm row}} E^{\rm row}_k(A,B) + \sum_{k=1}^{N_{\rm row}-1} E_k^{\rm \color{black} inter \color{black}}(A,B) \\ & = E(A,B). \end{align*} In view of Lemma \ref{lem:horizontal}, we obtain strict inequality if one of the sets $A^{\rm row}_k$, $B^{\rm row}_k$ or $(A \cup B)_k^{\rm row}$ is not connected. In a similar fashion, we get strict inequality whenever one of the properties (1)--(3) in Lemma \ref{lem:vertical} does not hold. \end{proof} In particular, for optimal configurations $(A,B)$, all sets $A^{\rm row}_k$, $B^{\rm row}_k$, and $(A \cup B)_k^{\rm row}$ are connected. In other words, inside any row we have first all points of one type and then all points of the other type without any gaps in between. Moreover, we may make use of this procedure (and prove an analogue of Lemma \ref{lem:horizontal}--Proposition \ref{prop:algorithmdecreasesenergy}) for columns in place of rows. Hence, the sets $A^{\rm col}_k$, $B^{\rm col}_k$, and $(A \cup B)_k^{\rm col}$ are connected. In other words, given an optimal configuration, in each column there are first all points of one type and then all points of the other type without any gaps in between. In particular, as a consequence, we get an important property of any minimising configuration. \begin{theorem}\label{thm:connected} Suppose that $(A,B)$ is an optimal configuration. Then $A$ and $B$ are connected. \end{theorem} \begin{proof} Suppose by contradiction that $A$ is not connected (we proceed similarly for $B$). First of all, let us notice that for each $k = 1,...,N_{\rm row}$ the set $A^{\rm row}_k$ is connected. Otherwise, by Proposition \ref{prop:algorithmdecreasesenergy} we find that $(A,B)$ was not an optimal configuration. Let us first suppose that $n^{\rm row}_k > 0$ for all $k \in 1,...,N_{\rm row}$ (i.e., $A^{\rm row}_k \neq \emptyset$). Since every $A^{\rm row}_k$ is connected, if $A$ is not connected, it means that there is no connection between $A^{\rm row}_k$ and $A^{\rm row}_{k+1}$ for some choice of $k$. In this case, by Lemma~\ref{lem:vertical} and by Proposition \ref{prop:algorithmdecreasesenergy} we find that $(A,B)$ was not an optimal configuration. Hence, the only remaining possibility that $A$ is not connected is that there exist $k_1 < k_2 < k_3$ such that $n^{\rm row}_{k_1}, n^{\rm row}_{k_3} > 0$ and $n^{\rm row}_{k_2} = 0$ (i.e., $A^{\rm row}_{k_1}, A^{\rm row}_{k_3} \neq \emptyset$ and $A^{\rm row}_{k_2} = \emptyset$). Without loss of generality, we may require that for every $k = k_1 + 1,...,k_3 - 1$ the set $A^{\rm row}_{k}$ is empty. Let us apply the reorganisation $(A,B) \rightarrow (\hat{A},\hat{B})$ using the procedure described above. Clearly, $(\hat{A},\hat{B})$ is still optimal by Proposition \ref{prop:algorithmdecreasesenergy}. Then, for $k = k_1$ we are either in Case 2 or in Case 4 of the procedure. We distinguish these two cases. In the first one, suppose that for $k = k_1$ Case~2 of the procedure applies. Then, the leftmost point of $\hat{B}_{k_1+1}$ lies directly below the leftmost point of $\hat{B}_{k_1}$. Then, since for every $k = k_1 + 1,...,k_3 - 1$ the set $A^{\rm row}_{k}$ is empty, Case 1 or 3 of the procedure shows that also the leftmost point of $\hat{B}_k$ lies below the leftmost point of $\hat{B}_{k_1+1}$ (hence below the leftmost point of $\hat{B}_{k_1}$). Now, for $k = k_3 - 1$, when we place the sets $\hat{A}_{k_3}$ and $\hat{B}_{k_3}$, we either fall into Case 1 or Case 3 in the description of the procedure. In Case 1, the leftmost point of $\hat{B}_{k_3}$ is again placed below the leftmost point of $\hat{B}_{k_1}$. Then, the rightmost point of $\hat{A}_{k_3}$ is placed below the rightmost point of $\hat{A}_{k_1}$. \color{black} Now, one reaches a contradiction by following the same construction of Proposition~\ref{prop:algorithmdecreasesenergy} by exchanging the role of rows and columns. \color{black} In Case 3, we either have that a point of $\hat{A}_{k_3}$ is placed below a point of $\hat{A}_{k_1}$, which as above is a contradiction to Proposition~\ref{prop:algorithmdecreasesenergy}, or the leftmost point of $\hat{A}_{k_3}$ is placed below the leftmost point of $\hat{B}_{k_1}$. In particular, the leftmost point of $\hat{A}_{k_3}$ is placed one point to the right of the rightmost point of $\hat{A}_{k_1}$. Then, by Lemma~\ref{lem:vertical}(1) and Proposition~\ref{prop:algorithmdecreasesenergy} for columns in place of rows we again see that the energy of $(\hat{A},\hat{B})$ was not minimal, a contradiction. In the second case, we have that for $k = k_1$ Case 4 of the algorithm applies. Then, the leftmost point of $\hat{B}_{k_1+1}$ does not lie directly below the leftmost point of $\hat{B}_{k_1}$, but it lies to its left (but no further than the leftmost point of $\hat{A}_{k_1}$). Again, for every $k = k_1 + 1,...,k_3 - 1$ the leftmost point of $\hat{B}_k$ lies below the leftmost point of $\hat{B}_{k_1+1}$. Again, when we place the sets $\hat{A}_{k_3}$ and $\hat{B}_{k_3}$, Case~1 or Case~3 of the procedure applies. In Case 1, the leftmost point of $\hat{B}_{k_3}$ is again placed below the leftmost point of $\hat{B}_{k_1+1}$. Hence, the rightmost point of $\hat{A}_{k_3}$ is placed either below a point in $\hat{A}_{k_1}$ or, in view of the definition of $\hat{B}_{k_1+1}$, one point to the left from the leftmost point of $\hat{A}_{k_1}$. As before, by Lemma~\ref{lem:vertical}(1) and Proposition \ref{prop:algorithmdecreasesenergy} for columns in place of rows, we see that the energy of $(\hat{A},\hat{B})$ was not minimal, a contradiction. In Case 3, a point of $\hat{A}_{k_3}$ is placed below the leftmost point of $\hat{B}_{k_3-1}$. This shows that the leftmost point of $\hat{A}_{k_3}$ is placed either below a point in $\hat{A}_{k_1}$ or one point to the right from the rightmost point of $\hat{A}_{k_1}$. As before, we obtain a contradiction to the minimality of $(\hat{A},\hat{B})$, and the proof is concluded. \end{proof} A careful inspection of the proofs of Proposition \ref{prop:algorithmdecreasesenergy} and Theorem \ref{thm:connected} provides some more information about the structure of any minimising configuration, collected in the following corollaries. \begin{corollary}\label{cor:rows} Let $(A,B)$ be an optimal configuration. Then, for any row $R_k$, the sets $A^{\rm row}_k$, $B^{\rm row}_k$ and $(A \cup B)_k^{\rm row}$ are connected. The same claim holds for columns. \hfill$\Box$ \vskip.3cm \end{corollary} \begin{corollary}\label{cor:nomissingrows} Let $(A,B)$ be an optimal configuration. If for some $1 \leq k_1 < k_2 \leq N_{\rm row}$ we have $A^{\rm row}_{k_1}, A^{\rm row}_{k_2} \neq \emptyset$, then also $A^{\rm row}_k \neq \emptyset$ for all $k_1 \leq k \leq k_2$. The same claim holds for columns and the set $B$. \hfill$\Box$ \vskip.3cm \end{corollary} \begin{corollary}\label{cor:allononeside} Let $(A,B)$ be an optimal configuration. Suppose that there exists a row $R_{k_0}$ such that $A^{\rm row}_{k_0}, B^{\rm row}_{k_0} \neq \emptyset$ and $A^{\rm row}_{k_0}$ lies to the left of $B^{\rm row}_{k_0}$. Then, for every row $R_k$ either $A^{\rm row}_k$ lies to the left of $B^{\rm row}_k$ or one of these sets is empty. The same claim holds for columns and if we interchange the roles of $A$ and $B$. \hfill$\Box$ \vskip.3cm \end{corollary} We observe that Theorem \ref{thm:connected} and Corollary \ref{cor:rows} imply Theorem \ref{thm:main}.i and that Theorem \ref{thm:main}.ii follows from Proposition~\ref{prop:algorithmdecreasesenergy} and Corollaries~\ref{cor:nomissingrows}--\ref{cor:allononeside}. Corollary \ref{cor:interface} implies Theorem~\ref{thm:main}.iii and will be crucial for our later considerations. To this end, we introduce the following definition. \begin{definition} The interface $I_{AB}$ (between $A$ and $B$) is the set of midpoints of edges connecting a point in $A$ with a point in $B$. We say that there is an edge between two points $p,q \in I_{AB}$ if $|p - q| \in \lbrace 1/\sqrt{2}, 1\rbrace$ and the line segment between $p$ and $q$ does not intersect any point in $\mathbb{Z}^2$. We say that the interface is connected if it is connected as a graph. \end{definition} In other words, a point $p \in \mathbb{R}^2$ lies in the interface $I_{AB}$ between $A$ and $B$ if there exist points $p_1 \in A$ and $p_2 \in B$ such that $|p - p_1| = |p - p_2| = 1/2$. Necessarily, the interface is a subset of the lattice $\{ (k + \frac{1}{2},l)\colon k,l \in \mathbb{Z} \} \cup \{ (k, l + \frac{1}{2})\colon k,l \in \mathbb{Z} \}$. An example is presented in Figure \ref{fig:interface}. \begin{figure}[h] \includegraphics[scale=0.09]{interface-eps-converted-to.pdf} \caption{Definition of the interface} \label{fig:interface} \end{figure} Notice that Corollary \ref{cor:rows} implies that there is at most one point in $I_{AB}$ which is a midpoint of an edge between a point in $A^{\rm row}_k$ and a point in $B^{\rm row}_k$. Similarly, there is at most one point in $I_{AB}$ which is a midpoint of an edge between a point in $A^{\rm col}_k$ and a point in $B^{\rm col}_k$. Hence, we get the following result. \begin{corollary}\label{cor:interface} For any optimal configuration $(A,B)$, the interface $I_{AB}$ is connected. Moreover, it is monotone: up to reflections, it goes only upwards and to the right, i.e., given $p,q \in I_{AB}$, if $p_1 > q_1$, then $p_2 \ge q_2$. \hfill$\Box$ \vskip.3cm \end{corollary} We will use this result to study the minimal configurations in the following way: we will identify all possible shapes of the interface, collected in different classes. Analysing the different classes in detail, we will show that there always exists an optimal configuration in the most natural class (called Class~$\mathcal{I}$). For this class, we are able to directly compute the minimal energy, explicitly exhibit a minimiser, and provide a sharp estimate of the possible mismatch of ground states in terms of their size, see \eqref{eq:fluct}. Let us also note that the introduction of $I_{AB}$ enables us to write a convenient formula for the energy associated to an optimal configuration $(A,B)$. Namely, denote by $E_A$ the energy inside $A$, i.e., minus the number of bonds between $A$-points. In a similar fashion, we define $E_B$. Eventually, by $E_{AB} := - \# I_{AB} \beta$ we denote the interfacial energy, i.e., minus the number of bonds between $A$- and $B$-points weighted by the coefficient $\beta$. Then, \begin{equation}\label{eq:formulafortheenergy} E(A,B) = E_A + E_B + E_{AB}. \end{equation} This simple formula has a very important consequence. Namely, if we separate the sets $A$ and $B$ and reattach them in a different way (i.e., apply an isometry to one or both sets), then $E_A$ and $E_B$ do not change, but $E_{AB}$ possibly might. Therefore, if a configuration is optimal, it has the longest possible interface with respect to this operation. We will use variants of this argument on multiple occasions in Section \ref{sec:regularisation}. \section{A collection of examples}\label{sec:info} In this short section, we consider a few examples of minimisers that will serve as a motivation for the discussion about possible shapes of the interface in the next section. By Theorem~\ref{thm:connected}, for any optimal configuration, both sets $A$ and $B$ are connected. The properties of an optimal configuration are further restricted by Corollaries \ref{cor:rows}--\ref{cor:interface}. The following configurations are optimal for the choices of $N_A, N_B > 0$ and $\beta \in (0,1)$ described below. Even though some of them are irregular, the main effort in this paper will be to prove that actually for $N_A = N_B$ and $\beta \le 1/2$ one may find an optimal configuration which is very regular, in the sense that they roughly consist of two rectangles as given in Theorem \ref{thm:main}.v. The first example consists of only three points: we have $N_A = 2$, $N_B = 1$, for any $\beta \in (0,1)$. Even then, the minimiser may fail to be unique: up to isometries, we have two minimisers, both presented in Figure \ref{fig:threepointexample}. \begin{figure}[h!] \includegraphics[scale=0.4]{example_three_points.png} \caption{Minimisers for $N_A = 2, N_B = 1$} \label{fig:threepointexample} \end{figure} The second example consists of six points: we have $N_A = N_B = 3$ for any $\beta \in (0,1)$. The numbers of $A$- and $B$-points are equal. The minimiser may fail to be unique: up to isometries, we have two minimisers, both presented in Figure \ref{fig:sixpointexample}. Note that the interface is not necessarily straight. However, there is a minimiser which has a straight interface. \begin{figure}[h!] \includegraphics[scale=0.35]{example_six_points_new.png} \caption{Minimisers for $N_A = 3, N_B = 3$} \label{fig:sixpointexample} \end{figure} The third example consists of eight points: we have $N_A = N_B = 4$ for any $\beta \in (0,1)$. In this case, the minimiser is unique. Up to isometries, the only solution is presented in Figure \ref{fig:fourpointexample}. Note that the interface is straight and both rectangles are ``full''. This situation is very special, and in a generic case we do not expect uniqueness. \begin{figure}[h!] \includegraphics[scale=0.55]{example_four_points.png} \caption{Unique minimiser for $N_A = 4, N_B = 4$} \label{fig:fourpointexample} \end{figure} The fourth example consists of seven points: we have $N_A = 3$, $N_B = 4$, for any $\beta \in (0,1)$. Up to isometries, we have three minimisers, presented in Figure \ref{fig:sevenpointexample}. As in the second example of Figure~\ref{fig:sixpointexample}, in the configuration on the right the interface is ``L-shaped''. \begin{figure}[h!] \includegraphics[scale=0.23]{example_seven_points.png} \caption{Minimisers for $N_A = 3, N_B = 4$} \label{fig:sevenpointexample} \end{figure} The fifth example consists of ten points: we have $N_A = N_B = 5$ for any $\beta \in (0,1)$. Up to isometries, we have five possible minimisers, presented in Figure \ref{fig:fivepoints}. Notice that the heights of the two types may differ and that the interface may fail to be straight. Furthermore, the two configurations on the left differ even though the interface is straight. \begin{figure}[h!] \centering \includegraphics[scale=0.34]{fivepoints_new.png} \caption{Minimisers for $N_A = 5, N_B = 5$} \label{fig:fivepoints} \end{figure} The final example consists of sixteen points: we have $N_A = 12$ and $N_B = 4$. Then, the situation may differ with $\beta$. For $\beta \in (1/2,1)$, up to isometries we have two possible minimisers (with energy $-20-4\beta$), presented in Figure \ref{fig:sixteenpointexample}. In one case, we have a straight interface, while in the other it is L-shaped. \begin{figure}[h!] \includegraphics[scale=0.31]{example_sixteen_points.png} \caption{Minimisers for $N_A = 12, N_B = 4$, large $\beta$} \label{fig:sixteenpointexample} \end{figure} For $\beta \in (0,1/2)$, up to isometries, we have three possible minimisers (with energy $-21-2\beta$), presented in Figure \ref{fig:sixteenpointexamplebigalpha}. Here, the structure of sets $A$ and $B$ is fixed, but we may attach them in a few different ways. \begin{figure}[h!] \includegraphics[scale=0.26]{example_sixteen_points_v2.png} \caption{Minimisers for $N_A = 12, N_B = 4$, small $\beta$} \label{fig:sixteenpointexamplebigalpha} \end{figure} For $\beta = 1/2$, all configurations presented in Figures \ref{fig:sixteenpointexample} and \ref{fig:sixteenpointexamplebigalpha} are minimal. \section{Classification of admissible configurations}\label{sec:classi} For simplicity, we will call the configurations which satisfy the statement of Theorem \ref{thm:connected} and of the corollaries below it {\it admissible}. In particular, these results show that optimal configurations are admissible. In this section, we collect admissible configurations in different classes. These classes will be analysed in more detail in the subsequent sections. The starting point is the observation that by Corollary \ref{cor:nomissingrows} we have that there cannot be a row $R_{k_0}$ such that $n^{\rm row}_k > 0$ above and below this row (for some $k > k_0$ and some other $k < k_0$), while $n^{\rm row}_{k_0} = 0$. The same result holds for columns. Therefore, we may cluster the minimisers into several classes which are easier to handle and are described using this property. Let us start from the top and suppose that $ n^{\rm row}_1 > 0$ (otherwise, we exchange the roles of the two types). Denote by $R_{k_0}$ the last row such that $n^{\rm row}_{k_0} > 0$. Then, we have the two possibilities \begin{equation}\label{eq:k0=Nr} {\rm (i)} \ \ k_0 = N_{\rm row} \quad \quad \quad \text{and} \quad \quad \quad {\rm (ii)} \ \ k_0 < N_{\rm row}. \end{equation} In case (i), we differ four possibilities, depending on whether $B_1$ and $B^{\rm row}_{N_{\rm row}}$ are empty or not: if $ m^{\rm row}_1, m^{\rm row}_{N_{\rm row}} > 0$, then each row contains points from both types. If $ m^{\rm row}_1$ or $m^{\rm row}_{N_{\rm row}}$ equals zero, then the $B$-part of the configuration has a smaller height. In case (ii), we differ two possibilities, depending on whether $B_1$ is empty or not. If $B_1$ is not empty, then $m^{\rm row}_k > 0$ for all $k = 1,...,N_{\rm row}$. If $B_1$ is empty, then there exists $k_1 > 0$ such that we have $m^{\rm row}_k = 0$ for $k \leq k_1$ and $m^{\rm row}_k > 0$ for $k = k_1 + 1, \ldots ,N_{\rm row}$. By performing the same analysis for columns, and recalling the corollaries after Theorem \ref{thm:connected}, we end up with a number of possibilities which we list below, where without restriction we assume that $ n_1^{\rm col} >0$. This list is complete up to isometries and changing roles of the types. For the sake of the presentation, by applying Corollary~\ref{cor:interface} we can without restriction (possibly up to isometry and changing the roles of the types) assume that the interface is going upwards and to the right. We divide all admissible configurations into five main \textit{classes}, the first three being quite regular and the last two a bit more difficult to handle. In this section, we list all classes and introduce appropriate notation for each of them. In the next section we advance a regularisation procedure for all configurations. This has the aim of proving that for $N_A = N_B$ and $\beta \le 1/2$ all minimal configurations belong to Class $\mathcal{I}$, $\mathcal{IV}$, or $ \mathcal{V}$, as well as checking some fine geometrical properties of such minimisers. \subsection{Class $\mathcal{I}$} The first possibility is the reference case: we say that an admissible configuration $(A,B)$ belongs to Class $\mathcal{I}$ if for each $k = 1,...,N_{\rm row}$ we have $n^{\rm row}_k > 0$ and $m^{\rm row}_k > 0$. In other words, \eqref{eq:k0=Nr}(i) holds with $m^{\rm row}_1, m^{\rm row}_{N_{\rm row}} > 0$. The situation is presented in Figure \ref{fig:ClassIbefore}. Examples of optimal configurations in Class $\mathcal{I}$ can be found in Figure \ref{fig:threepointexample} (on the right), in Figure \ref{fig:sixpointexample} (both), in Figure \ref{fig:fourpointexample}, in Figure \ref{fig:sevenpointexample} (in the middle), in Figure \ref{fig:fivepoints} (all but the two middle ones), and in Figure \ref{fig:sixteenpointexample} (on the right). The abundance of examples in Class $\mathcal{I}$ is in some sense expected. Indeed, we will prove that for many choices of $N_A$, $N_B$, and $\beta$ existence of an optimal configuration in Class $\mathcal{I}$ is guaranteed. \begin{figure}[h] \includegraphics[scale=0.08]{ClassIbefore-eps-converted-to.pdf} \caption{Class $\mathcal{I}$} \label{fig:ClassIbefore} \end{figure} Let us introduce the following notation. Let $h$ denote the number of rows (which in this case corresponds to the number of rows of both $A$ and $B$). Let $l_1$ denote the number of columns such that $A^{\rm col}_k \neq \emptyset$ and $B^{\rm col}_k = \emptyset$. Let $l_2$ denote the number of columns such that $A^{\rm col}_k \neq \emptyset$ and $B^{\rm col}_k \neq \emptyset$. Finally, let $l_3$ denote the number of columns such that $A^{\rm col}_k = \emptyset$ and $B^{\rm col}_k \neq \emptyset$. This notation is also presented in Figure \ref{fig:ClassIbefore}. Then, in view of \eqref{eq:eq}, the energy \eqref{eq: basic eneg} may be expressed as \begin{equation}\label{eq:formulaforenergyclassI} E(A,B) = - 2 (N_A + N_B) + (l_1 + l_2 + l_3) + h + (1-\beta) (l_2 + h). \end{equation} In particular, the energy splits into the \textit{bulk energy} $- 2 (N_A + N_B)$ and, up to a factor $1/2$, into the \textit{lattice perimeter} introduced in \eqref{eq:dbp3}. Clearly, only the latter is relevant for identifying optimal configurations. For convenience, we will frequently refer to it as the surface energy. In the next section, we will simplify the structure of configurations in Class $\mathcal{I}$, without increasing the energy, in order to compute the minimal energy in this class. After such regularisation, it will turn out that we have two possibilities: either $l_2 = 0$ or $l_2 = 1$, i.e., either the interface is a straight line or it has one horizontal jump, see Proposition \ref{prop:classIregularisationstep1}. \subsection{Class $\mathcal{II}$} We say that an admissible configuration $(A,B)$ belongs to Class~$\mathcal{II}$ if there exists a column $C_{k_0}$ such that for all $k \leq k_0$ we have $n^{\rm col}_k > 0$ and $m^{\rm col}_k = 0$, for all $k > k_0$ we have $n^{\rm col}_k = 0$ and $m^{\rm col}_k > 0$, and $(A,B)$ does not lie in Class~$\mathcal{I}$. In other words, the interface is a straight vertical line, and there exists at least one row which contains only one type (as otherwise $(A,B) \in \mathcal{I}$). Examples of optimal configurations in this class can be found in Figure \ref{fig:threepointexample} (on the left), in Figure \ref{fig:sevenpointexample} (on the left), and in Figure \ref{fig:sixteenpointexamplebigalpha} (all of them). Notice that in all these examples we have $N_A \neq N_B$. Indeed, in Section \ref{sec:regularisation} we will show that, if $N_A$ and $N_B$ are equal, such a configuration cannot be optimal. A priori, this set of configurations may arise from both cases in \eqref{eq:k0=Nr}. Up to changing the roles the two types, however, we may assume that we are in situation \eqref{eq:k0=Nr}(i), as we can see in the following simple observation. \begin{lemma}\label{lem:classIIregularisation} Fix $N_A, N_B > 0$ and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{II}$ is a minimal configuration. Then, there exists a minimal configuration $(\hat{A},\hat{B}) \in \mathcal{II}$ such that the last rows align, i.e., $n^{\rm row}_{N_{\rm row}} > 0$ and $m^{\rm row}_{N_{\rm row}} > 0$. \end{lemma} \begin{proof} Without loss of generality, suppose that $n^{\rm row}_{N_{\rm row}} > 0$ and that $r_0 < N_{\rm row}$ is the biggest number such that $m^{\rm row}_{r_0} > 0$. Notice that, since the interface is a straight line, we may move the set $B$ by the vector $(0,r_0 - N_{\rm row})$ so that the last two rows align and this procedure does not increase the energy. The resulting configuration $(\hat{A},\hat{B})$ also lies in Class $\mathcal{II}$: if after this procedure we had also $n^{\rm row}_{1} > 0$ and $m^{\rm row}_{1} > 0$, i.e., $(\hat{A},\hat{B})$ lies in Class $\mathcal{I}$, then we would have added at least one bond. This induces a drop in the energy, a contradiction to the fact that $(A,B)$ is a minimal configuration. \end{proof} After applying this regularisation argument, we introduce the following notation. Up to reflection along the (straight) interface and interchanging the roles of the types, we may assume that $A$ is on the left-hand side and that it has more nonempty rows than $B$. Then, let $h_1$ denote the number of rows such that $A^{\rm row}_k \neq \emptyset$ and $B^{\rm row}_k = \emptyset$, and let $h_2$ be the number of rows such that $A^{\rm row}_k \neq \emptyset$ and $B^{\rm row}_k \neq \emptyset$. Moreover, let $l_1$ denote the number of columns such that $A^{\rm col}_k \neq \emptyset$ and $l_3$ denote the number of columns such that $B^{\rm col}_k \neq \emptyset$ (the notation $l_2$ is omitted on purpose to simplify some later regularisation arguments). Then, arguing as in the justification of formula \eqref{eq:formulaforenergyclassI}, see also \eqref{eq:eq}, the energy \eqref{eq: basic eneg} may be expressed as \begin{equation}\label{eq: energy,class2} E(A,B) = - 2(N_A + N_B) + (l_1 + l_3) + (h_1 + h_2) + (1-\beta) h_2. \end{equation} The situation is presented in Figure \ref{fig:classIInotation}. \begin{figure}[h] \includegraphics[scale=0.08]{ClassIIbefore-eps-converted-to.pdf} \caption{Class $\mathcal{II}$} \label{fig:classIInotation} \end{figure} \subsection{Class $\mathcal{III}$} We say that an admissible configuration $(A,B)$ belongs to Class~$\mathcal{III}$ if for each $k = 1,...,N_{\rm row}$ we have $n^{\rm row}_k > 0$ and for each $l = 1,...,N_{\rm col}$ we have $n^{\rm col}_l > 0$. In other words, each row and each column of $(A,B)$ contains at least one $A$-point (or equivalently, for every $B$-point there is a $A$-point above it and another one to its left). An example of an optimal configuration in this class can be found in Figure~\ref{fig:sixteenpointexample}. Note that in this example the ratio $N_A / N_B$ is far away from $1$. Indeed, in Section \ref{sec:regularisation} we will show that for $N_A = N_B$ configurations in this class cannot be optimal. Counting from the left, let $l_1$ denote the number of columns such that $A^{\rm col}_k \neq \emptyset$ and $B^{\rm col}_k = \emptyset$, let $l_2$ denote the number of columns such that $A^{\rm col}_k \neq \emptyset$ and $B^{\rm col}_k \neq \emptyset$, and let $l_3$ be the number of columns such that $A^{\rm col}_k \neq \emptyset$ and $B^{\rm col}_k = \emptyset$. Similarly, counting from the top, denote by $h_1$ the number of rows such that $A^{\rm row}_k \neq \emptyset$ and $B^{\rm row}_k = \emptyset$, let $h_2$ be the number of rows such that $A^{\rm row}_k \neq \emptyset$ and $B^{\rm row}_k \neq \emptyset$, and finally let $h_3$ be the number of rows such that $A^{\rm row}_k \neq \emptyset$ and $B^{\rm row}_k = \emptyset$. Similarly to previous classes, the energy may be expressed as \begin{equation}\label{eq: nerg3} E(A,B) = - 2(N_A + N_B) + (l_1 + l_2 + l_3) + (h_1 + h_2 + h_3) + (1-\beta) (l_2 + h_2). \end{equation} The situation is presented in Figure \ref{fig:classIII}. \begin{figure}[h] \includegraphics[scale=0.6]{ClassIIIbefore.png} \caption{Class $\mathcal{III}$} \label{fig:classIII} \end{figure} \subsection{Class $\mathcal{IV}$} We say that an admissible configuration $(A,B)$ belongs to Class~$\mathcal{IV}$ if there exist $l_1, l_2, h_1, h_2 > 0$ such that $N_{\rm row} + N_{\rm col} - (l_1+l_2+h_1+h_2)>0$ and the following conditions hold: for each $k = 1,...,l_1$ we have $n^{\rm col}_k > 0$ and $m^{\rm col}_k = 0$. For each $k = l_1+1,...,l_1 + l_2$ we have $n^{\rm col}_k > 0$ and $m^{\rm col}_k > 0$. Finally, for all $k = l_1+l_2+1,...,N_{\rm row}$ (this may possibly be empty) we have $n^{\rm col}_k = 0$ and $m^{\rm col}_k > 0$. Similarly, for each $l = 1,...,h_1$ we have $n^{\rm row}_l > 0$ and $m^{\rm row}_l = 0$. For each $l = h_1+1,...,h_1+h_2$ we have $n^{\rm row}_l > 0$ and $m^{\rm row}_l > 0$. Finally, for all $l = h_1+h_2+1,...,N_{\rm col}$ (this may possibly be empty) we have $n^{\rm row}_l = 0$ and $m^{\rm row}_l > 0$. Setting $l_3 = N_{\rm col} - l_1 - l_2$ and $h_3 = N_{\rm row} - h_1 - h_2$ we observe $l_3>0$ or $h_3>0$, i.e., the configuration does not lie in Class $\mathcal{III}$. The energy may be expressed as \begin{equation}\label{eq:classIVformula} E(A,B) = - 2(N_A + N_B) + (l_1 + l_2 + l_3) + (h_1 + h_2 + h_3) + (1-\beta) (l_2 + h_2). \end{equation} The situation is presented in Figure \ref{fig:ClassIVbefore}. Examples of optimal configurations in this class can be found in Figure~\ref{fig:sevenpointexample} (on the right) and in Figure \ref{fig:fivepoints} (both in the middle). \begin{figure}[h] \includegraphics[scale=0.08]{ClassIVbefore-eps-converted-to.pdf} \caption{Class $\mathcal{IV}$} \label{fig:ClassIVbefore} \end{figure} \subsection{Class $\mathcal{V}$} We say that an admissible configuration $(A,B)$ belongs to Class~$\mathcal{V}$ if there exist $l_1, l_2, l_3, h_1, h_2, h_3 > 0$ such that $l_1 + l_2 + l_3 = N_{\rm col}$, $h_1+h_2+h_3 = N_{\rm row}$ and the following conditions hold: for each $k = 1,...,l_1$ we have $n^{\rm col}_k > 0$ and $m^{\rm col}_k = 0$. For each $k = l_1+1,...,l_1 + l_2$ we have $n^{\rm col}_k > 0$ and $m^{\rm col}_k > 0$. Finally, for all $k = l_1+l_2+1,...,N_{\rm row}$ we have $n^{\rm col}_k > 0$ and $m^{\rm col}_k = 0$. On the other hand, for each $l = 1,...,h_1$ we have $n^{\rm row}_l > 0$ and $m^{\rm row}_l = 0$. For each $l = h_1+1,...,h_1+h_2$ we have $n^{\rm row}_l > 0$ and $m^{\rm row}_l > 0$. Finally, for all $l = h_1+h_2+1,...,N_{\rm col}$ we have $n^{\rm row}_l = 0$ and $m^{\rm row}_l > 0$. The energy may be expressed as \begin{equation*} E(A,B) = - 2(N_A + N_B) + (l_1 + l_2 + l_3) + (h_1 + h_2 + h_3) + (1-\beta) (l_2 + h_2). \end{equation*} The situation is presented in Figure \ref{fig:classVbefore}. \begin{figure}[h] \includegraphics[scale=0.24]{ClassVbefore.png} \caption{Class $\mathcal{V}$} \label{fig:classVbefore} \end{figure} We close this section with the observation that the five classes cover all possible cases up to isometries, reflections, and changing roles of the types. \section{Analysis of Class $\mathcal{I}$}\label{sec:reg1} \subsection{Regularisation inside Class $\mathcal{I}$} The goal of this section is to make the configuration in Class~$\mathcal{I}$ more regular without increasing the energy. This regularisation will facilitate the computation of the minimal energy. We keep the notation as in the previous section, and begin with the following observation. \begin{proposition}\label{prop:classIregularisationstep1} Fix $N_A, N_B > 0$ and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{I}$ is an optimal configuration. Then, we either have $l_2 = 0$ or $l_2 = 1$. \end{proposition} Both cases can happen: take $N_A = N_B = 3$ and $ \beta \in (0,1)$. Then, there are two optimal configurations, one with $l_2 = 0$ and the other one with $l_2 = 1$, see Figure \ref{fig:sixpointexample}. \begin{proof} The idea of the proof is the following: we suppose by contradiction that $l_2 \geq 2$. We add more points to the configuration $(A,B)$, so that it becomes a full rectangle, keeping track of the change of the energy in the process. Then, we exchange a number of points, making the interface shorter and causing a drop in the energy. Finally, we remove the added points, again keeping track of the energy. This yields strictly smaller total energy, a contradiction. The argument is presented in Figure \ref{fig:regularisationClassI}. To be exact, let us modify the configuration $(A,B)$ as follows. We add $N'_A$ $A$-points on the left and $N'_B$ $B$-points on the right such that that $(A,B)$ becomes a full rectangle with sides $l_1 + l_2 + l_3$ and $h$. Notice that in this way we do not alter the surface energy. Meanwhile, the bulk energy changes by $- 2 (N_A' + N_B')$. Now, look at the rectangle in the middle with sides $l_2$ and $h$. If we exchange $A$-points from its rightmost column and $B$-points from its leftmost column (as many as we can), we will make one column (or two) full of points of one type. Hence, in the formula for the energy, see \eqref{eq:formulaforenergyclassI}, we replace $l_2$ by $l_2 - 1$ (respectively $l_2 - 2$), and $l_1 + l_3$ by $l_1 + l_3 + 1$ (respectively $l_1 + l_3 + 2$). This causes a drop in the surface energy by $ (1-\beta)$ or $2(1-\beta)$. Finally, we take care of the added points. We remove $N_A'$ $A$-points, starting from the leftmost column, going from top to bottom. In the process, the surface energy decreases or remains the same (since $l_1$ may decrease or remain the same). Similarly, we remove $N_B'$ $B$-points, starting from the rightmost column and going from top to bottom. In this way, we have obtained a configuration $(\hat{A},\hat{B})$ with the same number of $A$- and $B$-points as $(A,B)$, but with energy lower at least by $ (1-\beta) $. After this operation, we possibly end up with a shape of the interface different from the one in Class $\mathcal{I}$, but this does not matter since we only wanted to show that $(A,B)$ was not optimal. Hence, if $(A,B)$ is an optimal configuration, then $l_2 = 0$ or $l_2 = 1$. \end{proof} \begin{figure}[h] \includegraphics[scale=0.5]{ClassIregularisation.png} \caption{Regularisation of Class $\mathcal{I}$} \label{fig:regularisationClassI} \end{figure} By performing the modification described in the proof, we get that we may assume that the configuration is as compact as possible: given $h$, the values of $l_1$ and $l_3$ are as small as possible, and all the columns except for the leftmost and rightmost ones are full (i.e., have $h$ points). This is a property that we will use several times in the sequel. Now, let us focus on the case $N_A = N_B$. We will give an exact formula for the minimal energy. For this purpose, let us first prove that we may assume that $l_2 = 0$. To this end, let us first state the following technical lemma. \begin{lemma}\label{lem:classIstructurelemma} \color{black} Fix $N := N_A = N_B>0$ \color{black} and $\beta \in (0,1)$ . Suppose that $(A,B) \in \mathcal{I}$ is an optimal configuration such that $l_1 = l_3$ and $l_2 = 1$. Then, we have $l_1 = l_3 \geq h/2$. \end{lemma} \begin{proof} Without restriction we assume that $(A,B)$ has the form described before the statement of the lemma, see also the last picture in Figure \ref{fig:regularisationClassI}. Let $k = \lceil h/2 \rceil$. Suppose by contradiction that the statement does not hold, i.e., $l_1 = l_3 < k$ (in particular, $k \geq 2$). Consider two cases: first, assume that $h$ is even, so that $h = 2k$. Then, the whole configuration fits into a rectangle with height $2k$ and width $2l_1 + 1$, where $l_1 \leq k-1$. Let us rearrange all the points so that the resulting configuration lies in a rectangle with height $2k-1$ and width $2l_1+2$. We place the points by filling the columns from left to right, first with $A$-points and then with $B$-points, so that the resulting configuration lies in Class $\mathcal{I}$ and has $l_2 \leq 1$. In fact, all points may be placed in this rectangle since the assumption $l_1 \leq k-1$ implies $$ (2k-1)(2l_1+2) \geq 2k(2l_1+1).$$ But then the new configuration has strictly smaller energy since $h$ decreased by $1$, $l_2 \leq 1$, and $l_1+l_3$ grew by at most $1$. Hence, the original configuration was not optimal, a contradiction. In the second case, $h$ is odd, so that $h = 2k-1$. Then, the whole configuration fits into a rectangle with height $2k -1$ and width $2l_1 + 1$, where $l_1 \leq k-1$. Let us again rearrange all the points using the procedure from the previous paragraph, so that the resulting configuration lies in a rectangle with height $2k-2$ and width $2l_1+2$ and satisfies $l_2 \leq 1$. Indeed, if $l_1 \leq k-2$, all points may be placed in this rectangle since in this case we have \begin{align}\label{inequili} (2k-2)(2l_1+2) \geq (2k-1)(2l_1+1). \end{align} On the other hand, if $l_1 = k-1$, we have $$(2k-2)(2l_1+2) = (2k-1)(2l_1+1) - 1,$$ so the inequality \eqref{inequili} is not satisfied. In this case, however, $(2k-1)(2l_1+1)$ is odd. Thus, since the total number of points $\color{black} 2 N \color{black}$ is even, it is not possible that the entire rectangle with height $2k$ and width $2l_1 + 1$ was full in the original configuration. Therefore, we can still place all the points in the rectangle with height $2k-2$ and width $2l_1+2$. As before, the new configuration has strictly smaller energy since $h$ decreased by $1$, $l_2 \leq 1$, and $l_1+l_3$ grew by at most $1$: a contradiction. \end{proof} Now, we proceed to prove the main result for Class $\mathcal{I}$, namely that for the purpose of the computation of the minimal energy we may assume that $l_2 = 0$. \begin{proposition}\label{prop:classIregularisationstep2} \color{black} Fix $N := N_A = N_B>0$ and \color{black} $ \beta \in (0,1)$. Then, if $(A,B) \in \mathcal{I}$ is an optimal configuration, then there exists an optimal configuration $(\hat{A},\hat{B}) \in \mathcal{I}$ with $l_2 = 0$. \end{proposition} \begin{proof} If $(A,B) \in \mathcal{I}$ is such that $l_2 = 0$, there is nothing to prove. Suppose to the contrary that $l_2 > 0$. Then, by Proposition \ref{prop:classIregularisationstep1} we have that $l_2 = 1$. We introduce the following notation: again, $l_1$ is the number of columns with only $A$-points and $l_3$ is the number of columns with only $B$-points. We can assume that all columns except for the leftmost and rightmost ones are full, cf.\ last picture in Figure \ref{fig:regularisationClassI}. By $r_1 \in \{ 1,...,h \}$ we denote the number of $A$-points in the leftmost column, and $r_4 \in \{ 1,...,h \}$ is the number of $B$-points in the rightmost column. By $r_2,r_3 \in \{ 1,...,h-1 \}$ we denote the numbers of $A$- and $B$-points, respectively, in the single column which contains points of both types. Since $N_A = N_B$, we compute the number of points of each type and we get \begin{equation*} (l_1 - 1) h + r_1 + r_2 = (l_3 - 1) h + r_3 + r_4, \end{equation*} so \begin{equation}\label{eq: numbering} (l_1 - l_3) h = r_3 + r_4 - r_1 - r_2. \end{equation} Due to the range of $r_1,\ldots,r_4$, the left-hand side can take only values between $-2h+3$ and $2h-3$, so it needs to take values in the set $\{ -h,0,h\}$. Hence, up to exchanging the roles of the two types, we either have $l_1 = l_3$ or $l_1 = l_3 + 1$. First, suppose that $l_1 = l_3 + 1$. Then, by \eqref{eq: numbering} we have $r_1 + r_2 + h = r_3 + r_4$. In particular, $r_1 + r_2 < h$ as $r_3+r_4 \le 2h-1$. Hence, we may move the $r_2$ $A$-points from the single column with both types to the leftmost column, and replace them by $r_2$ $B$-points from the rightmost column. In this way, the double-type column disappeared altogether. This process strictly decreases the energy \eqref{eq:formulaforenergyclassI} since $l_1$ stays the same, $l_2$ decreases by $1$, and $l_3$ increases by $1$ or stays the same. This is a contradiction. Now, suppose that $l_1 = l_3$. Then, by \eqref{eq: numbering} we have $r_1 + r_2 = r_3 + r_4$. If $r_1 + r_2 \leq h$, we proceed as in the previous paragraph. Suppose otherwise, i.e., $r_1 + r_2 = r_3+r_4 > h$. Without restriction we can suppose that $r_3 \ge r_2$. Let $k \in \mathbb{N}$ such that $k = \lceil h/2 \rceil$. Notice that we may modify the configuration so that $r_2 = \lfloor h/2 \rfloor$ and $r_3 = k$. Indeed, otherwise we move $\lfloor h/2 \rfloor - r_2$ ($=r_3 - k$) $B$-points from the double-type column to the rightmost column and move $\lfloor h/2 \rfloor - r_2$ $A$-points from the leftmost column to the double-type column, so that both types have $\lfloor h/2 \rfloor$ and $k$ points, respectively, in the double-type column. In this way, since $$r_4 + \lfloor h/2 \rfloor - r_2 = (r_1+r_2-r_3) + \lfloor h/2 \rfloor - r_2 = r_1 + \lfloor h/2 \rfloor -r_3 \le h,$$ where we used $r_1 \le h$ and $r_3 \ge \lfloor h/2 \rfloor$, we did not add any additional column on the right. Thus, the total energy did not increase. As $l_1=l_3$ and $l_2 = 1$, by Lemma \ref{lem:classIstructurelemma} we have that $l_1 = l_3 \geq k$. Now, remove all the points in the double-type column and place them directly above the first row, $ \lfloor h/2 \rfloor $ $A$-points directly above the $l_1$ $A$-points (starting from the right) and $k$ $B$-points directly above the $l_3$ $B$-points (starting from the left). Finally, we merge the two connected components of the resulting configuration by moving the connected component on the left by $(1,0)$. In this way, $h$ increased by 1, $l_2$ decreased by 1, and $l_1$ and $l_3$ remain unchanged, so that the energy remains the same, see \eqref{eq:formulaforenergyclassI}. Hence, the resulting configuration $(\hat{A},\hat{B})$ is minimal, lies in Class $\mathcal{I}$, and satisfies $l_2 = 0$. This concludes the proof. \end{proof} \subsection{Exact calculation for Class $\mathcal{I}$} The regularisation procedure presented in the previous subsection enables us to compute directly the minimal energy for configurations in Class $\mathcal{I}$ for any $ \beta \in (0,1)$. In this subsection, we suppose that $N_A = N_B$ and denote the common value by \color{black} $N$. \color{black} Later, in Section \ref{sec:regularisation} we will show that there exists always a minimiser in Class~$\mathcal{I}$ which induces that the energy computed below coincides with the minimal energy. \begin{theorem}\label{thm:classIexact} Fix $\color{black} N := N_A = N_B>0$ and \color{black} $ \beta \in (0,1)$. Suppose that a minimal configuration $(A,B)$ is in Class $\mathcal{I}$. Then, its energy is equal to the smaller of the two numbers \begin{align}\label{eq: E*} E_*=-4 \color{black} N \color{black} +2\left\lceil \frac{\color{black} N \color{black}}{ \left\lfloor \sqrt{\frac{2 \color{black} N \color{black}}{ 2 - \beta }}\right\rfloor }\right\rceil + \left\lfloor \sqrt{\frac{2 \color{black} N \color{black}}{ 2 - \beta }}\right\rfloor ( 2 - \beta ) \end{align} and \begin{align}\label{eq: E**} E^*=-4 \color{black} N \color{black}+2\left\lceil \frac{\color{black} N \color{black}}{ \left\lceil \sqrt{\frac{2 \color{black} N \color{black}}{ 2 - \beta }}\right\rceil }\right\rceil+ \left\lceil \sqrt{\frac{2 \color{black} N \color{black}}{ 2 - \beta }}\right\rceil ( 2 - \beta ) \end{align} depending on $\color{black} N \color{black}$ and $ \beta$. \end{theorem} \begin{proof} By Proposition \ref{prop:classIregularisationstep2}, for the purpose of the computation of the minimal energy, we may assume that $l_2 = 0$. Hence, we also have $l_1 = l_3$, and denote the common value by $\ell$. Notice that we may minimise the energy under the constraint $${ h,\,\ell \in \mathbb{N}, \quad \color{black} N \color{black} =h \ell + r \quad \text{with}\ r\in \mathbb{N}, \ 0\leq r\leq h-1.}$$ This constraint is natural since for fixed $h$, the length $\ell$ is minimal whenever all the columns except for the leftmost and rightmost ones are full (i.e., have $h$ points). We also refer to the configuration given in Theorem~\ref{thm:main}.v. Under these assumptions, we may rewrite the energy \eqref{eq:formulaforenergyclassI} as \begin{equation*} E(A,B) = -4 \color{black} N \color{black} + 2 (\ell+\min\lbrace r,1 \rbrace) + h( 2- \beta ). \end{equation*} In particular, one can express $E$ solely in terms of $h\in \mathbb{N}$ as \begin{equation}\label{eq:E} E(h):=-4 \color{black} N \color{black}+ 2\left\lceil \frac{n}{h}\right\rceil + h( 2 - \beta ). \end{equation} Since the function $h \in (0,\infty) \mapsto -4 \color{black} N \color{black} + 2 \color{black} N \color{black}/h + h ( 2 - \beta ) $ is strictly convex and attains its minimum in $\sqrt{2 \color{black} N \color{black}/( 2 - \beta )}$, the minimiser $h$ of $E$ from \eqref{eq:E} is either \begin{align}\label{eq: h***} h_*= \left\lfloor \sqrt{\frac{2 \color{black} N \color{black}}{ 2 - \beta }}\right\rfloor \quad \text{or} \quad h^* = \left\lceil \sqrt{\frac{2 \color{black} N \color{black}}{ 2 - \beta }}\right\rceil. \end{align} In the first case, the minimal value of $E$ equals \eqref{eq: E*} and in the second case it equals \eqref{eq: E**}. Whether $E^*$ or $E_*$ is smaller depends on $n$ and $ \beta $. However, unless $h^* = h_*$, these two numbers cannot be equal for any $ \beta \in (0,1)$. \end{proof} We close this section with the observation that, once we have guaranteed the existence of a minimiser in Class~$\mathcal{I}$ (see Theorem \ref{thm:classIexistence} below), Theorem \ref{thm:main}.iv follows from Theorem \ref{thm:classIexact} and \eqref{eq:eq}. The construction of the configuration in the previous proof, in particular \eqref{eq: h***}, also yields the explicit solution in Theorem \ref{thm:main}.v. \section{Analysis and regularisation of other classes}\label{sec:regularisation}\label{sec:reg2} In this section, we show how to regularise configurations related to classes $\mathcal{II}$--$\mathcal{V}$. Our main goal is to show that for $N_A = N_B$, it is not possible that a minimiser lies in Class~$\mathcal{II}$ or Class~$\mathcal{III}$. While it is possible that a minimiser lies in Class~$\mathcal{IV}$, see Proposition \ref{prop:largeminimisersiv} below, we will show that under the constraint $ \beta \leq 1/2$ we can modify an optimal configuration so that it lies in Class~$\mathcal{I}$. \subsection{Class $\mathcal{II}$} Since the definition of Class $\mathcal{II}$ already involved a very regular interface, namely a straight line, the situation here is much simpler with respect to Class~$\mathcal{I}$. In fact, the whole analysis of the problem boils down to the following simple result. \begin{proposition}\label{prop:classIIregularisation} \color{black} Fix $N:= N_A = N_B>0$ \color{black} and $ \beta \in (0,1)$. If $(A,B)$ is an optimal configuration, then $(A,B) \notin \mathcal{II}$. \end{proposition} \begin{proof} Suppose otherwise. Then, recalling \eqref{eq: energy,class2}, notice that we may rewrite the energy as \begin{equation*} E(A,B) = -4 \color{black} N \color{black} + E_A + E_B - \beta h_2, \end{equation*} where $E_A = l_1 + h_1 + h_2$ and $E_B = l_3 + h_2$ are the energy between the void and $A$ and $B$, respectively, and the last term corresponds to the interface energy. Suppose first that $E_A > E_B$. Then, we modify the configuration as follows: set $\hat{B} = B$ and let $\hat{A}$ be the symmetric image of $B$ under the reflection along the interface. In this way, we obtain \begin{equation*} E(\hat{A},\hat{B}) = -4 \color{black} N \color{black}+ 2E_B - \beta h_2 < -4 \color{black} N \color{black} + E_A + E_B - \beta h_2 = E(A,B), \end{equation*} a contradiction to minimality of $(A,B)$. Now, we suppose $E_A \leq E_B$ instead. We modify the configuration as follows: set $\hat{A} = A$ and let $\hat{B}$ be the symmetric image of $A$ under the reflection along the interface. In this way, the part of the energy corresponding to the shape of $A$ stays the same, the part corresponding to $B$ drops or stays the same, and the length $h_2$ of the interface increases at least by $1$. Hence, the total energy decreases, so $(A,B)$ was not a minimal configuration. \end{proof} \subsection{Class $\mathcal{III}$} Using again the notation introduced in the previous section, our first goal is to show that we can modify an admissible configuration in Class~$\mathcal{III}$ such that we remain in Class~$\mathcal{III}$ and $l_3 = h_3 = 0$ without increasing the energy. Then, we will prove that such a configuration cannot be optimal if $N_A = N_B$. \begin{proposition}\label{prop:classIIIregularisationstep1} Fix $N_A, N_B > 0$ and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{III}$ is a minimal configuration and $l_3 > 0$ (respectively $h_3 > 0$). Then, there exists a minimal configuration $(\hat{A},\hat{B}) \in \mathcal{III}$ with $l_3 = 0$ (respectively $h_3 = 0$). \end{proposition} \begin{proof} Assume that $l_3 > 0$ (the proof in the case $h_3 > 0$ is analogous). Our construction is presented in Figure~\ref{fig:classIIIb}. We will modify the top $h_1$ rows of the configuration $(A,B)$ in the following way: for every $1 \leq k \leq N_{\rm row}$, denote by $x_k$ the first coordinate in the rightmost point of $(A \cup B)_k^{\rm row}$. Then, for $k \leq h_1$, we set $\hat{A}_k := A^{\rm row}_k + (\min\{x_{h_1+1} - x_k,0\},0)$, i.e., each row which has points further to the right than the rightmost point of $B_{h_1 + 1}$ is translated to the left, in such a way that its rightmost point aligns with the rightmost point of $B_{h_1 + 1}$. As we made no modifications inside rows, $E^{\rm row}_k(\hat{A},\hat{B}) = E^{\rm row}_k(A,B)$ for all $k = 1,\ldots,N_{\rm row}$, see \eqref{eq: row energy}. Regarding $E_k^{\rm \color{black} inter \color{black}}$, observe that for $k \geq h_1 + 1$ nothing changed in the configuration, so $E_k^{\rm \color{black} inter \color{black}}(\hat{A},\hat{B}) = E_k^{\rm \color{black} inter \color{black}}(A,B)$. On the other hand, for $k < h_1$, we either left two adjacent rows intact (so the number of connections between them stayed the same); moved both of them to the left so that their rightmost points align (so the number of connections between them stayed the same or increased); or moved only one of them to the left, but because the rightmost point of the other one has first coordinate smaller or equal to the first coordinate of $B_{h_1+1}$, this shift did not destroy any bonds and possibly created new ones. In every case, all these connections are of type $A$-$A$, so we have $E_k^{\rm \color{black} inter \color{black}}(\hat{A},\hat{B}) \leq E_k^{\rm \color{black} inter \color{black}}(A,B)$. Finally, for $k = h_1$, we did not change the number of $A$-$B$ connections and possibly added some $A$-$A$ connections. Thus, $E_k^{\rm \color{black} inter \color{black}}(\hat{A},\hat{B}) \leq E_k^{\rm \color{black} inter \color{black}}(A,B)$. Note that after this procedure all columns $A_l^{\rm col}$ for $l \leq l_1$ are still connected, as otherwise this would contradict Theorem \ref{thm:connected} and the minimality of the original configuration. Hence, the resulting configuration lies in Class $\mathcal{III}$. \end{proof} \begin{figure}[h] \includegraphics[scale=0.2]{ClassIIIpartone_new.png} \caption{Regularisation of Class $\mathcal{III}$: part one} \label{fig:classIIIb} \end{figure} In order to facilitate the proof that configurations in Class~$\mathcal{III}$ cannot be optimal, we further modify the configuration without increasing the energy. \begin{lemma}\label{lem:classIIIregularisationstep2} Fix $N_A, N_B > 0$ and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{III}$ is a minimal configuration. Then, there exists a minimal configuration $(\hat{A},\hat{B}) \in \mathcal{III}$ such that for every $k = 1,...,N_{\rm row}$ the rightmost point of $(\hat{A} \cup \hat{B})_k^{\rm row}$ has the same first coordinate and for every $k = 1,...,N_{\rm col}$ the lowest point of $(\hat{A} \cup \hat{B})_k^{\rm col}$ has the same second coordinate. \end{lemma} \begin{proof} By the previous proposition, we may assume that $l_3 = h_3 = 0$. We will use a version of the technique used for Class~$\mathcal{I}$, and refer to Figure \ref{fig:regularisationClassIII} for an illustration of the construction. Note that if we add $N_A'$ $A$-points on the top and on the left and $N_B'$ $B$-points on in the bottom right corner, so that the configuration $(A,B)$ becomes a full rectangle with sides $l_1 + l_2$ and $h_1 + h_2$, we do not alter the surface energy, but the bulk energy changes by $- 2 (N_A' + N_B')$. Having fixed $N_B' > 0$, let us remove the topmost $A$-point in the leftmost column and change the type of the topmost $B$-point in the leftmost column to $A$. In this way, we removed a $B$-point, without increasing the energy \eqref{eq: nerg3}. We repeat this procedure until we removed $N_B'$ $B$-points. Then, we remove $N_A'$ $A$-points, starting from the top of the leftmost column. Again, this cannot increase the energy. Moreover, the resulting configuration lies in Class $\mathcal{III}$ because if in this last step we removed a whole column or a point which lies next to the interface, we would decrease the energy. Hence, the resulting configuration is also minimal and satisfies the desired property. \end{proof} \begin{figure}[h] \includegraphics[scale=0.18]{ClassIIIregularisation.png} \caption{Regularisation of Class $\mathcal{III}$: part two} \label{fig:regularisationClassIII} \end{figure} These regularisation results imply that in the case when the numbers of points in the two types are equal, then the minimising configuration cannot lie in Class $\mathcal{III}$. \begin{proposition}\label{prop:classIIIexcluded} \color{black} Fix $ N_A = N_B>0$ \color{black} and $ \beta \in (0,1)$. Then, if $(A,B)$ is a minimal configuration, $(A,B) \notin \mathcal{III}$. \end{proposition} \begin{proof} Suppose otherwise and let $(A,B) \in \mathcal{III}$ be a minimal configuration. Apply the regularisation procedure described in Proposition \ref{prop:classIIIregularisationstep1} and Lemma \ref{lem:classIIIregularisationstep2}. After these operations, $(A,B)$ lies in a rectangle $R$ with sides $h_1 + h_2$ and $l_1 + l_2$. Then, the length of the interface equals $l_2 + h_2$. Without loss of generality $h_1 + h_2 \leq l_1 + l_2$ (otherwise, this is true after applying a symmetry with respect to the line $\mathbb{R}(-1,1)$). Then, we compare $(A,B)$ with a configuration $(\hat{A},\hat{B}) \in \mathcal{I}$ which fits into the rectangle $R$, with $A$-points on the left and $B$-points on the right such that the length of the interface is either $h_1 + h_2$ or $h_1 + h_2 + 1$, depending on whether $l_2 =0$ or $l_2 = 1$. Hence, by minimality of $(A,B)$, we have $l_2 + h_2 \leq h_1 + h_2 + 1$, i.e., \begin{align}\label{eq: l2h1} l_2 \leq h_1 + 1. \end{align} This gives a contradiction with the assumption $N_A = N_B$. To see this, first recall that the configuration is {\it full}, in the sense that the construction in Lemma \ref{lem:classIIIregularisationstep2} ensures that all the columns except for the leftmost one have the same number of points. Therefore, we may first estimate from above the number of $B$-points by \begin{equation*} N_B \leq l_2 h_2 \leq h_1 h_2 + h_2 \end{equation*} and the number of $A$-points from below by \begin{align*} N_A & \geq h_1 l_2 + h_1 (l_1 - 1) + h_2 (l_1 - 1) = h_1 (l_1 + l_2) - h_1 + h_2 (l_1 - 1)\\ &\geq h_1 (h_1 + h_2) - h_1 + h_2 (l_1 - 1) = h_1 h_2 + h_1 (h_1 - 1) + h_2 (l_1 - 1), \end{align*} where we used the assumption that $h_1 + h_2 \leq l_1 + l_2$. Hence, whenever $h_1, l_1 \geq 2$ or $l_1 \geq 3$, we have $N_A > N_B$, which would contradict the assumption $N_A = N_B$. Moreover, we get that necessarily $h_1 \leq h_2$. Finally, we have to take into consideration the case when $l_1 = 1$ (with $h_1$ arbitrary) or when $h_1 = 1$ and $l_1 = 2$. In the first case, by \eqref{eq: l2h1} we have $h_1 + h_2 \leq l_2 + 1 \leq h_1 + 2$, so $h_2 \leq 2$. But then $h_1 \leq h_2 \leq 2$, and thus $l_2 \leq h_1 + 1 \leq 3$. This leaves us with a finite (and small) number of configurations to consider separately and it may be checked that none of them is optimal. In the second case, again by \eqref{eq: l2h1} we have $l_2 \leq h_1 + 1 = 2$. Furthermore, $l_1 + l_2 \geq h_1 + h_2$, so $h_1 + h_2 \leq 4$, and hence $h_2 \leq 3$. Again, we end up with a small number of configurations, and it is easy to see that none of them is optimal. \end{proof} \subsection{Class $\mathcal{IV}$, part one}\label{sec:classIVpartone} The situation in Class~$\mathcal{IV}$ is not as clear-cut as in Classes $\mathcal{II}$ and $\mathcal{III}$: whereas configurations in Classes $\mathcal{II}$ and $\mathcal{III}$ are never optimal, the problem is that, even for $N_A = N_B$ and $ \beta = 1/2$, an optimal configuration may actually lie in Class~$\mathcal{IV}$, see Figure~\ref{fig:fivepoints}. Hence, the goal in this subsection is a bit different: we will prove that even though minimal configurations in Class $\mathcal{IV}$ may exist, there also exists an optimal configuration in Class $\mathcal{I}$. Moreover, the reasoning will also provide some further properties of optimal configurations in Class $\mathcal{IV}$. In particular, a careful inspection of the forthcoming constructions will show a fluctuation estimate for minimisers in Class~$\mathcal{IV}$, see Section~\ref{sec:law} below. This goal is achieved as follows: in the first part, we regularise our configuration such that $h_3 = 0$ and $h_1\le l_1$. This is achieved in Proposition \ref{prop:classIVregularisationstep3}, with the key part of the reasoning proved in Proposition \ref{prop:classIVregularisationstep2}. These arguments are valid for any $ \beta \in (0,1)$. Then, in the second part, under the restriction $ \beta \leq 1/2$, we regularise a configuration with $h_3 = 0$ and $h_1\le l_1$ to obtain a configuration in Class $\mathcal{I}$. This is achieved in Propositions \ref{prop:regularisationofclassIVpart4}--\ref{prop:wemayrequireclassI}. We break the reasoning into smaller pieces in order to highlight different techniques and different assumptions required at each point. \begin{lemma}\label{lem:classIVregularisationstep1} Fix $N_A, N_B > 0$ and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{IV}$ is a minimal configuration. Then, there exists a minimal configuration $(\hat{A},\hat{B}) \in \mathcal{I} \cup \mathcal{IV}$ such that $l_2 \leq h_2$, $\min \lbrace h_1, h_1+h_2 -l_1-l_2\rbrace \le 0$, and $\min \lbrace h_3, h_2+h_3 -l_2-l_3\rbrace \le 0$. \end{lemma} \begin{proof} Choose a minimal configuration $(A,B)$ in Class~$\mathcal{IV}$. Without loss of generality, we may assume that $l_2 \leq h_2$. Otherwise, consider a reflection of the original configuration with respect to the line $\mathbb{R}(-1,1)$. Then, we end up with a configuration of the same type with the roles of $h_i$ and $l_i$ reversed. We suppose that $h_1 \ge 1$ as otherwise the second condition in the statement of the lemma is satisfied. We modify the configuration without increasing the energy such that $h_1=0$ or $l_1 + l_2 \ge h_1 + h_2$. To see this, suppose that $l_1 + l_2 < h_1 + h_2$. Then, we remove all the points in the first row, and place them on the left-hand side starting from the second row, one in each row, possibly forming one additional column. The assumption guarantees that there was enough space to place all the points. In this way, $h_1$ decreases by 1 and $l_1$ increases possibly by 1, so the total energy decreases (in which case $(A,B)$ was not a minimal configuration) or stays the same, cf.\ \eqref{eq:classIVformula}. We repeat this procedure until $h_1 = 0$ or $l_1 + l_2 \geq h_1 + h_2$. In a similar fashion, we modify the configuration to obtain $\min \lbrace h_3, h_2+h_3 -l_2-l_3\rbrace \le 0$. Finally, if $h_1=h_3=0$, the configuration is in Class~$\mathcal{I}$. Otherwise, if $h_1 \ge 1$, the configuration is in Class~$\mathcal{IV}$, and if $h_1=0$, $h_3 \ge 1$, after a rotation by $\pi$ and interchanging the roles of the two types we obtain a configuration in Class~$\mathcal{IV}$. \end{proof} We continue the regularisation in the following proposition. \begin{proposition}\label{prop:classIVregularisationstep2} Fix $N_A, N_B > 0$ and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{IV}$ is a minimal configuration. Then, there exists a minimal configuration $(\hat{A},\hat{B}) \in \mathcal{I} \cup \mathcal{IV}$ which satisfies $l_2 \leq h_2$, $\min \lbrace h_1, h_1+h_2 -l_1-l_2\rbrace \le 0$, $\min \lbrace h_3, h_2+h_3 -l_2-l_3\rbrace \le 0$, and at least one of the following two properties: $$(1) \ \ l_2 = 1, \quad \quad \quad (2) \ \ h_3 = 0.$$ \end{proposition} For the proof, we introduce the following notation specific for Class $\mathcal{IV}$. With the notation of Figure \ref{fig:ClassIVbefore}, we will refer to the nine rectangles with sides $l_i$ and $h_j$ as $l_i:h_j$. For instance, the rectangle in the middle with sides $l_2$ and $h_2$ will be referred to as rectangle $l_2:h_2$. A priori, some of these rectangles may be not full or even empty, for instance the rectangle $l_3:h_1$. \begin{proof} Let $(A,B) \in \mathcal{IV}$ be a minimal configuration from Lemma \ref{lem:classIVregularisationstep1} which does not satisfy the desired properties, i.e., $l_2 >1$ and $h_1,h_3 >0$ ($l_2=0$ is not possible as it would imply $(A,B) \in \mathcal{II}$). Then, we first make a similar regularisation as we did for Class~$\mathcal{I}$. We add $N_A'$ $A$-points to the configuration $(A,B)$, so that the interface between $A$ and the void consists of four line segments (of lengths $l_1$, $h_1 + h_2$, $l_1 + l_2$ and $h_1$). This does not increase the surface energy. Then, we remove $N_A'$ $A$-points, column by column, starting from the leftmost column in $(A,B)$. If we removed a whole column, or if we removed a point which lies at the interface, the energy drops, so the original configuration $(A,B)$ was not minimal. Hence, the resulting configuration lies in Class~$\mathcal{IV}$. We proceed in a similar fashion for the $B$-points. In particular, the rectangle $l_2:h_2$ (in the middle) is full. Now, let us look at the (full) rectangle $l_2:h_2$. It contains exactly $l_2 h_2$ points, $N_A''$ of them of type $A$ and $N_B''$ of them of type $B$. We rearrange them (i.e., remove all the points in $l_2:h_2$ and place them back in $l_2:h_2$) in the following way: we start with the leftmost column and we fill the columns one by one with $A$-points until we end up with less than $h_2$ points to place. Then, we place the remaining points in the next column, starting from the top. Similarly, we place the $B$-points starting from the rightmost column and we fill the columns one by one until we end up with less than $h_2$ points. We place the remaining points on the bottom of the next column. In this way, the resulting configuration has an interface with at most one step in $l_2:h_2$, and we did not change the energy. By Lemma \ref{lem:classIVregularisationstep1} we also have \begin{equation}\label{eq:tripleassumption} {\rm (i)} \ \ l_2 \leq h_2, \quad \quad {\rm (ii)} \ \ l_1 \geq h_1, \quad \quad {\rm (iii) } \ \ l_3 \geq h_3. \end{equation} Indeed, (i) is clear. If $h_1 = 0$, (ii) is obvious. Otherwise we have $h_1+h_2 - l_1-l_2 \le 0$ which along with (i) shows (ii). The proof of (iii) is similar. As $l_2 \ge 2$ and the interface has at most one step, we observe that at least one of the following cases holds true: (a) The rightmost column of $l_2:h_2$ consists only of points of type $B$. (b) The leftmost column of $l_2:h_2$ consists only of points of type $A$. Then, we do one of the two following procedures: (a) We move the $A$-points from the rightmost column of the rectangle $l_2:h_1$ (in the upper right corner) to the rectangle $l_1:h_3$ (in the bottom left corner) and place them in its highest row (starting from the right). Here, we use \eqref{eq:tripleassumption}(ii) and $h_3 \ge 1$. In this way, we do not increase the surface energy, see \eqref{eq:classIVformula}, since we have $h_2 \rightarrow h_2 + 1$, $l_2 \rightarrow l_2-1$, $h_3 \rightarrow h_3 - 1$, $l_3 \rightarrow l_3 + 1$, and $h_1$ and $l_1$ remain unchanged. Finally, we perform a rearrangement in the new rectangle $l_2:h_2$ as above. (b) We move the $B$-points from the leftmost column of the rectangle $l_2:h_3$ (in the bottom left corner) to the rectangle $l_3:h_1$ (in the upper right corner) and place them in its lowest row (starting from the left). Here, we use \eqref{eq:tripleassumption}(iii) and $h_1 \ge 1$. In this way, we do not increase the surface energy since we have $h_2 \rightarrow h_2 + 1$, $l_2 \rightarrow l_2-1$, $h_1 \rightarrow h_1 - 1$, $l_1 \rightarrow l_1 + 1$, and $h_3$ and $l_3$ remain unchanged. Finally, we perform a rearrangement in the new rectangle $l_2:h_2$ as above. In both cases, after applying the procedure, the condition \eqref{eq:tripleassumption} is still satisfied, so we may repeat it. We repeat it until $l_2 = 1$, $h_1 = 0$, or $h_3 = 0$. Indeed, this follows after a finite number of steps since in each step $l_2$ decreases. If $l_2 = 1$ or $h_3 = 0$ hold, the proof is concluded. Otherwise, $h_3 = 0$ holds after a rotation by $\pi$ and interchanging the roles of the two types. \end{proof} We now come to the main result of this subsection. \begin{proposition}\label{prop:classIVregularisationstep3} Fix \color{black} $ N_A = N_B>0$ \color{black} and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{IV}$ is a minimal configuration. Then, there exists a minimal configuration $({A},{B}) \in \mathcal{I} \cup \mathcal{IV}$ such that $l_2 \leq h_2$, $h_1 \leq l_1$, and $h_3 = 0$. \end{proposition} \begin{proof} Let $(A,B)$ be a configuration from Proposition \ref{prop:classIVregularisationstep2}. Suppose by contradiction that $(A,B)$ (up to a rotation by $\pi$ and interchanging the roles of the two types) does not have the desired properties. Since \eqref{eq:tripleassumption} holds, we thus get that $l_2=1$ and $h_1, h_3 > 0$. By Proposition~\ref{prop:classIVregularisationstep2} and $h_1, h_3 > 0$ we also have \begin{equation}\label{eq:squareboundinthespecialcase} l_1 + 1 \geq h_1 + h_2 \qquad \mbox{and} \qquad l_3 + 1 \geq h_2 + h_3. \end{equation} As $h_2 \ge 1$, this particularly implies $h_1 \le l_1$. We can thus move the single column $l_2: h_1$ to the empty rectangle $l_1: h_3$ without increasing the energy. Note that $l_1 \ge h_1$ guarantees that there was enough space to place all the points. The resulting configuration has a straight interface with $h_1 >0$, i.e., lies in Class~$\mathcal{II}$. In view of Proposition~\ref{prop:classIIregularisation}, however, this contradicts optimality of the original configuration. \end{proof} Hence, for $N_A = N_B$ and any $ \beta \in (0,1)$, we may require that $h_3 = 0$ and $h_1 \le l_1$. We continue the analysis in the next subsection, with an additional requirement on $ \beta$. \subsection{Class $\mathcal{IV}$, part two}\label{sec:classIVparttwo} From now on, we will work with configurations which satisfy the statement of Proposition \ref{prop:classIVregularisationstep3}, i.e., $h_1 \leq l_1$ and $h_3 = 0$. Our goal is to perform a further modification such that configurations lie in Class $\mathcal{I}$. To this end, we assume without restriction that configurations from Proposition \ref{prop:classIVregularisationstep3} lie in Class~$\mathcal{IV}$ and that $\color{black} N \color{black} := N_A = N_B$. In due course, we will introduce an additional assumption on $ \beta \in (0,1)$. As a first step of the regularisation procedure, we again straighten the interface such that it has at most one step. \begin{lemma}\label{lemma: step lemma} Fix $\color{black} N_A = N_B>0 \color{black}$ and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{IV}$ is an optimal configuration with $h_3 = 0$. Then, there exists a minimal configuration with the same properties and at most one step in the interface. \end{lemma} \begin{proof} We proceed similarly to our reasoning in Class $\mathcal{I}$, i.e., as in the proof of Proposition \ref{prop:classIregularisationstep1}. We add points to the configuration such that the rectangles $l_i:h_j$ for $i=1,2,3$ and $j=1,2$, except for $l_3:h_1$ are full. In this way, the surface part of the energy did not change. Then, we remove the same number of $A$- and $B$-points that we added, starting with the leftmost and rightmost column. If we removed a full column, then the energy would drop and the original configuration would not be minimal. Hence, the rectangle $l_2:h_2$ is necessarily full. Let us now reorganise it in the following way: we put all the $A$-points to the left and all the $B$-points to the right, so that the interface between them (inside $l_2:h_2$) is vertical except for a single possible step to the right. Its length did not change, so the resulting configuration is optimal. \end{proof} \begin{lemma}\label{lem:h1smallerthanh2} Fix $\color{black} N_A = N_B>0 \color{black}$ and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{IV}$ is an optimal configuration such that $h_1 \leq l_1$ and $h_3 = 0$. Then, $h_1 \le h_2$. \end{lemma} \begin{proof} Suppose otherwise, i.e., $h_1 > h_2$. First, we can assume that $l_3 \leq h_2$. Indeed, if not, we can remove the whole rectangle $l_3:h_2$, rotate it by $\pi/2$ and reattach it to the configuration, adding at least one additional bond: a contradiction to minimality of $(A,B)$. Moreover, we can assume that $l_1 \ge 2$ as $l_1 = 1$ implies also $h_1 = 1$, and the inequality $h_1 \leq h_2$ is automatically satisfied. Finally, we can suppose that the interface has at most one step, see Lemma \ref{lemma: step lemma}. The main step of the proof is to show that $l_1 < l_3$. Indeed, then we obtain the contradiction \begin{equation*} h_2 \leq h_1 \leq l_1 < l_3 \leq h_2. \end{equation*} Let us now prove $l_1 < l_3$. To this end, we will calculate the total number of points in two ways. Denote by $r_1$ the number of $A$-points in the leftmost column, by $h_1 + r_2$ the number of $A$-points in the leftmost double-type column, by $r_3$ the number of $B$-points in the leftmost double-type column, and by $r_4$ the number of $B$-points in the rightmost column. Then, we have \begin{equation} N_A = (l_1-1)(h_1 + h_2) + l_2 h_1 + r_1 + r_2 \end{equation} and \begin{equation} N_B = (l_2 + l_3 - 2) h_2 + r_3 + r_4. \end{equation} Now, we subtract one of these equations from the other. Since $r_1> 0, r_2 \geq 0$, and $r_3, r_4 \leq h_2$ we get \begin{align*} 0 & = N_A - N_B = l_1 h_1 + l_1 h_2 + l_2 h_1 - h_1 - h_2 + r_1 + r_2 - l_2 h_2 - l_3 h_2 + 2h_2 - r_3 - r_4 \\ &> (l_1 - 1) h_1 - h_2 + (l_1 - l_3) h_2 + l_2(h_1 - h_2) \geq (l_1 - l_3) h_2, \end{align*} where in the last step we used $l_1 \ge 2$ and the assumption (by contradiction) that $h_1 \geq h_2$. This shows $l_1 < l_3$ and concludes the proof. \end{proof} \begin{proposition}\label{prop:regularisationofclassIVpart4} Fix $\color{black} N_A = N_B>0 \color{black}$ and $ \beta \leq 1/2$. Suppose that $(A,B) \in \mathcal{IV}$ is an optimal configuration such that $h_1 \leq l_1$ and $h_3 = 0$. Then, there exists an optimal configuration $(\hat{A},\hat{B})$ such that $(\hat{A},\hat{B}) \in \mathcal{IV}$ with $h_3 = 0$ and $l_2 \in \{ 1,2 \}$. \end{proposition} \begin{proof} Suppose that $(A,B)$ satisfies $l_2 \geq 3$. By Lemma \ref{lem:h1smallerthanh2} we have $h_1 \leq h_2$. Then, let us remove the rightmost two layers in $l_2:h_1$, and place the (at most $2h_1$) $A$-points on the left of the configuration, at most one point in every row. Since $h_1 \leq h_2$, there is enough space to place all the points. In this way, since the configuration can assumed to have only one step in the interface (see Lemma~\ref{lemma: step lemma}), $l_1$ increases by at most 1, $l_2$ decreases by 2, $l_3$ increases by 2, and all $h_i$ stay the same. Hence, by formula \eqref{eq:classIVformula} we see that the energy stays the same (for $ \beta = 1/2$), so the resulting configuration is optimal, or decreases (for $ \beta < 1/2$), so the original configuration was not optimal. We repeat this procedure until $l_2 \in \lbrace 1,2 \rbrace$. \end{proof} Hence, in order to prove existence of an optimal configuration in Class $\mathcal{I}$, we have two special cases to consider, depending on the value of $l_2$. We start with the case $l_2 = 1$. \begin{proposition}\label{prop:regularisationofclassIVpart5} Fix $\color{black} N_A = N_B>0 \color{black}$ and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{IV}$ is an optimal configuration such that $h_3 = 0$ and $l_2 = 1$. Then, $h_1 = 1$. Furthermore, there exists an optimal configuration $(\hat{A},\hat{B}) \in \mathcal{I}$. \end{proposition} \begin{proof} As in \eqref{eq:formulafortheenergy}, let us write the energy as \begin{equation} E(A,B) = E_A + E_B - (h_2+1)\beta, \end{equation} where $E_A$ is minus the number of bonds between points in $A$ and $E_B$ is minus the number of bonds between points in $B$. We consider two cases. First, suppose that $E_B < E_A$. We do the following rearrangement of points: we separate $A$ and $B$ and suppose without restriction that the leftmost column of $B$ is full as otherwise we can move the points in this column to the right-hand side of $B$, without changing the $E_B$. We replace $A$ by $\hat{A}$, a reflection of $B$ along the vertical axis. Then we reconnect $\hat{A}$ and $B$ along the vertical line segment of length $h_2$. In this way, the resulting configuration has energy \begin{equation} E(\hat{A},B) = E_B + E_B - h_2 \beta. \end{equation} Hence, as $E_B \le E_A -1$, the energy drops by at least $1-\beta$, so the original configuration was not optimal, a contradiction. Now, suppose that $E_A \leq E_B$. We do the following: we keep $A$ fixed (or, as above, we make $A$ flat on one side without changing $E_A$) and replace $B$ by $\hat{B}$, a reflection of $A$ along the vertical axis. Then, we join $A$ and $\hat{B}$ along the vertical line segment of length $h_1 + h_2$. In this way, the resulting configuration lies in Class $\mathcal{I}$, has a flat interface, and the energy is given by \begin{equation} E(A,\hat{B}) = E_A + E_A - (h_1+h_2) \beta. \end{equation} Therefore, the only way in which the energy does not decrease is that $E_A = E_B$ and $h_1 = 1$. \end{proof} We will employ another variant of the reflection argument to deal with the case $l_2 = 2$. This is formalised in the next proposition. \begin{proposition}\label{prop:regularisationofclassIVstep6} Fix $\color{black} N_A = N_B>0 \color{black}$ and $ \beta \leq 1/2$. Suppose that $(A,B) \in \mathcal{IV}$ is an optimal configuration such that $h_3 = 0$ and $l_2 = 2$. Then, $h_1 \le 2+1/\beta$ and there exists an optimal configuration $(\hat{A},\hat{B}) \in \mathcal{I}$. \end{proposition} \begin{proof} Again, as in \eqref{eq:formulafortheenergy}, we write the energy as \begin{equation*} E(A,B) = E_A + E_B - (h_2 + 2)\beta. \end{equation*} We consider three cases: first, suppose that either $E_B \le E_A -2$ or $E_B = E_A -1$ and $h_1 \le 2+1/\beta$. We do the following rearrangement of points: we keep $B$ fixed (up to making one side flat, as in the previous proof) and replace $A$ by $\hat{A}$, a reflection of $B$ along the vertical axis. Then, we join $\hat{A}$ and $B$ along the vertical line segment of length $h_2$. The resulting configuration lies in Class~$\mathcal{I}$ and satisfies \begin{equation} E(\hat{A},B) = E_B + E_B - h_2 \beta. \end{equation} Hence, the energy drops by $k -2\beta$, where $k = E_A - E_B \ge 1$. Thus, either the original configuration was not optimal (for $k \ge 2$ or $k=1$ and $\beta < 1/2$) or the resulting configuration is optimal (for $k=1$ and $\beta = 1/2$). Moreover, the resulting configuration lies in Class $\mathcal{I}$. Now, suppose that either $E_B = E_A - 1$ and $h_1 > 2+1/\beta$ or $E_B = E_A$ and $h_1 \ge 2$ or $E_A < E_B$. This time, we keep $A$ fixed (up to making one side flat) and replace $B$ by $\hat{B}$, a reflection of $A$ along the vertical axis. Then, we join $A$ and $\hat{B}$ along the vertical line segment of length $h_1 + h_2$. In this way, the resulting configuration lies in Class $\mathcal{I}$, has a flat interface, and the energy is given by \begin{equation} E(A,\hat{B}) = E_A + E_A - (h_1+h_2) \beta. \end{equation} Thus, the energy decreases by $k + (h_1-2)\beta$, where $k = E_B-E_A$. In particular, for $k=-1$ and $h_1 > 2+1/\beta$ or $k=0$ and $h_1 \ge 3$ or $k>0$ the energy drops. For $k=0$ and $h_1=2$ it stays the same, so the resulting configuration is optimal and lies in Class $\mathcal{I}$. The only case left to consider is when $E_A = E_B$ and $h_1 = 1$. We proceed as follows: we exchange the rightmost $A$-point (i.e., the rightmost point of the rectangle $l_2:h_1$) with the top $B$-point from column $C_{l_1+1}$, i.e., the point with two connections to points of type $A$ and two connections to points of type $B$. If $C_{l_1+1}$ contains only one $B$-point, then the interface became shorter (without changing the overall shape of the configuration) and the energy actually drops. If it contains more then one $B$-point, this procedure did not change the energy. Moreover, the resulting configuration is in Class $\mathcal{I}$. The construction is presented in Figure \ref{fig:exchangingonepoint}. \begin{figure}[h] \includegraphics[scale=0.23]{ClassIVlaststep.png} \caption{Final step of modification into Class $\mathcal{I}$} \label{fig:exchangingonepoint} \end{figure} Summarising, we have shown that $h_1 \le 2+1/\beta$ and that there exists an optimal configuration in Class~$\mathcal{I}$. \end{proof} We summarise the reasoning from this subsection in the following result. \begin{proposition}\label{prop:wemayrequireclassI} Fix $\color{black} N_A = N_B>0 \color{black}$ and $ \beta \leq 1/2$. Suppose that $(A,B) \in \mathcal{IV}$ is an optimal configuration. Then, there exists an optimal configuration $(\hat{A},\hat{B}) \in \mathcal{I}$. \end{proposition} \begin{proof} By Proposition \ref{prop:classIVregularisationstep3}, there exists an optimal configuration with $h_1 \leq l_1$ and $h_3 = 0$. Since $ \beta \leq 1/2$, by Proposition \ref{prop:regularisationofclassIVpart4} one may require additionally that $l_2 = 1$ or $l_2 = 2$. In both cases, existence of an optimal configuration in Class $\mathcal{I}$ is guaranteed by Proposition \ref{prop:regularisationofclassIVpart5} and by Proposition \ref{prop:regularisationofclassIVstep6}, respectively. \end{proof} \subsection{Class $\mathcal{V}$} Finally, we show that we can modify optimal configurations in Class~$\mathcal{V}$ to optimal configuration in Class $\mathcal{IV}$. Along with Proposition~\ref{prop:wemayrequireclassI} this shows that there always exists a minimiser in Class~$\mathcal{I}$. This is done in the following proposition which employs a similar technique to the one used for Class $\mathcal{III}$. \begin{proposition}\label{prop:classVregularisation} Fix $N_A, N_B > 0$ and $ \beta \in (0,1)$. Suppose that $(A,B) \in \mathcal{V}$. Then, there exists $(\hat{A},\hat{B}) \in \mathcal{IV}$ with $E(\hat{A},\hat{B}) \leq E(A,B)$. \end{proposition} \begin{proof} We will modify the top $h_1$ rows of the configuration $(A,B)$ in a similar fashion to the proof of Proposition \ref{prop:classIIIregularisationstep1}. For every $k \leq h_1$, we set $\hat{A}_k := A^{\rm row}_k + (-1,0)$. This translation implies $E^{\rm row}_k(\hat{A},\hat{B}) = E^{\rm row}_k(A,B)$ for all $k = 1,...,N_{\rm row}$. Regarding $E_k^{\rm \color{black} inter \color{black}}$, a change is possible at most for $k = h_1$, where we did not change the number of $A$-$B$ connections and added zero or one $A$-$A$ connections, so $E_k^{\rm \color{black} inter \color{black}}(\hat{A},\hat{B}) \leq E_k^{\rm \color{black} inter \color{black}}(A,B)$. Hence, the total energy did not increase. We repeat this procedure for all rows wit index $k \le h_1$ until the rightmost point of all $A^{\rm row}_{k}$ with $k \leq h_1$ does not lie right to the rightmost point of $B^{\rm row}_{h_1+1}$. We thus get a configuration which lies in Class~$\mathcal{IV}$. \end{proof} \subsection{Conclusion} Finally, we are in the position to state another of the main results, which together with Theorem \ref{thm:classIexact} gives the exact formula for the minimal energy, see Theorem \ref{thm:main}.iv. \begin{theorem}\label{thm:classIexistence} Fix $\color{black} N_A = N_B>0 \color{black}$ and $ \beta \leq 1/2$. Then, there exists an optimal configuration $(A,B)$ which lies in Class $\mathcal{I}$ and has a straight interface. \end{theorem} \begin{proof} Since the number of points is finite, there exists an optimal configuration. By Theorem \ref{thm:connected} and the discussion below it, it lies in one of the five classes. However, it cannot lie in Class $\mathcal{II}$ by Proposition \ref{prop:classIIregularisation}. It also cannot lie in Class $\mathcal{III}$ by Proposition \ref{prop:classIIIexcluded}. If it lies in Class $\mathcal{V}$, then there exists a minimal configuration in Class $\mathcal{IV}$ by virtue of Proposition \ref{prop:classVregularisation}. If it lies in Class $\mathcal{IV}$, then by Proposition \ref{prop:wemayrequireclassI} there exists a minimal configuration in Class $\mathcal{I}$. Finally, since there is an optimal configuration in Class $\mathcal{I}$, by Proposition \ref{prop:classIregularisationstep2} we may suppose that it has a flat interface. \end{proof} Let us note that in the above theorem we only state that a solution in Class $\mathcal{I}$ exists and that we cannot fully exclude existence of solutions in other classes. In particular, the following result shows that there exist arbitrarily large optimal configurations in Class $\mathcal{IV}$. \begin{proposition}\label{prop:largeminimisersiv} Let $\beta \in (0,1/2]\cap {\mathbb Q}$, $r,\, s \in {\mathbb N}$ with $r/s=1-\beta/2$, and $k \in {\mathbb N}$. Then, the Class-${\mathcal IV}$ configuration $(A,B)$ with \begin{align*} A&=\{(x,y)\in {\mathbb Z}^2 \, \colon \, x \in [-kr+1,0], \ y \in [1,ks]\}\} \cup (1,ks),\\ B&=\{(x,y)\in {\mathbb Z}^2 \, \colon \, x \in [1,kr], \ y \in [0,ks-1]\}\cup (0,0) \end{align*} is optimal. \end{proposition} \begin{proof} Using \eqref{eq:eq} and formula \eqref{eq:classIVformula}, one can directly compute \begin{equation} P(A,B) = 4kr + 2 (ks+1) + 2(1-\beta)(ks+1).\label{eq:to_compare} \end{equation} To prove optimality, it hence suffices to check that $P(A,B)=\min\{P_*,P^*\}$, where $P_*$ and $P^*$ are defined in Theorem \ref{thm:main}.iv for $\color{black} N := \color{black} N_A=N_B=k^2rs+1$. From $\beta \in (0,1/2]$ we get that $s/r = 2/(2-\beta) \in (1,4/3]$. This in particular entails that $s>r \geq 2$, which in turn allows to prove that \begin{align*} \sqrt{\frac{2 \color{black} N \color{black}}{2-\beta}} = \sqrt{\frac{k^2rs + 1}{r/s}} = \sqrt{{k^2s^2 + s/r}} \in (ks,ks+1). \end{align*} In particular, we have checked that \begin{align*} \left\lfloor\sqrt{\frac{2 \color{black} N \color{black}}{2-\beta}}\right\rfloor=ks \quad\text{and}\quad \left\lceil \sqrt{\frac{2 \color{black} N \color{black}}{2-\beta}} \right\rceil=ks+1. \end{align*} One can hence compute \begin{align*} P_*&= 4 \left\lceil \frac{ \color{black} N \color{black}}{\left\lfloor\sqrt{\frac{2 \color{black} N \color{black}}{2-\beta}}\right\rfloor } \right\rceil + 2 \left\lfloor\sqrt{\frac{2 \color{black} N \color{black}}{2-\beta}}\right\rfloor (2-\beta)\\ &=4 \left\lceil \frac{k^2rs +1}{ks } \right\rceil + 2 ks (2-\beta) = 4 \left\lceil kr + 1/ks \right\rceil + 2 ks (2-\beta) =4 kr + 4 + 2 ks (2-\beta). \end{align*} On the other hand, using again the fact that for $s> r \geq 2$ we get that $$\frac{k^2rs +1}{ks+1} \in (kr-1,kr]$$ and we can compute \begin{align*} P^*&= 4 \left\lceil \frac{ \color{black} N \color{black}}{\left\lceil\sqrt{\frac{2 \color{black} N \color{black}}{2-\beta}}\right\rceil } \right\rceil + 2 \left\lceil\sqrt{\frac{2 \color{black} N \color{black}}{2-\beta}}\right\rceil (2-\beta)\\ &=4 \left\lceil \frac{k^2rs +1}{ks+1} \right\rceil + 2(ks+1) (2-\beta) = 4kr + 2(ks+1) (2-\beta). \end{align*} We conclude that $$\min\{P_*,P^*\} = P^* \stackrel{\eqref{eq:to_compare}}{=}P(A,B)$$ which proves that $(A,B)$ is optimal. \end{proof} \section{$N^{1/2}$-law for minimisers}\label{sec:law} In this section, we give a quantitative upper bound on the difference of two optimal configurations, see Theorem \ref{thm:main}.vi. The goal is to prove that, even though in general there is no uniqueness of the optimal configurations and some of them may even not be in Class $\mathcal{I}$, they all have the same approximate shape. In the following, an isometry $T\colon \mathbb{Z}^2 \rightarrow \mathbb{Z}^2$ indicates a composition of the translations $x \mapsto x + \tau$ for $\tau \in \mathbb{Z}^2$, the rotation $(x_1,x_2) \mapsto (-x_2, x_1) $ by the angle $\pi/2$, and the reflections $(x_1,x_2) \mapsto (x_1,-x_2)$, $(x_1,x_2) \mapsto (-x_1,x_2)$. \begin{theorem}[$N^{1/2}$-law]\label{thm:nonehalf} \color{black} Fix $N := N_A = N_B>0$ \color{black} and $ \beta \leq \frac12$. Then, there exists a constant $ C_\beta$ only depending on $ \beta $ such that for each two optimal configurations $(A,B)$ and $(A',B')$ it holds that \begin{equation}\label{eq: N12} \min \bigg\{ \#(A \triangle T(A')) + \#(B \triangle T(B')) \colon \, T\colon \mathbb{Z}^2 \rightarrow \mathbb{Z}^2 \mbox{ is an isometry} \bigg\} \leq C_\beta N^{1/2}. \end{equation} \end{theorem} \begin{proof} Throughout the proof, $ C_\beta $ is a constant which depends only on $ \beta $ whose value may vary from line to line. We start the proof by mentioning that it suffices to check the assertion only for $N \ge N_0$ for some $N_0 \in \mathbb{N}$ depending only on $ \beta $. As observed in the proof of Theorem~\ref{thm:classIexistence}, every optimal configuration lies in the Classes $\mathcal{I}$, $\mathcal{IV}$, $\mathcal{V}$. In Step 1, we show \eqref{eq: N12} for two optimal configurations in Class $\mathcal{I}$. Afterwards, in Step 2 we show that for each optimal configuration $(A,B)$ in Class $\mathcal{IV}$ there exists $(A',B')$ in Class $\mathcal{I}$ such that \eqref{eq: N12} holds. Eventually, in Step 3 we check that for each optimal configuration $(A,B)$ in Class $\mathcal{V}$ there exists $(A',B')$ in Class $\mathcal{IV}$ such that \eqref{eq: N12} holds. The combination of these three steps yields the statement. {\bf Step 1: Class $\mathcal{I}$.} Let first $(A,B)$ be an optimal configuration in Class $\mathcal{I}$ such that $l_2 = 0$. Then by Theorem~\ref{thm:classIexact}, in particular by \eqref{eq: h***} we find \begin{equation}\label{eq: fixing hhh} h \sim \sqrt{\frac{\color{black} 2 \color{black} N}{ 2- \beta }}, \quad \quad \quad l_1=l_3 \sim \sqrt{\frac{N( 2- \beta )}{\color{black} 2 \color{black}}}, \end{equation} where here and in the following $\sim$ indicates that equality holds up to a constant depending only on $ \beta$. Consequently, two optimal configurations in Class $\mathcal{I}$ with $l_2= 0$ clearly satisfy \eqref{eq: N12}. Also, notice that since the interface is straight, reflection along the interface exchanges the roles of the sets $A$ and $B$. Now, consider an optimal configuration $(A,B)$ in Class~$\mathcal{I}$ with $l_2>0$. Then we get $l_2=1$ by Proposition \ref{prop:classIregularisationstep1}. The regularisation of Proposition \ref{prop:classIregularisationstep2} shows that $(A,B)$ can be modified to a configuration $(A',B')$ in Class $\mathcal{I}$ with $l_2 = 0$ such that \eqref{eq: N12} holds. Indeed, in this regularisation we only alter the configurations involving the single column containing points of both types and possibly merge two connected components by moving one connected component by $(1,0)$. This concludes Step 1 of the proof. {\bf Step 2: Class $\mathcal{IV}$.} We now consider an optimal configuration in Class $\mathcal{IV}$ and show that it can be modified to a configuration in Class $\mathcal{I}$ such that \eqref{eq: N12} holds. We will work through the proofs in Subsections \ref{sec:classIVpartone} and \ref{sec:classIVparttwo} in reverse order. Our strategy is as follows: we use the knowledge of the structure of the final step of the regularisation procedure, obtain some a posteriori bounds on the size of $l_i$ and $h_i$, and go back to see how these can change at every step of the regularisation procedure. Eventually, this will allow us to show that already after the first modification described in Lemma \ref{lem:classIVregularisationstep1} we obtain an optimal configuration in Class~$\mathcal{I}$, by moving at most $ C_\beta N^{1/2}$ many points. This will conclude Step 2 of the proof. {\bf Step 2.1.} Our starting points are Propositions \ref{prop:regularisationofclassIVpart5} and \ref{prop:regularisationofclassIVstep6}: recall that applying all the intermediate steps, in the end we have $h_3 = 0$ and we land with an alternative $l_1 = 1$ (which is covered in Proposition \ref{prop:regularisationofclassIVpart5}) or $l_2 = 2$ (which is covered by Proposition \ref{prop:regularisationofclassIVstep6}). In both cases, before applying these propositions, we have \begin{align}\label{eq: 2.1} l_2 \leq 2, \quad h_1 \leq 2 + \frac{1}{\beta}, \quad h_3 = 0, \quad h_2 \sim \sqrt{\frac{\color{black} 2 \color{black} N}{ 2 - \beta }}, \quad l_1,l_3 \sim \sqrt{\frac{N( 2 - \beta )}{\color{black} 2 \color{black}}}. \end{align} In fact, the last conditions follow from \eqref{eq: fixing hhh} (for $h=h_2$) and the reflection procedure described in the propositions. {\bf Step 2.2.} Now, we go a step back in the regularisation procedure. In Proposition~\ref{prop:regularisationofclassIVpart4}, for $ \beta < 1/2$ nothing changes and the same bounds hold. For $ \beta = 1/2$, \eqref{eq: 2.1} yields that $\frac{h_1}{h_2}\to 0$ as $N \rightarrow \infty$. This implies that in Proposition~\ref{prop:regularisationofclassIVpart4}, for sufficiently large $N$, we move at most two layers. In fact, if we moved at least three layers, the energy would strictly decrease since all of them fit into a single column. Hence, for sufficiently big $N$ (depending only on $ \beta$), we have the following bounds \begin{align}\label{eq: 22} l_2 \leq 4, \quad h_1 \leq 2 + \frac{1}{\beta}, \quad h_3 = 0, \quad h_2 \sim \sqrt{\frac{\color{black} 2 \color{black} N}{ 2 - \beta }}, \quad l_1,l_3 \sim \sqrt{\frac{N( 2 - \beta )}{\color{black} 2 \color{black}}}. \end{align} Finally, let us take one more step back in the regularisation procedure. In Lemma~\ref{lemma: step lemma}, we actually modify the configuration only slightly inside the rectangle $l_2:h_2$. In this way, $h_i$ and $l_i$ were not altered, so that the bounds \eqref{eq: 22} still holds. {\bf Step 2.3.} Now we come to the main part of the regularisation procedure, i.e., Proposition~\ref{prop:classIVregularisationstep2}. In its proof, we apply an iterative procedure, and at every step one of the following changes happens: \begin{align*} {\rm (a)} \ \ \ h_2 \rightarrow h_2 + 1, \quad l_2 \rightarrow l_2-1, \quad h_3 \rightarrow h_3 - 1, \quad l_3 \rightarrow l_3 + 1, \quad h_1 \rightarrow h_1, \quad l_1 \rightarrow l_1 \end{align*} or \begin{align*} {\rm (b)} \ \ \ h_2 \rightarrow h_2 + 1, \quad l_2 \rightarrow l_2-1, \quad h_1 \rightarrow h_1 - 1, \quad l_1 \rightarrow l_1 + 1, \quad h_3 \rightarrow h_3, \quad l_3 \rightarrow l_3. \end{align*} Notice that in both cases $l_1$ and $l_3$ cannot decrease during this procedure, and exactly one of them increases at every step. The procedure can end in two ways: $h_3 = 0$ (or equivalently $h_1 = 0$) or $l_2 = 1$. In the latter case, however, the proof of Proposition \ref{prop:classIVregularisationstep3} implies that the original configuration was not optimal, so we only need to examine the former case. Consider the last step of the regularisation procedure in the proof of Proposition~\ref{prop:classIVregularisationstep2}, i.e., the one before we reach $h_3 = 0$. Denote by $\hat{h}_1$ the value of $h_1$ at the end of the regularisation procedure, and note that $\hat{h}_1 \leq 2 + \frac{1}{\beta}$ by \eqref{eq: 22}. There are two possible situations: either \begin{equation} \hat{l}_1 \leq 2 \hat{h}_1 \qquad \mbox{or} \qquad \hat{l}_1 > 2 \hat{h}_1. \end{equation} In the second case, notice that we cannot have applied the construction from case (a) twice as otherwise a slightly modified procedure would give the following: we move the $A$-points from the rightmost two columns of the rectangle $l_2:h_1$ to the rectangle $l_1:h_3$, but we place them in a single row. In this way, we have \begin{equation} h_2 \rightarrow h_2 + 1, \quad l_2 \rightarrow l_2-2, \quad h_3 \rightarrow h_3 - 1, \quad l_3 \rightarrow l_3 + 2, \quad h_1 \rightarrow h_1, \quad l_1 \rightarrow l_1. \end{equation} This shows that the energy \eqref{eq:classIVformula} strictly decreases as the length of the interface is decreased. Hence, the original configuration was not optimal, so either $ \hat{l}_1 \leq 2 \hat{h}_1$ or we have applied a step of type (a) at most once. Similarly, since $\hat{h}_3$ at the end of the procedure equals zero, we consider the alternative \begin{equation*} \hat{l}_3 \leq 2 \qquad \mbox{or} \qquad \hat{l}_3 > 2. \end{equation*} We apply a similar argument to conclude that either $ \hat{l}_3 \leq 2 $ or that we have applied a step of type (b) at most once. In view of \eqref{eq: 22}, and because $l_1$ and $l_3$ can only increase during the regularisation procedure, we see that $l_1 \leq 2 \hat{h}_1$ and $l_3 \leq 2$ lead to contradictions for $N$ sufficiently large depending only $ \beta $. This implies that there can be at most one step of type (a) and (b), respectively. Therefore, using again \eqref{eq: 22} we see that before the application of Proposition~\ref{prop:classIVregularisationstep2} it holds that \begin{align}\label{eq: 23} l_2 \leq 6, \quad h_1 \leq 4 + \frac{1}{\beta}, \quad h_3 \le 2, \quad h_2 \sim \sqrt{\frac{\color{black} 2 \color{black} N}{ 2 - \beta }}, \qquad l_1,l_3 \sim \sqrt{\frac{N( 2 - \beta )}{\color{black} 2 \color{black}}}. \end{align} {\bf Step 2.4.} Finally, we consider the modification in Lemma \ref{lem:classIVregularisationstep1}. For simplicity, we only address the modification leading to $\min \lbrace h_1, h_1+h_2 -l_1-l_2\rbrace \le 0$. Note that each step of the procedure consists in $h_1 \rightarrow h_1- 1$ and $l_1 \rightarrow l_1 +1$. As after the application of Lemma \ref{lem:classIVregularisationstep1} we have $\hat{h}_2/\hat{l}_1 \ge 2/( 2 - \beta ) + {\rm O}(1/\sqrt{N})$, see \eqref{eq: 23}, and during its application $h_2$ does not change and $l_1$ can only increase, at each step of the procedure it holds that ${h}_2/{l}_1 \ge 2/( 2 - \beta ) + {\rm O}(1/\sqrt{N})$. In view of \eqref{eq: 23}, in particular the fact that $l_2 \leq 6$, for $N$ sufficiently large depending only on $\beta$ we have \begin{align}\label{eq: good con} (h_1+ h_2) / (l_1+l_2) \geq h_2/(l_1 + l_2) \ge c_\beta \end{align} at each step of the procedure, for some constant $ c_\beta >1$ only depending on $ \beta $. This ensures that at the beginning we have $h_1 \le M$ for $M \in \mathbb{N}$ such that $(M+1)/M < c_\beta $ since otherwise $M+1$ rows could be moved to $M$ columns leading to a strictly smaller energy. This along with \eqref{eq: 23} shows that at most $ C_\beta N^{1/2}$ are moved. Moreover, the modifications stops once $h_1=0$ or $h_1 + h_2 \le l_1 + l_2$ as been obtained. By \eqref{eq: good con} we see that it necessarily holds $h_1 = 0$. In a similar fashion, one gets $h_3 = 0$. This shows that directly after the application of Lemma \ref{lem:classIVregularisationstep1} we obtain a configuration in Class~$\mathcal{I}$. This concludes the proof as we have seen that in the modification of Lemma \ref{lem:classIVregularisationstep1} only $ C_\beta N^{1/2}$ points are moved. {\bf Step 3: Class $\mathcal{V}$.} We now consider an optimal configuration in Class $\mathcal{V}$ and show that it can be modified to a configuration in Class $\mathcal{IV}$ such that \eqref{eq: N12} holds. The modification in Proposition~\ref{prop:classVregularisation} consists in moving at most $h_1$ rows to the left. By Step 2 we know that $h_1\le C_\beta $ which implies that we have moved at most $ C_\beta N^{1/2}$ many points. This concludes the proof of Step 3. \end{proof} Let us highlight that in the proof of Theorem \ref{thm:nonehalf} we have not only shown the $N^{1/2}$-law for minimisers, but we also get explicit estimates on the shape of the configuration, written as a separate corollary. This is a consequence of equations \eqref{eq: 23}, \eqref{eq: good con}, and the procedure from Step 3 of the proof of Theorem \ref{thm:nonehalf}. \begin{corollary} Suppose that $(A,B) \in \mathcal{IV} \cup \mathcal{V} $ is an optimal configuration. Then, \begin{align*} l_2, h_1, h_3 \leq C_\beta , \quad h_2 \sim \sqrt{\frac{\color{black} 2 \color{black} N}{ 2- \beta }}, \quad l_1,l_3 \sim \sqrt{\frac{N( 2- \beta )}{\color{black} 2 \color{black}}}. \end{align*} \end{corollary} Finally, let us note that the quantitative bound given in Theorem \ref{thm:nonehalf} is sharp: the optimal configuration in Class $\mathcal{IV}$ given by Proposition \ref{prop:largeminimisersiv} differs from the one given in Theorem \ref{thm:main}.v by a number of points of exactly this order. \section{Proofs in the continuum setting}\label{sec:wulff} We conclude by providing the proofs of Corollaries \ref{cor: wulff} and \ref{cor: cryst-db} from the Introduction. \begin{proof}[Proof of Corollary \ref{cor: wulff}] For the explicit solution $(A'_{\color{black} N \color{black}},B'_{\color{black} N \color{black}})$ in Theorem \ref{thm:main}.v with $N_A=N_B=: \color{black} N \color{black}$, one can directly verify that $\mu_{A'_{\color{black} N \color{black}}} \stackrel{\ast}{\rightharpoonup} {\mathcal L} \mres {\mathcal A}$ and $\mu_{B'_{\color{black} N \color{black}}} \stackrel{\ast}{\rightharpoonup} {\mathcal L} \mres {\mathcal B}$, where $\mathcal{A}$ and $\mathcal{B}$ are given in \eqref{eq:wulff}. For a general sequence of solutions $(A_{\color{black} N \color{black}},B_{\color{black} N \color{black}})$ of \eqref{eq:dbp}, the statement follows from the fluctuation estimate in Theorem \ref{thm:main}.vi. \end{proof} \begin{proof}[Proof of Corollary \ref{cor: cryst-db}] We start by relating point configurations with sets of finite perimeter: given $(A_{\color{black} N \color{black}},B_{\color{black} N \color{black}})$ with $N_A = N_B =: \color{black} N \color{black} $, we define the sets \begin{align}\label{eq: sofp} A^{\color{black} N \color{black}}:= \frac{1}{\sqrt{\color{black} N \color{black}}} {\rm int}\Big(\bigcup_{p \in A_{\color{black} N \color{black}}} p + [-\tfrac{1}{2}, \tfrac{1}{2}]^2\Big), \quad \quad \quad B^{\color{black} N \color{black}} := \frac{1}{\sqrt{\color{black} N \color{black}}} {\rm int}\Big( \bigcup_{p \in B_{\color{black} N \color{black}}} p + [-\tfrac{1}{2}, \tfrac{1}{2}]^2\Big). \end{align} Clearly, $A^{\color{black} N \color{black}}$ and $B^{\color{black} N \color{black}}$ satisfy $A^{\color{black} N \color{black}} \cap B^{\color{black} N \color{black}} = \emptyset$ and $\mathcal L(A^{\color{black} N \color{black}})=\mathcal L(B^{N})=1$. It is an elementary matter to check that \eqref{eq:dbp} and \eqref{eq:dbp2} coincide in this case up to normalisation, i.e., \begin{align}\label{eq: scal-en-es} \color{black} N \color{black}^{-1/2} P(A_{\color{black} N \color{black}},B_{\color{black} N \color{black}}) = P_{\rm cont}(A^{\color{black} N \color{black}},B^{\color{black} N \color{black}}) := {\rm Per} (A^{\color{black} N \color{black}}) + {\rm Per} (B^{\color{black} N \color{black}}) - 2\beta {\rm L} (\partial^* A^{\color{black} N \color{black}} \cap \partial^* B^{\color{black} N \color{black}}). \end{align} Now, consider any pair of sets of finite perimeter with $A \cap B = \emptyset$ and $\mathcal L(A)=\mathcal L(B)=1$. Given $\varepsilon>0$, by the density result \cite[Theorem 2.1 and Corollary 2.4]{BraidesDensity} (for $\mathcal{Z}$ consisting of three values representing $A$, $B$, and the emptyset) we can find $A'$ and $B'$ with polygonal boundary such that $A' \cap B' = \emptyset$, $\mathcal L(A')=\mathcal L(B')=1$, and $$P_{\rm cont}(A',B') \le P_{\rm cont}(A,B) + \varepsilon. $$ (Strictly speaking, the constraint $\mathcal L(A')=\mathcal L(B')=1$ has not been addressed there. However, possibly after scaling one can assume that $\mathcal L(A')\le 1$, $\mathcal L(B') \le 1$, and then it suffices to add a disjoint squares of small volume and surface to satisfy the constraint.) We define a point configuration related to $A'$ and $B'$ by setting $$A_{\color{black} N \color{black}} = \lbrace p \in \mathbb{Z}^2 \colon \, p/\sqrt{\color{black} N \color{black}} \in A' \rbrace, \quad \quad \quad B_{\color{black} N \color{black}} = \lbrace p \in \mathbb{Z}^2 \colon \, p/\sqrt{\color{black} N \color{black}} \in B' \rbrace. $$ By $A^{\color{black} N \color{black}}$ and $B^{\color{black} N \color{black}}$ we denote the corresponding sets of finite perimeter defined in \eqref{eq: sofp}. Note that the sets $A^{\color{black} N \color{black}}$ and $B^{\color{black} N \color{black}}$ may have different cardinalities, although $\mathcal L(A')=\mathcal L(B')=1$. Still, equal cardinalities can be restored by adding points to one of the two sets. This can be achieved at the price of making a small error in the perimeter, which goes to $0$ with $\color{black} N \color{black}$ after rescaling. The fact that $(A', B')$ have polygonal boundary along with the properties of $\Vert \cdot \Vert_1$ implies that $$\lim_{\color{black} N \color{black} \to \infty} P_{\rm cont}(A^{\color{black} N \color{black}},B^{\color{black} N \color{black}}) = P_{\rm cont}(A',B').$$ This along with \eqref{eq: scal-en-es} and Theorem \ref{thm:main}.iv yields \begin{align*} P_{\rm cont}(A,B) & \ge \lim_{\color{black} N \color{black}\to \infty} P_{\rm cont}(A^{\color{black} N \color{black}},B^{\color{black} N \color{black}}) - \varepsilon \ge \lim_{\color{black} N \color{black} \to \infty} \color{black} N \color{black}^{-1/2}\min\{P_*,P^*\} -\varepsilon \\ & = 4 \frac{1}{\sqrt{\frac{2}{2-\beta}} } +2 \sqrt{\frac{2}{2-\beta}} (2-\beta) -\varepsilon = 4\sqrt{2}\sqrt{2-\beta} -\varepsilon. \end{align*} We directly compute $P_{\rm cont}(\mathcal A,\mathcal B) = 4\sqrt{2}\sqrt{2-\beta}$. As $\varepsilon>0$ is arbitrary, we conclude that the pair $(\mathcal A,\mathcal B)$ is a solution of \eqref{eq:dbp2}. \end{proof} \section*{Acknowledgements} MF acknowledges support of the DFG project FR 4083/3-1. This work was supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany's Excellence Strategy EXC 2044 -390685587, Mathematics M\"unster: Dynamics--Geometry--Structure. WG acknowledges support of the FWF grant I4354, the OeAD-WTZ project CZ 01/2021, and the grant 2017/27/N/ST1/02418 funded by the National Science Centre, Poland. US acknowledges support of the FWF grants I4354, F65, I5149, and P\,32788, and by the OeAD-WTZ project CZ 01/2021. The authors are indebted to Frank Morgan for pointing out many relevant references.
{ "timestamp": "2022-03-29T02:44:44", "yymm": "2109", "arxiv_id": "2109.01697", "language": "en", "url": "https://arxiv.org/abs/2109.01697", "abstract": "We investigate minimal-perimeter configurations of two finite sets of points on the square lattice. This corresponds to a lattice version of the classical double-bubble problem. We give a detailed description of the fine geometry of minimisers and, in some parameter regime, we compute the optimal perimeter as a function of the size of the point sets. Moreover, we provide a sharp bound on the difference between two minimisers, which are generally not unique, and use it to rigorously identify their Wulff shape, as the size of the point sets scales up.", "subjects": "Metric Geometry (math.MG); Mathematical Physics (math-ph); Combinatorics (math.CO)", "title": "The double-bubble problem on the square lattice", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9854964207345893, "lm_q2_score": 0.8128673201042492, "lm_q1q2_score": 0.8010778344948553 }
https://arxiv.org/abs/1612.00214
A remark on local fractional calculus and ordinary derivatives
In this short note we present a new general definition of local fractional derivative, that depends on an unknown kernel. For some appropriate choices of the kernel we obtain some known cases. We establish a relation between this new concept and ordinary differentiation. Using such formula, most of the fundamental properties of the fractional derivative can be derived directly.
\section{Introduction} Fractional calculus is a generalization of ordinary calculus, where derivatives and integrals of arbitrary real or complex order are defined. These fractional operators may model more efficiently certain real world phenomena, specially when the dynamics is affected by constraints inherent to the system. There exist several definitions for fractional derivatives and fractional integrals, like the Riemann--Liouville, Caputo, Hadamard, Riesz, Gr\"{u}nwald--Letnikov, Marchaud, etc. (see e.g., \cite{Kilbas,Podlubny} and references therein). Although most of them are already well-studied, some of the usual features for what concerns the differentiation of functions fails, like the Leibniz rule, the chain rule, the semigroup property, among others. As it was mentioned in \cite{Babakhani}, ``These definitions, however, are non-local in nature, which makes them unsuitable for investigating properties related to local scaling or fractional differentiability". Recently, the concept of local fractional derivative have gained relevance, namely because they kept some of the properties of ordinary derivatives, although they loss the memory condition inherent to the usual fractional derivatives \cite{Anderson,Chen,Katumgapola,Kolwankar,Kolwankar2}. One question is what is the best local fractional derivative definition that we should consider, and the answer is not unique. Similarly to what happens to the classical definitions of fractional operators, the best choice depends on the experimental data that fits better in the theoretical model, and because of this we find already a vast number of definitions for local fractional derivatives. \section{Local fractional derivative} We present a definition of local fractional derivative using kernels. \begin{definition} Let $k:[a,b]\to\mathbb R$ be a continuous nonnegative map such that $k(t)\not=0$, whenever $t>a$. Given a function $f:[a,b]\to\mathbb R$ and $\alpha\in(0,1)$ a real, we say that $f$ is $\alpha$-differentiable at $t>a$, with respect to kernel $k$, if the limit \begin{equation}\label{def}f^{(\a)}(t):=\lim_{\epsilon\to0}\frac{f(t+\epsilon k(t)^{1-\alpha})-f(t)}{\epsilon}\end{equation} exists. The $\alpha$-derivative at $t=a$ is defined by $$f^{(\a)}(a):=\lim_{t\to a^+}f^{(\a)}(t),$$ if the limit exists. \end{definition} Consider the limit $\alpha\to1^-$. In this case, for $t>a$, we obtain the classical definition for derivative of a function, $f^{(\a)}(t)=f'(t)$. Our definition is a more general concept, compared to others that we find in the literature. For example, taking $k(t)=t$ and $a=0$, we get the definition from \cite{Batarfi,Cenesiz,Hammad,Hesameddini,Khalil} (also called conformable fractional derivative); when $k(t)=t-a$, the one from \cite{Abdeljawad,Anderson2,Unal}; for $k(t)=t+1/\Gamma(\alpha)$, the definition in \cite{Atangana,Atangana2}. The following result is trivial, and we omit the proof. \begin{theorem} Let $f:[a,b]\to\mathbb R$ be a differentiable function and $t>a$. Then, $f$ is $\alpha$-differentiable at $t$ and $$f^{(\a)}(t)=k(t)^{1-\alpha}f'(t), \quad t>a.$$ Also, if $f'$ is continuous at $t=a$, then $$f^{(\a)}(a)=k(a)^{1-\alpha}f'(a).$$ \end{theorem} However, there exist $\alpha$-differentiable functions which are not differentiable in the usual sense. For example, consider the function $f(t)=\sqrt t$, with $t\geq0$. If we take the kernel $k(t)=t$, then $f^{(\a)}(t)=1/2 \, t^{1/2-\alpha}$. Thus, for $\alpha\in(0,1/2)$, $f^{(\a)}(0)=0$ and for $\alpha=1/2$, $f^{(\a)}(0)=1/2$. In general, if we consider the function $f(t)=\sqrt[n]{t}$, with $t\geq0$ and $n\in\mathbb N\setminus \{1\}$, we have $f^{(\a)}(t)=1/n \, t^{1/n-\alpha}$ and so $f^{(\a)}(0)=0$ if $\alpha\in(0,1/n)$ and for $\alpha=1/n$, $f^{(\a)}(0)=1/n$. \begin{theorem} If $f^{(\a)}(t)$ exists for $t>a$, then $f$ is differentiable at $t$ and $$f'(t)=k(t)^{\alpha-1} f^{(\a)}(t).$$ \end{theorem} \begin{proof} It follows from $$\begin{array}{ll} f'(t)&=\displaystyle \lim_{\delta\to0}\frac{f(t+\delta)-f(t)}{\delta}\\ &=\displaystyle k(t)^{\alpha-1} \lim_{\epsilon\to0}\frac{f(t+\epsilon k(t)^{1-\alpha})-f(t)}{\epsilon}\\ &=\displaystyle k(t)^{\alpha-1} f^{(\a)}(t). \end{array}$$ \end{proof} Of course we can not conclude anything at the initial point $t=a$, as was discussed before. Combining the two previous results, we have the main result of our paper. \begin{theorem}\label{MainT} A function $f:[a,b]\to\mathbb R$ is $\alpha$-differentiable at $t>a$ if and only if it is differentiable at $t$. In that case, we have the relation \begin{equation}\label{MainF}f^{(\a)}(t)=k(t)^{1-\alpha}f'(t), \quad t>a.\end{equation} \end{theorem} \section{Conclusion} In this short note we show that some of the existent notions about local fractional derivative are very close related to the usual derivative function. In fact, the $\alpha$-derivative of a function is equal to the first-order derivative, multiplied by a continuous function. Also, using formula \eqref{MainF}, most of the results concerning $\alpha$-differentiation can be deduced trivially from the ordinary ones. In the authors opinion, local fractional calculus is an interesting idea and deserves further research, but definitions like \eqref{def} are not the best ones and a different path should be followed. \section*{Acknowledgments} Research supported by Portuguese funds through the CIDMA - Center for Research and Development in Mathematics and Applications, and the Portuguese Foundation for Science and Technology (FCT-Funda\c{c}\~ao para a Ci\^encia e a Tecnologia), within project UID/MAT/04106/2013 (R. Almeida) and by the Warsaw School of Economics grant KAE/S15/35/15 (T. Odzijewicz).
{ "timestamp": "2016-12-02T02:04:31", "yymm": "1612", "arxiv_id": "1612.00214", "language": "en", "url": "https://arxiv.org/abs/1612.00214", "abstract": "In this short note we present a new general definition of local fractional derivative, that depends on an unknown kernel. For some appropriate choices of the kernel we obtain some known cases. We establish a relation between this new concept and ordinary differentiation. Using such formula, most of the fundamental properties of the fractional derivative can be derived directly.", "subjects": "Classical Analysis and ODEs (math.CA)", "title": "A remark on local fractional calculus and ordinary derivatives", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.985496421586532, "lm_q2_score": 0.8128673110375457, "lm_q1q2_score": 0.8010778262521677 }
https://arxiv.org/abs/2207.02060
A sharp Korn's inequality for piecewise $H^1$ space and its application
In this paper, we revisit Korn's inequality for the piecewise $H^1$ space based on general polygonal or polyhedral decompositions of the domain. Our Korn's inequality is expressed with minimal jump terms. These minimal jump terms are identified by characterizing the restriction of rigid body mode to edge/face of the partitions. Such minimal jump conditions are shown to be sharp for achieving the Korn's inequality as well. The sharpness of our result and explicitly given minimal conditions can be used to test whether any given finite element spaces satisfy Korn's inequality, immediately as well as to build or modify nonconforming finite elements for Korn's inequality to hold.
\section{Introduction}\Label{Intro} The Korn's inequality played a fundamental role in the development of linear elasticity. There is a work reviewing Korn's inequality and its applications in continuum mechanics \cite{horgan1995korn}. In fact, there are a lot of works on proving classical Korn's inequality \cite{nitsche1981korn,wang2003korn,ciarlet2010korn} for functions in $H^1$ vector fields. In \cite{brenner2004korn}, Korn’s inequalities for piecewise $H^1$ vector fields are established. We note that a strengthened version of Korn’s inequality for piecewise $H^1$ vector fields is presented in \cite{mardal2006observation}. By the Korn's inequality established in \cite{mardal2006observation}, the nonconforming finite element for 2D introduced in \cite{mardal2002robust} is shown to satisfy the Korn's inequality. In this paper, we revisit Korn's inequality for the piecewise $H^1$ space based on general polygonal or polyhedral decompositions of the domain. Our Korn's inequality is expressed with minimal jump terms, which can readily be used to design degree of freedoms. Namely, we move one step further from the result presented in \cite{mardal2006observation}. In particular, for 3D, we identify that the tangential component of rigid body mode restricted to the face of triangle is indeed the lowest Raviart-Thomas element on face, which is the rotated version discussed in \cite{mardal2006observation}. Additionally, with the minimal jump terms, we show that our Korn's inequality is sharp. We emphasize that with the help of sharpness of the Korn's inequality and explicit form of jump conditions, it is very easy to test whether any given finite element spaces satisfy the Korn's inequality. The proposed minimal jump conditions can also provide a guidance to modify some existing nonconforming finite elements so that the Korn's inequality is satisfied. Throughout the paper, we shall use the standard notation for Sobolev spaces. Namely, $H^k(\Omega)$ denotes the Sobolev space of scalar functions in $\Omega$ whose derivatives up to order $k$ are square integrable, with the norm $\|\cdot\|_k$. The notation $|\cdot |_k$ denotes the semi-norm derived from the partial derivatives of order equal to $k$. Furthermore, $\|\cdot\|_{k,T}$ and $|\cdot|_{k,T}$ denote the norm $\|\cdot\|_k$ and the semi-norm $|\cdot|_k$ restricted to the domain $T$. Given a partitions $\mathcal{T}_h$ of the domain $\Omega$, where $\Omega$ is a bounded connected open polyhedral domain in $\Reals{d}$ with $d = 2$ or $3$. , we shall also use $H^1(\Omega;\mathcal{T}_h)$ to denote the element-wise $H^1$ functions. We denote the vectors of size $d$ whose components are in $H^1(\Omega;\mathcal{T}_h)$ by $(H^1(\Omega;\mathcal{T}_h))^d$. We also denote $H({\rm div};\Omega)$ by the space consisting of vectors, whose divergence belongs to $L^2(\Omega)$. We use $P_\ell(T)$ for the space of polynomials of degree upto $\ell$ on the domain $T$ while $(P_\ell(T))^d$ denotes the vectors of size $d$ whose components are polynomials of degree at most $\ell$. We recall that the operator ${\bm{curl}}$ on a scalar function $q$ in $2D$ is defined by \begin{equation} {\bm{curl}} q = \left ( -\frac{\partial q}{\partial y}, \frac{\partial q}{\partial x} \right )^T, \end{equation} while on a vector function $\bm{q} = (q_i)_{i=1,2,3}^T$ in $3D$, it is defined by \begin{equation} {\bm{curl}} \bm{q} = \left ( \frac{\partial q_2}{\partial z} - \frac{\partial q_3}{\partial y}, \frac{\partial q_3}{\partial x} - \frac{\partial q_1}{\partial z}, \frac{\partial q_1}{\partial y} - \frac{\partial q_2}{\partial x} \right )^T, \end{equation} For any given vector space $\bm{V}$, by ${\rm dim} \bm{V}$, we mean the dimension of $\bm{V}$. The rest of our paper is organized as follows. In \S \ref{korncd}, we prove the Korn's inequality for piecewise $H^1$ space with minimal jump terms. \S \ref{sharp} considers the sharpness of the inequality, which indicates that the minimal jump terms presented in our Korn's inequality are necessary and can not be reduced any more. With the Korn's inequality for piecewise $H^1$ space with minimal jump terms and sharpness discussion, in \S \ref{appl}, some of existing finite elements have been discussed. And some application of our theory is provided to modify nonconforming finite elements which do not satisfy the Korn's inequality so that they satisfy the Korn's inequality. Lastly, we provide a concluding remark. \section{Korn's inequality for the piecewise $H^1$ vector functions}\label{korncd} Let $\Omega$ be a bounded connected open polyhedral domain in $\Reals{d}$ with $d = 2$ or $3$ and $\partial \Omega$ be the boundary of the domain $\Omega$. Then the classical Korn's inequality reads as follows: \begin{equation} |{\bf{u}} |_{H^1(\Omega)} \leq C_\Omega \left ( \| \mathcal{D}({\bf{u}})\|_0 + \|{\bf{u}}\|_0 \right ), \quad \forall {\bf{u}} \in [H^1(\Omega)]^d, \end{equation} where the strain tensor $\mathcal{D}({\bf{u}}) \in \Reals{d\times d}$ is given as follows: \begin{equation} \mathcal{D}_{ij}({\bf{u}}) = \frac{1}{2} \left ( \frac{\partial u_i}{\partial x_j} + \frac{\partial u_j}{\partial x_i} \right ) \quad \mbox{ for } 1 \leq i,j \leq d. \end{equation} Let ${\bm{RM}}(\Omega)$ be the space of rigid motions on $\Omega$ defined by \begin{equation} \bm{RM} (\Omega) = \{ \bm {a} + \bm{A} \bm{x} \in \Reals{d} : \bm{a} \in \Reals{d} \,\, \mbox{ and } \,\, \bm{A} \in {\textsf{Skw}}^{d\times d} \}, \end{equation} where $\bm{x} = (x_1, \cdots, x_d)^T \in \Reals{d}$ is the position vector function on $\Omega$ and ${\textsf{Skw}^{d\times d}}$ is the set of anti-symmetric $d\times d$ matrices. We would like to remark that if $\bm{v} = \bm{v}(\bm{x}) \in \bm{RM}(\Omega)$, then $\bm{v} = \bm{a} + \bm{A}\bm{x}$ for some constant vector $\bm{a}$ and a homogeneous polynomial of degree one $\bm{A}\bm{x}$ such that $\bm{x} \cdot \bm{A} \bm{x} = 0$. This is relevant to the lowest N\'ed\'elec first kind finite element \cite{monk2003finite}. It is easy to verify that the space $\bm{RM}(\Omega)$ is the kernel of the strain tensor, i.e., it holds that \begin{equation} \bm{RM}(\Omega) = \{\bm{v} \in (H^1(\Omega))^d : \mathcal{D}(\bm{v}) = 0 \}. \end{equation} \subsection{Korn's inequality with explicit upper bounds} Given $\Omega \subset \Reals{d}$ with $d = 2$ or $3$, we consider the mesh generation $\mathcal{T}_h$, which is a shape-regular polygonal and/or polyhedral partition for $\Omega$. Let $\mathcal{T}_h = \cup \{T\}$ denote the collection of decompositions and $h = \max_{T \in \mathcal{T}_h} {\rm diam} (T)$ denote the mesh size. We further denote by $\mathcal{E}_h$ the set of all edges/faces for $\mathcal{T}_h$ and \begin{equation} \mathcal{E}_h = \mathcal{E}_h^{o} \cup \mathcal{E}_h^\partial, \end{equation} where $\mathcal{E}_h^o$ denotes the set of all interior edges/faces of $\mathcal{T}_h$ and $\mathcal{E}_h^{\partial}$ denotes the set of all boundary edges/faces, respectively. Let $\mathcal{V}(T)$ be the set of all vertices for $T \in \mathcal{T}_h$ while $\mathcal{V} (\mathcal{T}_h)$ denotes the set of all vertices for the partition $\mathcal{T}_h$. Given two adjacent polygon/polyhedron $T^+$ and $T^-$ in $\mathcal{T}_h$, let $f = \partial T^+ \cap \partial T^-$ be the common boundary (interface) between $T^+$ and $T^-$ in $\mathcal{T}_h$, and $n^+$ and $n^-$ be unit normal vectors to $f$ pointing to the exterior of $T^+$ and $T^-$, respectively. For any edge (or face) $f \in \mathcal{E}_h^{o}$, and a scalar $q$ and vector $\bm{v}$, we define the jumps \begin{subeqnarray} \jump{q}_f &=& q|_{\partial T^+\cap f} - q|_{\partial T^-\cap f}, \\ \jump{\bm{v}}_f &=& \bm{v}|_{\partial T^+\cap f} - \bm{v}|_{\partial T^-\cap f}. \end{subeqnarray} When $f \in \mathcal{E}_h^{\partial}$ then the above quantities are defined as \begin{equation} \jump{q}_f = q|_{f}, \quad \mbox{ and } \quad \jump{\bm{v}} = \bm{v}|_{f}. \end{equation} Throughout the paper, we shall also consider the subspace of $\bm {RM}(\Omega)$, denoted by $\bm {RM}^\partial (\Omega)$ defined as follows: \begin{equation} \bm {RM}^{\partial}(\Omega) = \left \{\bm m \in \bm {RM}(\Omega): ~ \|\bm m\|_{L^2(\partial \Omega)}=1,\int_{\partial \Omega}\bm mds=0 \right \}. \end{equation} We also consider the following space of piecewise linear vector fields: \begin{eqnarray*} \bm{V}_h &=& \{\bm v\in (L^2(\Omega))^d: \bm v_T = \bm v|_T \in (P_1(T))^d, \quad \forall T \in \mathcal {T}_h\} \end{eqnarray*} and space of continuous piecewise linear vector fields: \begin{eqnarray*} \bm{W}_h &=& \{\bm v\in (H^1(\Omega))^d: \bm v_T =\bm v|_T \in (P_1(T))^d, \quad \forall T \in \mathcal{T}_h\}. \end{eqnarray*} Consider a linear map $E: \bm{V}_h \longrightarrow \bm{W}_h$ defined as follows: \begin{equation} E (\bm v)(p) = \frac{1}{|\chi_p|}\sum_{T \in \chi_p}\bm v|_T(p), \quad \forall p \in \mathcal{V}(\mathcal{T}_h), \end{equation} where $\chi_p = \{ T \in \mathcal{T}_h: p \in \mathcal{V}(T) \}$, the patch of the vertex $p$, i.e., the collection simplexes that contain $p$ as its vertex, and $|\chi_p|$ is the cardinality of the set $\chi_p$. We note that it holds true \begin{equation} |\chi_p| \lesssim 1, \quad \forall p \in \mathcal{V}(\mathcal{T}_h). \end{equation} We recall the following approximation property, which can be found in \cite{brenner2004korn}: \begin{Lemma} Let $\bm v \in \bm{V}_h$ and $T \in \mathcal{T}_h$. Then, with $\bm v_{T} = \bm v|_T$, for all $p \in \mathcal{V}(\mathcal{T}_h)$, we have the following estimate: \begin{equation} |\bm v_{T} - E(\bm v)(p)|^2 \lesssim \sum_{f \in \mathcal{E}_p} |\jump{\bm v}_f(p)|^2, \quad T \in \mathcal{T}_h \, \mbox{ and } \, p \in \mathcal{V}(T), \end{equation} where \begin{equation} \mathcal{E}_p = \{ e \in \mathcal{E}_h^o : p \in \partial f \}, \end{equation} is the set of interior sides sharing $p$ as a common vertex, and $\jump{\bm v}_f$ is the jump of $\bm v$ across the edge/face $f$. \end{Lemma} The main observation that will lead us to construct the minimal jump conditions or refined Korn's inequality will be presented at the following simple but important theorem: \begin{Theorem}\label{kornpre} For any $T \in \mathcal{T}_h$. Let $f \in \partial T$ and $\bm{c}_{f}$ be the barycenter of $f$. For some constant $c, c_1, c_2 \in \Reals{}$, we have for 2D and 3D, respectively, with $\bm{t}_f$ and $\bm{n}_f$ being the tangent vector to the edge and the normal vector to the face, \begin{subeqnarray*} {\bm{RM}}(f)^\perp &=& \{ (\bm{v} \cdot \bm{t}_f) |_f: \forall \bm{v} \in \bm{RM}(T) \} = {\rm span} \{ 1\} \quad \mbox{ for } d = 2 \\ &=& {\rm RT}_0(f) := \{ (\bm{v} \times \bm{n}_f)|_f : \forall \bm{v} \in \bm{RM}(T) \} \quad \mbox{ for } d = 3, \end{subeqnarray*} where ${\rm RT}_0(f)$ can be characterized as follows: \begin{equation} {\rm RT}_0(f) = \{ (c_1 + c x_1) \bm{t}_1 + (c_2 + c x_2) \bm{t}_2 : x_1 , x_2 \in \Reals{} \}, \end{equation} where $f = {\rm span} \{\bm{t}_1, \bm{t}_2 \}$ such that $\bm{t}_1 \cdot \bm{t}_2 = 0$. \end{Theorem} \begin{proof} We begin our proof for the {\rm 2D} case. Choose $\bm{v} \in {\bm{RM}}(T)$. Then, $\bm{v}$ is of the following form: \begin{equation} \bm{v} = \bm{a} + b (y, -x)^T, \end{equation} for some constant vector $\bm{a}$ and a constant $b$. Without loss of generality, we assume that $f \in \partial T$ can be expressed as a linear function $y = mx + n$. Then, we have that $\bm{t}_f =\frac{1}{\sqrt{1+m^2}} (1,m)^T$ and \begin{equation} (\bm{v} \cdot \bm{t}_f) |_f= c_0, \end{equation} where $c_0$ is some constant. We shall now consider {\rm 3D} case. For any $\bm{v} \in {\bm{RM}}(T)$ can be given as $\bm{v} = \bm{a} + \bm{A} \bm{x}$, where $\bm{A}$, $\bm{x}$, and $\bm{n}_f$ can be denoted by the followings: \begin{eqnarray*} \bm{A} = \left ( \begin{array}{ccc} 0 & a_1 & a_2 \\ -a_1 & 0 & a_3 \\ -a_2 & -a_3 & 0 \end{array} \right ), \quad \bm{x} = \left ( \begin{array}{c} x_1 \\ x_2 \\ x_3 \end{array} \right ) , \quad \mbox{ and } \quad \bm{n}_{f} = \left ( \begin{array}{c} n_1 \\ n_2 \\ n_3 \end{array} \right ). \end{eqnarray*} Then a simple calculation leads to \begin{eqnarray*} \bm{v} \times \bm{n}_{f} = \bm{a} \times \bm{n}_f + \left ( \begin{array}{c} a_1 x_2 + a_2 x_3 \\ -a_1 x_1 + a_3x_3 \\ -a_2x_1 - a_3x_2 \end{array} \right ) \times \left ( \begin{array}{c} n_1 \\ n_2 \\ n_3 \end{array} \right ). \end{eqnarray*} Particularly, we pay attention to the following quantity: \begin{eqnarray*} \bm{A} \bm{x} \times \bm{n}_{f} = \left ( \begin{array}{c} \eta x_1 + a_3 \bm{n}_f \cdot \bm{x} \\ \eta x_2 - a_2 \bm{n}_f \cdot \bm{x} \\ \eta x_3 + a_1 \bm{n}_f \cdot \bm{x} \end{array} \right ) = \left ( \begin{array}{c} \eta x_1 + a_3 \bm{n}_f \cdot (\bm{x} - \bm{c}_f) + a_3 \bm{n}_f \cdot \bm{c}_f \\ \eta x_2 - a_2 \bm{n}_f \cdot ( \bm{x} -\bm{c}_f ) - a_2 \bm{n}_f \cdot \bm{c}_f \\ \eta x_3 + a_1 \bm{n}_f \cdot ( \bm{x} - \bm{c}_f ) + a_1 \bm{n}_f \cdot \bm{c}_f \end{array} \right ) \end{eqnarray*} where $\eta = \bm{n}_f \cdot (-a_3, a_2, -a_1) = -a_1 n_3 + a_2 n_2 - a_3 n_1$ and $\bm{c}_f$ is the barycenter of the face $f$. In case $\eta = 0$, $(\bm{A} \bm{x} \times \bm{n}_f)|_f = \bm{n}_f \cdot \bm{c}_f (a_3, -a_2, a_1)^T.$ Note that $(a_3, -a_2, a_1)^T \in f$ and so, it can be expressed as $\bm{\mu} \times \bm{n}_f$ for some vector $\bm{\mu}$. Therefore, $(\bm{v} \times \bm{n}_f)|_f = \bm{b} \times \bm{n}_f$ with $\bm{b} = \bm{a} + \bm{\mu}$. In case $\eta \neq 0$, we observe that for $\bm{x} \in f$, \begin{eqnarray*} \bm{A} \bm{x} \times \bm{n}_{f} = \eta (\bm{x} - \bm{c}), \end{eqnarray*} where \begin{equation} \bm{c} = \frac{1}{\eta} \left ( \begin{array}{c} -a_3 \bm{n}_{f} \cdot \bm{c}_f \\ a_2 \bm{n}_{f} \cdot \bm{c}_f \\ - a_1 {\bm{n}}_{f} \cdot \bm{c}_f \end{array} \right ) \quad \mbox{ and } \quad (\bm{x} - \bm{c})\cdot \bm{n}_f = 0. \end{equation} This means that we can modify $\bm{A}\bm{x} \times \bm{n}_f$ further into the following, with $c = \eta$, \begin{equation} \bm{A} \bm{x} \times \bm{n}_{f} = c (\bm{x} - \bm{c}) = c (\bm{x} - \bm{c}_ f + \bm{c}_f - \bm{c}) = \bm{d} + c (\bm{x} - \bm{c}_f), \end{equation} where $\bm{d}$ is a constant vector such that $\bm{d} \cdot \bm{n}_f = 0$. Therefore, we arrive at the conclusion and this completes the proof. \end{proof} For 3D, we remark that for any given $T \in \mathcal{T}_h$ with $f$ being a face of $T$, the dimension of the space $\{\bm{q}(\bm{x}) \times \bm{n}_f : \bm{q} \in \bm{RM}(T), \,\, \forall \bm{x} \in f\}$ is three and we denote the space $\{ (c_1 + c x_1) \bm{t}_1 + (c_2 + c x_2) \bm{t}_2 : x_1, x_2 \in \Reals{} \}$ by ${\rm RT}_0(f)$ \cite{xie2008uniformly}. Now, we can prove the Korn's inequality for functions in $(H^1(\Omega,\mathcal{T}_h))^d$. We note that the spaces of projection onto the face of each $T$ are different for {\rm 2D} and {\rm{3D}}, the detailed jump terms shall be stated accordingly. We first define $\pi_1$ by the $L^2$ orthogonal projection from $L^2(f)$ onto $P_1(f)$ and $\pi_{RM^\perp(f)}$ is the $L^2$ orthogonal projection form $L^2(f)$ onto $P_0(f)$ for $d = 2$, but it is from $L^2(f)$ to ${{\rm RT}}_0(f)$ for $d = 3$, respectively. To state the discrete Korn's inequality, we shall first introduce on each $T \in \mathcal{T}_h$, a projection operator $\Pi_T$ from $(H^1(T))^d$ onto the $\bm{RM}(T)$ by the following conditions: \begin{equation} \left|\int_{T}(\bm v - \Pi_{T}\bm v)dx\right|=0, \quad \forall \bm v \in (H^1(T))^d, \end{equation} \begin{equation} \left|\int_{T}\nabla \times(\bm v-\Pi_{T}\bm v) dx\right| = 0, \quad \forall \bm v \in (H^1(T))^d. \end{equation} Hence, by the definition of $\Pi_T$, we have (see (3.3) in \cite{brenner2004korn}) \begin{equation} \label{3.3} | \bm v-\Pi_{T}\bm v |_{H^1(T)} \lesssim \| \mathcal{D} (\bm v - \Pi_T \bm v) \|_0 = \|\mathcal{D} (\bm v) \|_0, \quad \forall \bm v \in (H^1(T))^d. \end{equation} \begin{equation} \label{3.4} \|\bm v-\Pi_{T}\bm v\|_0\lesssim (\hbox{diam} T)|\bm v-\Pi_{T}\bm v|_{H^1(T)}, \quad \forall \bm v\in (H^1(T))^d. \end{equation} Using this local projection $\Pi_T$, we define $\Pi: (H^1(\Omega,\mathcal{T}_h))^d\longrightarrow \boldsymbol{V}_h$ by \begin{equation} (\Pi \bm u) = \Pi_T\bm u_T, \quad \forall T \in \mathcal{T}_h. \end{equation} Next, following \cite{brenner2004korn}, we also introduce a seminorm on $(H^1(\Omega;\mathcal{T}_h))^d$, denoted by $\Phi$ satisfying the following Assumptions: \begin{itemize} \item[(C1)] $|\Phi(\bm w)| \lesssim \|\bm w\|_{1}, \quad \forall \bm{w} \in (H^1(\Omega))^d,$ \item[(C2)] $\Phi(\bm{m}) = 0$ and $\bm{m} \in {\bf{RM}}(\Omega)$ if and only if $\bm{m}$ is constant. \item[(C3)] $(\Phi(\bm v - E \bm v))^2 \lesssim \sum_{f \in \mathcal{E}_h^o}(\hbox{diam}(f))^{d-2}\sum_{p\in \mathcal{V}(f)}|\jump{\bm v}_f(p)|^2, \, \forall \bm v \in \bm{V}_h$, where $\mathcal{V}(f)$ is the set of the vertices of $f$. \end{itemize} The first estimate is well-known for the piecewise $H^1$ functions, see \cite{brenner2004korn} : \begin{Lemma} Let $\Phi$ be the seminorm on $H^1(\Omega;\mathcal{T}_h)$ satisfying the Assumptios (C1), (C2), and (C3). Then, the following estimate holds: \begin{equation}\label{2.9} |\bm v|^2_{H^1(\Omega,\mathcal{T}_h)} \lesssim \|\mathcal{D}_{\mathcal {T}}(\bm v)\|_0^2+ (\Phi(\bm v))^2+\sum_{f \in \mathcal{E}^o_h}(\hbox{diam}~ f)^{d-2}\sum_{p\in \mathcal{V}(f)}|\jump{\bm v}_f(p)|^2 \end{equation} for all $\bm v \in \bm{V}_h$, where $\mathcal{D}_{\mathcal {T}}(\bm v)|_T=\mathcal{D}(\bm v|_T)$ for all $T \in \mathcal T_h$. \end{Lemma} Now, in order to refine the discrete Korn's inequality, we further impose an addition Assumption for $\Phi$ as follows: \begin{itemize} \item[(C4)] $|\Phi(\bm u-\Pi\bm u)| \lesssim \|\mathcal D_\mathcal{T}(\bm u)\|_0, \quad \forall \bm u \in (H^1(\Omega,\mathcal{T}_h))^d$. \end{itemize} \begin{remark} For $u\in (H^1(\Omega))^d$, Assumption (C1) implies Assumption (C4), but for $u \in (H^1(\Omega,\mathcal{T}_h))^d$, Assumption (C4) can not be derived from Assumption (C1). \end{remark} We then obtain the following main result in this paper: \begin{Theorem}\label{main} Let $\Phi:(H^1(\Omega, \mathcal{T}_h))^d\longrightarrow \Reals{}$ be a seminorm satisfying the Assumptions (C1), (C2), (C3) and (C4). We have the following results for 2D and 3D, respectively. For $2D$, we have \begin{eqnarray*} && |\bm u|^2_{H^1(\Omega,\mathcal{T}_h)} \lesssim \|\mathcal{D}_{\mathcal {T}}(\bm u)\|_0^2 + (\Phi(\bm u))^2 \\ &&+ \sum_{f \in \mathcal{E}^o_h}(\hbox{diam}(f))^{-1} \Big(\|\left [\pi_1(\jump{\bm u}_f \cdot \bm n_f) \right ] \bm{n}_f \|_{0,f}^2 + \| \left [\pi_{RM^\perp(f)}(\jump{\bm u}_f \cdot \bm t_f) \right ] \bm t_f \|_{0,f}^2\Big). \end{eqnarray*} For $3D$, we have \begin{eqnarray*} && |\bm u|^2_{H^1(\Omega,\mathcal{T}_h)} \lesssim \|\mathcal{D}_{\mathcal {T}}(\bm u)\|_0^2 + (\Phi(\bm u))^2 \\ &&+ \sum_{f \in \mathcal{E}^o_h}(\hbox{diam}(f))^{-1}\Big(\|\left [ \pi_1(\jump{\bm u}_f\cdot \bm n_f) \right ] \bm{n}_f \|_{0,f}^2+\| \left [\pi_{RM^\perp(f)}(\jump{\bm u}_f \times \bm n_f) \right ] \times \bm{n}_f\|_{0,f}^2\Big). \end{eqnarray*} \end{Theorem} \begin{proof} Let $\bm u\in (H^1(\Omega,\mathcal{T}_h))^d$, from \eqref{2.9} and \eqref{3.3}, we have \begin{equation*}\label{3.8} \begin{split} |\bm u|^2_{H^1(\Omega,\mathcal{T}_h)}&\lesssim |\bm u-\Pi \bm u|^2_{H^1(\Omega,\mathcal{T}_h)}+ |\Pi \bm u|^2_{H^1(\Omega,\mathcal{T}_h)}\\ &\lesssim \|\mathcal{D}_{\mathcal {T}}(\bm u)\|_0^2+ (\Phi(\Pi\bm u))^2+\sum_{f \in \mathcal{E}^o_h}(\hbox{diam}(f))^{d-2}\sum_{p\in \mathcal{V}(f)}|\jump{\Pi \bm u}_f(p)|^2. \end{split} \end{equation*} Using condition (C4), we find \begin{equation}\label{3.8} \Phi(\Pi \bm u)\leq \Phi(\bm u-\Pi\bm u)+\Phi(\bm u)\lesssim \|\mathcal{D}_{\mathcal {T}}(\bm u)\|_0+\Phi(\bm u). \end{equation} Let $f \in \mathcal{E}^o_h$ be arbitrary and $p \in \mathcal V(f)$, we have, by inverse estimate \begin{equation}\label{3.9} |\jump{\Pi\bm u}_f(p)|^2\lesssim (\hbox{diam}(f))^{1-d} \| \jump{\Pi\bm u}_f\|_{0,f}^2. \end{equation} We first prove the $d = 2$ case. Let $f = T^{+} \cap T^{-}$ and choose $\bm{n}_f = \bm n^{-}$ as the unit normal vector and $\bm{t}_f = \bm t^{-}$ as the unit tangential vector of $f$. Then we have that \begin{equation} \begin{split} \|\jump{\Pi\bm u}_f\|_{0,f}^2 &= \int_f |(\jump{\Pi\bm u}_f \cdot \bm{n}_f )\bm{n}_f |^2ds + \int_f|(\jump{\Pi\bm u}_f \cdot \bm{t}_f)\bm{t}_f |^2 ds \\ &= \int_f (\jump{\Pi\bm u}_f \cdot \bm{n}_f)^2 +(\jump{\Pi\bm u}_f \cdot \bm{t}_f)^2 ds. \end{split} \end{equation} Using Theorem \ref{kornpre}, we see that \begin{equation} \jump{\Pi\bm u}_f \cdot \bm{t}_f = c \end{equation} for some constant $c$. Therefore, we have that \begin{equation}\label{3.10} \begin{split} \|\jump{\Pi\bm u}_f\|_{0,f}^2&=\int_f \Big(\pi_1(\jump{\Pi\bm u}_f \cdot \bm{n}_f )\Big)^2 \, ds + \int_f \Big(\pi_{RM^\perp(f)}(\jump{\Pi\bm u}_f\cdot \bm{t}_f)\Big)^2\, ds \\ &\leq \int_f \Big(\pi_1(\jump{\Pi \bm u-\bm u}_f\cdot \bm{n}_f )\Big)^2 \, ds +\int_f \Big(\pi_{RM^\perp(f)}(\jump{\Pi \bm u-\bm u}_f \cdot \bm{t}_f )\Big)^2 \, ds\\ &+ \int_f \Big(\pi_1(\jump{\bm u}_f \cdot \bm{n}_f )\Big)^2 \, ds +\int_f \Big(\pi_{RM^\perp(f)}(\jump{\bm u}_f \cdot \bm{t}_f )\Big)^2 \, ds. \end{split} \end{equation} Let $\mathcal T_f = T^{+}\cup T^{-}$, it follows from \eqref{3.3}, \eqref{3.4} and trace theorem that \begin{equation}\label{3.11} \begin{aligned} &\int_f \Big(\pi_1(\jump{\Pi\bm u-\bm u}_f \cdot \bm{n}_f )\Big)^2 \, ds +\int_f \Big(\pi_{RM^\perp(f)}(\jump{\Pi\bm u-\bm u}_f \cdot \bm{t}_f )\Big)^2 ds \\ \leq&\|\jump{\Pi\bm u-\bm u}\|^2_{0,f} \\ \lesssim & \sum_{T \in \mathcal T_f} \Big((\hbox{diam}(T))|\bm u_T - \Pi_T\bm u_T|_{H^1(T)}^2 + (\hbox{diam}(T))^{-1}|\bm u_T-\Pi_T\bm u_T|_{0,T}^2\Big) \\ \lesssim &\sum_{T \in \mathcal{T}_f} (\hbox{diam}(T))\|\mathcal D(\bm u_T)\|_{0,T}^2. \end{aligned} \end{equation} Combining \eqref{3.9} \eqref{3.10} and \eqref{3.11}, and noting that diameter of $T$ is equivalrent to diameter of $f$, we find \begin{equation}\label{3.12} \begin{split} &\sum_{f \in \mathcal{E}^o_h}(\hbox{diam}(f))^{d-2} \sum_{p \in \mathcal{V}(f)}|\jump{\bm u}_f(p)|^2\\ &\lesssim \|\mathcal D_{\mathcal T}(\bm u)\|_{0}^2+\sum_{f \in \mathcal{E}^o_h} (\hbox{diam}(f))^{-1}\Big(\|\pi_1(\jump{\bm u}_f \cdot \bm{n}_f)\|_{0,f}+\|\pi_{RM^\perp(f)}(\jump{\bm u}_f \cdot \bm{t}_f)\|_{0,f}\Big). \end{split} \end{equation} For $d=3$, we have that \begin{equation} \jump{\Pi \bm u}_f = -(\jump{\Pi\bm u}_f \times \bm{n}_f ) \times \bm{n}_f + (\jump{\Pi\bm u}_f \cdot \bm{n}_f ) \bm{n}_f. \end{equation} Furthermore, similar to the proof of Theorem \ref{kornpre}, we can show that \begin{equation} \jump{\Pi\bm u}_f \times \bm{n}_f = \bm{a} \times \bm{n}_f + c (\bm{x} - \bm{c}_f), \quad \forall \bm{x} \in f, \end{equation} where $\bm{c}_f$ is a barycenter of $f$. Using the same argument for $2D$, we can establish the estimate for 3D case in Theorem \ref{main}. This completes the proof. \end{proof} \begin{definition} A subspace of rigid body mode will be denoted and defined by \begin{equation} \bm {RM}^{\partial}(\Omega) = \left \{\bm m \in \bm {RM}(\Omega): ~ \|\bm m\|_{L^2(\partial \Omega)}=1,\int_{\partial \Omega}\bm mds=0 \right \}, \end{equation} \end{definition} We shall set $\Phi(\bm u)$ as follows: \begin{equation} \Phi(\bm u) : = \sup \limits_{\bm m \in \bm {RM}^{\partial} (\Omega)} \int_{\partial \Omega}\bm u \cdot \bm m \, ds. \end{equation} Under this setting, we can show that $\Phi$ satisfies the Aummptions (C1), (C2), (C3) and (C4). Additionally, we have \begin{equation} \Phi(\bm u) \leq \sum_{f \subset \partial \Omega}\Big(\|\pi_1(\jump{\bm u}_f \cdot \bm n_f)\|_{0,f}^2+\|\pi_{RM^\perp(f)}(\jump{\bm u}_f \cdot \bm t_f)\|_{0,f}^2\Big)~\hbox{for}~ d=2, \end{equation} and \begin{equation} \Phi(\bm u)\leq \sum_{f \subset \partial \Omega}\Big(\|\pi_1(\jump{\bm u}_f\cdot \bm n_f)\|_{0,f}^2+\|\pi_{RM^\perp(f)}(f)(\jump{\bm u}_f \times \bm n_f)\|_{0,f}^2\Big)~\hbox{for}~ d=3. \end{equation} Let $P_{1/0}(f) = P_1(f)/P_0(f)$ and $\pi_{1/0}$ be the orthogonal projection from $L^2(f)$ onto $P_{1/0}(f)$. \begin{Corollary}\label{DisKorn2d} For any $\bm u\in (H^1(\Omega, \mathcal{T}_h))^2$, and if $\bm u$ satisfies the following continuity conditions across the edges $f\in \mathcal{E}_h^o$ \begin{enumerate} \item $\int_{f} \jump{\bm u}_f \cdot \bm n_f \, p \, ds = 0, \quad \forall p \in P_1(f)$; \item $\int_{f} \jump{\bm u}_f \cdot \bm t_f ds = 0$; \end{enumerate} then the following estimate holds: \begin{equation} |\bm u|_{H^1(\Omega,\mathcal{T}_h)}\lesssim \|\mathcal{D}_{\mathcal {T}}(\bm u)\|_0+\sup\limits_{\bm m\in \bm {RM}^\partial(\Omega)}\int_{\partial \Omega}\bm u\cdot \bm mds. \end{equation} \end{Corollary} \begin{proof} This is immediate from Theorem \ref{main}. This completes the proof. \end{proof} \begin{Corollary}\label{DisKorn3d} For any $\bm u \in (H^1(\Omega, \mathcal{T}_h))^3$, and if $\bm u$ satisfies the following continuity conditions across the edges $f \in \mathcal{E}_h^o$ \begin{enumerate} \item $\int_{f} \jump{\bm u}_f \cdot \bm n_f \, p \, ds = 0, \quad \forall p \in P_1(f)$; \item $\int_{f} (\jump{\bm u}_f \times \bm n_f ) \cdot \bm {p}\, ds = 0, \quad \forall \bm{p} \in {\rm RT}_0(f)$; \end{enumerate} then the following estimate holds: \begin{equation} |\bm u|_{H^1(\Omega,\mathcal{T}_h)}\lesssim \|\mathcal{D}_{\mathcal {T}}(\bm u)\|_0+\sup\limits_{\bm m\in \bm {RM}^\partial(\Omega)}\int_{\partial \Omega}\bm u\cdot \bm m\, ds. \end{equation} \end{Corollary} \begin{proof} This is immediate from Theorem \ref{main}. This completes the proof. \end{proof} \section{On sharpness of the Korn's inequality}\label{sharp} In this subsection, we establish that the proposed Korn's inequality is sharp. We first start with 2D case and move on to 3D case. The goal is to establish that the proposed Korn's inequality is sharp, see Corollary \ref{DisKorn2d} and Corollary \ref{DisKorn3d}. More precisely, the above continuity conditions in Corollary \ref{DisKorn2d} and Corollary \ref{DisKorn3d} minimize the conditions to obtain the classical Korn's inequality for piecewise $H^1$ space. To put it in another way, if one of the conditions is violated, the classical Korn's inequality does not hold, but the inequality can be made to hold by adding an appropriate jump terms. For a sharpness proof, we shall take examples showing that if any one of the conditions is missing, then the inequality does not hold. These examples are constructed in special 2D and 3D domains as given in Figure \ref{ex}. \begin{figure}[h] \includegraphics[width=12cm, height=5cm]{image/domain.png} \caption{Domain 2D and 3D}\label{ex} \end{figure} For simplicity, we shall divide our discussion for 2D and 3D. For 2D, we consider a special $\Omega=\mathcal T_h = T_1 \cup T_2$, which consists of the vertexes $(-1,0), (0,0), (0,1),(-1,1)$ and $ (0,0),(1,0), (1,1),(0,1)$. We denote $f = T_1 \cap T_2$ and \begin{subeqnarray*} E_1 &=& \{\bm u\in (H^1(\Omega, \mathcal{T}_h))^2: \pi_{1/0}(\jump{\bm u}_f \cdot \bm n_f)=0~\hbox{and}~\pi_0(\jump{\bm u}_f \cdot \bm t_f)=0\}, \\ E_2 &=& \{\bm u\in (H^1(\Omega, \mathcal{T}_h))^2: \pi_0(\jump{\bm u}_f \cdot \bm n_f)=0~\hbox{and}~\pi_0(\jump{\bm u}_f \cdot \bm t_f)=0\}, \\ E_3 &=& \{\bm u \in (H^1(\Omega, \mathcal{T}_h))^2: \pi_0(\jump{\bm u}_f \cdot \bm n_f)=0~\hbox{and}~\pi_{1/0}(\jump{\bm u}_f \cdot \bm n_f)=0\}. \end{subeqnarray*} We first note that a simple calculation leads that \begin{equation}\label{RM0} \bm {RM}^\partial(\Omega) \subset \hbox{span}\{(1-2y, 2x)^t\}. \end{equation} Furthermore, we shall note that the definitions of $\pi_0$ and $\pi_{1/0}$ mean \begin{eqnarray} \pi_0 (u) &=& \frac{1}{|f|} \int_f u ds \quad \mbox{ and } \quad \pi_{1/0}(u) = \frac{1}{|f|} \int_f u s ds. \end{eqnarray} \begin{Theorem} There exists $\bm u\in E_k,$ for $k=1,2,$ or $3$, such that \begin{equation}\label{notkorn1} |\bm u|^2_{H^1(\Omega,\mathcal{T}_h)} \neq 0, \quad \mbox{ but } \quad \|\mathcal{D}_{\mathcal {T}}(\bm u)\|_0^2 + \left ( \sup\limits_{\bm m\in \bm {RM}^\partial (\Omega)}\int_{\partial \Omega}\bm u\cdot \bm mds \right )^2=0. \end{equation} \end{Theorem} \begin{proof} Let us consider $\bm u$ such that $\bm u_i = \bm u|_{T_i}$ for $i = 1,2$ given as follows: \begin{equation}\label{form} \bm u_i = \left ( \begin{array}{c} a_i \\ b_i \end{array} \right ) + c_i\left ( \begin{array}{c} -y \\ x \end{array} \right ) \in \bm {RM}(T_i), \quad\forall i=1,2. \end{equation} Here six coefficients $(a_i,b_i,c_i)_{i=1,2}$ shall be appropriately determined so that the Korn's inequality does not hold. We observe that $\|\mathcal{D}_{\mathcal {T}}(\bm u)\|_0 = 0$ by construction for any choice of coefficients. We first note that from \eqref{RM0}, i.e., $\bm {RM}^\partial(\Omega) \subset\hbox{span}\{(1-2y, 2x)^t\}$, we can calculate that with $\bm m = c(1 - 2y, 2x)^t$ and $c \neq 0$, \begin{equation}\label{cal} \int_{\partial \Omega}\bm u\cdot \bm m \, ds = c\left ( 4 (b_2 - b_1) + \frac{9}{2}(c_1 + c_2) \right ). \end{equation} We also observe that with $\overline{n} = n_2 - n_1$, \begin{eqnarray*} \jump{\bm{u}}_f \cdot \bm{n}_f &=& \overline{a} - \overline{c}y \\ \jump{\bm{u}}_f \cdot \bm{t}_f &=& \overline{b} + \overline{c} x. \end{eqnarray*} A simple computation leads that \begin{eqnarray*} \int_f \jump{\bm{u}}_f \cdot \bm{n}_f ds &=& \int_0^1 \overline{a} - \overline{c} y \, dy = \overline{a} - \frac{1}{2} \overline{c} \\ \int_f (\jump{\bm{u}}_f \cdot \bm{n}_f) y ds &=& \int_0^1 \overline{a}y - \overline{c} y^2 \, dy = \frac{1}{2} \overline{a} - \frac{1}{3} \overline{c} \\ \int_f \jump{\bm{u}}_f \cdot \bm{t}_f \, ds &=& \int_0^1 \overline{b} \, dy = \overline{b}. \end{eqnarray*} Now, we begin our search of $\bm{u}$ of the aforementioned form \eqref{form} that belongs to $E_k$, which satisfies \eqref{notkorn1} for all $k=1,2,$ or $3$. This shall be discussed case by case as follows. \begin{itemize}[leftmargin=*] \item For $k=1$, we shall choose $\overline{a} = \frac{2}{3} \overline{c} \neq 0$ and $\overline{b} = 0$. This way, we can make $\bm u \in E_1$. On the other hand, for \eqref{cal}, we must choose $c_1 + c_2 = 0$. Then, $\bm u$ satisfies \eqref{notkorn1}. \item For $k=2$, we shall choose $\overline{a} = \frac{1}{2} \overline{c} \neq 0$. Then by choosing $\overline{b} = 0$, we can make $\bm u \in E_2$. Again, we choose $c_1 + c_2 = 0$ for \eqref{cal}. Namely, $c_1 = c/2$ and $c_2 = -c/2$ for arbitrary $c \neq 0$ to guarantee the above conditions. We note that for $c \neq 0$, $\bm u$ satisfies \eqref{notkorn1}. \item For $k=3$, we shall choose $\overline{a} = \overline{c} = 0$. Then by choosing $\overline{b} \neq 0$, we can make $\bm u \in E_3$. On the other hand, for \eqref{cal}, we choose $c_1 =- c_2 = c/2\neq 0$ so that it holds, we see that since $c \neq 0$, $\bm u$ satisfies \eqref{notkorn1}. \end{itemize} This completes the proof. \end{proof} We shall now turn our attention to three dimensional case. For 3D case, we consider two cubes as given in Figure \ref{ex} (b). Namely, $\Omega=\mathcal T_h = T_1 \cup T_2$, whose coordinates are $(1, -1, 0), (1, 0, 0)$, $(0, 0, 0),(0, -1, 0),(1, -1, 1), (1, 0, 1),(0, 0, 1),(0, -1, 1)$ and $(1, 0, 0), (1, 1, 0),(0, 1, 0)$, $(0, 0, 0), (1, 0, 1), (1, 1, 1),(0, 1, 1),(0, 0, 1)$. We denote $f = T_1\cap T_2$ by the face and expand the two conditions in Corollary \ref{DisKorn3d} as following six conditions \begin{enumerate} \item [(A1)] $\int_{f} \jump{\bm u}_f \cdot \bm n_f ds = 0$; \item [(A2)] $\int_{f} \jump{\bm u}_f \cdot \bm n_f x ds = 0$; \item [(A3)] $\int_{f} \jump{\bm u}_f \cdot \bm n_f z ds = 0$; \item [(A4)] $\int_{f} (\jump{\bm u}_f \times \bm n_f) \cdot (0,0,1)^T ds = 0$; \item [(A5)] $\int_{f} (\jump{\bm u}_f \times \bm n_f) \cdot (1,0,0)^T ds = 0$; \item [(A6)] $\int_{f} (\jump{\bm u}_f \times \bm n_f) \cdot (x,0,z)^T ds = 0$. \end{enumerate} We now list a total of six subsets of $(H^1(\Omega;\mathcal{T}_h))^3$ as follows: \begin{subeqnarray*} F_1 &=& \{\bm u\in (H^1(\Omega, \mathcal{T}_h))^3: \bm u ~~ \hbox{satisfies (A2), (A3), (A4), (A5), and (A6)} \} \\ F_2 &=& \{\bm u\in (H^1(\Omega, \mathcal{T}_h))^3: \bm u ~~ \hbox{satisfies (A1), (A3), (A4), (A5), and (A6)} \} \\ F_3 &=& \{\bm u\in (H^1(\Omega, \mathcal{T}_h))^3: \bm u ~~ \hbox{satisfies (A1), (A2), (A4), (A5), and (A6)} \} \\ F_4 &=& \{\bm u\in (H^1(\Omega, \mathcal{T}_h))^3: \bm u ~~ \hbox{satisfies (A1), (A2), (A3), (A5), and (A6)} \} \\ F_5 &=& \{\bm u\in (H^1(\Omega, \mathcal{T}_h))^3: \bm u ~~ \hbox{satisfies (A1), (A2), (A3), (A4), and (A6)} \} \\ F_6 &=& \{\bm u\in (H^1(\Omega, \mathcal{T}_h))^3: \bm u ~~ \hbox{satisfies (A1), (A2), (A3), (A4), and (A5)} \}. \end{subeqnarray*} We first establish that with $\Omega = T_1\cup T_2$, the space of $\bm {RM}^\partial(\Omega)$. A tedious but simple calculation shows that \begin{equation}\label{RM1} \bm{RM}^\partial(\Omega) \subset \hbox{span}\{\bm m_1, \bm m_2, \bm m_3\}, \end{equation} where $\bm m_i$ with $i=1,2,3$ are given as follows: \begin{eqnarray*} \bm m_1 = \left ( \begin{array}{c} -1 + 2z \\ 0 \\ 1 - 2x \end{array} \right ), \,\, \bm m_2 = \left ( \begin{array}{c} 2y \\ 1 - 2x \\ 0 \end{array} \right ), \,\, \mbox{ and } \,\, \bm m_3 = \left (\begin{array}{c} y \\ -x + z \\ -y \end{array} \right ). \end{eqnarray*} We shall now state and prove the main result in this section. \begin{Theorem} There exists $\bm u \in F_k,$ for $k=1,2,3,4,5$ or $6$, such that \begin{equation}\label{notkorn2} |\bm u|^2_{H^1(\Omega,\mathcal{T}_h)} \neq 0, \quad \mbox{ but } \quad \|\mathcal{D}_{\mathcal {T}}(\bm u)\|_0^2 + \left ( \sup\limits_{\bm m\in \bm {RM}^\partial (\Omega)}\int_{\partial \Omega}\bm u\cdot \bm mds \right )^2=0. \end{equation} \end{Theorem} \begin{proof} Let us consider $\bm u$ such that $\bm u_i = \bm u|_{T_i}$ for $i=1,2$ given as follows: \begin{equation}\label{form} \bm u_i = \left ( \begin{array}{c} a_i \\ b_i \\ c_i \end{array} \right ) + \left ( \begin{array}{ccc} 0 & d_i & e_i \\ -d_i & 0 & f_i \\ -e_i & -f_i & 0 \end{array} \right ) \left ( \begin{array}{c} x \\ y \\ z \end{array} \right ) \in \bm {RM}(T_i), \quad\forall i=1,2. \end{equation} Here twelve coefficients $(a_i,b_i,c_i,d_i,e_i,f_i)_{i=1,2}$ shall be appropriately determined so that the Korn's inequality does not hold under the condition that $\bm u \in F_k$ for any $k$. We observe that $\|\mathcal{D}_{\mathcal {T}}(\bm u)\|_0 = 0$ by construction for any choice of coefficients. We first investigate what constraints are given from the conditions: \begin{equation} \int_{\partial \Omega} \bm u \cdot \bm m_i \, ds = 0, \quad \forall i=1,2,3, \end{equation} where $\bm {RM}^\partial(\Omega) \subset\hbox{span}\{ \bm m_1, \bm m_2, \bm m_3\}$. We observe that \begin{eqnarray*} \bm u \cdot \bm m_1|_{T_i} &=& (a_i + d_i y + e_i z) ( -1 + 2z ) + (c_i - e_i x - f_i y) (1 - 2x); \\ \bm u \cdot \bm m_2|_{T_i} &=& (a_i + d_i y + e_i z) ( 2y ) + (b_i - d_i x + f_i z) ( 1 - 2x ); \\ \bm u \cdot \bm m_3|_{T_i} &=& (a_i + d_i y + e_i z) ( y ) + (b_i - d_i x + f_i z) ( -x + z ) + (c_i - e_i x - f_i y) (- y). \end{eqnarray*} A simple but tedious computation leads that \begin{subeqnarray}\label{rot} \int_{\partial \Omega} \bm u \cdot \bm m_1 \, ds &=& \left ( \frac{9}{3} (e_1 + e_2) \right ) ;\\ \int_{\partial \Omega} \bm u \cdot \bm m_2 \, ds &=& 6(a_2 - a_1) + 3(e_2 - e_1) + \frac{37}{6}(d_1 + d_2) \slabel{eq2};\\ \int_{\partial \Omega} \bm u \cdot \bm m_3 \, ds &=& 3(a_2 - a_1) - 3(c_2 - c_1) + 3 (e_2 - e_1) \slabel{eq3} \\ && + \frac{37}{12} \left [ (d_2 + d_1) + (f_2 + f_1) \right ]. \nonumber \end{subeqnarray} We also observe that with $\overline{n} = n_2 - n_1$, \begin{eqnarray*} \jump{\bm{u}}_f \cdot \bm{n}_f &=& \overline{b} - \overline{d} x + \overline{f}z; \\ \jump{\bm{u}}_f \times \bm{n}_f &=& (- (\overline{c} - \overline{e} x - \overline{f} y), 0, \overline{a} + \overline{d} y + \overline{e} z)^t|_f = (- (\overline{c} - \overline{e} x), 0, \overline{a} + \overline{e} z)^t. \end{eqnarray*} A simple computation leads that \begin{eqnarray*} \int_f \jump{\bm{u}}_f \cdot \bm{n}_f ds &=& \int_0^1 \int_0^1 \overline{b} - \overline{d} x + \overline{f} z \, dxdz = \overline{b} - \frac{1}{2} \overline{d} + \frac{1}{2} \overline{f}; \\ \int_f \jump{\bm{u}}_f \cdot \bm{n}_f x ds &=& \int_0^1 \int_0^1 \overline{b}x - \overline{d} x^2 + \overline{f} xz \, dxdz = \frac{1}{2} \overline{b} - \frac{1}{3} \overline{d} + \frac{1}{4} \overline{f}; \\ \int_f \jump{\bm{u}}_f \cdot \bm{n}_f z ds &=& \int_0^1 \int_0^1 \overline{b} z - \overline{d} xz + \overline{f} z^2 \, dxdz = \frac{1}{2} \overline{b} - \frac{1}{4} \overline{d} + \frac{1}{3} \overline{f}. \end{eqnarray*} Furthermore, we have that \begin{eqnarray*} \int_f [\jump{\bm{u}}_f \times \bm{n}_f ] \cdot (0,0,1)^T ds &=& \int_0^1 \int_0^1 \overline{a} + \overline{e} z \, dxdz = \overline{a} + \frac{1}{2} \overline{e};\\ \int_f [\jump{\bm{u}}_f \times \bm{n}_f ] \cdot (1,0,0)^T ds &=& \int_0^1 \int_0^1 -\overline{c} + \overline{e} x \, dxdz = -\overline{c} + \frac{1}{2} \overline{e};\\ \int_f [\jump{\bm{u}}_f \times \bm{n}_f ] \cdot (x,0,z)^T ds &=& \int_0^1 \int_0^1 -\overline{c} x + \overline{e} x^2 + \overline{a} z + \overline{e} z^2 \, dxdz \\ &=& -\frac{1}{2} \overline{c} + \frac{2}{3} \overline{e} + \frac{1}{2} \overline{a}. \end{eqnarray*} Now, we begin our search of $\bm{u}$ of the aforementioned form \eqref{form} that belongs to $F_k$, which satisfies \eqref{notkorn2} for all $k=1,2,3,4,5,$ or $6$. This shall be discussed case by case as follows. \begin{itemize}[leftmargin=*] \item For $k=1$, we shall choose $\overline{d} = - \overline{f} \neq 0$. Then by choosing $\overline{b} = -\frac{7}{6} \overline{f}$, we can make $(A2)$ and $(A3)$ hold, but $(A1)$ does not. For $(A4), (A5)$ and $(A6)$, we can simply choose $\overline{a} = \overline{c} = \overline{e} = 0$. On the other hand, for \eqref{rot}, we must choose $d_i, e_i$ and $f_i$ so that $d_1 + d_2 = f_1 + f_2 = e_1 + e_2 = 0$. This means, $e_1 = e_2 = 0$. On the other hand, we can choose $d_2 = c/2, d_1 = -c/2$ and $f_2 = -c/2$ and $f_1 = c/2$ for arbitrary $c \neq 0$ to guarantee the above conditions. We note that for $c \neq 0$, $\bm u$ satisfies \eqref{notkorn2}. \item For $k=2$, we shall choose $\overline{b} = \frac{1}{2} \overline{d} \neq 0$. Then by choosing $\overline{f} = 0$, we can make $(A1)$ and $(A3)$ hold, but $(A2)$ does not. For $(A4), (A5)$ and $(A6)$, we can simply choose $\overline{a} = \overline{c} = \overline{e} = 0$. On the other hand, for \eqref{rot}, we must choose $d_i, e_i$ and $f_i$ so that $d_1 + d_2 = f_1 + f_2 = e_1 + e_2 = 0$. This means, $e_1 = e_2 = 0$. On the other hand, we can choose $d_2 = c/2, d_1 = -c/2$ for arbitrary $c \neq 0$ to guarantee the above conditions. We note that for $c \neq 0$, $\bm u$ satisfies \eqref{notkorn2}. \item For $k=3$, we shall choose $\overline{b} = \frac{1}{2} \overline{f} \neq 0$. Then by choosing $\overline{d} = 0$, we can make $(A1)$ and $(A2)$ hold, but $(A3)$ does not. For $(A4), (A5)$ and $(A6)$, we can simply choose $\overline{a} = \overline{c} = \overline{e} = 0$. On the other hand, for \eqref{rot}, we must choose $d_i, e_i$ and $f_i$ so that $d_1 + d_2 = f_1 + f_2 = e_1 + e_2 = 0$. This means, $e_1 = e_2 = 0$. On the other hand, we can choose $f_2 = c/2, f_1 = -c/2$ for arbitrary $c \neq 0$ to guarantee the above conditions. We note that for $c \neq 0$, $\bm u$ satisfies \eqref{notkorn2}. \item For $k=4$, we shall choose $\overline{c} = \frac{1}{2} \overline{e} \neq 0$ and $\overline{a} = -\frac{1}{6} \overline{e} \neq 0$. Then by choosing $\overline{d} = 0$, we can make $(A5)$ and $(A6)$ hold, but $(A4)$ does not. For $(A1), (A2)$ and $(A3)$, we can simply choose $\overline{b} = \overline{d} = \overline{f} = 0$. On the other hand, for \eqref{rot}, we must choose $d_i, e_i$ and $f_i$ so that $e_1 + e_2 = 0$ and both $d_1 + d_2$ and $f_1 + f_2$ appropriately to guarantee \eqref{eq2} and \eqref{eq3}. With such a choice, $\bm u$ satisfies \eqref{notkorn2}. \item For $k=5$, we shall choose $\overline{c} = \frac{1}{6} \overline{e} \neq 0$ and $\overline{a} = -\frac{1}{2} \overline{e} \neq 0$. Then by choosing $\overline{d} = 0$, we can make $(A4)$ and $(A6)$ hold, but $(A5)$ does not. For $(A1), (A2)$ and $(A3)$, we can simply choose $\overline{b} = \overline{d} = \overline{f} = 0$. On the other hand, for \eqref{rot}, we must choose $d_i, e_i$ and $f_i$ so that $e_1 + e_2 = 0$ and both $d_1 + d_2$ and $f_1 + f_2$ appropriately to guarantee \eqref{eq2} and \eqref{eq3}. With such a choice, $\bm u$ satisfies \eqref{notkorn2}. \item For $k=6$, we shall choose $\overline{c} = \frac{1}{2} \overline{e} = -\overline{a} \neq 0$. Then by choosing $\overline{d} = 0$, we can make $(A4)$ and $(A5)$ hold, but $(A6)$ does not. For $(A1), (A2)$ and $(A3)$, we can simply choose $\overline{b} = \overline{d} = \overline{f} = 0$. On the other hand, for \eqref{rot}, we must choose $d_i, e_i$ and $f_i$ so that $e_1 + e_2 = 0$ and both $d_1 + d_2$ and $f_1 + f_2$ appropriately to guarantee \eqref{eq2} and \eqref{eq3}. With such a choice, $\bm u$ satisfies \eqref{notkorn2}. \end{itemize} This completes the proof. \end{proof} \section{Some Applications}\label{appl} We begin this section with a simple example of the nonconforming finite element space on triangular partition introduced in Mardal et al., \cite{mardal2002robust} for the 2D case. Note that the space is composed of functions which are cubic vector fields with constant divergence on each element and with linear normal component on each edge. Namely, the local space is given as follows: \begin{equation} \bm V(T) = \{\bm v \in (P_3(T))^2 : \, {\rm div} \bm v \in P_0(T),\,\, \bm v\cdot\bm n_f |_f \in P_1(f), \,\, \forall f \in \partial T \}. \end{equation} Note that the degrees of freedom consist of \begin{equation} \int_f (\bm v \cdot \bm n_f) q \, ds, \quad \forall q \in P_1(f) \quad \mbox{ and } \quad \int_f \bm v \cdot \bm t_f \, ds. \end{equation} This is exactly what is required minimally for the Korn's inequality. In the next subsection, we shall consider enriched $H({\rm div})$ conforming finite element spaces due to Xue et al. \cite{xie2008uniformly}. We then slightly modify these spaces so that we obtain enriched Crouzeix-Raviart element spaces that satisfy the Korn's inequality. Finally, we shall discuss the lowest enrichment for the $H({\rm div})$ conforming finite element space \cite{johnny2012family} and its remedy by modifying the enrichment. Throughout the section, we consider triangular (for 2D) or tetrahedral (for 3D) partitions, for $T \in \mathcal{T}_h$, ${\rm{BDM}}_\ell(T)$ denotes the local Brezzi-Douglas-Marini (BDM) space of order $\ell \geq 1$, namely \begin{equation} {\rm{BDM}}_\ell(T) = (P_{\ell}(T))^d, \quad \ell \geq 1, \end{equation} and ${\rm{RT}}_\ell(T)$ denotes the Raviart-Thomas space of order $\ell \geq 0$, \begin{equation} {\rm{RT}}_\ell(T) = (P_{\ell}(T))^d + \widetilde{P}_{\ell}(T) \bm x, \quad \ell \geq 0, \end{equation} where $\widetilde{P}_\ell(T)$ denotes the homogeneous polynomial space of degree $\ell$. Also, for any given $T \in \mathcal{T}_h$, we denote $\lambda_i$ with $i = 1,\cdots,d+1$ the barycentric coordinates of $T$. We shall frequently use the following standard bubble functions. Namely, for $T \in \mathcal{T}_h$, we denote $b_T$ the bubble function defined by \begin{equation}\label{tbubble} b_T = \Pi_{i=1}^{d+1} \lambda_i. \end{equation} Similarly, we can define edge/face bubble functions denoted by $b_f$ for any $f \in \partial T$. \subsection{Enriched $H({\rm div})$ conforming finite elements that satisfy the Korn's inequality} In this section, we shall recall some of finite elements, which can be shown to satisfy the Korn's inequality within our framework. The finite element spaces listed in this subsection are constructed based on ${H}({\rm div};\Omega)$ finite element spaces plus divergence free functions \cite{xie2008uniformly}. This subsection is an example indicating the powerfulness of our framework to show the Korn's inequality. The local space of the finite elements introduced in \cite{xie2008uniformly} would be of the following form: for $T \in \mathcal{T}_h$, \begin{equation} \bm{V}_1(T) = \bm{V}_{1,T}^0 + \bm{curl} (b_T \bm{Y}), \end{equation} where $\bm{V}^0_{1,T}$ is the following well-known $H({\rm div})$-conforming finite element spaces, either ${\rm{BDM}}_1(T)$ or ${\rm{RT}}_1(T)$. For the space $\bm{Y}$, we choose the following polynomial spaces: \begin{equation} \bm{Y} = \left \{ \begin{array}{ll} \bm{Y}_1 = P_1(T) & \mbox{ for } d = 2; \\ \bm{Y}_2 = (P_1(T))^3 & \mbox{ for } d = 3;\\ \bm{Y}_3 = (P_1(T))^3/ {\rm span} \left \{ \left( \lambda_i - \frac{1}{3} \right )\nabla \lambda_i \right \}_{i=1}^4 & \mbox{ for } d = 3. \end{array} \right. \end{equation} Total of six elements can be listed in the following Table \ref{xutable} with degrees of freedom. \begin{table} \begin{tabular}{ |p{2.17cm}||p{1.5cm}|p{0.5cm}|p{5.cm}|p{1cm}| } \hline \multicolumn{5}{|c|}{The six modified $H({\rm div})$ elements} \\ \hline\hline {Elements} & $\bm{V}_{1,T}^0$ & \bm{Y} & {\rm DOF} & Korn \\ \hline $1^{\rm st}$ FEM (2D) & ${\rm RT}_1(T)$ & $\bm{Y}_1$ & $ {\small{\begin{array}{l} \int_f \bm{v} \cdot \bm{n} q\, ds, \quad \forall q \in P_1(f), \\ \int_T \bm{v} \cdot \bm{q} \, dx, \quad \bm{q} \in (P_0(T))^2 \\ \int_f \bm{v} \cdot \bm{t} \, ds \end{array}}} $ & Yes \\ \hline $2^{\rm nd}$ FEM (2D) & ${\rm BDM}_1(T)$ & $\bm{Y}_1$ & $ {\small{\begin{array}{l} \int_f \bm{v} \cdot \bm{n} q\, ds, \quad \forall q \in P_1(f), \\ \int_f \bm{v} \cdot \bm{t} \, ds \end{array}}} $ & Yes \\ \hline $1^{\rm st}$ FEM (3D) & ${\rm RT}_1(T)$ & $\bm{Y}_2$ & $ {\small{\begin{array}{l} \int_f \bm{v} \cdot \bm{n} q\, ds, \quad \forall q \in P_1(f), \\ \int_T \bm{v} \cdot \bm{q} \, dx, \quad \bm{q} \in (P_0(T))^3 \\ \int_f (\bm{v} \times \bm{n})\cdot \bm{r} \, ds, \quad \forall \bm{r} \in {\rm RT}_0(f) \end{array}}} $ & Yes \\ \hline $2^{\rm nd}$ FEM (3D) & ${\rm BDM}_1(T)$ & $\bm{Y}_2$ & $ {\small{\begin{array}{l} \int_f \bm{v} \cdot \bm{n} q\, ds, \quad \forall q \in P_1(f), \\ \int_f (\bm{v} \times \bm{n}) \cdot \bm{r} \, ds, \quad \forall \bm{r} \in {\rm RT}_0(f) \end{array}}} $ & Yes \\ \hline $3^{\rm rd}$ FEM (3D) & ${\rm RT}_1(T)$ & $\bm{Y}_3$ &$ {\small{\begin{array}{l} \int_f \bm{v} \cdot \bm{n} q\, ds, \quad \forall q \in P_1(f), \\ \int_T \bm{v} \cdot \bm{q} \, dx, \quad \bm{q} \in (P_0(T))^3 \\ \int_f (\bm{v} \times \bm{n})\cdot\bm{r} \, ds, \quad \forall \bm{r} \in (P_0(f))^2 \end{array}}} $ & No \\ \hline $4^{\rm th}$ FEM (3D) & ${\rm BDM}_1(T)$ & $\bm{Y}_3$ &$ {\small{\begin{array}{l} \int_f \bm{v} \cdot \bm{n} q\, ds, \quad \forall q \in P_1(f), \\ \int_f (\bm{v} \times \bm{n})\cdot\bm{r} \, ds, \quad \bm{r} \in (P_0(f))^2 \end{array}}} $ & No \\ \hline \end{tabular}\caption{FEMs introduced in \cite{xie2008uniformly}}\label{xutable} \end{table} Due to our framework, we easily notice that the first four elements are the ones that satisfy the Korn's inequality while the last two do not. \subsection{Enriched ${\rm CR}$ finite elements that satisfy the Korn's inequality} In this subsection, we shall consider enriched {\rm CR} elements that satisfy the Korn's inequality. Our discussion will be done for 2D and 3D separately. We note that in \cite{falk1991nonconforming}, Falk analyzed Korn's inequality for some nonconforming two dimensional finite element spaces. In particular, it was shown that the Crouzeix-Raviart element \cite{crouzeix1973conforming} does not satisfy such a Korn's inequality. A simple remedy can be accomplished. Namely, we add the same enrichment for $\bm{V}_1(T) = {\rm BDM}_1(T)$ replaced with $({\rm CR}(T))^d$ for both $d = 2$ and $d = 3$ and therefore, we propose the enriched $({\rm{CR}}(T))^d$ finite element spaces so that for $T \in \mathcal{T}_h$, we define the local space and degrees of freedom by \begin{equation} \bm{V}(T) := ({\rm CR}(T))^d + {\bm{curl}}(b_T \bm{Y}), \end{equation} where $\bm{Y}$ is either $\bm{Y}_1$ or $\bm{Y}_2$, depending on $d = 2$ or $d= 3$, respectively. \begin{Lemma} The space $\bm{V}(T)$ with the degrees of freedom given as in Table \ref{xutable} is unisolvent. \end{Lemma} \begin{proof} We let $\bm{v} \in \bm{V}(T)$ be decomposed into two parts: \begin{equation} \bm{v} = \bm{v}_0 + {\bm{curl}}(b_T \bm{q}), \quad \bm{v}_0 \in ({\rm{CR}}(T))^d, \quad \bm{q} \in \bm{Y}. \end{equation} In $3D$, we have that \begin{equation} {\bm{curl}}(b_T \bm{q})\cdot\bm{n} = {\bm{curl}}_f (b_T \bm{q})_f = 0, \end{equation} where $(b_T \bm{q})_f$ is the tangential component of $b_T \bm{q}$ on $f$. In 2D, we have that ${\bm{curl}}(b_T q)\cdot\bm{n} = \partial_t (b_T q) = 0$. Thus $\bm{v} \cdot \bm{n} = \bm{v}^0 \cdot \bm{n} \in P_1(f), \quad \forall f \in \partial T$. Since for $({\rm CR}(T))^d$ is unisolvent with the degrees of freedom, $\int_f \bm{v}^0 \cdot \bm{n} q \, ds$ for $q \in P_1(f)$, we arrive at $\bm{v}^0 = \bm{0}$. Therefore, it is sufficient to show that $\bm{curl}(b_T \bm{q}) = \bm{0}$ using the tangential component of the degrees of freedom. However, this part of the proof is essentially presented in \cite{xie2008uniformly} and also in \cite{tai2006discrete}. Therefore, we shall skip the proof and completes the proof. \end{proof} In passing to next subsection, we propose another enrichment for ${\rm CR}^2$ finite element space that satisfies the Korn's inequality. \subsubsection{Enriched {\rm CR} for 2D} Let $T \in \mathcal{T}_h$ be a triangle in 2D with $\{a_i, a_j, a_k\}$ vertexes of $T$. We denote the barycentric coordinates by $\{\lambda_i, \lambda_j, \lambda_k\}$. Further, $\bm n_{ij},\bm n_{jk},\bm n_{ki} $ are the normals to the edges $\{e_{ij} = [a_i,a_j], e_{jk} = [a_j,a_k], e_{ki} = [a_k,a_i]\}$. We let ${\rm CR}(T)$ be the standard {\rm CR} element on $T$. We then define an edge bubble function $b$ by \begin{equation} b = \lambda_i \lambda_j + \lambda_j \lambda_k + \lambda_k \lambda_i - \frac{1}{6}, \end{equation} where the constant $1/6$ comes from the fact that \begin{equation} \frac{1}{6} = \frac{1}{|f_{\nu\mu}|} \int_{f_{\nu\mu}} \lambda_\nu \lambda_\mu \, ds, \quad \nu\mu = ij, jk, ki. \end{equation} We are now in a position to define an enriched {\rm CR} element on $T$ by \begin{equation} \bm{E_{CR}}(T) = ({\rm{CR}}(T))^2 + \bm{V_{EC}}(T), \end{equation} where ${\bm{V_{EC}}}(T)$ is the space with the following functions as bases: \begin{equation} \bm \psi_{ij} = b (\lambda_i-\lambda_j)\bm n_{ij},~~~\bm \psi_{jk}=b (\lambda_j-\lambda_k)\bm n_{jk},~~~\bm \psi_{ki} = b(\lambda_k-\lambda_i)\bm n_{ki}. \end{equation} The degrees of freedom are given as follows: \begin{equation}\label{dof} \int_{f} \bm v \cdot \bm{t}_f ds \quad \hbox{and} \quad \int_{f} \bm v\cdot \bm{n}_f q ds, \quad \forall q \in P_1(f). \end{equation} It is simple to prove that the aforementioned degrees of freedom can be equivalently formulated as \begin{equation}\label{dof} \int_{f} \bm v ds \quad \hbox{and} \quad \int_{f} \bm v \cdot \bm{n}_f q ds, \quad \forall q \in P_1(f). \end{equation} We shall now prove that the space $\bm{E_{CR}}(T)$ with the degrees of freedom \eqref{dof} is unisolvent. \begin{Lemma} The space $\bm {E_{CR}}(T)$ with the degrees of freedom \eqref{dof} is unisolvent. \end{Lemma} \begin{proof} We let $\bm{v} \in \bm{E_{CR}}(T)$ be decomposed into two parts: \begin{equation} \bm{v} = \bm{v}_0 + \bm{v}_1, \quad \bm{v}_0 \in ({\rm{CR}}(T))^2, \quad \bm{v}_1 \in \bm{V_{EC}}(T). \end{equation} We shall assume that $\int_f \bm{v} \, ds = 0$ and $\int_{f} \bm v \cdot \bm{n}_f q \, ds=0, \,\, \forall q \in P_1(f)$, for all $f \in \partial T$ and shall show that $\bm v = \bm{0}$. We first note that for any $\bm {v}_1 \in \bm {V_{EC}}(T)$, we have $\int_{f} \bm {v}_1 \, ds = 0$, for any $f \in \partial T$, since for all $f \in \partial T$, it holds that $\int_{f} b q \, ds = 0, \forall q \in P_1(f)$. This measn that $\int_{f} \bm {v}_0\, ds = 0$ for any $f \in \partial T$ and so, $\bm {v}_0 = 0$. Therefore, it is sufficient to show that $\bm{v}_1 = \bm{0}$. By the definition of $\bm{V_{CR}}(T)$, we can write $\bm {v}_1$ as follows for some constants $c_1, c_2, c_3$: \begin{equation} \bm{v}_1 = c_1\bm \psi_{ij} + c_2 \bm \psi_{jk} + c_3 \bm \psi_{ki}. \end{equation} The fact that $\int_{f} \bm{v}_1 \cdot \bm{n} \, q ds = 0, \,\, \forall q \in P_1(f)$ leads to the following linear system: \begin{subeqnarray*} \int_{f_{ij}} (c_1\bm \psi_{ij}+c_2\bm \psi_{jk}+c_3 \bm \psi_{ki}) \cdot \bm n_{ij} \lambda_i ds &=& 0, \\ \int_{f_{jk}}(c_1\bm \psi_{ij}+c_2\bm \psi_{jk}+c_3 \bm \psi_{ki})\cdot \bm n_{jk} \lambda_j ds &=& 0, \\ \int_{f_{ki}} (c_1\bm \psi_{ij}+c_2\bm \psi_{jk}+c_3 \bm \psi_{ki}) \cdot \bm n_{ki} \lambda_k ds &=& 0. \end{subeqnarray*} Or equivalently, we have that \begin{equation}\label{system} \left ( \begin{array}{ccc} -2 & \bm{n}_{jk} \cdot \bm{n}_{ij} & \bm{n}_{ki} \cdot \bm{n}_{ij} \\ \bm{n}_{jk} \cdot \bm{n}_{ij} & -2 & \bm{n}_{ki} \cdot \bm{n}_{jk} \\ \bm{n}_{ij} \cdot \bm{n}_{ki} & \bm{n}_{jk} \cdot \bm{n}_{ki} & - 2 \end{array} \right ) \left ( \begin{array}{c} c_1 \\ c_2 \\ c_3 \end{array} \right ) = \left ( \begin{array}{c} 0 \\ 0 \\ 0 \end{array} \right ). \end{equation} Since $|\bm n_{ij}\cdot \bm n_{ki}|+|\bm n_{jk}\cdot \bm n_{ki}| < 2$, the coefficient matrix in the equation \eqref{system} is diagonally dominant and so it is invertible, which implies $\bm {v}_1 = \bm 0$. This completes the proof. \end{proof} \subsection{A new finite element space that satisfies the Korn's inequality} In this subsection, we shall investigate the finite elements introduced in \cite{johnny2012family}. We observe that some of their finite element for the lowest order case do not satisfy the Korn's inequality and provide modification, which satisfies the Korn's inequality. For a given $T \in \mathcal{T}_h$, we let \begin{equation} Q_f^{0}(T) = P_{0}(T)~ \hbox{for}~ d=2. \end{equation} and \begin{equation} \bm{Q}_f^{0}(T) = (P_{0}(T))^3 \times \bm{n}_f~ \hbox{for}~ d=3, \end{equation} and define $Q^{0}(T) = \sum_{f \in \partial T} b_f {Q}_f^{0}(T)$ for $d = 2$ and $\bm Q^{0}(T)= \sum_{f \in \partial T} b_f \bm{Q}_f^{0}(T)$ for $d=3$. We now consider the lowest order case of finite elements introduced in \cite{johnny2012family}. We recall that the local space was given as follows: for $T \in \mathcal{T}_h$, \begin{equation}\label{guzman} \bm{V}_1(T) = \bm{V}_{1,T}^0 + \bm{curl} (b_T \bm{Y}), \end{equation} where $\bm{V}^0_{1,T}$ is the well-known $H({\rm div})$-conforming finite element spaces, either ${\rm{BDM}}_1(T)$ or ${\rm{RT}}_1(T)$. The space $\bm{Y}$ is given as following: \begin{equation} \bm{Y} = \left \{ \begin{array}{ll} \bm{Y}_4 = Q^0(T) & \mbox{ for } d = 2; \\ \bm{Y}_5 = \bm Q^0(T) & \mbox{ for } d = 3. \end{array} \right. \end{equation} From our sharp Korn's inequality, we can easily check if the above four finite elements given in \eqref{guzman} from different choices of $\bm{V}_{1,T}^0$ and $\bm{Y}$ satisfy the Korn's inequality immediately. This is presented in the following Table \ref{guzmantable}: \begin{table}[ht] \begin{tabular}{ |p{2.17cm}||p{1.5cm}|p{0.5cm}|p{5.cm}|p{1cm}| } \hline \multicolumn{5}{|c|}{The six modified $H({\rm div})$ elements} \\ \hline\hline {Elements} & $\bm{V}_{1,T}^0$ & \bm{Y} & {\rm DOF} & Korn \\ \hline $1^{\rm st}$ FEM (2D) & ${\rm RT}_1(T)$ & $\bm{Y}_4$ & $ {\small{\begin{array}{l} \int_f \bm{v} \cdot \bm{n} q\, ds, \quad \forall q \in P_1(f), \\ \int_T \bm{v} \cdot \bm{q} \, dx, \quad \bm{q} \in (P_0(T))^2 \\ \int_f \bm{v} \cdot \bm{t} \, ds \end{array}}} $ & Yes \\ \hline $2^{\rm nd}$ FEM (2D) & ${\rm BDM}_1(T)$ & $\bm{Y}_4$ & $ {\small{\begin{array}{l} \int_f \bm{v} \cdot \bm{n} q\, ds, \quad \forall q \in P_1(f), \\ \int_f \bm{v} \cdot \bm{t} \, ds \end{array}}} $ & Yes \\ \hline $3^{\rm rd}$ FEM (3D) & ${\rm RT}_1(T)$ & $\bm{Y}_5$ &$ {\small{\begin{array}{l} \int_f \bm{v} \cdot \bm{n} q\, ds, \quad \forall q \in P_1(f), \\ \int_T \bm{v} \cdot \bm{q} \, dx, \quad \bm{q} \in (P_0(T))^3 \\ \int_f (\bm{v} \times \bm{n})\cdot\bm{r} \, ds, \quad \forall \bm{r} \in (P_0(f))^2 \end{array}}} $ & No \\ \hline $4^{\rm th}$ FEM (3D) & ${\rm BDM}_1(T)$ & $\bm{Y}_5$ &$ {\small{\begin{array}{l} \int_f \bm{v} \cdot \bm{n} q\, ds, \quad \forall q \in P_1(f), \\ \int_f (\bm{v} \times \bm{n})\cdot\bm{r} \, ds, \quad \bm{r} \in (P_0(f))^2 \end{array}}} $ & No \\ \hline \end{tabular} \caption{Lowest FEMs introduced in \cite{johnny2012family}}\label{guzmantable} \end{table} We note that particularly, for the local space \begin{equation} \bm{V}_1(T) = \bm{V}_{1,T}^0 + \bm{curl} (b_T \bm{Y}_5)= \bm{V}_{1,T}^0 + \bm{curl} (b_T \bm{Q}^0(T)), \end{equation} the degrees of freedom are not sufficient since the constant projected tangential component on the face being zero is not sufficient for the Korn's inequality. Therefore, we investigate how to remedy this situation. Namely, we shall modify the local space $\bm V_1(T)$ and the definition of degrees of freedom so that the Korn's inequality can be satisfied in the remainder of this subsection. We define the local space for the nonconforming finite element that satisfies the Korn's inequality : \begin{equation} \bm{V}^1(T) = \bm{V}_{1,T}^0 + {\bm{curl}} \left ( b_T \bm{Q}^*(T) \right ), \end{equation} where \begin{eqnarray*} \bm{Q}^{*}(T) &=& \sum_{f \in \partial T} b_f \bm{Q}_f^*(T), \end{eqnarray*} with \begin{eqnarray*} \bm{Q}_f^*(T) = \left \{ \bm{q} \times \bm{n}_f : \bm{q} \in \bm{RM}(T), \int_T (\bm{q} \times {\bm{n}}_f) \cdot (\bm{w} \times \bm{n}_f) b_T b_f \, dx = 0, \,\, \bm{w} \in (P_0(T))^3 \right \}. \end{eqnarray*} It is easy to see that the dimension of $\bm{Q}_f^*(T)$ is three. Therefore, the dimension of the space $\bm{Q}^*(T)$ is 12. Note that the space $ \bm{V}_{1,T}^0$ can be either $BDM_1$ or $RT_1$. But, we shall only consider $BDM_1$ for simplicity. Under this choice, it is well-known that a function $\bm{v} \in \bm{V}_{1,T}^0$ is uniquely determined by the following degrees of freedom: for all faces $f \in \partial T$, \begin{equation} \langle \bm{v}\cdot \bm{n}_f, \bm{\mu} \rangle_f, \quad \forall \bm{\mu} \in ({P}_1(f))^3. \end{equation} With this in mind, we can provide the following degrees of freedom that define a function $\bm{v} \in \bm{V}^1(T)$ uniquely: for all $f\in\partial T$, \begin{eqnarray} \langle \bm{v}\cdot \bm{n}_f, \bm{\mu} \rangle_f, && \quad \forall \bm{\mu} \in ({P}_1(f))^3 \label{dof1}; \\ \langle \bm{v}\times \bm{n}_f, \bm{\kappa} \rangle_f, && \quad \forall \bm{\kappa} \in {\rm{RT}}_0(f). \label{dof2} \end{eqnarray} We can show that the following holds true: \begin{Lemma} For any $T \in \mathcal{T}_h$, we have \begin{equation} {\rm dim} \,{\bm{curl}} \left ( b_T \bm{Q}^*(T) \right ) = 4 \, {\rm dim} \, \bm{Q}_f^*(T). \end{equation} \end{Lemma} \begin{proof} The proof can be done similarly to the argument of Lemma 3.2 in \cite{johnny2012family}. Therefore, we complete the proof. \end{proof} \begin{Theorem} We have the following relation that \begin{equation} \bm{V}^1(T) = \bm{V}_{1,T}^0 \oplus {\bm{curl}} \left ( b_T \bm{Q}^*(T) \right ) \end{equation} and \begin{equation} {\rm dim} \bm{V}^1(T) = {\rm dim} \bm{V}_{1,T}^0 + {\rm dim} {\bm{curl}} \left ( b_T \bm{Q}^*(T) \right ). \end{equation} Furthermore, any function $\bm{v} \in \bm{V}^1(T)$ is uniquely determined by the degrees of freedom \eqref{dof1} and \eqref{dof2}. \end{Theorem} \begin{proof} The proof can be done similarly to the argument of Theorem 3.3 in \cite{johnny2012family}. Therefore, we complete the proof. \end{proof} \section{Conclusion}\label{con} We have proven a Korn's inequality for the piecewise $H^1$ space. Our characterization of the Korn's inequality considers to be the rotated version of what is characterized in \cite{mardal2006observation}. We further have shown some sharpness of the Korn's inequality. This result has been applied to a number of finite elements to check that they satisfy the Korn's inequality or not. Also the minimal jump terms that appear in the Korn's inequality are explicit enough so that they can be used for construction of finite element spaces that satisfy the Korn's inequality. \bibliographystyle{plain}
{ "timestamp": "2022-07-06T02:17:20", "yymm": "2207", "arxiv_id": "2207.02060", "language": "en", "url": "https://arxiv.org/abs/2207.02060", "abstract": "In this paper, we revisit Korn's inequality for the piecewise $H^1$ space based on general polygonal or polyhedral decompositions of the domain. Our Korn's inequality is expressed with minimal jump terms. These minimal jump terms are identified by characterizing the restriction of rigid body mode to edge/face of the partitions. Such minimal jump conditions are shown to be sharp for achieving the Korn's inequality as well. The sharpness of our result and explicitly given minimal conditions can be used to test whether any given finite element spaces satisfy Korn's inequality, immediately as well as to build or modify nonconforming finite elements for Korn's inequality to hold.", "subjects": "Numerical Analysis (math.NA)", "title": "A sharp Korn's inequality for piecewise $H^1$ space and its application", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9854964173268185, "lm_q2_score": 0.8128673133042217, "lm_q1q2_score": 0.801077825023387 }
https://arxiv.org/abs/2011.12449
Iterations for the Unitary Sign Decomposition and the Unitary Eigendecomposition
We construct fast, structure-preserving iterations for computing the sign decomposition of a unitary matrix $A$ with no eigenvalues equal to $\pm i$. This decomposition factorizes $A$ as the product of an involutory matrix $S = \operatorname{sign}(A) = A(A^2)^{-1/2}$ times a matrix $N = (A^2)^{1/2}$ with spectrum contained in the open right half of the complex plane. Our iterations rely on a recently discovered formula for the best (in the minimax sense) unimodular rational approximant of the scalar function $\operatorname{sign}(z) = z/\sqrt{z^2}$ on subsets of the unit circle. When $A$ has eigenvalues near $\pm i$, the iterations converge significantly faster than Padé iterations. Numerical evidence indicates that the iterations are backward stable, with backward errors often smaller than those obtained with direct methods. This contrasts with other iterations like the scaled Newton iteration, which suffers from numerical instabilities if $A$ has eigenvalues near $\pm i$. As an application, we use our iterations to construct a stable spectral divide-and-conquer algorithm for the unitary eigendecomposition.
\section{Introduction} \label{sec:intro} Every matrix $A \in \mathbb{C}^{n \times n}$ with no purely imaginary eigenvalues can be written uniquely as a product \[ A = SN, \] where $S \in \mathbb{C}^{n \times n}$ is involutory ($S^2=I$), $N \in \mathbb{C}^{n \times n}$ has spectrum in the open right half of the complex plane, and $S$ commutes with $N$. This is the celebrated matrix sign decomposition~\cite{higham1994matrix}, whose applications are widespread~\cite{denman1976matrix,kenney1995matrix}. In terms of the principal square root $(\cdot)^{1/2}$, we have $S=A(A^2)^{-1/2} =: \mathrm{sign}(A)$ and $N=(A^2)^{1/2}$. When $A$ is unitary, so too are $S$ and $N$. It follows that $S=S^{-1}=S^*$, so we may write, for any unitary $A$ with $\Lambda(A) \cap i\mathbb{R} = \emptyset$, \begin{equation} \label{unitarysign} A = SN, \quad S^2=I, \; S=S^*, \; N^2=A^2, \; N^*N=I, \, \Lambda(N) \subset \mathbb{C}_+, \end{equation} where $\Lambda(N)$ denotes the spectrum of $N$ and $\mathbb{C}_+ = \{z \in \mathbb{C} \mid \Re(z)>0\}$. We refer to this decomposition as the \emph{unitary sign decomposition}. We say that an algorithm for computing the decomposition~(\ref{unitarysign}) is backward stable if it computes matrices $\widehat{S}$ and $\widehat{N}$ with the property that the quantities \begin{equation} \label{backwarderrors} \|A-\widehat{S}\widehat{N}\|, \, \|\widehat{S}^2-I\|, \, \|\widehat{S}-\widehat{S}^*\|, \, \|\widehat{N}^*\widehat{N}-I\|, \, \|\widehat{N}^2-A^2\|, \, \max\{0,-\min_{\lambda \in \Lambda(\widehat{N})} \Re\lambda\} \end{equation} are each a small multiple of the unit roundoff $u$ ($=2^{-53}$ in double-precision arithmetic).\footnote{Note that this property implies $\|\widehat{N}\widehat{S}-\widehat{S}\widehat{N}\|$ is small as well; see Lemma~\ref{lemma:SNcommute}.} Here, $\|\cdot\|$ denotes the 2-norm. The goal of this paper is to design backward stable iterations for computing the decomposition~(\ref{unitarysign}). To illustrate why this is challenging, let us point out the pitfalls of naive approaches. A widely used iteration for computing the sign of a general matrix $A \in \mathbb{C}^{n \times n}$ is the Newton iteration~\cite{roberts1980linear}~\cite[Section 5.3]{higham2008functions} \begin{equation} \label{newtonit} X_{k+1} = \frac{1}{2}(X_k+X_k^{-1}), \quad X_0 = A. \end{equation} If $A$ is unitary, then the first iteration is simply \begin{equation} \label{firstnewtonit} X_1 = \frac{1}{2}(A+A^*). \end{equation} In floating point arithmetic, this calculation is susceptible to catastrophic cancellation if $A$ has eigenvalues near $\pm i$. Indeed, if we carry out~(\ref{firstnewtonit}) followed by~(\ref{newtonit}) for $k=1,2,\dots$ on the $100 \times 100$ unitary matrix \verb$A = gallery('orthog',100,3)$ from the MATLAB matrix gallery, then the iteration diverges. Scaling the iterates with standard scaling heuristics~\cite{kenney1992scaling} leads to convergence, but the computed sign of $A$ satisfies $\|\widehat{S}-\widehat{S}^*\| > 0.1$ in typical experiments. This happens because $A$ has several eigenvalues lying near $\pm i$. The above algorithm can be reinterpreted in a different way: It is computing the unitary factor in the polar decomposition of $(A+A^*)/2$. Indeed, the Newton iteration $X_{k+1} = \frac{1}{2}(X_k+X_k^{-*})$ for the polar decomposition~\cite[Section 8.3]{higham2008functions} coincides with~(\ref{newtonit}) on Hermitian matrices. This suggests another family of potential algorithms: compute the polar decomposition of $(A+A^*)/2$ via iterative methods or other means. However, numerical experiments confirm that such algorithms are similarly inaccurate on matrices with eigenvalues near $\pm i$. This unstable behavior is also shared by the superdiagonal Pad\'e iterations for the matrix sign function~\cite{kenney1991rational}, all of which map eigenvalues $\lambda \approx \pm i$ of $A$ to a small real number (or the inverse thereof) in the first iteration. One way to overcome these difficulties is to adopt structure-preserving iterations. Here, we say that an iteration $X_{k+1} = g_k(X_k)$ for the unitary sign decomposition is structure-preserving if the iterates $X_k$ are unitary for every $k$. Examples include the diagonal family of Pad\'e iterations~\cite{higham2004computing}, whose lowest-order member is the iteration \begin{equation} \label{padelow} X_{k+1} = X_k(3I+X_k^2)(I+3X_k^2)^{-1}, \quad X_0 = A. \end{equation} By keeping $X_k$ unitary, a structure-preserving iteration ensures that the eigenvalues of $X_k$ remain on the unit circle, ostensibly skirting the dangers of catastrophic cancellation. We observe numerically that, if implemented in a clever way (described in Section~\ref{sec:algorithm}), the diagonal Pad\'e iterations are backward stable. However, they can take excessively long to converge on matrices with eigenvalues near $\pm i$. For example, when \verb$A = gallery('orthog',100,3)$, the iteration~(\ref{padelow}) takes 34 iterations to converge. We construct in this paper a family of structure-preserving iterations for the unitary sign decomposition that converge more rapidly---sometimes dramatically more so---than the diagonal Pad\'e iterations. Numerical evidence indicates that these iterations are backward stable, with backward errors often smaller than those obtained with direct methods. The key ingredient that we use to construct our iterations is a recently discovered formula for the best (in the minimax sense) unimodular rational approximant of the scalar function $\sign(z) = z/\sqrt{z^2}$ on subsets of the unit circle~\cite{gawlik2020zolotarev}. Remarkably, it can be shown that composing two such approximants yields a best approximant of higher degree~\cite{gawlik2020zolotarev}, laying the foundations for an iteration. When applied to matrices, the iteration produces a sequence of unitary matrices $X_0=A$, $X_1$, $X_2$, $\dots$ that converges rapidly to $S=\sign(A)$, often significantly faster than the corresponding diagonal Pad\'e iteration. When \verb$A = gallery('orthog',100,3)$, for example, the lowest-order iteration converges within 6 iterations, which is about 6 times faster than the corresponding diagonal Pad\'e iteration~(\ref{padelow}). \paragraph{Prior work} Matrix iterations constructed from rational minimax approximants have attracted growing interest in recent years. Early examples include the optimal scaling heuristic proposed by Byers and Xu~\cite{byers2008new} for the Newton iteration for the polar decomposition, as well as an analogous scaling heuristic for the matrix square root proposed by Wachspress~\cite{wachspress1962} and Beckermann~\cite{beckermann2013optimally}. Nakatsukasa, Bai, and Gygi~\cite{nakatsukasa2010optimizing} designed an optimal scaling heuristic for the Halley iteration for the polar decompsition, and their strategy was generalized to higher order by Nakatsukasa and Freund~\cite{nakatsukasa2016computing}. The latter work elucidated the link between these scaling heuristics and the seminal work of Zolotarev~\cite{zolotarev1877applications} on rational minimax approximation. The iterations derived in~\cite{nakatsukasa2016computing} have a variety of applications, including algorithms for the symmetric eigendecomposition, singular value decomposition, polar decomposition, and CS decomposition~\cite{nakatsukasa2016computing,gawlik2018backward}. All of the aforementioned algorithms rely crucially on the following fact: if two rational minimax approximants of the scalar function $\sign(x)$ on suitable real intervals are composed with one another, then their composition is a best approximant of higher degree~\cite{nakatsukasa2016computing}. A related composition law for rational minimax approximants of $\sqrt{z}$ has been used to construct iterations for the matrix square root~\cite{gawlik2018zolotarev}. These iterations were generalized to the matrix $p$th root in~\cite{gawlik2020rational} and used to derive approximation theoretic results in~\cite{gawlik2019approximating}. An even more recent advancement---a composition law for rational minimax approximants of $\sign(z)$ on subsets of the unit circle~\cite{gawlik2020zolotarev}---is what inspired the present paper. \paragraph{Connections to other iterations} The iterations we derive in this paper are intimately connected to several existing iterations for the matrix sign function and the polar decomposition. When applied to a unitary matrix $A$, our iterations produce a sequence of unitary matrices whose Hermitian part coincides with the sequence of matrices generated by Nakatsuka and Freund's iterations~\cite{nakatsukasa2016computing} for the polar decomposition of $(A+A^*)/2$. A special case of this result is a connection between our lowest-order iteration for $\sign(A)$ and the optimally scaled Halley iteration for the polar decomposition of $(A+A^*)/2$~\cite{nakatsukasa2010optimizing}. It is important to note that these equivalences hold only in exact arithmetic. In floating-point arithmetic, our iterations behave very differently from the aforementioned algorithms. There is also a link between our iterations and the diagonal Pad\'e iterations. Roughly speaking, our iterations are designed using rational minimax approximants of $\sign(z)$ on two circular arcs containing $\pm 1$. If these arcs are each shrunk to a point, then the diagonal Pad\'e iterations are recovered. This helps to explain the slow convergence of the diagonal Pad\'e iterations on unitary matrices with eigenvalues near $\pm i$: The iterations need to approximate $\sign(z)$ near $z = \pm i$, but they use rational functions that are designed to approximate $\sign(z)$ near $z = \pm 1$. \paragraph{Unitary eigendecomposition} Our emphasis on handling eigenvalues near $\pm i$ is not merely pedantic. It is precisely the sort of situation that one often encounters if the unitary sign decomposition is used as part of a spectral divide-and-conquer algorithm for the unitary eigendecomposition. Indeed, consider a unitary matrix $A \in \mathbb{C}^{m \times m}$ with eigendecomposition $A=V\Lambda V^*$. The matrix $(I+\sign(A))/2$ is a spectral projector onto the invariant subspace $\mathcal{V}_+$ of $A$ associated with eigenvalues having positive real part. A spectral divide-and-conquer algorithm uses this projector to find orthonormal bases $U_1 \in \mathbb{C}^{m \times m_1}$, $U_2 \in \mathbb{C}^{m \times m_2}$, $m_1+m_2=m$, for $\mathcal{V}_+$ and its orthogonal complement. Then $\begin{pmatrix} U_1 & U_2 \end{pmatrix}^* A \begin{pmatrix} U_1 & U_2 \end{pmatrix}$ is block diagonal, so recursion can be used to determine $V$ and $\Lambda$. At each step, scalar multiplication by complex numbers with unit modulus can be used to rotate the spectrum so that it is distributed approximately evenly between the left and right half-planes. If $A$ has a cluster of nearby eigenvalues, then it is reasonable to expect this process to center the cluster near $\pm i$ at some step. This is precisely what we observe in practice, and the ability to compute the unitary sign decomposition quickly and accurately in the presence of eigenvalues near $\pm i$ becomes paramount. \paragraph{Organization} This paper is organized as follows. We begin in Section~\ref{sec:scalarsign} by studying rational minimax approximants of $\sign(z)$ on the unit circle. This material is largely drawn from~\cite{gawlik2020zolotarev}, but we add some additional results and insights to relate these approximants to Pad\'e approximants. Next, we use these approximants to construct matrix iterations for the unitary sign decomposition in Section~\ref{sec:algorithm}. We illustrate their utility by constructing a spectral divide-and-conquer algorithm for the unitary eigendecomposition in Section~\ref{sec:eig}. We conclude with numerical examples in Section~\ref{sec:numerical}. \section{Rational Approximation of the Sign Function on the Unit Circle} \label{sec:scalarsign} In this section, we study rational approximants of the scalar function $\sign(z) = z/\sqrt{z^2}$ on the set \[ \mathbb{S}_\Theta = \{z \in \mathbb{C} \mid |z|=1, \, \arg z \notin (\Theta,\pi-\Theta) \cup (-\pi+\Theta,-\Theta) \}, \] where $\Theta \in (0,\pi/2)$. Since our ultimate interest is in constructing structure-preserving iterations for the unitary sign decomposition, we focus on rational functions $r$ satisfying $|r(z)|=1$ for $|z|=1$. We call such rational functions unimodular. Unimodular rational functions have the property that $r(A)$ is unitary for any unitary matrix $A$. The problem of determining the best (in the minimax sense) unimodular rational approximant of $\sign(z)$ on $\mathbb{S}_\Theta$ has recently been solved in~\cite{gawlik2020zolotarev}. To describe the solution, let us introduce some notation. We use $\mathrm{sn}(\cdot,\ell)$, $\mathrm{cn}(\cdot,\ell)$, and $\mathrm{dn}(\cdot,\ell)$ to denote Jacobi's elliptic functions with modulus $\ell$, and we use $\ell' = \sqrt{1-\ell^2}$ to denote the modulus complementary to $\ell$. We denote the complete elliptic integral of the first kind by $K(\ell) = \int_0^{\pi/2} (1-\ell^2 \sin^2\theta)^{-1/2} \, d\theta$. We say that a rational function $r(z)=p(z)/q(z)$ has type $(m,n)$ if $p$ and $q$ are polynomials of degree at most $m$ and $n$, respectively. \begin{theorem} Let $\Theta \in (0,\pi/2)$ and $n \in \mathbb{N}_0$. Among all rational functions $r$ of type $(2n+1,2n+1)$ that satisfy $|r(z)|=1$ for $|z|=1$, the ones which minimize \[ \max_{z \in \mathbb{S}_\Theta} \left| \arg\left(\frac{r(z)}{\sign(z)}\right) \right| \] are \[ r(z) = r_{2n+1}(z;\Theta) = z \prod_{j=1}^n \frac{z^2+a_j}{1+a_j z^2} \] and its reciprocal, where \[ a_j = a_j(\Theta) = \left( \frac{\ell \sn(v_j,\ell') + \dn(v_j,\ell')}{\cn(v_j,\ell')} \right)^{2(-1)^{j+n}}, \] $v_j = \frac{2j-1}{2n+1}K(\ell')$, $\ell = \cos\Theta$, and $\ell' = \sqrt{1-\ell^2} = \sin\Theta$. \end{theorem} \begin{proof} See~\cite[Theorem 2.1 and Remark 2.2]{gawlik2020zolotarev}. \end{proof} \begin{remark} For simplicity, we have chosen to focus only on best unimodular rational approximants of $\sign(z)$ on $\mathbb{S}_\Theta$ of type $(2n+1,2n+1)$ in this paper. Best approximants of type $(2n,2n)$ can also be written down; see~\cite{gawlik2020zolotarev} for details. \end{remark} The rational function $r_{2n+1}(z;\Theta)$ has the following remarkable behavior under composition. \begin{theorem} \label{thm:composition} Let $\Theta \in (0,\pi/2)$, $m,n \in \mathbb{N}_0$, and $\widetilde{\Theta} = \left| \arg(r_{2n+1}(e^{i\Theta}; \Theta)) \right|$. Then \[ r_{2m+1}(r_{2n+1}(z;\Theta); \widetilde{\Theta}) = r_{(2m+1)(2n+1)}(z;\Theta). \] \end{theorem} \begin{proof} See~\cite[Theorem 3.3 and Remark 3.6]{gawlik2020zolotarev}. \end{proof} We also have the following error estimate. \begin{theorem} \label{thm:error} Let $\Theta \in (0,\pi/2)$ and $n \in \mathbb{N}_0$. We have \[ \max_{z \in \mathbb{S}_\Theta} \left| \arg\left(\frac{r_{2n+1}(z;\Theta)}{\sign(z)}\right) \right| \le 4 \rho^{-(2n+1)}, \] where \begin{equation} \label{rho} \rho = \rho(\Theta) = \exp\left( \frac{ \pi K(\cos\Theta) }{ 2K(\sin\Theta) } \right). \end{equation} \end{theorem} \begin{proof} See~\cite[Theorem 3.2]{gawlik2020zolotarev}, and note that their definition of $\rho$ differs from ours by a factor of 2 in the exponent. \end{proof} \begin{remark} \label{remark:theta0} Theorems~\ref{thm:composition} and~\ref{thm:error} continue to hold when $\Theta=0$ if we adopt the convention that $\rho(0) = \infty$, $\mathbb{S}_0 = \{-1,1\}$, and $r_{2n+1}(z;0) = z \prod_{j=1}^n \frac{z^2+a_j(0)}{1+a_j(0) z^2}$. We elaborate on this fact below. \end{remark} \subsection{Connections with Other Rational Approximants} \label{sec:connections} The rational function $r_{2n+1}(z;\Theta)$ is closely connected to several other well-known rational approximants of $\sign(z)$. \begin{proposition} \label{prop:pade} As $\Theta \rightarrow 0$, $r_{2n+1}(z;\Theta)$ converges coefficientwise to $z p_n(z^2)$, where $p_n(z)$ is the type-$(n,n)$ Pad\'e approximant of $z^{-1/2}$ at $z=1$. \end{proposition} \begin{proof} This is a consequence of~\cite[Proposition 3.9]{gawlik2020zolotarev}, where it is shown that $\sqrt{z}/r_{2n+1}(\sqrt{z};\Theta)$ converges coefficientwise to $1/p_n(z)$ as $\Theta \rightarrow 0$. \end{proof} In the notation of Remark~\ref{remark:theta0}, the above proposition states that \[ r_{2n+1}(z;0) = zp_n(z^2). \] This rational function has been studied extensively in~\cite{kenney1991rational,kenney1994hyperbolic,gomilko2012pade}~\cite[Theorem 5.9]{higham2008functions}. It satisfies~\cite[Theorem 5.9]{higham2008functions} \begin{equation} \label{tanh} z p_n(z^2) = \tanh((2n+1)\arctanh z) \end{equation} It also has the following properties. Both $p_n(z)$ and $zp_n(z^2)$ are unimodular~\cite{higham2004computing}; that is, for any $n \in \mathbb{N}_0$, \[ |zp_n(z^2)| = |p_n(z)|=1, \text{ if } |z|=1. \] Under composition, we have~\cite[Theorem 5.9)(c)]{higham2008functions} \begin{equation} \label{composition0} r_{2m+1}(r_{2n+1}(z;0);0) = r_{(2m+1)(2n+1)}(z;0) \end{equation} for any $m,n \in \mathbb{N}_0$. Finally, $r_{2n+1}(1;0)=-r_{2n+1}(-1;0)=1$ for all $n \in \mathbb{N}_0$. These last two facts justify Remark~\ref{remark:theta0}. The rational functions $r_{2n+1}(z;0)$, $n \in \mathbb{N}_0$, have been used in~\cite{kenney1991rational} to construct iterations for computing the matrix sign function. The iterations constitute the diagonal family of Pad\'e iterations. The first few diagonal Pad\'e approximants of $z^{-1/2}$ at $z=1$ are \[ p_0(z)=1, \; p_1(z) = \frac{3+z}{1+3z}, \; p_2(z) = \frac{5+10z+z^2}{1+10z+5z^2}, \; p_3(z) = \frac{7+35z+21z^2+z^3}{1+21z+35z^2+7z^3}. \] More generally, Pad\'e iterations can be constructed from rational functions of the form $zp_{m,n}(z^2)$, where $p_{m,n}(z)$ is the type-$(m,n)$ Pad\'e approximant of $z^{-1/2}$ at $z=1$. However, when $m\neq n$, the Pad\'e iterations are not structure-preserving, as $|p_{m,n}(z)| \not\equiv 1$ for $|z|=1$ and $m \neq n$. We now turn our attention back to the rational function $r_{2n+1}(z,\Theta)$ with positive $\Theta$. Interestingly, this function is intimately connected to the solution of another rational approximation problem: approximating $\sign(x)$ on the union of real intervals $[-1,-\ell] \cup [\ell,1]$. \begin{theorem} \label{thm:realpart} Let $\Theta \in [0,\pi/2)$ and $n \in \mathbb{N}_0$. For $z \in \mathbb{C}$ with $|z|=1$, we have \begin{equation} \label{realpart} \Re r_{2n+1}(z;\Theta) = \widehat{R}_{2n+1}(\Re z; \cos\Theta), \end{equation} where \[ \widehat{R}_m(x;\ell) = \begin{cases} \frac{R_m(x;\ell)}{\max_{y \in [\ell,1]} R_m(y;\ell)} &\mbox{ if } \ell \in (0,1), \\ x p_n(x^2), &\mbox{ if } \ell=1, \end{cases} \] and \[ R_m(\cdot;\ell) = \argmin_{R \in \mathcal{R}_{m,m}} \max_{x \in [-1,-\ell] \cup [\ell,1]} |R(x)-\sign(x)|. \] \end{theorem} \begin{proof} This identity is proven for $\Theta \in (0,\pi/2)$ in~\cite[Theorem 2.4]{gawlik2020zolotarev}. To see that it also holds when $\Theta = 0$, we must show that if $|z|=1$ and $x = \Re z = \frac{1}{2}(z+1/z)$, then \[ \frac{1}{2} \left( \tanh((2n+1)\arctanh z) + \frac{1}{\tanh((2n+1)\arctanh z)} \right) = \tanh((2n+1)\arctanh x). \] Since $\frac{1+x}{1-x} = -\left(\frac{1+z}{1-z}\right)^2$, we have $\arctanh x = \frac{1}{2}\log\left(\frac{1+x}{1-x}\right) = \log\left(i \frac{1+z}{1-z}\right)$. Thus, \begin{equation} \label{tanhleft} \tanh((2n+1)\arctanh x) = \tanh\left( (2n+1)\log\left(i \frac{1+z}{1-z}\right) \right). \end{equation} On the other hand, the identity $\tanh(2y) = \frac{2\tanh y}{1+\tanh^2 y}$ shows that \begin{align} \frac{1}{2} &\left( \tanh((2n+1)\arctanh z) + \frac{1}{\tanh((2n+1)\arctanh z)}) \right) \nonumber \\ &= \coth((4n+2)\arctanh z) \nonumber \\ &= \coth\left( (2n+1)\log\left(\frac{1+z}{1-z}\right)\right). \label{tanhright} \end{align} Since $(2n+1)\log\left(\frac{1+z}{1-z}\right)$ differs from $(2n+1)\log\left(i \frac{1+z}{1-z}\right)$ by an odd multiple of $\frac{\pi i}{2}$, it follows that~(\ref{tanhleft}) and~(\ref{tanhright}) are equal. \end{proof} Written another way, the lemma above states that \begin{equation} \label{rplusrinv} \frac{1}{2} \left( r_{2n+1}(z;\Theta) + \frac{1}{r_{2n+1}(z;\Theta)} \right) = \widehat{R}_{2n+1}\left( \frac{z+1/z}{2}; \cos\Theta \right) \end{equation} for all $z$ with $|z|=1$. In particular, \[ \frac{1}{2}\left( zp_n(z^2) + \frac{1}{z p_n(z^2)} \right) = \left( \frac{z+1/z}{2}\right) p_n\left( \left(\frac{z+1/z}{2}\right)^2 \right). \] Since these equalities hold on the unit circle, they hold on all of $\mathbb{C}$. By combining~(\ref{composition0}),~(\ref{realpart}), and Theorem~\ref{thm:composition}, one sees that the function $\widehat{R}_{2n+1}(x;\ell)$ satisfies \begin{equation} \label{Rcomp} \widehat{R}_{2m+1}(\widehat{R}_{2n+1}(x,\ell),\widetilde{\ell}) = \widehat{R}_{(2m+1)(2n+1)}(x,\ell), \quad \text{ if } \widetilde{\ell} = \widehat{R}_{2n+1}(\ell,\ell) \end{equation} for all $m,n \in \mathbb{N}_0$ and all $\ell \in [0,1)$. This equality was derived in~\cite{nakatsukasa2016computing} for $\ell \in (0,1)$ by counting extrema of $\widehat{R}_{2m+1}(\widehat{R}_{2n+1}(x,\ell),\widetilde{\ell})-\sign(x)$. It can be leveraged to construct iterations for the matrix sign function, and such iterations are particularly well-suited for computing the sign of a Hermitian matrix $B$ (which coincides with the unitary factor in the polar decomposition of $B$); see~(\ref{zolopd1}-\ref{zolopd2}) below. \section{Algorithm} \label{sec:algorithm} \subsection{Matrix Iteration} Theorem~\ref{thm:composition} suggests the following iteration for computing the sign of a unitary matrix $A$ with spectrum contained in $\mathbb{S}_\Theta$, $\Theta \in [0,\pi/2)$: \begin{align} X_{k+1} &= r_{2n+1}(X_k; \Theta_k), & X_0 &= A,\label{zolo1} \\ \Theta_{k+1} &= |\arg r_{2n+1}(e^{i\Theta_k};\Theta_k)|, & \Theta_0 &= \Theta. \label{zolo2} \end{align} Below we summarize the properties of the iteration~(\ref{zolo1}-\ref{zolo2}). \begin{proposition} The iteration~(\ref{zolo1}-\ref{zolo2}) is structure-preserving. That is, if $A$ is unitary, then $X_k$ is unitary for every $k \ge 0$. \end{proposition} \begin{proof} Since $|r_{2n+1}(z;\Theta_k)|=1$ for every scalar $z$ with unit modulus, $r_{2n+1}(X; \Theta_k)$ is unitary for every unitary matrix $X$. \end{proof} \begin{theorem} Let $A$ be a unitary matrix with spectrum contained in $\mathbb{S}_\Theta$ for some $\Theta \in (0,\pi/2)$. For any $n \in \mathbb{N}$, the iteration~(\ref{zolo1}-\ref{zolo2}) converges to $\sign(A)$ with order of convergence $2n+1$. In fact, \begin{equation} \label{errorbound} \|\log(X_k \sign(A)^{-1})\| \le 4 \rho^{-(2n+1)^k}, \end{equation} for every $k \ge 0$, where $\rho$ is given by~(\ref{rho}). \end{theorem} \begin{proof} By Theorem~\ref{thm:composition}, we have \[ X_k = r_{(2n+1)^k}(A; \Theta) \] for every $k \ge 0$. Thus, every eigenvalue of $X_k \sign(A)^{-1}$ is of the form $r_{(2n+1)^k}(\lambda;\Theta) / \sign(\lambda)$ for some eigenvalue $\lambda$ of $A$. By Theorem~\ref{thm:error}, \begin{align*} \|\log(X_k \sign(A)^{-1})\| &= \max_{\lambda \in \Lambda(X_k \sign(A)^{-1})} |\arg\lambda| \\ &= \max_{\lambda \in \Lambda(A)} \left| \arg\left( \frac{r_{(2n+1)^k}(\lambda;\Theta)}{\sign(\lambda)} \right)\right| \\ &\le \max_{z \in \mathbb{S}_\Theta} \left| \arg\left(\frac{r_{(2n+1)^k}(z;\Theta)}{\sign(z)}\right) \right| \\ &\le 4 \rho^{-(2n+1)^k}. \end{align*} \end{proof} \subsection{Connections with Other Iterations} There is an intimate connection between the iteration~(\ref{zolo1}-\ref{zolo2}) and several existing iterations for the matrix sign function. First, Proposition~\ref{prop:pade} implies that~(\ref{zolo1}-\ref{zolo2}) reduces to the diagonal Pad\'e iteration when we set $\Theta=0$: \begin{equation} \label{pade} X_{k+1} = X_k p_n(X_k^2), \quad X_0 = A. \end{equation} Second, there is a link between the iteration~(\ref{zolo1}-\ref{zolo2}) and the iteration \begin{align} Y_{k+1} &= \widehat{R}_{2n+1}(Y_k; \ell_k), & Y_0 &= B,\label{zolopd1} \\ \ell_{k+1} &= \widehat{R}_{2n+1}(\ell_k;\ell_k), & \ell_0 &= \ell, \label{zolopd2} \end{align} which was introduced in~\cite{nakatsukasa2016computing} to compute the sign of a Hermitian matrix $B$ with spectrum contained in $[-1,-\ell] \cup [\ell,1]$. Note that~(\ref{zolopd1}-\ref{zolopd2}) reduces to \begin{align} Y_{k+1} &= Y_k p_n(Y_k^2), &\quad Y_0 &= B \label{padepd} \end{align} when we set $\ell=1$ and ignore the spectrum of $B$. This is the same iteration as~(\ref{pade}), but with a starting matrix labelled $B$ rather than $A$. \begin{proposition} Let $A$ be a unitary matrix with no eigenvalues equal to $\pm i$. Let $n \in \mathbb{N}$ and $\Theta \in [0,\pi/2)$. If $B = (A+A^*)/2$ and $\ell = \cos\Theta$, then the iterations~(\ref{zolo1}-\ref{zolo2}) and~(\ref{zolopd1}-\ref{zolopd2}) generate sequences satisfying \[ Y_k = \frac{1}{2}(X_k+X_k^*), \text{ and } \ell_k = \cos\Theta_k \] for every $k \ge 0$. \end{proposition} \begin{proof} It follows from Theorem~\ref{thm:composition} that in the iteration~(\ref{zolo1}-\ref{zolo2}), we have \[ X_k = r_{(2n+1)^k}(A;\Theta), \quad \Theta_k = |\arg r_{(2n+1)^k}(e^{i\Theta};\Theta)|, \] for each $k \ge 0$. On the other hand, the composition law~(\ref{Rcomp}) implies that in the iteration~(\ref{zolopd1}-\ref{zolopd2}), we have \[ Y_k = \widehat{R}_{(2n+1)^k}(B;\ell), \quad \ell_k = \widehat{R}_{(2n+1)^k}(\ell;\ell), \] for each $k \ge 0$. Thus, by~(\ref{rplusrinv}), \begin{align*} \frac{1}{2}(X_k+X_k^*) &= \frac{1}{2}(X_k+X_k^{-1}) \\ &= \frac{1}{2}\left(r_{(2n+1)^k}(A;\Theta) + r_{(2n+1)^k}(A;\Theta)^{-1}\right) \\ &= \widehat{R}_{(2n+1)^k}((A+A^{-1})/2;\cos\Theta) \\ &= \widehat{R}_{(2n+1)^k}(B;\ell) \\ &= Y_k. \end{align*} Also, by Theorem~\ref{thm:realpart}, \[ \cos\Theta_k = \Re e^{i\Theta_k} = \Re r_{(2n+1)^k}(e^{i\Theta};\Theta) = \widehat{R}_{(2n+1)^k}(\Re e^{i\Theta};\cos\Theta) = \widehat{R}_{(2n+1)^k}(\ell;\ell) = \ell_k. \] \end{proof} In the case that $\Theta=0$, the above result implies a connection between the diagonal Pad\'e iterations~(\ref{pade}) and~(\ref{padepd}). \begin{corollary} Let $A$ be a unitary matrix with no eigenvalues equal to $\pm i$, and let $n \in \mathbb{N}$. If $B = (A+A^*)/2$, then the diagonal Pad\'e iterations~(\ref{pade}) and~(\ref{padepd}) generate sequences satisfying \[ Y_k = \frac{1}{2}(X_k+X_k^*) \] for every $k \ge 0$. \end{corollary} \subsection{Implementation} To implement the $k$th step of the iteration~(\ref{zolo1}-\ref{zolo2}), one must compute products of unitary matrices of the form \begin{equation} \label{XaX} V_j = (X_k^2 + a_j I) (I+a_j X_k^2)^{-1} = (X_k + a_j X_k^*) (X_k^*+a_j X_k)^{-1}, \quad j=1,2,\dots,n, \end{equation} where $X_k$ is unitary. The following lemma describes a method for computing~(\ref{XaX}) that is guaranteed to produce a matrix that is unitary to machine precision. \begin{lemma} Let $B \in \mathbb{C}^{m \times m}$ be a nonsingular normal matrix. Let $Q_1 R_1 = B$ and $Q_2 R_2 = B^*$ be the QR factorizations of $B$ and $B^*$, respectively. Then \[ B B^{-*} = Q_1 Q_2^*. \] \end{lemma} \begin{proof} Since $R_1$ is the Cholesky factor of $B^*B$ and $R_2$ is the Cholesky factor of $BB^* = B^*B$, we have $R_1=R_2$. Hence, $BB^{-*} = Q_1 R_1 R_2^{-1} Q_2^* = Q_1 Q_2^*$. \end{proof} Once~(\ref{XaX}) has been computed for each $j$, one must decide in what order to multiply the matrices $V_1, V_2, \dots, V_n$, and $X_k$. Our numerical experience suggests that this decision has a strong influence on the backward stability of the algorithm. We find that the choice \begin{equation} \label{Xkp1} X_{k+1} = \frac{1}{2}(X_k V_1 V_2 \cdots V_n + V_n V_{n-1} \cdots V_1 X_k) \end{equation} is preferable to, for instance, $X_{k+1} = X_k V_1 V_2 \cdots V_n$ or $X_{k+1} = V_n V_{n-1} \cdots V_1 X_k$. This choice appears to guarantee that $\|X_k A - A X_k\|=O(u)$ for each $k$, which is essential for backward stability; see Lemma~\ref{lemma:backwardstability} for details. A proof that $\|X_k A - A X_k\|=O(u)$ when~(\ref{Xkp1}) is used remains an open problem. \paragraph{Termination} We must also decide how to terminate the iteration. Here we suggest terminating slightly early and applying two post-processing steps---symmetrization followed by one step of the Newton-Schulz iteration~\cite[Equation 8.20]{higham2008functions} for the polar decomposition---to ensure that the computed matrix $\widehat{S} \approx \sign(A)$ is Hermitian and unitary to machine precision. These post-processing steps have the following effect. Let $\{\sigma_j \cos\theta_j + i\sin\theta_j\}_{j=1}^m$ be the eigenvalues of $X_k$, where $\sigma_j \in \{-1,1\}$ and $|\theta_j|<\pi/2$ for each $j$. Then \begin{equation} \label{symmetrize} Y = \frac{1}{2}(X_k + X_k^*) \end{equation} has eigenvalues $\{\sigma_j \cos\theta_j \}_{j=1}^m$, and \begin{equation} \label{newtonschulz} Z = \frac{1}{2} Y(3I-Y^*Y) = \frac{1}{2} Y(3I-Y^2) \end{equation} has eigenvalues $\{ \frac{1}{2}\sigma_j \cos\theta_j (3 - \cos^2\theta_j) \}_{j=1}^m$. For small $\theta_j$, we have \[ \frac{1}{2}\sigma_j \cos\theta_j (3 - \cos^2\theta_j) = \sigma_j \left(1-\frac{3}{8}\theta_j^4\right) + O(\theta_j^6). \] This number will lie within a tolerance $\delta$ of $\pm 1$ if \begin{equation} \label{thetaconverged} \theta_j \lesssim \left( \frac{8\delta}{3} \right)^{1/4}. \end{equation} The above calculations suggest the following termination criterion. Since the eigenvalues of $X_k-X_k^*$ are $\{2i\sin\theta_j\}_{j=1}^m \approx \{2i\theta_j\}_{j=1}^m$, we terminate the iteration and carry out the post-processing steps~(\ref{symmetrize}-\ref{newtonschulz}) as soon as \[ \|X_k-X_k^*\| \le 2 \left( \frac{8\delta}{3} \right)^{1/4}. \] Note that since the Frobenius norm $\|\cdot\|_F$ is an upper bound for the $2$-norm $\|\cdot\|$, we may safely replace $\|X_k-X_k^*\|$ by $\|X_k-X_k^*\|_F$ in the criterion above. If desired, a second symmetrization can be performed after the Newton-Schulz step. This has virtually no effect on the eigenvalues' distance to $\pm 1$, but it may be desirable if an exactly Hermitian matrix is sought. \paragraph{Spectral angle} Let us also mention how to determine $\Theta$ so that $\Lambda(A) \subset \mathbb{S}_\Theta$. We hereafter refer to the smallest such $\Theta$ as the \emph{spectral angle} of $A$, denoted $\Theta(A)$. A simple heuristic is to estimate the eigenvalues $\lambda_+$ and $\lambda_-$ of $A$ that lie closest to $i$ and $-i$, respectively. Then one can set \[ \Theta = \max\{ \pi/2 - |\arg(i\lambda_-)|, |\arg(i\lambda_+)| - \pi/2 \}. \] In practice, it is not necessary to determine the spectral angle of $A$ precisely. Our experience suggests that underestimates and overestimates of $\Theta$ can be used without significant harm, unless $\Theta$ is very close to $\pi/2$. \paragraph{Spectral angles close to $\pi/2$} There are a few delicate numerical issues that arise when the spectral angle of $A$ is close to $\pi/2$. First, as noted in~\cite[Section 4.3]{nakatsukasa2016computing}, the built-in MATLAB functions \verb$ellipj$ and \verb$ellipke$ cannot be used to reliably compute $\sn(\cdot,\ell')$, $\cn(\cdot,\ell')$, $\dn(\cdot,\ell')$, and $K(\ell')$ when $\Theta = \arccos\ell$ is close to $\pi/2$. Instead, the code described in~\cite[Section 4.3]{nakatsukasa2016computing} is preferred. In addition, the lowest-order iteration ($n=1$) appears to be more reliable than the higher-order iterations when $\Theta>\pi/2-u^{1/2}$, so we advocate using the lowest-order iteration until $\Theta_k$ falls below $\pi/2-u^{1/2}$ (recall that $u=2^{-53}$ denotes the unit roundoff). Typically this takes two or fewer iterations, after which one can switch to a higher-order iteration if desired. To implement the lowest-order iteration ($n=1$) when $\Theta>\pi/2-u^{1/2}$, we have found the following heuristic to be useful for ensuring rapid convergence. If, at the $k$th iteration, $\Theta_k$ lies above $\pi/2-u^{1/2}$, we compute $\Theta_{k+1}$ as $\Theta_{k+1}=\Theta(X_{k+1})$ (the spectral angle of $X_{k+1}$) rather than via~(\ref{zolo2}). This tends to speed up the iteration. To improve stability, we have also found it prudent to replace $\Theta_k$ by $\pi/2-10u$ if $\Theta_k > \pi/2-10u$. A summary of our proposed algorithm for computing the unitary sign decomposition is presented in Algorithm~\ref{alg:zolosign}. \begin{algorithm} \caption{Order-$(2n+1)$ iteration for the unitary sign decomposition\newline \textit{Inputs}: Unitary matrix $A \in \mathbb{C}^{m \times m}$, tolerance $\delta>0$, degree $n \in \mathbb{N}$\newline \textit{Outputs}: Matrices $S,N \in \mathbb{C}^{m \times m}$ satisfying~(\ref{unitarysign})} \label{alg:zolosign} \begin{algorithmic}[1] \STATE{$\Theta_0 = \min\{\Theta(A),\pi/2-10u$\}} \label{line:theta0} \STATE{$X_0=A$, $n_0=n$, $k=0$} \WHILE{$\|X_k-X_k^*\|_F > 2(8\delta/3)^{1/4}$} \LINEIFELSE{$\Theta_k>\pi/2-u^{1/2}$}{$n=1$}{$n=n_0$} \STATE{$Y=X_k$, $Z=X_k$} \FOR{$j=1$ \TO $n$} \STATE{$Q_1 R_1 = X_k+a_j(\Theta_k) X_k^*$ (QR factorization)} \STATE{$Q_2 R_2 = X_k^*+a_j(\Theta_k) X_k$ (QR factorization)} \STATE{$Y= Y Q_1 Q_2^* $} \STATE{$Z = Q_1 Q_2^* Z$} \ENDFOR \STATE{$X_{k+1} = \frac{1}{2}(Y+Z)$} \IF{$\Theta_k>\pi/2-u^{1/2}$} \STATE{$\Theta_{k+1} = \min\{\Theta(X_{k+1}),\pi/2-10u\}$} \ELSE \STATE{$\Theta_{k+1} = |\arg r_{2n+1}(e^{i\Theta_k};\Theta_k)|$} \ENDIF \STATE{$k = k+1$} \ENDWHILE \STATE{$S = (X_k+X_k^*)/2$} \STATE{$S = S(3I-S^2)/2$} \STATE{$S = (S+S^*)/2$} \STATE{$N = S A$} \RETURN $S$, $N$ \end{algorithmic} \end{algorithm} \subsection{Backward Stability} We now discuss how some of the choices made above are inspired by backward stability considerations. We first address a remark that was made in the footnote of this paper's introduction concerning the list of backward errors~(\ref{backwarderrors}). At first glance, this list may appear to be incomplete because the norm of $\widehat{N}\widehat{S}-\widehat{S}\widehat{N}$ is absent. The following lemma shows that if $\widehat{S}$ and $\widehat{N}$ are well-conditioned matrices and $\|\widehat{N}^2-A^2\|$, $\|A-\widehat{S}\widehat{N}\|$, and $\|\widehat{S}^2-I\|$ are small, then $\|\widehat{N}\widehat{S}-\widehat{S}\widehat{N}\|$ is automatically small as well. \begin{lemma} \label{lemma:SNcommute} Let $A \in \mathbb{C}^{m \times m}$ be a unitary matrix. For any invertible matrices $\widehat{S},\widehat{N} \in \mathbb{C}^{m \times m}$, we have \begin{equation*} \|\widehat{N}\widehat{S}-\widehat{S}\widehat{N}\| \le \left( \|\widehat{N}^2-A^2\| + (1 + \|\widehat{S}\|\|\widehat{N}\|) \|A-\widehat{S}\widehat{N}\| + \|\widehat{N}\|^2 \|\widehat{S}^2-I\| \right) \|\widehat{N}^{-1}\| \|\widehat{S}^{-1}\|. \end{equation*} \end{lemma} \begin{proof} This follows from the identity \[ (\widehat{N}\widehat{S}-\widehat{S}\widehat{N})\widehat{S}\widehat{N} = \widehat{N}^2-A^2 + A(A-\widehat{S}\widehat{N}) + (A-\widehat{S}\widehat{N})\widehat{S}\widehat{N} + \widehat{N}(\widehat{S}^2-I)\widehat{N}. \] \end{proof} The next lemma shows that in order to achieve backward stability, it is prudent to compute a Hermitian matrix $\widehat{S}$ such that $\|\widehat{S}^2-I\|$ and $\|A\widehat{S}-\widehat{S}A\|$ are small, and then set $\widehat{N} = \widehat{S}A$. This highlights the importance of ensuring the smallness of $\|AX_k-X_kA\|$ in Algorithm~\ref{alg:zolosign}. \begin{lemma} \label{lemma:backwardstability} Let $A \in \mathbb{C}^{m \times m}$ be a unitary matrix, let $\widehat{S}$ be an invertible Hermitian matrix, and let $\widehat{N} = \widehat{S}A$. Then \begin{align} \|\widehat{N}^*\widehat{N}-I\| &\le \|\widehat{S}^2-I\|, \label{ineq1} \\ \|A-\widehat{S}\widehat{N}\| &\le \|\widehat{S}^2-I\|, \label{ineq2} \\ \|\widehat{N}^2-A^2\| &\le \|\widehat{S}\| \|A\widehat{S}-\widehat{S}A\| + \|\widehat{S}^2-I\|, \label{ineq3} \\ \|\widehat{N}\widehat{S}-\widehat{S}\widehat{N}\| &\le \|\widehat{S}\| \|A\widehat{S}-\widehat{S}A\|. \label{ineq4} \end{align} \end{lemma} \begin{proof} Since $A^*A=I$, $\widehat{N}=\widehat{S}A$, and $\widehat{S}=\widehat{S}^*$, we have \[ \widehat{N}^*\widehat{N}-I = A^*\widehat{S}^2 A - I = A^* (\widehat{S}^2-I)A. \] Taking the norm of both sides proves~(\ref{ineq1}). Similarly, the equalities \begin{align*} A-\widehat{S}\widehat{N} &= (I-\widehat{S}^2)A, \\ \widehat{N}^2-A^2 &= \widehat{S}A\widehat{S}A - A^2 = \widehat{S}(A\widehat{S}-\widehat{S}A)A + (\widehat{S}^2-I)A^2, \\ \widehat{N}\widehat{S}-\widehat{S}\widehat{N} &= \widehat{S}(A\widehat{S}-\widehat{S}A) \end{align*} yield~(\ref{ineq2}-\ref{ineq4}). \end{proof} \section{A Spectral Divide-and-Conquer Algorithm for the Unitary Eigendecomposition} \label{sec:eig} The iteration we have proposed for computing the unitary sign decomposition can be used to construct a spectral divide-and-conquer algorithm for the unitary eigendecomposition, following~\cite{nakatsukasa2013stable,nakatsukasa2016computing}. The idea is as follows. Given a unitary matrix $A \in \mathbb{C}^{m \times m}$, we scale $A$ by a complex number $e^{i\phi}$ so that roughly half (say, $m_1$) of the eigenvalues of $e^{i\phi}A$ lie in the right half of the complex plane, and roughly half (say, $m_2$) lie in the left half of complex plane. We then compute $S = \sign(e^{i\phi}A)$ using Algorithm~\ref{alg:zolosign}. The matrix $P=(I+S)/2$ is a spectral projector onto the invariant subspace $\mathcal{V}_+$ of $e^{i\phi}A$ associated with the eigenvalues of $e^{i\phi}A$ having positive real part. Using subspace iteration, we can compute orthonormal bases $U_1 \in \mathbb{C}^{m \times m_1}$ and $U_2 \in \mathbb{C}^{m \times m_2}$ (where $m_1+m_2=m$) for $\mathcal{V}_+$ and its orthogonal complement. Then \[ \begin{pmatrix} U_1^* \\ U_2^* \end{pmatrix} A \begin{pmatrix} U_1 & U_2 \end{pmatrix} = \begin{pmatrix} A_1 & 0 \\ 0 & A_2 \end{pmatrix} \] is block diagonal, so we can recurse to find eigendecompositions $A_1 = V_1 \Lambda_1 V_1^*$ and $A_2 = V_2 \Lambda_2 V_2^*$. The eigendecomposition of $A$ is then $A = V\Lambda V^*$, where \[ V = \begin{pmatrix} U_1 V_1 & U_2 V_2 \end{pmatrix} \] and \[ \Lambda = \begin{pmatrix} \Lambda_1 & 0 \\ 0 & \Lambda_2 \end{pmatrix}. \] Since every eigenvalue of $P$ is either $0$ and $1$, subspace iteration with $P$ typically converges in one iteration, or, in rare cases, two. To choose the scalar $e^{i\phi}$, a simple heuristic is to compute the median $\mu$ of the arguments of the diagonal entries of $A$ and set $\phi=\pi/2-\mu$. When $A$ is nearly diagonal, this has the effect of centering the eigenvalues around $i$. A summary of the algorithm just described is presented in Algorithm~\ref{alg:eig}. \begin{algorithm} \caption{Divide-and-conquer algorithm for the unitary eigendecomposition\newline \textit{Inputs}: Unitary matrix $A \in \mathbb{C}^{m \times m}$ \newline \textit{Outputs}: Matrices $V,\Lambda \in \mathbb{C}^{m \times m}$ satisfying $V\Lambda V^* = A$, $V^*V = I$, and $\Lambda$ diagonal} \label{alg:eig} \begin{algorithmic}[1] \STATE{$\phi = \frac{\pi}{2} - \operatorname{median} \{\arg A_{11},\dots,\arg A_{mm}\}$} \STATE{$S = \sign(e^{i\phi} A)$} \label{line:sign} \STATE{$P = (I+S)/2$} \STATE{Use subspace iteration to compute orthonormal bases $U_1 \in \mathbb{C}^{m \times m_1}$ and $U_2 \in \mathbb{C}^{m \times m_2}$ for the 0- and 1-eigenspaces of $P$.} \STATE{$A_1 = U_1^* A U_1$, $A_2 = U_2^* A U_2$} \STATE{Recurse to find eigendecompositions $V_1 \Lambda_1 V_1^* = A_1$ and $V_2 \Lambda_2 V_2^* = A_2$.} \STATE{$V = \begin{pmatrix} U_1 V_1 & U_2 V_2 \end{pmatrix}$} \STATE{$\Lambda = \begin{pmatrix} \Lambda_1 & 0 \\ 0 & \Lambda_2 \end{pmatrix}$} \RETURN $V$, $\Lambda$ \end{algorithmic} \end{algorithm} \section{Numerical Examples} \label{sec:numerical} In this section, we study the iteration~(\ref{zolo1}-\ref{zolo2}) numerically, and we test Algorithms~\ref{alg:zolosign} and~\ref{alg:eig} on a collection of unitary matrices. \subsection{Scalar Iteration} \begin{table} \centering \begin{tabularx}{\linewidth}{ Y*{11}{Y} } & \multicolumn{11}{c}{$\frac{\pi}{2}-\Theta$} \\ $n$ & 1.5 & 1 & 0.5 & $10^{-2}$ & $10^{-4}$ & $10^{-6}$ & $10^{-8}$ & $10^{-10}$ & $10^{-12}$ & $10^{-14}$ & $10^{-16}$ \\ \cmidrule(lr){1-1} \cmidrule(lr){2-12} 1 & 1 & 2 & 2 & 3 & 4 & 4 & 5 & 5 & 5 & 5 & 5 \\ 2 & 1 & 2 & 2 & 3 & 3 & 3 & 3 & 3 & 3 & 4 & 4 \\ 3 & 1 & 1 & 2 & 2 & 2 & 3 & 3 & 3 & 3 & 3 & 3 \\ 4 & 1 & 1 & 1 & 2 & 2 & 2 & 3 & 3 & 3 & 3 & 3 \\ 5 & 1 & 1 & 1 & 2 & 2 & 2 & 2 & 2 & 3 & 3 & 3 \\ 6 & 1 & 1 & 1 & 2 & 2 & 2 & 2 & 2 & 2 & 2 & 2 \\ 7 & 1 & 1 & 1 & 2 & 2 & 2 & 2 & 2 & 2 & 2 & 2 \\ 8 & 1 & 1 & 1 & 2 & 2 & 2 & 2 & 2 & 2 & 2 & 2 \\ \end{tabularx} \caption{Smallest integer $k$ for which $4\rho(\Theta)^{-(2n+1)^k} \le (8\delta/3)^{1/4}$, where $\delta = 10^{-16}$, for various values of $n$ and $\Theta$.} \label{tab:iter} \end{table} \begin{table} \centering \begin{tabularx}{\linewidth}{ Y*{11}{Y} } & \multicolumn{11}{c}{$\frac{\pi}{2}-\Theta$} \\ $n$ & 1.5 & 1 & 0.5 & $10^{-2}$ & $10^{-4}$ & $10^{-6}$ & $10^{-8}$ & $10^{-10}$ & $10^{-12}$ & $10^{-14}$ & $10^{-16}$ \\ \cmidrule(lr){1-1} \cmidrule(lr){2-12} 1 & 1 & 2 & 3 & 7 & 11 & 15 & 19 & 24 & 28 & 32 & 37 \\ 2 & 1 & 2 & 2 & 5 & 8 & 10 & 13 & 16 & 19 & 22 & 25 \\ 3 & 1 & 2 & 2 & 4 & 6 & 9 & 11 & 13 & 16 & 18 & 21 \\ 4 & 1 & 1 & 2 & 4 & 6 & 8 & 10 & 12 & 14 & 16 & 19 \\ 5 & 1 & 1 & 2 & 3 & 5 & 7 & 9 & 11 & 13 & 15 & 17 \\ 6 & 1 & 1 & 2 & 3 & 5 & 7 & 9 & 10 & 12 & 14 & 16 \\ 7 & 1 & 1 & 2 & 3 & 5 & 6 & 8 & 10 & 12 & 13 & 15 \\ 8 & 1 & 1 & 2 & 3 & 5 & 6 & 8 & 9 & 11 & 13 & 14 \end{tabularx} \caption{Smallest integer $k$ for which $|r_{(2n+1)^k}(e^{i\Theta};0)-1| \le (8\delta/3)^{1/4}$, where $\delta = 10^{-16}$, for various values of $n$ and $\Theta$.} \label{tab:iterpade} \end{table} To understand how rapidly the iteration~(\ref{zolo1}-\ref{zolo2}) can be expected to converge, let us study the upper bound~(\ref{errorbound}). Table~\ref{tab:iter} reports the smallest integer $k$ for which $4\rho(\Theta)^{-(2n+1)^k}$ falls below the number $(8\delta/3)^{1/4}$ appearing in the convergence criterion~(\ref{thetaconverged}). Here, we took $\delta = 10^{-16}$ and considered various choices of $n$ and $\Theta$. The integer $k$ so computed provides an estimate for the number of iterations one can expect~(\ref{zolo1}-\ref{zolo2}) to take to converge to $\sign(A)$ if $A$ has spectrum contained in $\mathbb{S}_\Theta$. For comparison, we computed the number of iterations needed for the scalar Pad\'e iteration \[ z_{k+1} = r_{2n+1}(z_k;0) = z_k p_n(z_k^2) \] to converge to $\sign z_0$, starting from $z_0 = e^{i\Theta}$. The results, reported in Table~\ref{tab:iterpade}, show that the Pad\'e iterations take significantly longer to converge if $\Theta$ is close to $\pi/2$. This suggests the matrix Pad\'e iteration~(\ref{pade}) will require a large number of iterations to converge to $\sign(A)$ if the spectral angle $\Theta(A)$ is close to $\pi/2$. \subsection{Matrix Iteration} To test Algorithm~\ref{alg:zolosign}, we computed the sign decomposition of four unitary matrices: \begin{enumerate} \item \label{mat1} A matrix sampled randomly from the Haar measure on the $m \times m$ unitary group. \item \label{mat2} \verb$A = gallery('orthog',m,3)$. This is the $m$-point discrete Fourier transform matrix with entries $A_{jk} = e^{2\pi i (j-1)(k-1)/m} / \sqrt{m}$. Its eigenvalues are $1,-1,i,-i$. The spectrum of the floating point representation of $A$ therefore includes $O(u)$-perturbations of $\pm i$, posing a challenge to numerical algorithms for the unitary sign decomposition. \item \label{mat3} \verb$A = circshift(eye(m),1)$. This is a permutation matrix with eigenvalues $e^{2\pi i j/m}$, $m=1,2,\dots,m$. For even $m$, the spectrum of $A$ includes $\pm i$. The same is true of the floating point representation of $A$, since the entries of $A$ are integers. \item \label{mat4} \verb$A = gallery('orthog',m,-2)$ (with columns normalized). The entries of $A$ (prior to normalizing columns) are $A_{jk} = \cos((k-1/2)(j-1)\pi/m)$. The spectrum of $A$ is clustered near $\pm 1$, making its sign decomposition somewhat easy to compute iteratively. \end{enumerate} \medskip In our numerical experiment, we used $m=100$. The computed spectral angles for the matrices above were $\pi/2-\Theta(A) = 0.026$, $4.4 \times 10^{-16}$, $0$, and $0.95$, respectively. On each of the matrices above, we compared 10 algorithms: \begin{itemize} \item Algorithm~\ref{alg:zolosign} with $n=1,4,8$. \item The diagonal Pad\'e iteration~(\ref{pade}) with $n=1,4,8$. We implemented this by running Algorithm~\ref{alg:zolosign} with line~\ref{line:theta0} replaced by $\Theta_0=0$. \item Three algorithms that compute the unitary factor $S$ in the polar decomposition of $B=(A+A^*)/2$. The first uses the Newton iteration with $1,\infty$-norm scaling, as described in~\cite[Section 8.6]{higham2008functions} and implemented in~\cite{Higham:MCT}. The second uses the Zolo-pd algorithm from~\cite{nakatsukasa2016computing}. The third computes $S$ as $S=UV^*$, where $B=U\Sigma V^*$ is the SVD of $B$. In all three cases, we applied post-processing to $S$ ($S=(S+S^*)/2$, followed by $S=S(3I-S^2)/2$, followed by $S=(S+S^*)/2$) and set $N=SA$. \item A direct method: computing the eigendecomposition $A=V\Lambda V^*$ of $A$ and setting $S = V\sign(\Lambda)V^*$. We computed the eigendecomposition by using the MATLAB command \verb$schur(A,'complex')$ and setting the off-diagonal entries of the triangular factor to zero. We applied post-processing to $S$ ($S=S(3I-S^2)/2$ followed by $S=(S+S^*)/2$) and set $N=SA$. \end{itemize} \medskip The results of the tests are reported in Table~\ref{tab:sign}. All of the algorithms under consideration performed in a backward stable way on the first and fourth matrices. On the second and third matrices (\verb$gallery('orthog',m,3)$ and \verb$circshift(eye(m),1)$), only the direct method and the structure-preserving iterations (Algorithm~\ref{alg:zolosign} and the Pad\'e iteration~(\ref{pade})) exhibited backward stability. Among the structure-preserving iterations, Algorithm~\ref{alg:zolosign} consistently converged more quickly than the Pad\'e iteration~(\ref{pade}) for each degree $n$. The reduction in iteration count was particularly noticeable for \verb$gallery('orthog',m,3)$ and \verb$circshift(eye(m),1)$. \begin{table}[t] \centering \pgfplotstabletypeset[ every head row/.style={ after row=\midrule}, every nth row={10}{before row=\midrule}, columns={leftcol2,0,1,2,3,4,5}, create on use/leftcol2/.style={create col/set list={ Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct, Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct, Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct, Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct}}, columns/leftcol/.style={string type,column type/.add={}{},column name={$\frac{\pi}{2}-\Theta(A)$}}, columns/leftcol/.style={string type,column type/.add={}{},column name={$\pi/2-\Theta(A)$}}, columns/leftcol2/.style={string type,column type/.add={}{},column name={Algorithm}}, columns/0/.style={column type/.add={}{},column name={$k$}}, columns/1/.style={sci,sci e,sci zerofill,precision=1,column type/.add={}{},column name={$\|A-\widehat{S}\widehat{N}\|$}}, columns/2/.style={sci,sci e,sci zerofill,precision=1,column type/.add={}{},column name={$\|\widehat{S}^2-I\|$}}, columns/3/.style={sci,sci e,sci zerofill,precision=1,column type/.add={}{},column name={$\|\widehat{N}^*\widehat{N}-I\|$}}, columns/4/.style={sci,sci e,sci zerofill,precision=1,column type/.add={}{},column name={$\|\widehat{N}^2-A^2\|$}}, columns/5/.style={sci,sci e,sci zerofill,precision=1,column type/.add={}{},column name={$\mu(\widehat{N})$}} ] {sign.dat} \caption{Performance of algorithms for computing the unitary sign decomposition of the matrices~\ref{mat1}-\ref{mat4}. The table reports the iteration count $k$ and backward errors $\|A-\widehat{S}\widehat{N}\|$, $\|\widehat{S}^2-I\|$, $\|\widehat{N}^*\widehat{N}-I\|$, $\|\widehat{N}^2-A^2\|$, $\mu(\widehat{N})=\max\{0,-\min_{\lambda \in \Lambda(\widehat{N})} \Re\lambda\}$ for each algorithm.} \label{tab:sign} \end{table} \subsection{Unitary eigendecomposition} Next, we tested our spectral divide-and-conquer algorithm~\ref{alg:eig} on the same four matrices. We implemented line~\ref{line:sign} of Algorithm~\ref{alg:eig} in nine different ways, namely, by using the nine indirect methods considered in the previous experiment. We compared the results with the following direct method: \verb$[V,Lambda]=schur(A,'complex'); Lambda = diag(diag(Lambda))$. The results are reported in Table~\ref{tab:eig}. All of the algorithms under consideration performed in a backward stable way on the first, second, and fourth matrices. On the third matrix \verb$circshift(eye(m),1)$, the algorithms that used Zolo-pd and the SVD did not. Curiously, the algorithm that used the Newton iteration succeeded, but this is an anomaly. Changing \verb$circshift(eye(m),1)$ to \verb$circshift(eye(m),1)+eps*randn(m)$ leads to a backward error $\|A-\widehat{V}\widehat{\Lambda}\widehat{V}^*\|$ close to 0.1 for the Newton-based algorithm, and it has a negligible effect on the other algorithms' backward errors. \begin{table}[t] \hspace{-0.25in} \pgfplotstabletypeset[ every head row/.style={after row=\midrule}, every nth row={10}{before row=\midrule}, columns={leftcol2,0,1}, create on use/leftcol2/.style={create col/set list={ Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct, Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct, Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct, Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct}}, columns/leftcol/.style={string type,column type/.add={}{},column name={$\frac{\pi}{2}-\Theta(A)$}}, columns/leftcol2/.style={string type,column type/.add={}{},column name={Algorithm}}, columns/0/.style={sci,sci e,sci zerofill,precision=1,column type/.add={}{},column name={$\|A-\widehat{V}\widehat{\Lambda}\widehat{V}^*\|$}}, columns/1/.style={sci,sci e,sci zerofill,precision=1,column type/.add={}{},column name={$\|\widehat{V}^*\widehat{V}-I\|$}} ] {eig12.dat} \hspace{0.1in} \pgfplotstabletypeset[ every head row/.style={ after row=\midrule}, every nth row={10}{before row=\midrule}, columns={leftcol2,0,1}, create on use/leftcol2/.style={create col/set list={ Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct, Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct, Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct, Alg.~\ref{alg:zolosign} ($n=1$),Alg.~\ref{alg:zolosign} ($n=4$),Alg.~\ref{alg:zolosign} ($n=8$), Pad\'e ($n=1$), Pad\'e ($n=4$), Pad\'e ($n=8$), Polar (Newton), Polar (Zolo-pd), Polar (SVD), Direct}}, columns/leftcol/.style={string type,column type/.add={}{},column name={$\frac{\pi}{2}-\Theta(A)$}}, columns/leftcol2/.style={string type,column type/.add={}{},column name={Algorithm}}, columns/0/.style={sci,sci e,sci zerofill,precision=1,column type/.add={}{},column name={$\|A-\widehat{V}\widehat{\Lambda}\widehat{V}^*\|$}}, columns/1/.style={sci,sci e,sci zerofill,precision=1,column type/.add={}{},column name={$\|\widehat{V}^*\widehat{V}-I\|$}} ] {eig34.dat} \caption{Performance of algorithms for computing the unitary eigendecomposition of the matrices~\ref{mat1}-\ref{mat2} (left) and~(\ref{mat3}-\ref{mat4}) (right). With the exception of the entries labeled ``Direct'', the entries reported in column 1 refer to the algorithms for the unitary sign decomposition used in line~\ref{line:sign} of Algorithm~\ref{alg:eig}.} \label{tab:eig} \end{table} \section{Conclusion} \label{sec:conclusion} This paper constructed structure-preserving iterations for computing the unitary sign decomposition using rational minimax approximants of the scalar function $\sign(z)$ on the unit circle. Relative to other structure-preserving iterations, they converge significantly faster, and relative to non-structure-preserving iterations, they exhibit much better numerical stability. We used our iterations to construct a spectral divide-and-conquer algorithm for the unitary eigendecomposition.
{ "timestamp": "2020-11-26T02:07:00", "yymm": "2011", "arxiv_id": "2011.12449", "language": "en", "url": "https://arxiv.org/abs/2011.12449", "abstract": "We construct fast, structure-preserving iterations for computing the sign decomposition of a unitary matrix $A$ with no eigenvalues equal to $\\pm i$. This decomposition factorizes $A$ as the product of an involutory matrix $S = \\operatorname{sign}(A) = A(A^2)^{-1/2}$ times a matrix $N = (A^2)^{1/2}$ with spectrum contained in the open right half of the complex plane. Our iterations rely on a recently discovered formula for the best (in the minimax sense) unimodular rational approximant of the scalar function $\\operatorname{sign}(z) = z/\\sqrt{z^2}$ on subsets of the unit circle. When $A$ has eigenvalues near $\\pm i$, the iterations converge significantly faster than Padé iterations. Numerical evidence indicates that the iterations are backward stable, with backward errors often smaller than those obtained with direct methods. This contrasts with other iterations like the scaled Newton iteration, which suffers from numerical instabilities if $A$ has eigenvalues near $\\pm i$. As an application, we use our iterations to construct a stable spectral divide-and-conquer algorithm for the unitary eigendecomposition.", "subjects": "Numerical Analysis (math.NA)", "title": "Iterations for the Unitary Sign Decomposition and the Unitary Eigendecomposition", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9854964190307041, "lm_q2_score": 0.8128673110375457, "lm_q1q2_score": 0.8010778241746188 }
https://arxiv.org/abs/1303.6012
Are the Snapshot Difference Quotients Needed in the Proper Orthogonal Decomposition?
This paper presents a theoretical and numerical investigation of the following practical question: Should the time difference quotients of the snapshots be used to generate the proper orthogonal decomposition basis functions? The answer to this question is important, since some published numerical studies use the time difference quotients, whereas other numerical studies do not. The criterion used in this paper to answer this question is the rate of convergence of the error of the reduced order model with respect to the number of proper orthogonal decomposition basis functions. Two cases are considered: the no_DQ case, in which the snapshot difference quotients are not used, and the DQ case, in which the snapshot difference quotients are used. The error estimates suggest that the convergence rates in the $C^0(L^2)$-norm and in the $C^0(H^1)$-norm are optimal for the DQ case, but suboptimal for the no_DQ case. The convergence rates in the $L^2(H^1)$-norm are optimal for both the DQ case and the no_DQ case. Numerical tests are conducted on the heat equation and on the Burgers equation. The numerical results support the conclusions drawn from the theoretical error estimates. Overall, the theoretical and numerical results strongly suggest that, in order to achieve optimal pointwise in time rates of convergence with respect to the number of proper orthogonal decomposition basis functions, one should use the snapshot difference quotients.
\section{Conclusions} \label{sec:conclusions} The effect of using or not the snapshot DQs in the generation of the POD basis (the $DQ$ and the $no\_DQ$ cases, respectively) was investigated theoretically and numerically. The criterion used in this theoretical and numerical investigation was the rate of convergence with respect to $r$ of the POD-G-ROM solution to the exact solution, where $r$ is the number of POD basis functions used in the POD-G-ROM. The error estimates in Section~\ref{sec:error} yielded the following conclusions: In the $DQ$ case, the convergence rates were optimal in all three norms considered (the $C^0(L^2)$-norm, the $C^0(H^1)$-norm and the $L^2(H^1)$-norm. In the $no\_DQ$ case, the convergence rates were suboptimal in the $C^0(L^2)$-norm and in the $C^0(H^1)$-norm, and optimal in the $L^2(H^1)$-norm. The numerical results in Section~\ref{sec:numerical} for the (linear) heat equation and the (nonlinear) Burgers equation confirmed the conclusions suggested by the theoretical error estimates in Section~\ref{sec:error}: In the $DQ$ case, the convergence rates were superoptimal in the $C^0(L^2)$-norm, the $C^0(H^1)$-norm, and the $L^2(H^1)$-norm. In the $no\_DQ$ case, the convergence rates were suboptimal in the $C^0(L^2)$-norm, the $C^0(H^1)$-norm, and the $L^2(H^1)$-norm. The only departure from the theoretical conclusions was that, in the $no\_DQ$ case, the convergence rate in the $L^2(H^1)$-norm was suboptimal. We emphasize that, for both the heat equation and the Burgers equation, {\it the convergence rates in the $DQ$ case in all three norms were much higher than (and usually at least twice as high as) the corresponding rates of convergence in the $no\_DQ$ case}. The theoretical error estimates in Section~\ref{sec:error} and the numerical results in Section~\ref{sec:numerical} strongly suggest the following conjecture: {\it ``The snapshot DQs should be used in the generation of the POD basis in order to achieve optimal rates of convergence with respect to $r$, the number of POD basis functions utilized in the POD-G-ROM. } We also conjecture that using the snapshot DQs in the generation of the POD basis could alleviate some of the degrading of convergence with respect to $r$ seen in, e.g., \cite{rowley2004model,carlberg2011efficient,baiges2012explicit,caiazzo2013numerical}. We intend to investigate this conjecture in a future study. \section{Error Estimates} \label{sec:error} In this section, we prove estimates for the error $ u - u_r $, where $u$ is the solution of the weak formulation of the heat equation \eqref{eqn:heat_weak} and $u_r$ is the solution of the POD-G-ROM~\eqref{eqn:pod_g_rom}. Error estimates for the POD reduced order modeling of general systems were derived in, e.g., \cite{KV01,KV02,veroy2005certified,HPS07,LCNY08,rozza2008reduced,kalashnikova2010stability,drohmann2012reduced,urban2012new,amsallem2013error,sachs2013priori,herkt2013convergence} In our theoretical analysis, we consider two cases, depending on the type of snapshots used in the derivation of the POD basis: {\bf Case I}: $V = V^{DQ}$ (i.e., with the DQs); and {\bf Case II}: $V = V^{no\_DQ}$ (i.e., without the DQs). The main goal of this paper is to investigate whether {\bf Case I}, {\bf Case II}, or both {\bf Case I} and {\bf Case II}, yield {\it error estimates that are optimal with respect to $r$}. The optimality with respect to $r$ is given by the following error estimates: \begin{eqnarray} \| u - u_r \| &\leq& C \, \| \eta^{interp} \| \, , \label{eqn:optimality_eta} \\ \| \nabla (u - u_r) \| &\leq& C \, \| \nabla \eta^{interp}\| \, , \label{eqn:optimality_nabla_eta} \end{eqnarray} where $\eta^{interp}$ is the POD interpolation error defined in~\eqref{eqn:definition_interpolation_error}. We emphasize that $\| \eta^{interp} \|$ and $\| \nabla \eta^{interp}\|$ scale differently with respect to $r$: The scaling of $\| \eta^{interp} \|$ is given by the POD approximation property~\eqref{pod_error_formula_nodq_pointwise}--\eqref{pod_error_formula_dq_pointwise} in Assumption~\ref{assumption_pod_error_formula_pointwise}. The scaling of $\| \nabla \eta^{interp} \|$ is {\it not} given by the POD approximation property~\eqref{pod_error_formula_nodq_pointwise}--\eqref{pod_error_formula_dq_pointwise} in Assumption~\ref{assumption_pod_error_formula_pointwise}. To derive such an estimate, we use the fact that the interpolation error lives in a finite dimensional space, i.e., the space spanned by the snapshots. Using an inverse estimate similar to that presented in Lemma~\ref{lemma_inverse_pod} but for the entire space of snapshots (of dimension $d$), we get the following estimate: \begin{eqnarray} \| \nabla \eta^{interp} \|_{L^2} \leq C_{inv}(d) \, \| \eta^{interp} \|_{L^2} \, , \label{eqn:inverse_estimate_interpolation_error} \end{eqnarray} where $C_{inv}(d)$ is the constant in the inverse estimate in Lemma~\ref{lemma_inverse_pod}. Following the discussion in Remark~\ref{remark_pod_inverse_estimate_scalings}, we conclude that the scaling of $\| \nabla \eta^{interp} \|$ is of lower order with respect to $r$ than the scaling of $\| \eta^{interp} \|$. Thus, if the error analysis yields estimates of the form \begin{eqnarray} \| u - u_r \| \leq C \, \| \nabla \eta^{interp}\| \, , \label{eqn:suboptimal_error_estimate} \end{eqnarray} these estimates will be called {\it suboptimal} with respect to $r$. In this section, we investigate the optimality question from a theoretical point of view, by monitoring the dependency of the error estimates on $r$. In Section~\ref{sec:numerical}, we investigate the same question from a numerical point of view. We note that we perform the error analysis only for the {\it semidiscretization} of the POD-G-ROM~\eqref{eqn:pod_g_rom}. In fact, in this semidiscretization, we only consider the error component corresponding to the POD truncation. Of course, in practical numerical simulations, the semidiscretization also has a spatial component (e.g., due to the finite element discretization -- see Section~\ref{sec:numerical}). Furthermore, when considering the full discretization, the error also has a time discretization component (e.g., due to the time stepping algorithm -- see Section~\ref{sec:numerical}). All these error components should be included in a rigorous error analysis of the discretization of the POD-G-ROM~\eqref{eqn:pod_g_rom} (see, e.g., \cite{LCNY08,iliescu2012variational,iliescu2013variational}). For clarity of presentation, however, we only consider the error component corresponding to the POD truncation. In what follows, we will show that this is sufficient for answering the question asked in the title of this paper. We start by introducing some notation and we list several results that will be used throughout this section. We note that, since the POD basis is computed in the $L^2$-norm (see~\eqref{pod_min}), the POD mass matrix $M_{r} \in \mathbb{R}^{r \times r}$ with $M_{i j} = (\varphi_j , \varphi_i)$ is the identity matrix. Thus, the POD inverse estimate that was proven in \cite{iliescu2012variational} (see also Lemma 2 and Remark 2 in~\cite{KV01}) becomes: \begin{lemma}[POD Inverse Estimate] Let $S_{r} \in \mathbb{R}^{r \times r}$ with $S_{i j} = (\nabla \varphi_j , \nabla \varphi_i)$ be the POD stiffness matrix and let $\| \cdot \|_2$ denote the matrix 2-norm. Then, for all $v_r \in X^r$, the following POD inverse estimate holds: \begin{eqnarray} \| \nabla v_r \|_{L^2} \leq C_{inv}(r) \, \| v_r \|_{L^2} \, , \label{lemma_inverse_pod_1} \end{eqnarray} where $C_{inv}(r) := \sqrt{\| S_r \|_2}$. \label{lemma_inverse_pod} \end{lemma} \begin{remark}[POD Inverse Estimate Scalings] \label{remark_pod_inverse_estimate_scalings} Since the $r$ dependency of the error estimates presented in this section will be carefully monitored, we try to get some insight into the scalings of the constant $C_{inv}(r)$ in \eqref{lemma_inverse_pod_1}, i.e., the scalings of $\| S_r \|_2$ with respect to $r$, the dimension of the POD basis. We note that, since the POD basis significantly varies from test case to test case, it would be difficult to derive a general scaling of $\| S_r \|_2$. We emphasize, however, that when the underlying system is {\it homogeneous} (i.e., invariant to spatial translations), the POD basis is identical to the Fourier basis (see, e.g., \cite{HLB96}). In this case, it is easy to derive the scalings of $\| S_r \|_2$ with respect to $r$. Without loss of generality, we assume that the computational domain is $[0,1] \subset \mathbb{R}^{1}$ and that the boundary conditions are homogeneous Dirichlet. In this case, the Fourier basis functions are given by the following formula: $\varphi_j(x) = \sin(j \, \pi \, x)$. It is a simple calculation to show that the matrix $S_r$ is diagonal and that its diagonal entries are given by the following formula: \begin{eqnarray} S_{j j} = \int_{0}^{1} ( j \, \pi)^2 \, \sin^2(j \, \pi \, x) \, dx = \frac{1}{2} \, ( j \, \pi)^2 \, . \label{eqn:pod_stiffness_matrix_entries_1d} \end{eqnarray} It is easy to see that, in the $n$-dimensional case, formula~\eqref{eqn:pod_stiffness_matrix_entries_1d} becomes \begin{eqnarray} S_{j j} = \int_{\Omega} ( j \, \pi)^{2 \, n} \, \sin^2(j \, \pi \, x) \, dx = \frac{1}{2} \, ( j \, \pi)^{2 \, n} \, . \label{eqn:pod_stiffness_matrix_entries_nd} \end{eqnarray} Since the POD stiffness matrix $S_r$ is symmetric, its matrix 2-norm is given by $\| S_r \|_2 = \lambda_{max}$, where $\lambda_{max}$ is the largest eigenvalue of $S_r$. Thus, in the $n$-dimensional case, we have \begin{eqnarray} \| S_r \|_2 = \frac{1}{2} \, ( r \, \pi)^{2 \, n} = \mathcal{O}(r^{2 \, n}) \, . \label{eqn:pod_stiffness_matrix_scaling_nd} \end{eqnarray} Thus, we conclude that, when the underlying system is homogeneous, the 2-norm of the POD stiffness matrix $S_r$ scales as $\mathcal{O}(r^{2 \, n})$, where $n$ is the spatial dimension. As mentioned at the beginning of the remark, for general (non-homogeneous) systems is would be hard to derive theoretical scalings. The numerical tests in Section~\ref{sec:numerical}, however, seem to confirm the theoretical scaling in \eqref{eqn:pod_stiffness_matrix_scaling_nd}. For the heat equation and the Burgers equation (see Section~\ref{sec:numerical} for details regarding the numerical simulations), we monitor the scaling of $\|S_r\|_2$ with respect to $r$ for the POD-G-ROM~\eqref{eqn:pod_g_rom}. We consider two cases: when the DQs are used in the generation of the POD basis (the corresponding results are denoted by DQ), and when the DQs are not used in the generation of the POD basis (the corresponding results are denoted by no-DQ). The scalings are plotted in Figure \ref{fig:test1_Sr}. It is seen that both no-DQ and DQ approaches yields scalings that are similar to the theoretical scaling~\eqref{eqn:pod_stiffness_matrix_scaling_nd} predicted for the homogeneous flow fields (i.e., when the POD basis reduces to the Fourier basis). The only exception seems to be for the Burgers equation in the DQ case. In all cases, however, the scaling $\| S_r \|_2 = \mathcal{O}(r^{\alpha})$, where $\alpha$ is a positive constant, seems to be valid. \begin{figure}[htp] \begin{minipage}[h]{.45\linewidth} \includegraphics[width=1.1\textwidth]{Heat_Sr_wrt_r.eps} \end{minipage} \hspace{.3cm} \begin{minipage}[h]{.45\linewidth} \includegraphics[width=1.1\textwidth]{Burg_Sr_wrt_r.eps} \end{minipage} \caption{ Heat equation (left); Burgers equation (right). Plots of the scalings of $\|S_r\|_2$ with respect to $r$ when the DQs are used (denoted by DQ) and when the DQs are not used (denoted by no-DQ). } \label{fig:test1_Sr} \end{figure} \end{remark} \medskip After these preliminaries, we are ready to derive the error estimates. The error analysis will proceed along the same lines as the error analysis of the finite element semidiscretization~\cite{GR86,layton2008introduction,thomee2006galerkin}. The main difference between the finite element and the POD settings is that the finite element approximation property is {\it global} \cite{GR86,layton2008introduction}, whereas the POD approximation property is {\it local}, i.e., it is only valid at the time instances at which the snapshots were taken (see~\eqref{pod_error_formula_nodq}-\eqref{pod_error_formula_dq_pointwise}). Thus, in order to be able to use the POD approximation property~\eqref{pod_error_formula_nodq}-\eqref{pod_error_formula_dq_pointwise}, in what follows we assume that the final error estimates for the semidiscretization are considered only at the time instances $t_i, \, i =1, \ldots, N$. We start by considering the error equation: \begin{eqnarray} ( e_t , v_r ) + \nu \, ( \nabla e , \nabla v_r ) = 0 \qquad \forall \, v_r \in X^r , \label{s_error_semi_eq_1} \end{eqnarray} where $e := u - u_r$ is the error. The error is split into two parts: \begin{eqnarray} e = u - u_r = ( u - w_r ) - ( u_r - w_r ) = \eta - \phi_r , \label{s_error_semi_eq_2} \end{eqnarray} where $w_r$ is an {\it arbitrary} function in $X^r$, $\eta := u - w_r$, and $\phi_r := u_r - w_r$. Using this decomposition in the error equation \eqref{s_error_semi_eq_1}, we get \begin{eqnarray} ( \phi_{r,t} , v_r ) + \nu \, ( \nabla \phi_r , \nabla v_r ) = ( \eta_t , v_r ) + \nu \, ( \nabla \eta , \nabla v_r ) . \label{s_error_semi_eq_2a} \end{eqnarray} The analysis proceeds by using \eqref{s_error_semi_eq_2a} to show that \begin{eqnarray} \| \phi_r \| \leq C \, \| \eta \| . \label{s_error_semi_eq_3} \end{eqnarray} Using the triangle inequality, one then gets \begin{eqnarray} \| e \| \leq \| \eta \| + \| \phi_r \| \leq (1 + C) \, \| \eta \| . \label{s_error_semi_eq_4} \end{eqnarray} Since $w_r$ was chosen arbitrarily and since the LHS does not depend on $w_r$, we can take the infimum over $w_r$ in \eqref{s_error_semi_eq_4} and get the following error estimate: \begin{eqnarray} \| e \| \leq (1 + C) \, \inf_{w_r \in X^r} \| u - w_r \| . \label{s_error_semi_eq_5} \end{eqnarray} At this stage, one invokes the approximability property of the space $X^r$ in \eqref{s_error_semi_eq_5} and concludes that the error estimate~\eqref{s_error_semi_eq_5} is {\it optimal with respect to $r$}. That is, the error in~\eqref{s_error_semi_eq_5} is, up to a constant, the interpolation error in $X^r$. There are a lot of variations of the error analysis sketched above, but most of the existing error analyses follow the above path. The main variations are the following: (i) the choice of the arbitrary function $w_r \in X^r$; (ii) the norm $\| \cdot \|$ used above; and (iii) the choice of the test function used in the error equation to get \eqref{s_error_semi_eq_3} \cite{thomee2006galerkin}. In the remainder of this section, we investigate whether error estimates that are {\it optimal with respect to $r$} can be obtained with or without including the DQs in the set of snapshots. To this end, in Section~\ref{sec:case_I} we consider the case in which the DQs are included in the set of snapshots (i.e., $V = V^{DQ}$). Then, in Section~\ref{sec:case_II} we consider the case in which the DQs are not included in the set of snapshots (i.e., $V = V^{no\_DQ}$). \subsection{Case I ($V = V^{DQ}$)} \label{sec:case_I} The standard approach used to prove error estimates in this case is to use the Ritz projection~\cite{KV01,KV02,iliescu2012variational,iliescu2013variational}. We note that this is the standard approach used in the finite element context~\cite{GR86,layton2008introduction}. As pointed out in \cite{thomee2006galerkin}, the Ritz projection was first used by Wheeler in \cite{wheeler1973priori} to obtain {\it optimal} error estimates for the finite element discretization of parabolic problems. We start by choosing $w_r := R_r(u)$ in~\eqref{s_error_semi_eq_2}, where $R_r(u)$ is the {\it Ritz projection} of $u$, given by: \begin{eqnarray} \bigl( \nabla ( u - R_r(u) ) , \nabla v_r \bigr) = 0 \qquad \forall \, v_r \in X^r . \label{s_error_semi_eq_6} \end{eqnarray} To emphasize that we are using the Ritz projection, in the remainder of Section~\ref{sec:case_I} we will use the notation $\eta = u - R_r(u) = \eta^{Ritz}$. Using \eqref{s_error_semi_eq_6}, \eqref{s_error_semi_eq_2a} becomes \begin{eqnarray} ( \phi_{r,t} , v_r ) + \nu \, ( \nabla \phi_r , \nabla v_r ) = ( \eta^{Ritz}_t , v_r ) + \nu \, \cancelto{0}{( \nabla \eta^{Ritz} , \nabla v_r )} . \label{s_error_semi_eq_7} \end{eqnarray} It is the cancelation of the last term on the RHS of \eqref{s_error_semi_eq_7} that yields {\it optimal} error estimates. We let $v_r := \phi_r$ in \eqref{s_error_semi_eq_7}, and then we apply the Cauchy-Schwarz inequality to the remaining term on the RHS: \begin{eqnarray} \frac{1}{2} \, \frac{d}{dt} \| \phi_r \|^2 + \nu \, \| \nabla \phi_r \|^2 \leq \| \eta^{Ritz}_t \| \, \| \phi_r \| . \label{s_error_semi_eq_8} \end{eqnarray} We rewrite the first term on the LHS of \eqref{s_error_semi_eq_8} as \begin{eqnarray} \frac{1}{2} \, \frac{d}{dt} \| \phi_r \|^2 = \| \phi_r \| \, \frac{d}{dt} \| \phi_r \| . \label{s_error_semi_eq_9} \end{eqnarray} We apply the Poincare-Friedrichs inequality to the second term on the LHS of \eqref{s_error_semi_eq_8}: \begin{eqnarray} \nu \, \| \nabla \phi_r \|^2 \geq C \, \nu \, \| \phi_r \|^2 . \label{s_error_semi_eq_10} \end{eqnarray} We note that the Poincare-Friedrichs inequality $C \| v \|^2 \leq \| \nabla v \|^2$ holds for every function $v$ in the {\it continuous} space $H_0^1(\Omega)$, and, in particular, for $\phi_r \in X^r \subset X = H_0^1(\Omega)$ (see equation (3) in \cite{KV01}). Thus, the constant $C$ in \eqref{s_error_semi_eq_10} does not depend on $r$. Using \eqref{s_error_semi_eq_9} and \eqref{s_error_semi_eq_10} in \eqref{s_error_semi_eq_8}, we get \begin{eqnarray} \frac{1}{2} \, \frac{d}{dt} \| \phi_r \| + C \, \nu \, \| \phi_r \| \leq \| \eta^{Ritz}_t \| . \label{s_error_semi_eq_11} \end{eqnarray} Using Gronwall's lemma in \eqref{s_error_semi_eq_11}, we get for $0 < t \leq T$ \begin{eqnarray} \| \phi_r (t) \| \leq e^{-2 \, C \, \nu \, t} \, \| \phi_r(0) \| + 2 \, \int_0^t e^{-2 \,C \, \nu \, (t-s)} \, \| \eta^{Ritz}_t(s) \| \, ds . \label{s_error_semi_eq_12} \end{eqnarray} Using~\eqref{s_error_semi_eq_4}, the first term on the RHS of \eqref{s_error_semi_eq_12} can be estimated as follows: \begin{eqnarray} \| \phi_r(0) \| \leq \| e(0) \| + \| \eta^{Ritz}(0) \| . \label{s_error_semi_eq_13} \end{eqnarray} Thus, \eqref{s_error_semi_eq_12} becomes \begin{eqnarray} \| \phi_r (t) \| \leq e^{-2 \, C \, \nu \, t} \, \biggl( \| e(0) \| + \| \eta^{Ritz}(0) \| \biggr) + 2 \, \int_0^t e^{-2 \, C \, \nu \, (t-s)} \, \| \eta^{Ritz}_t(s) \| \, ds . \label{s_error_semi_eq_14} \end{eqnarray} Applying the triangle inequality, just as in \eqref{s_error_semi_eq_4}, we get \begin{eqnarray} \| e(t) \| \leq \| \eta^{Ritz}(t) \| + e^{-2 \, C \, \nu \, t} \, \biggl( \| e(0) \| + \| \eta^{Ritz}(0) \| \biggr) + 2 \, \int_0^t e^{-2 \, C \, \nu \, (t-s)} \, \| \eta^{Ritz}_t(s) \| \, ds . \nonumber \\ \label{s_error_semi_eq_15} \end{eqnarray} Estimate~\eqref{s_error_semi_eq_15} shows that, as long as the Ritz projection~\eqref{s_error_semi_eq_6} yields estimates for $\| \eta^{Ritz}(t) \|$ (including for $t=0$) and $\| \eta^{Ritz}_t(t) \|$ that are {\it optimal} with respect to $r$, the estimates for the POD error $e$ are also optimal with respect to $r$. \begin{remark}[POD Ritz Projection] \label{remark_ritz} In the finite element context, both $\| \eta^{Ritz} \|$ and $\| \eta^{Ritz}_t \|$ are optimal with respect to the mesh size $h$. In the POD context, however, this is not that clear. To the best of the authors' knowledge, the state-of-the-art regarding the Ritz projection in a POD context is given in the pioneering paper of Kunisch and Volkwein~\cite{KV01}. Since the Ritz projection plays such an important role in this paper, we summarize below the results in~\cite{KV01}. The main result in \cite{KV01} regarding the Ritz projection is Lemma 3 (see also (10) and (11) in \cite{KV01}), which, in our notation, states the following: \begin{eqnarray} \| \nabla \eta^{Ritz} \|^ 2 \leq \| S_d \|_2 \, \sum \limits_{j=r+1}^{d} \lambda_j \, . \label{eqn:kv01_lemma3_2} \end{eqnarray} For clarity of presentation, we have not included in \eqref{eqn:kv01_lemma3_2} the constants that do not depend on $r$. Thus, the following relationship between the Ritz projection error and the POD interpolation error holds: \begin{eqnarray} \| \nabla \eta^{Ritz} \| \leq C \, \sqrt{\| S_d \|_2} \, \| \eta^{interp} \| \, . \label{eqn:relationship_ritz_interpolation_nabla_eta} \end{eqnarray} The scaling in \eqref{eqn:relationship_ritz_interpolation_nabla_eta} suggests that the Ritz projection yields optimal error estimates with respect to $r$ in the $H^1$-seminorm (see \eqref{eqn:inverse_estimate_interpolation_error}). We emphasize that Lemma 3 in \cite{KV01} does not include any bounds for the $L^2$-norm of $\eta^{Ritz}$, i.e., for $\| \eta^{Ritz} \|$. This is in clear contrast with the finite element context, in which $\| \eta^{Ritz} \|$ is estimated by the usual duality argument (the Aubin-Nitsche ``trick," see, e.g., \cite{thomee2006galerkin}). Using a duality argument, however, is challenging in the POD context, since any auxiliary dual problem would not necessarily inherit the POD approximation property~\eqref{pod_error_formula_nodq}--\eqref{pod_error_formula_dq_pointwise}. To the best of the authors' knowledge, such a duality argument has never been used in a POD context. We emphasize that not being able to use a duality argument in the Ritz projection to get error estimates that are optimal with respect to $r$ has significant consequences in the error analysis. Indeed, in the proof of Theorem 7 in \cite{KV01} (the error estimate for the backward Euler time discretization), to estimate the $\| \eta^{Ritz} \|$ error component in equations (27a) and (27b), the authors use the $\| \nabla \eta^{Ritz} \|$ estimate given in Lemma 3 and the Poincare-Friedrichs inequality given in equation (3). Since the Poincare-Friedrichs constant does not depend on $r$, we conclude that $\| \eta^{Ritz}\|$ and $\| \nabla \eta^{Ritz} \|$ have the same order. This, in turn, suggests that $\| \eta^{Ritz} \|$ is suboptimal with respect to $r$. We note that the same approach (i.e., Lemma 4 and the Poincare-Friedrichs inequality) is used in~\cite{KV01} to estimate the DQ approximation of $\| \eta^{Ritz}_t \|$ (see the two inequalities above equation (29a)). Thus, the analysis in \cite{KV01} suggests that the estimates for $\| \eta^{Ritz} \|$ and $\| \eta^{Ritz}_t \|$ are suboptimal with respect to $r$. The numerical tests in Section~\ref{sec:numerical} of this report, however, suggest that these estimates are actually optimal, just as in the finite element case. In the analysis that follows, we will use the insight from the numerical results in Section~\ref{sec:numerical} and make the following assumption: \begin{assumption} We assume that the POD Ritz projection error $\eta^{Ritz}$ satisfies optimal error estimates with respect to $r$ in the $L^2$-norm:\begin{eqnarray} \| \eta^{Ritz} \| &\leq& C \, \| \eta^{interp} \| \, , \label{eqn:relationship_ritz_interpolation_eta} \\ \| \eta^{Ritz}_t \| &\leq& C \, \| \eta^{interp} \| \, . \label{eqn:relationship_ritz_interpolation_eta_t} \end{eqnarray} \label{assumption_ritz} \end{assumption} \end{remark} \subsection{Case II ($V = V^{no\_DQ}$)} \label{sec:case_II} This approach was used in \cite{chapelle2012galerkin,singler2012new}. The motivation for this approach is the following: In Case I ($V = V^{DQ}$), the first term on the RHS of \eqref{s_error_semi_eq_7}, $( \eta_t , v_r)$, yields a term $\| \eta_t \|$ that stays in all the subsequent inequalities, including the final error estimate \eqref{s_error_semi_eq_15}. Chapelle et al. proposed in \cite{chapelle2012galerkin} a different approach that eliminated the $( \eta_t , v_r)$ term in \eqref{s_error_semi_eq_7}. Their approach was straightforward: Instead of using the Ritz projection (as in Case I), they used the $L^2$ projection. That is, they chose $w_r := P_r(u)$, where $P_r(u)$ is the {\it $L^2$ projection} of $u$, given by \begin{eqnarray} \bigl( u - P_r(u) , v_r \bigr) = 0 \qquad \forall \, v_r \in X^r . \label{s_error_semi_eq_16} \end{eqnarray} To emphasize that we are using the $L^2$ projection, in the remainder of Section~\ref{sec:case_II} we will use the notation $\eta = u - P_r(u) = \eta^{L^2}$. We note that the POD $L^2$ projection error $\eta^{L^2}$ is exactly the POD interpolation error defined in~\eqref{eqn:definition_interpolation_error}: \begin{eqnarray} \eta^{L^2} = \eta^{interp} \, . \label{s_error_semi_eq_16.5} \end{eqnarray} Next, we show how the error analysis in Case I changes with $w_r = P_r(u)$ as in \cite{chapelle2012galerkin} (see also \cite{singler2012new}). Using \eqref{s_error_semi_eq_16}, \eqref{s_error_semi_eq_2a} becomes \begin{eqnarray} ( \phi_{r,t} , v_r ) + \nu \, ( \nabla \phi_r , \nabla v_r ) = \cancelto{0}{( \eta^{L^2}_t , v_r )} + \nu \, ( \nabla \eta^{L^2} , \nabla v_r ) . \label{s_error_semi_eq_17} \end{eqnarray} We emphasize that it is the cancelation of the first term on the RHS of \eqref{s_error_semi_eq_17} that yields error estimates that do not require the DQs. We let $v_r := \phi_r$ in \eqref{s_error_semi_eq_17} and we apply Cauchy-Schwarz inequality to the remaining term on the RHS: \begin{eqnarray} \frac{1}{2} \, \frac{d}{dt} \| \phi_r \|^2 + \nu \, \| \nabla \phi_r \|^2 \leq \nu \, \| \nabla \eta^{L^2} \| \, \| \nabla \phi_r \| . \label{s_error_semi_eq_18} \end{eqnarray} The error analysis can then proceed in several directions. \subsubsection{Approach II.A} \label{sec:approach_II.A} One approach is to use Young's inequality in \eqref{s_error_semi_eq_18} to get \begin{eqnarray} \frac{1}{2} \, \frac{d}{dt} \| \phi_r \|^2 + \nu \, \| \nabla \phi_r \|^2 \leq \frac{\nu}{2} \, \| \nabla \eta^{L^2} \|^2 + \frac{\nu}{2} \, \| \nabla \phi_r \|^2 , \label{s_error_semi_eq_19} \end{eqnarray} which implies \begin{eqnarray} \frac{1}{2} \, \frac{d}{dt} \| \phi_r \|^2 + \frac{\nu}{2} \, \| \nabla \phi_r \|^2 \leq \frac{\nu}{2} \, \| \nabla \eta^{L^2} \|^2 . \label{s_error_semi_eq_20} \end{eqnarray} Noticing that the second term on the LHS of \eqref{s_error_semi_eq_20} is positive, we get \begin{eqnarray} \frac{d}{dt} \| \phi_r \|^2 \leq \nu \, \| \nabla \eta^{L^2} \|^2 . \label{s_error_semi_eq_21} \end{eqnarray} Using~\eqref{s_error_semi_eq_16.5} and~\eqref{eqn:suboptimal_error_estimate}, we conclude that Approach II.A will yield error estimates that are suboptimal with respect to $r$. We also note in passing that estimate~\eqref{s_error_semi_eq_21} suggests that Approach II.A will yield error estimates that are suboptimal with respect to $h$ as well. \subsubsection{Approach II.B} \label{sec:approach_II.B} The other way of continuing from \eqref{s_error_semi_eq_18} is to apply the POD inverse estimate~\eqref{lemma_inverse_pod_1}: \begin{eqnarray} \| \nabla \phi_r \| \leq C_{inv}(r) \, \| \phi_r \| , \label{s_error_semi_eq_22} \end{eqnarray} where $C_{inv}(r) := \sqrt{\| S_r \|_2 }$. Using~\eqref{s_error_semi_eq_22} in \eqref{s_error_semi_eq_18} yields \begin{eqnarray} \frac{1}{2} \, \frac{d}{dt} \| \phi_r \|^2 + \nu \, \| \nabla \phi_r \|^2 \leq C_{inv}(r) \, \nu \, \| \nabla \eta^{L^2} \| \, \| \phi_r \| . \label{s_error_semi_eq_23} \end{eqnarray} Dropping $\nu \, \| \nabla \phi_r \|^2$ in \eqref{s_error_semi_eq_23} and simplifying the resulting inequality by $\| \phi_r \|$ (as we did in \eqref{s_error_semi_eq_12}), we get \begin{eqnarray} \frac{1}{2} \, \frac{d}{dt} \| \phi_r \| \leq C_{inv}(r) \, \nu \, \| \nabla \eta^{L^2} \| . \label{s_error_semi_eq_24} \end{eqnarray} Comparing estimate \eqref{s_error_semi_eq_24} with estimate \eqref{s_error_semi_eq_21} in Approach II.A, we note that both estimates have $\| \nabla \eta \|$ on the RHS. In addition, estimate \eqref{s_error_semi_eq_24} has $C_{inv}(r)$ on the RHS, which increases the suboptimality with respect to $r$ (see Remark~\ref{remark_pod_inverse_estimate_scalings}). Thus, estimate~\eqref{s_error_semi_eq_24} suggests that Approach II.B yields estimates that are suboptimal with respect to $r$ (and $h$), just as Approach II.A. \subsubsection{Approach II.C} \label{sec:approach_II.C} Since both Approach II.A and Approach II.B yield error estimates that are suboptimal with respect to $r$ in the $L^2$-norm, one can try instead to prove optimal error estimates in the $H^1$-seminorm. To this end, we use the approach in \cite{thomee2006galerkin} and, instead of choosing $v_r := \phi_r$ in \eqref{s_error_semi_eq_17}, we choose $v_r := \phi_{r,t}$: \begin{eqnarray} \| \phi_{r,t} \|^2 + \frac{\nu}{2} \, \frac{d}{dt} \| \nabla \phi_r \|^2 \leq \nu \,\| \nabla \eta^{L^2} \| \, \| \nabla \phi_{r,t} \| . \label{s_error_semi_eq_25} \end{eqnarray} Applying Young's inequality and the POD inverse estimate~\eqref{s_error_semi_eq_22} in \eqref{s_error_semi_eq_25}, we get \begin{eqnarray} \| \phi_{r,t} \|^2 + \frac{\nu}{2} \, \frac{d}{dt} \| \nabla \phi_r \|^2 &\leq& \nu \, \| \nabla \eta^{L^2} \| \, \| \nabla \phi_{r,t} \| \nonumber \\ &\leq& \frac{\nu^2}{2} \, C_{inv}(r)^2 \, \| \nabla \eta^{L^2} \|^2 + \frac{1}{2 \, C_{inv}(r)^2} \, \| \nabla \phi_{r,t} \|^2 \nonumber \\ &\stackrel{\eqref{s_error_semi_eq_22}}{\leq}& \frac{\nu^2}{2} \, C_{inv}(r)^2 \, \| \nabla \eta^{L^2} \|^2 + \frac{1}{2} \, \| \phi_{r,t} \|^2 , \label{s_error_semi_eq_26} \end{eqnarray} which implies \begin{eqnarray} \frac{d}{dt} \| \nabla \phi_r \|^2 \leq \nu \, C_{inv}(r)^2 \, \| \nabla \eta^{L^2} \|^2 . \label{s_error_semi_eq_27} \end{eqnarray} In contrast with estimate~\eqref{s_error_semi_eq_21} in Approach II.A and estimate~\eqref{s_error_semi_eq_24} in Approach II.B, estimate~\eqref{s_error_semi_eq_27} seems to yield error estimates that are optimal with respect to $r$. As in estimates~\eqref{s_error_semi_eq_21} and \eqref{s_error_semi_eq_27}, estimate~\eqref{s_error_semi_eq_27} contains the term $\| \nabla \eta^{L^2} \|$ on the RHS. We note, however, that this term does not cause any problems, since now we are considering the $H^1$-seminorm of the error. The factor $C_{inv}(r)^2$ in~\eqref{s_error_semi_eq_27}, however, increases the suboptimality with respect to $r$ (see Remark~\ref{remark_pod_inverse_estimate_scalings}). Thus, estimate~\eqref{s_error_semi_eq_27} suggests that Approach II.C yields estimates that are suboptimal with respect to $r$, just as Approaches II.A and II.B. Since for {\bf Case I ($V = V^{DQ}$)} in Section~\ref{sec:case_I}, we did not prove error estimates in the $H^1$-norm, for a fair comparison with Approach II.C, we prove these error estimates below. To this end, we let $v_r := \phi_{r,t}$ in \eqref{s_error_semi_eq_7}: \begin{eqnarray} ( \phi_{r,t} , \phi_{r,t} ) + \nu \, ( \nabla \phi_r , \nabla \phi_{r,t} ) = ( \eta^{Ritz}_t , \phi_{r,t} ) . \label{s_error_semi_eq_28} \end{eqnarray} Applying Young's inequality on the RHS of \eqref{s_error_semi_eq_28}, we get \begin{eqnarray} \| \phi_{r,t} \|^2 + \frac{\nu}{2} \, \frac{d}{dt} \| \nabla \phi_r \|^2 &\leq& \| \eta^{Ritz}_t \| \, \| \phi_{r,t} \| \nonumber \\ &\leq& \frac{1}{2} \, \| \eta^{Ritz}_t \|^2 + \frac{1}{2} \, \| \phi_{r,t} \|^2 , \label{s_error_semi_eq_29} \end{eqnarray} which implies \begin{eqnarray} \frac{d}{dt} \| \nabla \phi_r \|^2 \leq \frac{1}{\nu} \, \| \eta^{Ritz}_t \|^2 . \label{s_error_semi_eq_30} \end{eqnarray} Comparing \eqref{s_error_semi_eq_30} with \eqref{s_error_semi_eq_27} in Approach II.C, we note that the latter does not contain the factor $C_{inv}(r)^2$. Thus, {\bf Case I ($V = V^{DQ}$)} in Section~\ref{sec:case_I} yields optimal error estimates with respect to $r$, as opposed to Approach II.C. We also note that, at first glance, estimate~\eqref{s_error_semi_eq_30} suggests that one can get superconvergence in the $H^1$-seminorm. As mentioned in~\cite{thomee2006galerkin}, however, when applying the triangle inequality one obtains the expected convergence rate: \begin{eqnarray} \| \nabla e \| \leq \| \nabla \eta^{Ritz} \| + \| \nabla \phi_r \| . \label{s_error_semi_eq_31} \end{eqnarray} \subsubsection{Approach II.D} \label{sec:approach_II.D} Approaches II.A, II.B, and II.C suggest that the pointwise in time (i.e., in the $C^0(0,T; L^2(\Omega))$-norm and the $C^0(0,T; H^1(\Omega))$-norm) error estimates are suboptimal with respect to $r$. Thus, we derive error estimates in the solution norm (i.e., in the $L^2(0,T; H^1(\Omega))$-norm). Integrating~\eqref{s_error_semi_eq_20} from $0$ to $T$, we get \begin{eqnarray} \| \phi_r(T) \|^2 + \frac{\nu}{2} \, \int_{0}^{T} \| \nabla \phi_r(s) \|^2 \, ds \leq \| \phi_r(0) \|^2 + \frac{\nu}{2} \, \int_{0}^{T} \| \nabla \eta^{L^2}(s) \|^2 \, ds . \label{s_error_semi_eq_23b} \end{eqnarray} Using \eqref{s_error_semi_eq_16.5}, \eqref{eqn:optimality_eta}, and \eqref{eqn:optimality_nabla_eta}, we conclude that estimate in \eqref{s_error_semi_eq_23b} is optimal with respect to $r$. We note that Proposition 3.3 in \cite{chapelle2012galerkin} yields a similar estimate. \bigskip Case II (i.e., $V = V^{no\_DQ}$) yields the following general conclusions when the $L^2$ projection is used instead of the standard Ritz projection: If the error is computed pointwise in time (i.e., in the $C^0(0,T; L^2(\Omega))$-norm and the $C^0(0,T; H^1(\Omega))$-norm), then the error estimates are suboptimal with respect to $r$. This is the main message of Approaches II.A, II.B, and II.C. If, however, the error is computed in the solution norm (i.e., in the $L^2(0,T; H^1(\Omega))$-norm), then the error estimates are optimal with respect to $r$. This is the main message of Approach II.D. \bigskip \section{Introduction} \label{s_introduction} This paper addresses the following question: {\it ``Should the time difference quotients (DQs) of the snapshots be used in the generation of the Proper Orthogonal Decomposition (POD) basis functions?"} We emphasize that this is an important question. There are two schools of thought: one uses the DQs (see, e.g. \cite{KV01,KV02,chaturantabut2010nonlinear,chaturantabut2012state}), the other does not (see, e.g. \cite{LCNY08,luo2009finite,chapelle2012galerkin,singler2012new}). To our knowledge, the first instance in which the snapshot DQs were incorporated in the generation of the POD basis functions was the pioneering paper of Kunisch and Volkwein \cite{KV01}. In that report, the authors considered two types of errors for a general parabolic equation: the time discretization errors and the POD truncation errors. They argued that one needs to include the temporal difference quotients in the set of snapshots; otherwise, the error will be suboptimal, containing an extra $\frac{1}{\Delta t^2}$ factor (see Remark 1 in \cite{KV01}). Thus, the motivation for using the temporal difference quotients was purely theoretical. In numerical investigations, however, the authors reported contradictory findings: in \cite{KV01}, the use of the DQs did not improve the quality of the reduced-order model; in \cite{homberg2003control}, however, it did. Kunisch and Volkwein used again the snapshot DQs when they considered the Navier-Stokes equations \cite{KV02}. The snapshot DQs were also used in the {\it Discrete Empirical Interpolation Method (DEIM)} of Chaturantabut and Sorensen \cite{chaturantabut2010nonlinear,chaturantabut2012state} (which is a discrete variant of the {\it Empirical Interpolation Method (EIM)} \cite{barrault2004eim}). The motivation in \cite{chaturantabut2010nonlinear,chaturantabut2012state}, however, was different from that in \cite{KV01}. Indeed, the authors considered in \cite{chaturantabut2010nonlinear,chaturantabut2012state} a general, nonlinear system of equations of the form ${\bf y}' = {\bf f}({\bf y},t)$. In the set of snapshots, they included not only the state variables ${\bf y}$, but also the {\it nonlinear} snapshots ${\bf f}({\bf y},t)$. They further noted (see page 48 in \cite{chaturantabut2012state}) that, since ${\bf f}({\bf y},t) = {\bf y}'$ and $( {\bf y}^{n+1} - {\bf y}^{n} ) / \Delta t \sim {\bf y}'$, this is similar to including the temporal difference quotients, as done in \cite{KV01,KV02}. To our knowledge, the first reports on POD analysis in which the DQs were not used were \cite{LCNY08} for the heat equation and \cite{luo2009finite} for the Navier-Stokes equations. Chapelle et al.~\cite{chapelle2012galerkin} used a different approach that did not use the DQs either. This approach employed the $L^2$ projection instead of the standard $H^1$ projection used in, e.g., \cite{KV01,KV02}. Further improvements to the approach used in~\cite{chapelle2012galerkin} (as well as that used in \cite{KV01,KV02}) were made by Singler in \cite{singler2012new}. From the above discussion, it is clear that the question whether the snapshot DQs should be included or not in the set of snapshots is important. To our knowledge, this question is still open. This report represents a first step in answering this question. All our discussion will be centered around the heat equation, although most (if not all) of it could be extended to general parabolic equations in a straightforward manner. From a theoretical point of view, the only motivation for using the snapshot DQs was given in Remark 1 in \cite{KV01}. The main point of this remark is the following: In the error analysis of the evolution equation, to approximate $u_t(t^n)$, the time derivative of the exact solution $u$ evaluated at time $t^n$, one usually uses the DQ $\overline{\partial} u(t^n) := \frac{u(t^{n+1}) - u(t^{n})}{\Delta t}$. To approximate the DQ $\overline{\partial} u(t^n)$ in the POD space, one naturally uses the POD DQ $\overline{\partial} u_r(t^n) := \frac{u_r(t^{n+1}) - u_r(t^{n})}{\Delta t}$, where $u_r$ is the POD reduced order model approximation. We assume that $u_r(t^{n+1})$ is an optimal approximation for $u(t^{n+1})$ and that $u_r(t^{n})$ is an optimal approximation for $u(t^{n})$, where the optimality is with respect to $r$ and $\Delta t$. Then, it would appear that, with respect to $\Delta t$, $\overline{\partial} u_r(t^n)$ is a {\it suboptimal} approximation for $\overline{\partial} u(t^n)$, because of the $\Delta t$ in the denominator of the two difference quotients. Although the argument above, used in Remark 1 in \cite{KV01} to motivate the inclusion of the snapshot DQs in the derivation of the POD basis, seems natural, we point out that this issue should be treated more carefully. Indeed, in the finite element approximation of parabolic equations, it is well known that the DQs $\overline{\partial} u_h(t^n) := \frac{u_h(t^{n+1}) - u_h(t^{n})}{\Delta t}$ are actually {\it optimal} (with respect to $\Delta t$) approximations of the DQs $\overline{\partial} u(t^n)$ (see, e.g., \cite{labovschii2009defect,shan2012decoupling}). Thus, since the POD and finite element approximations are similar (both use a Galerkin projection in the spatial discretization), one could question the validity of the argument used in Remark 1 in \cite{KV01}. We emphasize that we are not claiming that the above argument is not valid in a POD setting; we are merely pointing out that a rigorous numerical analysis is needed before drawing any conclusions. \begin{figure}[htp] \begin{center} \begin{minipage}[h]{.6\linewidth} \includegraphics[width=0.9\textwidth]{Heat_eL2_dt.eps} \end{minipage} \end{center} \caption{ Heat equation. Plots of the errors in the $L^2(L^2)$-norm with respect to the time step $\Delta t$ when the DQs are used (denoted by DQ) and when the DQs are not used (denoted by no-DQ). } \label{fig:test1_dt} \end{figure} Our preliminary numerical studies indicate that not using the DQs does {\it not} yield suboptimal (with respect to $\Delta t$) error estimates. For the heat equation (see Section~\ref{sec:numerical} for details regarding the numerical simulation), we monitor the rates of convergence with respect to $\Delta t$ for the POD reduced order model. We consider two cases: when the DQs are used in the generation of the POD basis (the corresponding results are denoted by DQ), and when the DQs are not used in the generation of the POD basis (the corresponding results are denoted by no-DQ). The errors (defined in Section~\ref{sec:numerical}) are listed in Table \ref{tab:test1_dt} and plotted in Figure \ref{fig:test1_dt} with associated {\it linear regressions (LR)}. Both no-DQ and DQ approaches yield an optimal approximation order $\mathcal{O}(\Delta t^2)$ in the $L^2$-norm. \begin{table}[h] \centering \tabcaption{Errors of the $no\_DQ$ and $DQ$ approaches when $\Delta t$ varies.} \label{tab:test1_dt} \begin{tabular}{c|ccc|ccc} \hline \multirow{2}{*}{$\Delta t$}&\multicolumn{3}{c|}{$no\_DQ$}&\multicolumn{3}{c}{$DQ$}\\ \cline{2-7} {} & r & $\mathcal{E}_{L^2(L^2)}$ & $\mathcal{E}_{L^2(H_1)}$ & r & $\mathcal{E}_{L^2(L^2)}$ & $\mathcal{E}_{L^2(H_1)}$ \\ \hline 2.00e-01 & 6 & 3.71e-02 & 9.26e-01 & 6 & 3.71e-02 & 9.26e-01 \\ 1.00e-01 & 11 & 1.27e-02 & 5.81e-01 & 11 & 1.27e-02 & 5.81e-01 \\ 5.00e-02 & 21 & 2.99e-03 & 1.97e-01 & 21 & 2.99e-03 & 1.97e-01 \\ 2.50e-02 & 41 & 6.53e-04 & 3.81e-02 & 41 & 6.53e-04 & 3.81e-02 \\ 1.00e-02 & 59 & 1.03e-04 & 1.15e-02 & 88 & 1.03e-04 & 1.15e-02 \\ \hline \end{tabular} \end{table} The rest of the paper is organized as follows: In Section~\ref{sec:pod}, we sketch the derivation of the POD reduced order model. In Section~\ref{sec:error}, we carefully derive the error estimates for the POD reduced order model. We focus on the rates of convergence with respect to $r$, the number of POD basis functions. In Section~\ref{sec:numerical}, we present numerical results for two test problems: the heat equation and the Burgers equation. Finally, in Section~\ref{sec:conclusions}, we draw several conclusions. \section{Numerical Results} \section{Numerical Results} \label{sec:numerical} The main goal of this section is to numerically investigate the rates of convergence with respect to $r$ of the POD-G-ROM~\eqref{eqn:pod_g_rom} in the two cases considered in Section~\ref{sec:error}: {\bf Case I} ($V = V^{DQ}$) and {\bf Case II} ($V = V^{no\_DQ}$). Although the error analysis in Section~\ref{sec:error} has been centered around the (linear) heat equation, in this section we consider both the heat equation (Section~\ref{sec:heat}) and the nonlinear Burgers equation (Section~\ref{sec:burgers}). To measure the errors in the two cases (i.e., $V = V^{DQ}$ and $V = V^{no\_DQ}$), the same norms as those used in Section~\ref{sec:error} are used in this section. Denoting the error at time $t_j$ by $e_j := u^r_h(\cdot, t_j) - u(\cdot, t_j)$, the following norms are considered: the error in the $C^0(0,T; L^2(\Omega))$-norm, approximated by $\mathcal{E}_{C^0(L^2)} = \max\limits_{0\leq j \leq N} \|e_j\|_{{L^2}(\Omega)}$; the error in the $C^0(0,T; H^1(\Omega))$-norm, approximated by $\mathcal{E}_{C^0(H^1)} = \max\limits_{0\leq j \leq N} \|e_j\|_{{H^1}(\Omega)}$; and the error in the $L^2(0,T; H^1(\Omega))$-norm, approximated by $\mathcal{E}_{L^2(H^1)}=\sqrt{\frac{1}{N+1} \sum\limits_{0\leq j \leq N} \| e_j \|_{{H^1}(\Omega)}^2}$. For clarity, we also use the following notation: $\Lambda_r= \sqrt{\sum\limits_{j=r+1}^d \lambda_j}$. As mentioned at the beginning of Section~\ref{sec:error}, the POD-G-ROM~\eqref{eqn:pod_g_rom} error estimates are optimal if the following statements hold: (i) The $L^2$-norm of the error scales as $L^2$-norm of the POD interpolation error. Using \eqref{eqn:optimality_eta} and \eqref{pod_error_formula_nodq}--\eqref{pod_error_formula_dq_pointwise}, this statement is equivalent to \begin{eqnarray} \mathcal{E}_{C^0(L^2)} = \mathcal{O}\left(\sqrt{\sum\limits_{j=r+1}^d \lambda_j} \ \right) \, , \label{eqn:numerical_1} \end{eqnarray} (ii) The $H^1$-norm of the error scales as $H^1$-norm of the POD interpolation error. Using \eqref{eqn:optimality_nabla_eta}, \eqref{eqn:inverse_estimate_interpolation_error}, and \eqref{pod_error_formula_nodq}--\eqref{pod_error_formula_dq_pointwise}, this statement is equivalent to \begin{eqnarray} \mathcal{E}_{C^0(H^1)} \sim \mathcal{E}_{L^2(H^1)} = \mathcal{O}\left(\sqrt{ \| S_d \|_2 \, \sum\limits_{j=r+1}^d \lambda_j}\right) \, , \label{eqn:numerical_2} \end{eqnarray} Based on the error analysis in Section~\ref{sec:error}, we expect the following convergence rates with respect to $r$: \begin{table}[h] \centering \tabcaption{Theoretical convergence rates for the $no\_DQ$ and the $DQ$ cases.} \label{tab:numerical_1} \begin{tabular}{|c|c|c|} \hline & & \\[-0.2cm] & $no\_DQ$ & $DQ$ \\[0.2cm] \hline & & \\[-0.2cm] $\mathcal{E}_{C^0(L^2)}$ & suboptimal & optimal \\ & \eqref{s_error_semi_eq_21}; \eqref{s_error_semi_eq_24} & \eqref{s_error_semi_eq_15} \\[0.2cm] \hline & & \\[-0.2cm] $\mathcal{E}_{C^0(H^1)}$ & suboptimal & optimal \\ & \eqref{s_error_semi_eq_27} & \eqref{s_error_semi_eq_31} \\[0.2cm] \hline & & \\[-0.2cm] $\mathcal{E}_{L^2(H^1)}$ & optimal & optimal \\ & \eqref{s_error_semi_eq_23b} & \\[0.2cm] \hline \end{tabular} \end{table} \subsection{Heat Equation} \label{sec:heat} We consider the one-dimensional heat equation~\eqref{eqn:heat_weak} with a known exact solution that represents the propagation in time of a steep front: \begin{eqnarray} u(x,t) = \sin(\pi x)\left[\frac{1}{\pi}\arctan\left(\frac{c}{25}-c\left(x-\frac{t}{2}\right)^2\right) + \frac{1}{2}\right], \label{eqn:exact_solution_heat} \end{eqnarray} where $x\in [0, 1]$ and $t\in [0, 1]$. The constant $c$ in~\eqref{eqn:exact_solution_heat} controls the steepness of the front. In all the numerical tests in this section, we used the value $c = 100$. The value of the diffusion coefficient used in the heat equation~\eqref{eqn:heat_weak} is $\nu = 10^{-2}$. Piecewise linear finite elements are used to generate snapshots for the POD-G-ROM~\eqref{eqn:pod_g_rom}. A mesh size $h=1/1024$ and the Crank-Nicolson scheme with a time step $\Delta t = 10^{-3}$ are employed for the spatial and temporal discretizations. The time evolution of the finite element solution is shown in Figure \ref{fig:test1dns}. In total, 1001 snapshots are collected and used for generating POD basis functions in the $L^2$ space. The same numerical solver as that used in the finite element approximation is utilized in the POD-G-ROM. \begin{figure}[h] \centering \begin{minipage}[h]{.6\linewidth} \includegraphics[width=1\textwidth]{Heat_dns.eps} \end{minipage} \caption{ Heat equation. Fine resolution finite element solution used to generate the snapshots. } \label{fig:test1dns} \end{figure} By varying $r$, the number of basis functions used in the POD-G-ROM, we check the rates of convergence with respect to $r$ for the $no\_DQ$ and the $DQ$ cases. The errors are listed in Table \ref{tab:test1_DQ0} (in the $no\_DQ$ case) and in Table \ref{tab:test1_DQ1} (in the $DQ$ case). To visualize the rates of convergence with respect to $r$, these errors with their linear regression plots are drawn in Figure \ref{fig:test1_eC0L2_eC0H1_eL2H1}. The convergence rate of the error in the $C^0(L^2)$-norm, $\mathcal{E}_{C^0(L^2)}$, is superoptimal in the $DQ$ case and suboptimal in the $no\_DQ$ case. This supports the theoretical rates of convergence in Table~\ref{tab:numerical_1}, although the suboptimality in the $no\_DQ$ case is mild. The convergence rate of the error in the $C^0(H^1)$-norm, $\mathcal{E}_{C^0(H^1)}$, is slightly superoptimal in the $DQ$ case and strongly suboptimal in the $no\_DQ$ case. This again supports the theoretical rates of convergence in Table~\ref{tab:numerical_1}. The convergence rate of the error in the $L^2(H^1)$-norm, $\mathcal{E}_{L^2(H^1)}$, is optimal in the $DQ$ case and strongly suboptimal in the $no\_DQ$ case. This supports the theoretical rates of convergence in Table~\ref{tab:numerical_1} for the $DQ$ case, but not for the $no\_DQ$ case. Overall, the numerical results support the theoretical rates of convergence proved in Section~\ref{sec:error} and summarized in Table~\ref{tab:numerical_1}. We also emphasize that {\it the convergence rates in the $DQ$ case in all three norms are much higher than (and almost twice as high as) the corresponding rates of convergence in the $no\_DQ$ case}. \begin{table}[h] \centering \tabcaption{Heat equation. Errors in the $no\_DQ$ case.} \label{tab:test1_DQ0} \begin{tabular}{|c|c|c|c|c|c} \hline r & $\Lambda_r$ & $\mathcal{E}_{C^0(L^2)}$ & $\mathcal{E}_{C^0(H^1)}$ & $\mathcal{E}_{L^2(H^1)}$ \\ \hline 3 & 5.72e-02 & 9.46e-02 & 2.30e+00 & 1.59e+00 \\ 5 & 2.71e-02 & 4.70e-02 & 1.58e+00 & 1.14e+00 \\ 7 & 1.58e-02 & 3.69e-02 & 1.38e+00 & 8.22e-01 \\ 10 & 7.34e-03 & 1.57e-02 & 8.54e-01 & 5.31e-01 \\ 13 & 3.84e-03 & 7.78e-03 & 5.84e-01 & 3.50e-01 \\ \hline \end{tabular} \end{table} \begin{table}[h] \centering \tabcaption{Heat equation. Errors in the $DQ$ case.} \label{tab:test1_DQ1} \begin{tabular}{|c|c|c|c|c|c} \hline r & $\Lambda_r$ & $\mathcal{E}_{C^0(L^2)}$ & $\mathcal{E}_{C^0(H^1)}$ & $\mathcal{E}_{L^2(H^1)}$ \\ \hline 19 & 5.49e-02 & 7.15e-03 & 4.96e-01 & 3.19e-01 \\ 23 & 2.95e-02 & 2.03e-03 & 1.98e-01 & 1.19e-01 \\ 28 & 1.41e-02 & 6.52e-04 & 7.97e-02 & 4.91e-02 \\ 33 & 6.75e-03 & 2.41e-04 & 3.68e-02 & 2.80e-02 \\ 37 & 3.76e-03 & 8.60e-05 & 2.67e-02 & 2.29e-02 \\ \hline \end{tabular} \end{table} \begin{figure}[h] \begin{minipage}[h]{.45\linewidth} \includegraphics[width=1.1\textwidth]{Heat_eC0L2_wrt_r.eps} \end{minipage} \hspace{.3cm} \begin{minipage}[h]{.45\linewidth} \includegraphics[width=1.1\textwidth]{Heat_eC0H1_wrt_r_withSr.eps} \end{minipage} \hspace{.3cm} \begin{center} \begin{minipage}[h]{.45\linewidth} \includegraphics[width=1.1\textwidth]{Heat_eH1_wrt_r_withSr.eps} \end{minipage} \end{center} \caption{ Heat equation. Plots of errors in $C^0(L^2)$-norm (top, left), $C^0(H^1)$-norm (top, right), and $L^2(H^1)$-norm (bottom). } \label{fig:test1_eC0L2_eC0H1_eL2H1} \end{figure} \subsection{Burgers Equation} \label{sec:burgers} In this section, we consider the one-dimensional Burgers equation. As mentioned at the beginning of Section~\ref{sec:numerical}, the error estimates proved in Section~\ref{sec:error} are valid for the (linear) heat equation, but not necessarily valid for the nonlinear Burgers equation. Nevertheless, to gain some insight into the range of validity of the theoretical development in Section~\ref{sec:error}, we investigate the rates of convergence with respect to $r$ in the $no\_DQ$ and the $DQ$ cases for the nonlinear Burgers equation: \begin{eqnarray} \label {burgers} \left\{ \begin{array}{ll} u_t - \nu \, u_{xx} + u \, u_x = f & \qquad \text{ in } \Omega \times (0, T] \, , \\ u(x,0) = u_0(x) & \qquad \text{ in } \Omega \, , \\ u(x, t) = g(x, t) & \qquad \text{ on } \partial \Omega \times (0, T] \, . \end{array} \right. \end{eqnarray} The initial condition is \begin{eqnarray} u_0(x) = \left\{ \begin{array}{cc} 1 & \quad \text{ if } x \in \left( 0, \frac{1}{2} \right] \\ \\ 0 & \qquad \text{ if } x \in \left( \frac{1}{2} , 1 \right) \, , \end{array} \right. \label{ic_1} \end{eqnarray} which is similar to that used in \cite{KV01}. The diffusion parameter is $\nu=10^{-2}$, the forcing term is $f = 0$, $\Omega = [0, 1]$, and $T= 1$. The boundary conditions are homogeneous Dirichlet, that is, $u(0,t) = u(1,t) = 0$ for all $t\in [0, 1]$. To generate snapshots, we use piecewise linear finite elements with mesh size $h=1/1024$ and the backward Euler method with a time step $\Delta t = 10^{-4}$, and save data at each time instance. The time evolution of the finite element solution is shown in Figure \ref{fig:test2dns}. All snapshots are used for the POD basis generation and the same numerical solver is used in the POD-G-ROM. \begin{figure}[h] \centering \begin{minipage}[h]{.6\linewidth} \includegraphics[width=1\textwidth]{Burg_dns.eps} \end{minipage} \caption{ Burgers equation. Fine resolution finite element solution used to generate the snapshots. } \label{fig:test2dns} \end{figure} By varying $r$, we check the rates of convergence with respect to $r$ for the $no\_DQ$ and the $DQ$ cases. Since the exact solution of the Burgers equation~\eqref{burgers} with the initial condition~\eqref{ic_1} is not known, we consider the errors between the POD-G-ROM results and the snapshots. The errors are listed in Table \ref{tab:test2_DQ0} (in the $no\_DQ$ case) and in Table \ref{tab:test2_DQ1} (in the $DQ$ case). These errors with their linear regression plots are drawn in Figure~\ref{fig:test2_eC0L2_eC0H1_eL2H1}. The convergence rate of the error in the $C^0(L^2)$-norm, $\mathcal{E}_{C^0(L^2)}$, is superoptimal in the $DQ$ case and strongly suboptimal in the $no\_DQ$ case. This clearly supports the theoretical rates of convergence in Table~\ref{tab:numerical_1}. The convergence rate of the error in the $C^0(H^1)$-norm, $\mathcal{E}_{C^0(H^1)}$, is superoptimal in the $DQ$ case and extremely suboptimal in the $no\_DQ$ case. This strongly supports the theoretical rates of convergence in Table~\ref{tab:numerical_1}. Finally, the convergence rate of the error in the $L^2(H^1)$-norm, $\mathcal{E}_{L^2(H^1)}$, is optimal in the $DQ$ case and strongly suboptimal in the $no\_DQ$ case. This supports the theoretical rates of convergence in Table~\ref{tab:numerical_1} for the $DQ$ case, but not for the $no\_DQ$ case. Overall, the numerical results clearly support the theoretical rates of convergence proved in Section~\ref{sec:error} and summarized in Table~\ref{tab:numerical_1}. We also emphasize that {\it the convergence rates in the $DQ$ case in all three norms are much higher than (and at least twice as high as) the corresponding rates of convergence in the $no\_DQ$ case}. \begin{table}[h] \centering \tabcaption{Burgers equation. Errors in the $no\_DQ$ approach.} \label{tab:test2_DQ0} \begin{tabular}{|c|c|c|c|c|c} \hline r & $\Lambda_r$ & $\mathcal{E}_{C^0(L^2)}$ & $\mathcal{E}_{C^0(H^1)}$ & $\mathcal{E}_{L^2(H^1)}$ \\ \hline 3 & 8.74e-02 & 2.38e-01 & 4.48e+01 & 2.34e+00 \\ 5 & 3.95e-02 & 1.60e-01 & 4.44e+01 & 1.76e+00 \\ 7 & 1.97e-02 & 1.17e-01 & 4.37e+01 & 1.24e+00 \\ 9 & 1.02e-02 & 9.05e-02 & 4.28e+01 & 8.94e-01 \\ 11 & 5.47e-03 & 7.01e-02 & 4.14e+01 & 6.84e-01 \\ \hline \end{tabular} \end{table} \begin{table}[h] \centering \tabcaption{Burgers equation. Errors in the $DQ$ approach.} \label{tab:test2_DQ1} \begin{tabular}{|c|c|c|c|c|c} \hline r & $\Lambda_r$ & $\mathcal{E}_{C^0(L^2)}$ & $\mathcal{E}_{C^0(H^1)}$ & $\mathcal{E}_{L^2(H^1)}$ \\ \hline 18 & 8.55e-02 & 6.83e-03 & 5.60e-01 & 2.82e-01 \\ 21 & 4.56e-02 & 2.99e-03 & 2.49e-01 & 1.37e-01 \\ 24 & 2.39e-02 & 1.31e-03 & 1.23e-01 & 6.72e-02 \\ 28 & 9.81e-03 & 4.29e-04 & 4.73e-02 & 2.57e-02 \\ 31 & 4.97e-03 & 1.88e-04 & 2.28e-02 & 1.25e-02 \\ \hline \end{tabular} \end{table} \begin{figure}[h] \begin{minipage}[h]{.45\linewidth} \includegraphics[width=1.1\textwidth]{Burg_eC0L2_wrt_r.eps} \end{minipage} \hspace{.3cm} \begin{minipage}[h]{.45\linewidth} \includegraphics[width=1.1\textwidth]{Burg_eC0H1_wrt_r_withSr.eps} \end{minipage} \hspace{.3cm} \begin{center} \begin{minipage}[h]{.45\linewidth} \includegraphics[width=1.1\textwidth]{Burg_eH1_wrt_r_withSr.eps} \end{minipage} \end{center} \caption{ Burgers equation. Plots of errors in $C^0(L^2)$-norm (top, left), $C^0(H^1)$-norm (top, right), and $L^2(H^1)$-norm (bottom). } \label{fig:test2_eC0L2_eC0H1_eL2H1} \end{figure} \section{Proper Orthogonal Decomposition Reduced Order Modeling} \label{sec:pod} In this section, we sketch the derivation of the standard POD Galerkin reduced order model for the heat equation. For a detailed presentation of reduced order modeling in general settings, the reader is referred to, e.g., \cite{HLB96,KV99,Sir87abc,burkardt2006pod,AK04,bui2008model,wang2012proper,balajewicz2012novel}. For clarity, we will denote by $C$ a generic constant that can depend on all the parameters in the system, except on $M$ (the number of snapshots), $d$ (the dimension of the set of snapshots, $V$), and $r$ (the number of POD modes used in the POD reduced order model). Let $X := H_0^1(\Omega)$, where $\Omega$ is the computational domain. Let $u(\cdot,t)\in X, t\in[0,T]$ be the weak solution of the weak formulation of the heat equation: \begin{eqnarray} (u_t , v) + \nu \, (\nabla u , \nabla v) = (f , v) \, \quad \forall v \in X. \label{eqn:heat_weak} \end{eqnarray} Given the time instances $t_1, \ldots, t_N \in [0,T]$, we consider the following two ensembles of snapshots: \begin{eqnarray} V^{no\_DQ} &:=& \mbox{span}\left\{ u(\cdot,t_0), \ldots, u(\cdot,t_N) \right\}, \label{snapshots_nodq} \\ V^{DQ} &:=& \mbox{span}\left\{ u(\cdot , t_0), \ldots, u(\cdot , t_N), \overline{\partial} u(\cdot , t_1), \ldots, \overline{\partial} u(\cdot , t_N) \right\}, \label{snapshots_dq} \end{eqnarray} where $\overline{\partial} u(t_n) := \frac{u(t_n) - u(t_{n-1})}{\Delta t} , \ n = 1, \ldots, N$ are the time {\it difference quotients (DQs)}. The two ensembles of snapshots correspond to the two cases investigated in this paper: (i) with the DQs not included in the snapshots (i.e., $V^{no\_DQ}$); and (ii) with the DQs included in the snapshots (i.e., $V^{DQ}$). As pointed out in Remark 1 in~\cite{KV01}, the ensemble of snapshots $V^{no\_DQ}$ and $V^{DQ}$ yield {\it different} POD bases. This is clearly illustrated by Figures~\ref{fig:pod_basis_heat}-\ref{fig:pod_basis_burgers}, which display POD basis functions for the heat equation and the Burgers equation, respectively (see Section~\ref{sec:numerical} for details regarding the numerical simulations). \begin{figure}[h] \begin{center} \begin{minipage}[h]{.9\linewidth} \includegraphics[width=0.9\textwidth]{Heat_basis_cmp.eps} \end{minipage} \end{center} \caption{ Heat equation. Plots of the POD basis functions when the DQs are used (denoted by DQ) and when the DQs are not used (denoted by no-DQ). } \label{fig:pod_basis_heat} \end{figure} \begin{figure}[h] \begin{center} \begin{minipage}[h]{.9\linewidth} \includegraphics[width=0.9\textwidth]{Burg_basis_cmp.eps} \end{minipage} \end{center} \caption{ Burgers equation. Plots of the POD basis functions when the DQs are used (denoted by DQ) and when the DQs are not used (denoted by no-DQ). } \label{fig:pod_basis_burgers} \end{figure} To simplify the presentation, we denote both sets of snapshots (i.e., $V^{no\_DQ}$ and $V^{DQ}$) by \begin{equation*} V = \mbox{span}\left\{ s_1, s_2, \ldots, s_M \right\}, \end{equation*} where $M=N+1$ when $V^{no\_DQ}$ is considered and $M=2N+1$ when $V^{DQ}$ is considered. We use the specific notation (i.e., $V^{no\_DQ}$ or $V^{DQ}$) only when this is necessary. Let dim $V = d$. The POD seeks a low-dimensional basis $\{ \varphi_1, \ldots, \varphi_r \}$, with $r \leq d$, which optimally approximates the input collection. Specifically, the POD basis satisfies \begin{eqnarray} \min \frac{1}{M} \sum_{i=1}^M \left\| s_i - \sum_{j=1}^r \bigl( s_i \, , \, \varphi_j(\cdot) \bigr)_{L^2} \, \varphi_j(\cdot) \right\|_{L^2}^2 \, , \label{pod_min} \end{eqnarray} subject to the conditions that $(\varphi_i,\varphi_j)_{L^2} = \delta_{ij}, \ 1 \leq i, j \leq r$. In order to solve \eqref{pod_min}, we consider the eigenvalue problem \begin{eqnarray} K \, v = \lambda \, v \, , \label{pod_eigenvalue} \end{eqnarray} where $K \in \mathbb{R}^{M \times M}$, with $\displaystyle K_{ij} = \frac{1}{M} \, ( s_j, s_i )_{L^2} \,$, is the snapshot correlation matrix, $\lambda_1 \geq \lambda_2 \geq \ldots \geq \lambda_d >0$ are the positive eigenvalues, and $v_k, \, k = 1, \ldots, d,$ are the associated eigenvectors. It can then be shown (see, e.g., \cite{HLB96,KV99}), that the solution of \eqref{pod_min} is given by \begin{eqnarray} \varphi_{k}(\cdot) = \frac{1}{\sqrt{\lambda_k}} \, \sum_{j=1}^{M} (v_k)_j \, s_{j}, \quad 1 \leq k \leq r, \label{pod_basis_formula} \end{eqnarray} where $(v_k)_j$ is the $j$-th component of the eigenvector $v_k$. \begin{definition} The term \begin{eqnarray} \eta^{interp}(x,t) := u(x,t) - \sum_{j=1}^r \bigl( u(x,t) \, , \, \varphi_j(x) \bigr)_{L^2} \, \varphi_j(x) \label{eqn:definition_interpolation_error} \end{eqnarray} will be denoted as the {\it POD interpolation error}. \end{definition} \begin{remark} \label{remark:pod_norm} Note that the $H^1$-norm can also be used to generate the POD basis \cite{KV01,singler2012new}. In this case, $\| \eta^{interp} \|_{L^2} \sim \| \nabla \eta^{interp} \|_{L^2}$. Thus, the two cases considered in this paper (i.e., $V = V^{no\_DQ}$ and $V = V^{DQ}$) yield error estimates that have the same convergence rates with respect to $r$. \end{remark} It can also be shown~\cite{KV01} that the following {\it POD approximation property} holds: \begin{eqnarray} && \frac{1}{N+1} \sum_{i=0}^N \left\| u(\cdot,t_i) - \sum_{j=1}^r \bigl( u(\cdot,t_i) \, , \, \varphi_j(\cdot) \bigr)_{L^2} \, \varphi_j(\cdot) \right\|_{L^2}^2 = \sum_{j=r+1}^{d} \lambda_j \qquad \text{if } \ V = V^{no\_DQ} \, , \label{pod_error_formula_nodq} \\ && \frac{1}{2 N + 1} \, \sum_{i=0}^N \left\| u(\cdot,t_i) - \sum_{j=1}^r \bigl( u(\cdot,t_i) \, , \, \varphi_j(\cdot) \bigr)_{L^2} \, \varphi_j(\cdot) \right\|_{L^2}^2 \label{pod_error_formula_dq} \\ && + \ \frac{1}{2 N + 1} \, \sum_{i=1}^N \left\| \overline{\partial} u(\cdot,t_i) - \sum_{j=1}^r \bigl( \overline{\partial} u(\cdot,t_i) \, , \, \varphi_j(\cdot) \bigr)_{L^2} \, \varphi_j(\cdot) \right\|_{L^2}^2 = \sum_{j=r+1}^{d} \lambda_j \quad \text{if } \ V = V^{DQ} \, . \quad \nonumber \end{eqnarray} The approximation property~\eqref{pod_error_formula_nodq}-\eqref{pod_error_formula_dq} represents the relationship between the average of the square of the $L^2$-norm of the interpolation error and the sum of the eigenvalues of the POD modes that are not included in the POD reduced order model. In order to be able to prove pointwise in time error estimates in Section~\ref{sec:error}, we also make the following assumption: \begin{assumption} We assume that, for $i = 1, \ldots, N$, the interpolation error satisfies the the following estimates: \begin{eqnarray} && \left\| u(\cdot,t_i) - \sum_{j=1}^r \bigl( u(\cdot,t_i) \, , \, \varphi_j(\cdot) \bigr)_{L^2} \, \varphi_j(\cdot) \right\|_{L^2}^2 \leq C \, \sum_{j=r+1}^{d} \lambda_j \qquad \text{if } \ V = V^{no\_DQ} \, , \label{pod_error_formula_nodq_pointwise} \\ && \left\| u(\cdot,t_i) - \sum_{j=1}^r \bigl( u(\cdot,t_i) \, , \, \varphi_j(\cdot) \bigr)_{L^2} \, \varphi_j(\cdot) \right\|_{L^2}^2 + \ \left\| \overline{\partial} u(\cdot,t_i) - \sum_{j=1}^r \bigl( \overline{\partial} u(\cdot,t_i) \, , \, \varphi_j(\cdot) \bigr)_{L^2} \, \varphi_j(\cdot) \right\|_{L^2}^2 \nonumber \\ && \hspace*{6.2cm} \leq C \, \sum_{j=r+1}^{d} \lambda_j \quad \text{if } \ V = V^{DQ} \, . \quad \label{pod_error_formula_dq_pointwise} \end{eqnarray} \label{assumption_pod_error_formula_pointwise} \end{assumption} \begin{remark} Assumption~\ref{assumption_pod_error_formula_pointwise} is natural. It simply says that in the sums in \eqref{pod_error_formula_nodq} and \eqref{pod_error_formula_dq} no individual term is much larger than the other terms in this sum. We also note that Assumption~\ref{assumption_pod_error_formula_pointwise} does not play an essential role in the error analysis in Section~\ref{sec:error}, since we will exclusively consider the continuous in time formulation. We mention that Assumption~\ref{assumption_pod_error_formula_pointwise} would follow directly from the POD approximation property \eqref{pod_error_formula_nodq}--\eqref{pod_error_formula_dq} if we dropped the $\frac{1}{M}$ factor in the snapshot correlation matrix $K$. In fact, this approach is used in, e.g., \cite{KV02,volkwein2011model}. We note, however, that this would most likely increase the magnitudes of the eigenvalues on the RHS of the POD approximation property \eqref{pod_error_formula_nodq}--\eqref{pod_error_formula_dq}. \label{remark_pod_error_formula_pointwise} \end{remark} In what follows, we will use the notation $X^r = \mbox{span}\{ \varphi_1, \varphi_2, \ldots, \varphi_r \} \, .$ To derive the POD reduced-order model for the heat equation~\eqref{eqn:heat_weak}, we employ the Galerkin truncation, which yields the following approximation $u_r \in X^r$ of $u$: \begin{eqnarray} u_r(x, t) := \sum_{j=1}^{r} a_j(t) \, \varphi_j(x) . \label{pod_g_truncation} \end{eqnarray} Plugging \eqref{pod_g_truncation} into \eqref{eqn:heat_weak} and multiplying by test functions in $X^r \subset X$ yields the {\em POD Galerkin reduced order model (POD-G-ROM)}: \begin{eqnarray} (u_{r, t} , v_r) + \nu \, (\nabla u_r , \nabla v_r) = (f , v_r) \quad \forall \, v_r \in X^r . \label{eqn:pod_g_rom} \end{eqnarray} The main advantage of the POD-G-ROM \eqref{eqn:pod_g_rom} over a straightforward finite element discretization of \eqref{eqn:heat_weak} is clear -- the computational cost of the former is dramatically lower than that of the latter. \section*{Acknowledgments} \bibliographystyle{plain}
{ "timestamp": "2013-06-18T02:00:31", "yymm": "1303", "arxiv_id": "1303.6012", "language": "en", "url": "https://arxiv.org/abs/1303.6012", "abstract": "This paper presents a theoretical and numerical investigation of the following practical question: Should the time difference quotients of the snapshots be used to generate the proper orthogonal decomposition basis functions? The answer to this question is important, since some published numerical studies use the time difference quotients, whereas other numerical studies do not. The criterion used in this paper to answer this question is the rate of convergence of the error of the reduced order model with respect to the number of proper orthogonal decomposition basis functions. Two cases are considered: the no_DQ case, in which the snapshot difference quotients are not used, and the DQ case, in which the snapshot difference quotients are used. The error estimates suggest that the convergence rates in the $C^0(L^2)$-norm and in the $C^0(H^1)$-norm are optimal for the DQ case, but suboptimal for the no_DQ case. The convergence rates in the $L^2(H^1)$-norm are optimal for both the DQ case and the no_DQ case. Numerical tests are conducted on the heat equation and on the Burgers equation. The numerical results support the conclusions drawn from the theoretical error estimates. Overall, the theoretical and numerical results strongly suggest that, in order to achieve optimal pointwise in time rates of convergence with respect to the number of proper orthogonal decomposition basis functions, one should use the snapshot difference quotients.", "subjects": "Numerical Analysis (math.NA)", "title": "Are the Snapshot Difference Quotients Needed in the Proper Orthogonal Decomposition?", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9770226280828406, "lm_q2_score": 0.8198933425148214, "lm_q1q2_score": 0.8010543482514554 }
https://arxiv.org/abs/2012.00525
The algebraic classification of nilpotent algebras
We give the complete algebraic classification of all complex 4-dimensional nilpotent algebras. The final list has 234 (parametric families of) isomorphism classes of algebras, 66 of which are new in the literature.
\section*{Introduction} The description, up to isomorphism, of all the algebras of some fixed dimension satisfying certain properties (the so-called algebraic classification) is a classical problem in algebra. There are many papers devoted to algebraic classification of small-dimensional algebras in several varieties of associative and non-associative algebras \cite{ack, lisa, cfk18, degr3, usefi1, degr1, degr2, fkkv, kk20, kkl19, demir, kpv19, hac16, ha16, maz79, kkp19geo }. Another interesting direction in the classification of algebras is the geometric classification (see, \cite{ fkkv,kpv19,kkp19geo } and references therein). Restricting our consideration to subvarieties of complex nilpotent $4$-dimensional algebras, we mention here the results on associative \cite{degr1}, commutative \cite{fkkv}, bicommutative \cite{kpv19}, Leibniz \cite{demir}, terminal \cite{kkp19geo} and $\mathfrak{CD}$-algebras \cite{kk20}. In the present paper, we complete the algebraic classification of allcomplex $4$-dimensional nilpotent algebras. Namely, we find $66$ new algebras and parametric families of algebras completing the list of $4$-dimensional nilpotent algebras initiated in the above mentioned works. Our approach is based on the calculation of central extensions of nilpotent algebras of dimension less than $4$. This method was developed by Skjelbred and Sund for Lie algebras in \cite{ss78} and has been an important tool for years (see, for example, \cite{hac16, omirov, ss78} and references therein). In Section~\ref{SS:method} we give a brief description of the method of central extensions and its adaptation to our case. It turns out that most of the nilpotent $4$-dimensional algebras are $\frak{CD}$-algebras, which were classified in \cite{kk20}. Our classification will thus be carried out modulo this subvariety of $4$-dimensional nilpotent algebras, as explained in Section~\ref{sec-non-CD}. In Section~\ref{sec-spec-types} we discuss several other useful classes of nilpotent algebras of dimension at most $4$. In particular, we give a list of the $3$-dimensional nilpotent algebras (they are all, in fact, $\frak{CD}$-algebras). The remainder of Section~\ref{S:alg} (Sections~\ref{sec-CD^3_01}--\ref{sec-CD^3_04}) is devoted to the construction and classification of $1$-dimensional non-$\frak{CD}$ central extensions of $3$-dimensional nilpotent algebras. The work done in Section~\ref{S:alg} culminates in the full classification of $4$-dimensional nilpotent algebras, which is completed in Section~\ref{S:class}, where we present the full list of representatives of isomorphism classes. Finally, in Section~\ref{S:apps} we present a few applications of our results to the classes of Lie-admissible and Alia type algebras. \medskip \paragraph{\bf Main result} The main result of the paper is Theorem~\ref{teo-alg} giving a full classification of complex $4$-dimensional nilpotent algebras. The complete list of non-isomorphic algebras consists of four parts: \begin{enumerate} \item trivial $\mathfrak{CD}$-algebras, which were classified in \cite[Theorem 2.1, Theorem 2.3 and Theorem 2.5]{demir}, the only anticommutative trivial $\mathfrak{CD}$-algebra being $\mathfrak{CD}_{03}^{4*}$; \item terminal non-trivial extensions of the family $\cd {3*}{04}(\lambda)$, which were classified in \cite[1.4.5. 1-dimensional central extensions of ${\bf T}^{3*}_{04}$]{kkp19geo}; \item non-trivial non-terminal $\mathfrak{ CD}$-algebras, which were classified in \cite{kk20}; \item the nilpotent algebras found in Sections~\ref{sec-CD^3_01}--\ref{sec-CD^3_04} of the present paper. \end{enumerate} \section{The algebraic classification of nilpotent algebras}\label{S:alg} \subsection{Method of classification of nilpotent algebras}\label{SS:method} Let ${\bf A}$ be an algebra, ${\bf V}$ a vector space and ${\rm Z}^{2}\left( {\bf A},{\bf V}\right)\cong {\rm Hom}({\bf A}\otimes {\bf A},\bf V)$ denote the space of bilinear maps $\theta :{\bf A}\times {\bf A}\longrightarrow {\bf V}.$ For $f\in{\rm Hom}({\bf A},{\bf V})$, we introduce $\delta f\in {\rm Z}^{2}\left( {\bf A},{\bf V}\right)$ by the equality $\delta f\left( x,y\right) =f(xy)$ and define ${\rm B}^{2}\left( {\bf A},{\bf V}\right) =\left\{\delta f \mid f\in {\rm Hom}\left( {\bf A},{\bf V}\right) \right\} $. One can easily check that ${\rm B}^{2}({\bf A},{\bf V})$ is a linear subspace of ${\rm Z}^{2}\left( {\bf A},{\bf V}\right)$. Let us define $\rm {H}^{2}\left( {\bf A},{\bf V}\right) $ as the quotient space ${\rm Z}^{2}\left( {\bf A},{\bf V}\right) \big/{\rm B}^{2}\left( {\bf A},{\bf V}\right)$. The equivalence class of $\theta\in {\rm Z}^{2}\left( {\bf A% },{\bf V}\right)$ in $\rm {H}^{2}\left( {\bf A},{\bf V}\right)$ is denoted by $\left[ \theta \right]$. Suppose now that $\dim{\bf A}=m<n$ and $\dim{\bf V}=n-m$. For any bilinear map $\theta :{\bf A}\times {\bf A}\longrightarrow {\bf V% }$, one can define on the space ${\bf A}_{\theta }:={\bf A}\oplus {\bf V}$ the bilinear product $\left[ -,-\right] _{% {\bf A}_{\theta }}$ by the equality $\left[ x+x^{\prime },y+y^{\prime }\right] _{% {\bf A}_{\theta }}= xy +\theta \left( x,y\right) $ for $x,y\in {\bf A},x^{\prime },y^{\prime }\in {\bf V}$. The algebra ${\bf A}_{\theta }$ is called an $(n-m)$-{\it dimensional central extension} of ${\bf A}$ by ${\bf V}$. It is also clear that ${\bf A}_{\theta }$ is nilpotent if and only if ${\bf A}$ is. For a bilinear map $\theta :{\bf A}\times {\bf A}\longrightarrow {\bf V}$, the space $\theta ^{\bot }=\left\{ x\in {\bf A}\mid \theta \left( x,{\bf A}\right) =\theta \left( {\bf A},x\right) =0\right\} $ is called the {\it annihilator} of $\theta$. For an algebra ${\bf A}$, the ideal ${\rm Ann}\left( {\bf A}\right) =\left\{ x\in {\bf A}\mid x{\bf A} ={\bf A}x =0\right\}$ is called the {\it annihilator} of ${\bf A}$. One has \begin{equation*} {\rm Ann}\left( {\bf A}_{\theta }\right) =\left( \theta ^{\bot }\cap {\rm Ann}\left( {\bf A}\right) \right) \oplus {\bf V}. \end{equation*} Any $n$-dimensional algebra with non-trivial annihilator can be represented in the form ${\bf A}_{\theta }$ for some $m$-dimensional algebra ${\bf A}$, an $(n-m)$-dimensional vector space ${\bf V}$ and $\theta \in {\rm Z}^{2}\left( {\bf A},{\bf V}\right)$, where $m<n$ (see \cite[Lemma 5]{hac16}). Moreover, there is a unique such representation with $m=n-\dim{\rm Ann}({\bf A})$. Note also that the latter equality is equivalent to the condition $\theta ^{\bot }\cap {\rm Ann}\left( {\bf A}\right)=0$. Let us pick some $\phi\in {\rm Aut}\left( {\bf A}\right)$, where ${\rm Aut}\left( {\bf A}\right)$ is the automorphism group of ${\bf A}$. For $\theta\in {\rm Z}^{2}\left( {\bf A},{\bf V}\right)$, let us define $(\phi \theta) \left( x,y\right) =\theta \left( \phi \left( x\right) ,\phi \left( y\right) \right) $. Then we get an action of ${\rm Aut}\left( {\bf A}\right) $ on ${\rm Z}^{2}\left( {\bf A},{\bf V}\right)$ that induces an action of the same group on $\rm {H}^{2}\left( {\bf A},{\bf V}\right)$. \begin{definition} Let ${\bf A}$ be an algebra and $I$ be a subspace of ${\rm Ann}({\bf A})$. If ${\bf A}={\bf A}_0 \oplus I$, for some subalgebra ${\bf A}_0$ of ${\bf A}$, then $I$ is called an {\it annihilator component} of ${\bf A}$. We say that ${\bf A}$ is {\it split} if it has some non-zero annihilator component; otherwise we say that ${\bf A}$ is {\it non-split}. \end{definition} For a linear space $\bf U$, the {\it Grassmannian} $G_{s}\left( {\bf U}\right) $ is the set of all $s$-dimensional linear subspaces of ${\bf U}$. For any $s\ge 1$, the action of ${\rm Aut}\left( {\bf A}\right)$ on $\rm {H}^{2}\left( {\bf A},\mathbb{C}\right)$ induces an action of the same group on $G_{s}\left( \rm {H}^{2}\left( {\bf A},\mathbb{C}\right) \right)$. Let us define $$ {\bf T}_{s}\left( {\bf A}\right) =\left\{ {\bf W}\in G_{s}\left( \rm {H}^{2}\left( {\bf A},\mathbb{C}\right) \right)\left|\underset{[\theta]\in W}{\cap }\theta^{\bot }\cap {\rm Ann}\left( {\bf A}\right) =0\right.\right\}. $$ Note that ${\bf T}_{s}\left( {\bf A}\right)$ is stable under the action of ${\rm Aut}\left( {\bf A}\right) $. Let us fix a basis $e_{1},\ldots,e_{s} $ of ${\bf V}$ and $\theta \in {\rm Z}^{2}\left( {\bf A},{\bf V}\right) $. Then there are unique $\theta _{i}\in {\rm Z}^{2}\left( {\bf A},\mathbb{C}\right)$ ($1\le i\le s$) such that $\theta \left( x,y\right) =\underset{i=1}{\overset{s}{% \sum }}\theta _{i}\left( x,y\right) e_{i}$ for all $x,y\in{\bf A}$. Note that $\theta ^{\bot }=\theta^{\bot} _{1}\cap \theta^{\bot} _{2}\cdots \cap \theta^{\bot} _{s}$ in this case. If $\theta ^{\bot }\cap {\rm Ann}\left( {\bf A}\right) =0$, then by \cite[Lemma 13]{hac16} the algebra ${\bf A}_{\theta }$ is split if and only if $\left[ \theta _{1}\right],\left[\theta _{2}\right] ,\ldots ,\left[ \theta _{s}\right] $ are linearly dependent in $\rm {H}^{2}\left( {\bf A},\mathbb{C}\right)$. Thus, if $\theta ^{\bot }\cap {\rm Ann}\left( {\bf A}\right) =0$ and ${\bf A}_{\theta }$ is non-split, then $\left\langle \left[ \theta _{1}\right] , \ldots,% \left[ \theta _{s}\right] \right\rangle$ is an element of ${\bf T}_{s}\left( {\bf A}\right)$. Now, if $\vartheta\in {\rm Z}^{2}\left( {\bf A},\bf{V}\right)$ is such that $\vartheta ^{\bot }\cap {\rm Ann}\left( {\bf A}\right) =0$ and ${\bf A}_{\vartheta }$ is non-split, then by \cite[Lemma 17]{hac16} one has ${\bf A}_{\vartheta }\cong{\bf A}_{\theta }$ if and only if $\left\langle \left[ \theta _{1}\right] ,\left[ \theta _{2}% \right] ,\ldots ,\left[ \theta _{s}\right] \right\rangle, \left\langle \left[ \vartheta _{1}\right] ,\left[ \vartheta _{2}\right] ,\ldots,% \left[ \vartheta _{s}\right] \right\rangle\in {\bf T}_{s}\left( {\bf A}\right)$ belong to the same orbit under the action of ${\rm Aut}\left( {\bf A}\right) $, where $% \vartheta \left( x,y\right) =\underset{i=1}{\overset{s}{\sum }}\vartheta _{i}\left( x,y\right) e_{i}$. Hence, there is a one-to-one correspondence between the set of $% {\rm Aut}\left( {\bf A}\right) $-orbits on ${\bf T}_{s}\left( {\bf A}% \right) $ and the set of isomorphism classes of central extensions of $\bf{A}$ by $\bf{V}$ with $s$-dimensional annihilator and trivial annihilator component. Consequently, to construct all $s$-dimensional central extensions with trivial annihilator component of a given $(n-s)$-dimensional algebra ${\bf A}$ one has to describe ${\bf T}_{s}({\bf A})$, ${\rm Aut}({\bf A})$ and the action of ${\rm Aut}({\bf A})$ on ${\bf T}_{s}({\bf A})$ and then for each orbit under the action of ${\rm Aut}({\bf A})$ on ${\bf T}_{s}({\bf A})$ pick a representative and construct the algebra corresponding to it. \subsubsection{Reduction to non-$\mathfrak{CD}$-algebras}\label{sec-non-CD} The class of $\mathfrak{CD}$-algebras is defined by the property that the commutator of any pair of multiplication operators is a derivation \cite{ack, kk20}; namely, an algebra $\mathfrak{A}$ is a $\mathfrak{CD}$-algebra if and only if \[ [T_x,T_y] \in \mathfrak{Der} (\mathfrak{A}),\] for all $x,y \in \mathfrak{A}$, where $T_z \in \{ R_z,L_z\}$. Here we use the notation $R_z$ (resp. $L_z$) for the operator of right (resp. left) multiplication in $\mathfrak{A}$. We will denote the variety of $\mathfrak{CD}$-algebras by $\mathfrak{CD}$. In terms of identities, the class of $\mathfrak{CD}$-algebras is defined by the following three: \begin{align*} ((xy)a)b-((xy)b)a&=((xa)b-(xb)a)y+x((ya)b-(yb)a),\\ (a(xy))b-a((xy)b)&=((ax)b-a(xb))y+x((ay)b-a(yb)),\\ a(b(xy))-b(a(xy))&=(a(bx)-b(ax))y+x(a(by)-b(ay)). \end{align*} Our method of classification of nilpotent algebras will be based on a classification of $\mathfrak{CD}$-algebras~\cite{kk20}. Clearly, any central extension of a non-$\mathfrak{CD}$-algebra is a non-$\mathfrak{CD}$-algebra. But a $\mathfrak{CD}$-algebra may have extensions which are not $\mathfrak{CD}$-algebras. More precisely, let $\bf{A}$ be a $\mathfrak{CD}$-algebra and $\theta \in {\rm Z^2}(\bf{A}, {\mathbb C}).$ Then ${\bf{A}}_{\theta }$ is a $\mathfrak{CD}$-algebra if and only if \begin{align} \label{cd1} \theta((xy)a,b)-\theta((xy)b,a)&=\theta((xa)b-(xb)a,y)+\theta(x,(ya)b-(yb)a),\\ \label{cd2} \theta(a(xy),b)-\theta(a,(xy)b)&=\theta((ax)b-a(xb),y)+\theta(x,(ay)b-a(yb)),\\ \label{cd3} \theta(a,b(xy))-\theta(b,a(xy))&=\theta(a(bx)-b(ax),y)+\theta(x,a(by)-b(ay)). \end{align} for all $x,y,z\in \bf{A}.$ Define the subspace ${\rm Z_\mathfrak{CD}^2}(\bf{A},{\mathbb C})$ of ${\rm Z^2}(\bf{A},{\mathbb C})$ by \begin{equation*} {\rm Z_{\mathfrak{CD}}^2}(\bf{A},{\mathbb C}) =\left\{\begin{array}{ll} \theta \in {\rm Z^2}(\bf{A},{\mathbb C}): & \theta \mbox{ satisfies } (\ref{cd1}), (\ref{cd2}) \mbox{ and } (\ref{cd3}) \end{array}\right\}. \end{equation*} Observe that ${\rm B^2}(\bf{A},{\mathbb C})\subseteq{\rm Z_\mathfrak{CD}^2}(\bf{A},{\mathbb C}).$ Let ${\rm H_\mathfrak{CD}^2}(\bf{A},{\mathbb C}) =% {\rm Z_\mathfrak{CD}^2}(\bf{A},{\mathbb C}) \big/{\rm B^2}(\bf{A},{\mathbb C}).$ Then ${\rm H_\mathfrak{CD}^2}(\bf{A},{\mathbb C})$ is a subspace of $% {\rm H^2}(\bf{A},{\mathbb C}).$ Define \[{\bf R}_{s}(\bf{A}) =\left\{ {\bf W}\in {\bf T}_{s}(\bf{A}) :{\bf W}\in G_{s}({\rm H_\mathfrak{CD}^2}(\bf{A},{\mathbb C}) ) \right\}, \] \[ {\bf U}_{s}(\bf{A}) =\left\{ {\bf W}\in {\bf T}_{s}(\bf{A}) :{\bf W}\notin G_{s}({\rm H_\mathfrak{CD}^2}(\bf{A},{\mathbb C}) ) \right\}.\] Then ${\bf T}_{s}(\bf{A}) ={\bf R}_{s}(\bf{A})$ $\mathbin{\mathaccent\cdot\cup}$ ${\bf U}_{s}(\bf{A}).$ The sets ${\bf R}_{s}(\bf{A}) $ and ${\bf U}_{s}(\bf{A})$ are stable under the action of $\operatorname{Aut}(\bf{A}).$ Thus, the nilpotent algebras corresponding to the representatives of $\operatorname{Aut}(\bf{A})$-orbits on ${\bf R}_{s}(\bf{A})$ are $\mathfrak{CD}$-algebras, while those corresponding to the representatives of $\operatorname{Aut}(\bf{A}) $-orbits on ${\bf U}_{s}(\bf{A})$ are non-$\mathfrak{CD}$-algebras. Hence, we may construct all nilpotent non-split non-$\mathfrak{CD}$-algebras $\bf{A}$ of dimension $n$ with $s$-dimensional annihilator from a given nilpotent algebra $\bf{A}^{\prime }$ of dimension $n-s$ as follows: \begin{enumerate} \item If $\bf{A}^{\prime }$ is non-$\mathfrak{CD}$, then apply the procedure. \item Otherwise, do the following: \begin{enumerate} \item Determine ${\bf U}_{s}(\bf{A}^{\prime})$ and $\operatorname{Aut}(\bf{A}^{\prime }).$ \item Determine the set of $\operatorname{Aut}(\bf{A}^{\prime })$-orbits on ${\bf U}_{s}(\bf{A}^{\prime }).$ \item For each orbit, construct the nilpotent algebra corresponding to one of its representatives. \end{enumerate} \end{enumerate} We will use the following auxiliary notation during the construction of central extensions. Let ${\bf A}$ be an algebra with basis $e_{1},e_{2},\ldots,e_{n}$. Then $\Delta_{ij}:{\bf A}\times {\bf A}\longrightarrow \mathbb{C}$ denotes the bilinear form defined by the equalities $\Delta _{ij}\left( e_{i},e_{j}\right) =1$ and $\Delta _{ij}\left( e_{l},e_{m}\right) =0$ for $(l,m) \neq (i,j)$. In this case, the $\Delta_{ij}$ with $1\leq i , j\leq n $ form a basis of the space of bilinear forms on $\bf{A}$. We also denote by \begin{longtable}{lll} $\mathfrak{CD}^{i*}_j$& the $j$th $i$-dimensional nilpotent trivial $\mathfrak{CD}$-algebra, \\ $\mathfrak{CD}^i_j$& the $j$th $i$-dimensional nilpotent non-trivial $\mathfrak{CD}$-algebra, \\ $\D{i}{j}$& the $j$th $i$-dimensional nilpotent terminal algebra,\\ ${{\bf N}}^i_j$& the $j$th $i$-dimensional nilpotent non-$\mathfrak{CD}$-algebra. \end{longtable} \subsection{Some special types of nilpotent algebras}\label{sec-spec-types} \subsubsection{Trivial $\mathfrak{CD}$-algebras} Recall that the class of $n$-dimensional algebras defined by the identities $(xy)z=0$ and $x(yz)=0$ lies in the intersection of all well-known varieties of algebras defined by a family of polynomial identities of degree $3,$ such as Leibniz, Zinbiel, associative, Novikov and many other algebras. On the other hand, all algebras defined by the identities $(xy)z=0$ and $x(yz)=0$ are central extensions of some suitable algebra with zero product. The list of all non-anticommutative $4$-dimensional algebras defined by the identities $(xy)z=0$ and $x(yz)=0$ can be found in \cite{demir}. Note that there is only one $4$-dimensional nilpotent anticommutative algebra satisfying the identity $(xy)z=0.$ Obviously, all algebras from this list are $4$-dimensional nilpotent $\mathfrak{CD}$-algebras. \subsubsection{$2$-dimensional nilpotent algebras} There is only one non-zero $2$-dimensional nilpotent algebra. It is a $\mathfrak{CD}$-algebra: \begin{longtable}{ll llll} $\cd {2*}{01}$ & $:$ & $e_1 e_1 = e_2$ \end{longtable} \subsubsection{$3$-dimensional nilpotent algebras} Thanks to \cite{cfk18}, we have the classification of all $3$-dimensional nilpotent algebras. It is easy to see that each $3$-dimensional nilpotent algebra is a $\mathfrak{CD}$-algebra: \begin{longtable}{lllll llll} $\cd 3{01}$&$:$& $e_1 e_1 = e_2$ & $e_2 e_2=e_3$ \\ \hline $\cd 3{02}$&$:$& $e_1 e_1 = e_2$ & $e_2 e_1= e_3$ & $e_2 e_2=e_3$ \\ \hline$\cd 3{03}$&$:$& $e_1 e_1 = e_2$ & $e_2 e_1=e_3$ \\ \hline$\cd 3{04}(\lambda)$&$:$& $ e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$ \\ \hline$\cd {3*}{01}$&$:$& $e_1 e_1 = e_2$\\ \hline$\cd {3*}{02}$&$:$& $e_1 e_1 = e_3$ &$ e_2 e_2=e_3$ \\ \hline$\cd {3*}{03}$&$:$& $e_1 e_2=e_3$ & $e_2 e_1=-e_3$ \\ \hline$\cd {3*}{04}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_2 e_1=e_3$ & $e_2 e_2=e_3$ \end{longtable} \subsubsection{$4$-dimensional nilpotent algebras with $2$-dimensional annihilator } Thanks to \cite{cfk18}, we have the classification of all $4$-dimensional non-split nilpotent non-trivial algebras with $2$-dimensional annihilator. All of these algebras are $\mathfrak{CD}$-algebras: \begin{longtable}{lllll llll} $\cd 4{05}$&: & $e_1 e_1 = e_2$ & $e_2 e_1=e_4$ & $e_2 e_2=e_3$ \\ \hline$\cd 4{06}$ & $:$ & $ e_1 e_1 = e_2$ & $e_1 e_2=e_4$ & $e_2 e_1=e_3$ \\ \hline$\cd 4{07}(\lambda)$& $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_4$ & $e_2 e_1=\lambda e_4$ & $e_2 e_2=e_3$ \\ \end{longtable} \subsubsection{$4$-dimensional nilpotent algebras with $1$-dimensional annihilator } Thanks to \cite{kk20} we have the classification of all $4$-dimensional nilpotent $\mathfrak{CD}$-algebras with $1$-dimensional annihilator. The remaining nilpotent non-$\mathfrak{CD}$-algebras with $1$-dimensional annihilator will be found in the present paper. \subsubsection{Trivial central extensions.} Thanks to \cite{kk20} we know that all central extensions of $\mathfrak{CD}^{3*}_{01},$ $\mathfrak{CD}^{3*}_{02},$ $\mathfrak{CD}^{3*}_{03}$ and $\mathfrak{CD}^{3*}_{04}$ are $\mathfrak{CD}$-algebras. \bigskip\bigskip \subsection{$1$-dimensional central extensions of $\cd 3{01}$}\label{sec-CD^3_01} Here we collect all information about ${\mathfrak{CD}_{01}^{3}}$: \begin{longtable}{|l|l|l|} \hline Algebra & Multiplication & Cohomology \\ \hline ${\mathfrak{CD}}_{01}^{3}$ & $\begin{array}{l}e_1 e_1 = e_2\\ e_2 e_2=e_3 \end{array}$& $\begin{array}{lcl} {\rm H^2_{\mathfrak{CD}}}({\mathfrak{CD}}_{01}^{3})&=&\langle [\Delta_{12}], [\Delta_{21}] \rangle \\ {\rm H^2}({\mathfrak{CD}}_{01}^{3})&=&{\rm H^2_{\mathfrak{CD}}}({\mathfrak{CD}_{01}^3}) \oplus \langle[\Delta_{13}], [\Delta_{31}], [\Delta_{23}], [\Delta_{32}], [\Delta_{33}] \rangle \end{array}$\\ \hline \end{longtable} Let us use the following notations \begin{longtable}{llll} $\nb 1 = \Dl 12$&$ \nb 2 = \Dl 21$& $\nb 3 = \Dl 13$ &$\nb 4=\Dl 31$\\ $\nb 5 = \Dl 23$&$ \nb 6 = \Dl 32$& $\nb 7 = \Dl 33$. \end{longtable} Take $\theta=\sum_{i=1}^7\alpha_i\nb i\in {\rm H}^2 (\cd 3{01})$. If $$ \phi= \begin{pmatrix} x & 0 & 0\\ 0 & x^2 & 0\\ y & 0 & x^4 \end{pmatrix}\in\aut{\cd 3{01}}, $$ then $$ \phi^T\begin{pmatrix} 0 & \alpha_1 & \alpha_3\\ \alpha_2 & 0 & \alpha_5\\ \alpha_4 & \alpha_6 & \alpha_7 \end{pmatrix} \phi= \begin{pmatrix} \alpha^* & \alpha_1^* & \alpha_3^* \\ \alpha^*_2 & 0 & \alpha^*_5\\ \alpha^*_4 & \alpha^*_6 & \alpha^*_7 \end{pmatrix}, $$ where \begin{longtable}{llll} $\alpha^*_1=x^2 (x \alpha_1+y \alpha_6)$ & $\alpha^*_2=x^2 (x \alpha_2+y \alpha_5)$ & $\alpha^*_3=x^4 (x \alpha_3+y \alpha_7)$ & $\alpha^*_4=x^4 (x \alpha_4+y \alpha_7)$\\ $\alpha^*_5=x^6 \alpha_5$ & $\alpha^*_6=x^6 \alpha_6$& $\alpha^*_7=x^8 \alpha_7$.& \end{longtable} Hence, $\phi\langle\theta\rangle=\langle\theta^*\rangle$, where $\theta^*=\sum\limits_{i=1}^7 \alpha_i^* \nb i.$ We are only interested in elements with $(\alpha_3, \alpha_4, \alpha_5, \alpha_6, \alpha_7)\neq (0,0,0,0,0).$ Then \begin{enumerate} \item $\alpha_7\neq 0.$ \begin{enumerate} \item $\alpha_6\neq 0$. Then choosing $x=\sqrt{\alpha_6 \alpha_7^{-1}}$ and $y=- \alpha_4\alpha_7^{-1}x$, we have the family of representatives $\langle \alpha \nabla_1+\beta \nabla_2+\gamma \nabla_3+\delta \nb 5+\nb 6+\nb 7 \rangle.$ Observe that two distinct quadruples $(\alpha,\beta,\gamma, \delta)$ and $(\alpha',\beta',\gamma', \delta')$ determine the same orbit if and only if $(\alpha,\beta,\gamma, \delta)=(-\alpha',-\beta',-\gamma', \delta').$ \item $\alpha_6= 0$ and $\alpha_5\neq0$. Choosing $x=\sqrt{\alpha_5 \alpha_7^{-1}}$ and $y=- \alpha_4\alpha_7^{-1}x$, we have the family of representatives $\langle \alpha \nabla_1+\beta \nabla_2+\gamma \nabla_3+ \nb 5 +\nb 7 \rangle.$ Observe that two distinct triples $(\alpha,\beta,\gamma)$ and $(\alpha',\beta',\gamma')$ determine the same orbit if and only if $(\alpha,\beta,\gamma)=(-\alpha',-\beta',-\gamma').$ \item $\alpha_6=\alpha_5=0$ and $\alpha_3 \neq \alpha_4$. Choosing $x=\sqrt[3]{(\alpha_3-\alpha_4) \alpha_7^{-1}}$ and $y=- \alpha_4\alpha_7^{-1}x$, we have the family of representatives $\langle \alpha \nabla_1+\beta \nabla_2+ \nabla_3+ \nb 7 \rangle.$ Observe that two pairs $(\alpha,\beta)$ and $(\alpha',\beta')$ determine the same orbit if and only if $(\alpha,\beta )=(\xi \alpha',\xi \beta'),$ where $\xi^3=1.$ \item $\alpha_6=\alpha_5= 0,$ $\alpha_3 = \alpha_4$ and $\alpha_2\ne 0$. Choosing $x=\sqrt[5]{\alpha_2\alpha_7^{-1}}$ and $y=-\alpha_4\alpha_7^{-1}x$, we have the family of representatives of distinct orbits $\langle \alpha \nabla_1+\nabla_2+\nb 7 \rangle.$ \item $\alpha_6=\alpha_5=0$, $\alpha_3 = \alpha_4$ and $\alpha_2=0$. Choosing $y=-\alpha_4\alpha_7^{-1}x$, we have two representatives $\langle \nb 7 \rangle$ and $\langle \nabla_1+ \nb 7 \rangle$ depending on whether $\alpha_1=0$ or not. \end{enumerate} \item $\alpha_7=0.$ \begin{enumerate} \item $\alpha_6\neq 0$ and $\alpha_4 \neq 0$. Choosing $x=\alpha_4 \alpha_6^{-1}$ and $y=-\alpha_1\alpha_4 \alpha_6^{-2},$ we have the family of representatives of distinct orbits $\langle \alpha \nabla_2+\beta \nabla_3+ \nabla_4+ \gamma \nb 5 +\nb 6 \rangle.$ \item $\alpha_6\neq 0,$ $\alpha_4 = 0$ and $\alpha_3\neq0$. Choosing $x=\alpha_3 \alpha_6^{-1}$ and $y=-\alpha_1\alpha_3 \alpha_6^{-2},$ we have the family of representatives of distinct orbits $\langle \alpha \nabla_2+ \nabla_3+ \beta \nb 5 +\nb 6 \rangle.$ \item $\alpha_6\neq 0$ and $\alpha_4=\alpha_3 = 0$. Choosing $y=-\alpha_1\alpha_6^{-1}x$ we have two families of representatives of distinct orbits $\langle \alpha \nb 5 +\nb 6 \rangle$ and $\langle \nabla_2+ \alpha \nb 5 +\nb 6 \rangle$ depending on whether $\alpha_1\alpha_5 - \alpha_2\alpha_6=0$ or not. \item $\alpha_6 = 0,$ $\alpha_5\neq0$ and $\alpha_4 \neq 0$. Choosing $x=\alpha_4 \alpha_5^{-1}$ and $y=-\alpha_2\alpha_4 \alpha_5^{-2},$ we have the family of representatives of distinct orbits $\langle \alpha \nb 1+\beta \nabla_3 +\nb 4+\nb 5 \rangle.$ \item $\alpha_6 = 0,$ $\alpha_5\neq0$ and $\alpha_4 = 0$. Choosing $y=-\alpha_2\alpha_5^{-1}x$, we have the following representatives of distinct orbits $\langle \alpha \nb 1 +\nb 3+\nb 5 \rangle$, $\langle \nb 1 +\nb 5 \rangle$ and $\langle \nb 5 \rangle,$ corresponding to the 3 cases: $\alpha_3\ne 0$; $\alpha_3=0$, $\alpha_1\ne 0$; $\alpha_3=\alpha_1=0$, respectively. \item $\alpha_6 = \alpha_5 = 0$ and $\alpha_4 \neq 0$. We have the following families of representatives of distinct orbits $\langle \nb 1+\alpha \nb 2+\beta\nb 3+\nb 4 \rangle$, $\langle \nb 2+\alpha\nb 3+\nb 4 \rangle$ and $\langle \alpha\nb 3+\nb 4 \rangle$ corresponding to the 3 cases: $\alpha_1\ne 0$; $\alpha_1=0$, $\alpha_2\ne 0$; $\alpha_1=\alpha_2=0$, respectively. \item $\alpha_6 = \alpha_5=\alpha_4 = 0$ and $\alpha_3\neq 0$. We have the following representatives of distinct orbits $\langle \alpha \nb 1+ \nabla_2 +\nb 3 \rangle,$ $\langle \nb 1+ \nb 3 \rangle$ and $\langle \nb 3 \rangle$ corresponding to the 3 cases: $\alpha_2\ne 0$; $\alpha_2=0$, $\alpha_1\ne 0$; $\alpha_2=\alpha_1=0$, respectively. \end{enumerate} \end{enumerate} Summarizing, we have the following distinct orbits: \begin{longtable}{lll} $\langle \alpha \nb 1+ \nabla_2 +\nb 3 \rangle$ & \multicolumn{2}{l}{$\langle \alpha \nabla_1+\beta \nabla_2+\gamma \nabla_3+\delta \nb 5+\nb 6+\nb 7 \rangle^{O(\alpha,\beta,\gamma, \delta)=O(-\alpha,-\beta,-\gamma, \delta)}$}\\ $\langle \alpha \nb 1+\nb 3+\nb 5 \rangle$ & \multicolumn{2}{l}{$\langle \alpha \nabla_1+\beta \nabla_2+\gamma \nabla_3+\nb 5+\nb 7 \rangle^{O(\alpha,\beta,\gamma)=O(-\alpha,-\beta,-\gamma)}$}\\ \multicolumn{2}{l}{$\langle \alpha \nabla_1+\beta \nabla_2+ \nabla_3 +\nb 7 \rangle^{O(\alpha,\beta )=O(\sqrt[3]1(\alpha,\beta ))}$}& $\langle \alpha \nabla_1+\nabla_2+ \nb 7 \rangle$ \\ $\langle \nb 1+ \nb 3 \rangle$ & $\langle \nb 1+\alpha \nb 2+\beta\nb 3+\nb 4 \rangle$ & $\langle \alpha \nb 1+\beta \nabla_3 +\nb 4+\nb 5 \rangle$ \\ $\langle \nb 1 +\nb 5 \rangle$ & $\langle \nabla_1+ \nb 7 \rangle$ & $\langle \nb 2+ \alpha \nabla_3 +\nb 4 \rangle$ \\ $\langle \alpha \nabla_2+\beta \nabla_3+ \nabla_4+ \gamma \nb 5 +\nb 6 \rangle$ & $\langle \alpha \nabla_2+ \nabla_3+ \beta \nb 5 +\nb 6 \rangle$ & $\langle \nb 2 + \alpha\nb 5+\nb 6 \rangle$ \\ $\langle \alpha \nabla_3 +\nb 4 \rangle$ & $\langle \nb 3 \rangle$ & $\langle \nb 5 \rangle$ \\ $\langle \alpha \nb 5 +\nb 6 \rangle$ & $\langle \nb 7 \rangle$ \end{longtable} They correspond to the following new algebras: {\tiny \begin{longtable}{lllllllllllllllllll} ${\bf N}^4_{01}(\alpha)$ &$:$& $e_1 e_1 = e_2$& $e_1e_2=\alpha e_4$& $e_1e_3=e_4$& $e_2e_1=e_4$& $e_2 e_2=e_3$ \\ \hline ${\bf N}^4_{02}(\alpha,\beta,\gamma,\delta)$ &$:$& $e_1 e_1 = e_2$& $e_1e_2=\alpha e_4$& $e_1e_3=\gamma e_4$& $e_2e_1=\beta e_4$\\&& $e_2 e_2=e_3$& $e_2e_3=\delta e_4$& $e_3e_2=e_4$& $e_3e_3=e_4$\\ \hline ${\bf N}^4_{03}(\alpha)$ &$:$& $e_1 e_1 = e_2$& $e_1e_2=\alpha e_4$& $e_1e_3=e_4$& $e_2 e_2=e_3$ & $e_2e_3=e_4$\\ \hline ${\bf N}^4_{04}(\alpha,\beta,\gamma)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=\gamma e_4$&$e_2e_1=\beta e_4$\\ && $e_2 e_2=e_3$&$e_2e_3=e_4$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{05}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=e_4$ &$e_2e_1=\beta e_4$& $e_2 e_2=e_3$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{06}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_1=e_4$& $e_2 e_2=e_3$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{07}$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$&$e_1e_3=e_4$& $e_2 e_2=e_3$ \\ \hline ${\bf N}^4_{08}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$&$e_1e_3=\beta e_4$&$e_2e_1=\alpha e_4$& $e_2 e_2=e_3$ &$e_3e_1=e_4$\\ \hline ${\bf N}^4_{09}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=\beta e_4$& $e_2 e_2=e_3$ &$e_2e_3=e_4$&$e_3e_1=e_4$ \\ \hline ${\bf N}^4_{10}$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$& $e_2 e_2=e_3$ &$e_2e_3=e_4$ \\ \hline ${\bf N}^4_{11}$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$& $e_2 e_2=e_3$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{12}(\alpha)$ &$:$& $e_1 e_1 = e_2$& $e_1e_3=\alpha e_4$&$e_2e_1=e_4$ & $e_2 e_2=e_3$ &$e_3e_1=e_4$ \\ \hline ${\bf N}^4_{13}(\alpha,\beta,\gamma)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=\beta e_4$&$e_2e_1=\alpha e_4$& $e_2 e_2=e_3$\\ &&$e_2e_3=\gamma e_4$&$e_3e_1=e_4$&$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{14}(\alpha, \beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3 =e_4$&$e_2e_1= \alpha e_4$& $e_2 e_2=e_3$ &$e_2e_3=\beta e_4$&$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{15}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_4$& $e_2 e_2=e_3$ &$e_2e_3=\alpha e_4$ &$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{16}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=\alpha e_4$ & $e_2 e_2=e_3$ &$e_3e_1=e_4$\\ \hline ${\bf N}^4_{17}$ &$:$& $e_1 e_1 = e_2$& $e_1e_3=e_4$ & $e_2 e_2=e_3$\\ \hline ${\bf N}^4_{18}$ &$:$& $e_1 e_1 = e_2$& $e_2 e_2=e_3$ &$e_2e_3=e_4$\\ \hline ${\bf N}^4_{19}(\alpha)$ &$:$& $e_1 e_1 = e_2$& $e_2 e_2=e_3$ &$e_2e_3=\alpha e_4$ &$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{20}$ &$:$& $e_1 e_1 = e_2$& $e_2 e_2=e_3$ &$e_3e_3=e_4$\\ \hline \end{longtable} } \begin{longtable}{cc} ${\bf N}^4_{02}(\alpha,\beta,\gamma,\delta) \cong {\bf N}^4_{02}(-(\alpha,\beta,\gamma),\delta)$ & ${\bf N}^4_{04}(\alpha,\beta,\gamma) \cong {\bf N}^4_{04}(-(\alpha,\beta,\gamma))$ \\ \multicolumn{2}{c}{${\bf N}^4_{05}(\alpha,\beta) \cong {\bf N}^4_{05}(\sqrt[3]{1}(\alpha,\beta))$} \end{longtable} \subsection{$1$-dimensional central extensions of $\cd 3{02}$} Here we collect all information about ${\mathfrak{CD}_{02}^{3}}$: \begin{longtable}{|l|l|l|} \hline Algebra & Multiplication & Cohomology \\ \hline ${\mathfrak{CD}}_{02}^{3}$ & $\begin{array}{l} e_1 e_1 = e_2\\ e_2 e_1= e_3\\ e_2 e_2=e_3 \end{array}$& $\begin{array}{lcl} {\rm H^2_{\mathfrak{CD}}}({\mathfrak{CD}}_{02}^{3})&=&\langle [\Delta_{12}], [\Delta_{21}] \rangle \\ {\rm H^2}({\mathfrak{CD}}_{02}^{3})&=&{\rm H^2_{\mathfrak{CD}}}({\mathfrak{CD}_{02}^3}) \oplus \langle[\Delta_{13}], [\Delta_{31}], [\Delta_{23}], [\Delta_{32}], [\Delta_{33}] \rangle \end{array}$\\ \hline \end{longtable} Let us use the following notations \begin{longtable}{llll} $\nb 1 = \Dl 12$&$ \nb 2 = \Dl 21$& $\nb 3 = \Dl 13$ &$\nb 4=\Dl 31$\\ $\nb 5 = \Dl 23$&$ \nb 6 = \Dl 32$& $\nb 7 = \Dl 33$. \end{longtable} Take $\theta=\sum_{i=1}^7\alpha_i\nb i\in {\rm H}^2(\cd 3{02})$. If $$ \phi= \begin{pmatrix} 1 & 0 & 0\\ 0 & 1 & 0\\ x & 0 & 1 \end{pmatrix}\in\aut{\cd 3{02}}, $$ then $$ \phi^T\begin{pmatrix} 0 & \alpha_1 & \alpha_3\\ \alpha_2 & 0 & \alpha_5\\ \alpha_4 & \alpha_6 & \alpha_7 \end{pmatrix} \phi= \begin{pmatrix} \alpha^* & \alpha_1^* & \alpha_3^* \\ \alpha^*_2 & 0 & \alpha^*_5\\ \alpha^*_4 & \alpha^*_6 & \alpha^*_7 \end{pmatrix}, $$ where \begin{longtable}{llll} $\alpha^*_1=\alpha_1+x \alpha_6$ & $\alpha^*_2=\alpha_2+x \alpha_5$ & $\alpha^*_3=\alpha_3+x \alpha_7$ & $\alpha^*_4=\alpha_4+x \alpha_7$\\ $\alpha^*_5= \alpha_5$ & $\alpha^*_6= \alpha_6$& $\alpha^*_7= \alpha_7$.& \end{longtable} Hence, $\phi\langle\theta\rangle=\langle\theta^*\rangle$, where $\theta^*=\sum\limits_{i=1}^7 \alpha_i^* \nb i.$ We are only interested in elements with $(\alpha_3, \alpha_4, \alpha_5, \alpha_6, \alpha_7)\neq (0,0,0,0,0).$ Then \begin{enumerate} \item $\alpha_7\ne 0$. Then choosing $x=-\alpha_3\alpha_7^{-1}$, we obtain the family of representatives of distinct orbits $\langle \alpha \nb 1+ \beta \nb 2+\gamma \nb 4+\delta \nb 5+\varepsilon \nb 6+\nb 7 \rangle$. \item $\alpha_7=0$ and $\alpha_6\ne 0$. Then choosing $x=-\alpha_1\alpha_6^{-1}$, we obtain the family of representatives of distinct orbits $\langle \alpha \nb 2+\beta \nb 3+\gamma \nb 4+ \delta \nb 5+ \nb 6 \rangle$. \item $\alpha_7=\alpha_6=0$ and $\alpha_5\ne 0$. Then choosing $x=-\alpha_2\alpha_5^{-1}$, we obtain the family of representatives of distinct orbits $\langle \alpha \nb 1+\beta \nb 3+\gamma \nb 4+ \nb 5 \rangle$. \item $\alpha_7=\alpha_6=\alpha_5=0$ and $\alpha_4\ne 0$. Then we obtain the family of representatives of distinct orbits $\langle \alpha \nb 1+\beta \nb 2+\gamma\nb 3+ \nb 4 \rangle$. \item $\alpha_7=\alpha_6=\alpha_5=\alpha_4=0$ and $\alpha_3\ne 0$. Then we obtain the family of representatives of distinct orbits $\langle \alpha \nb 1+\beta \nb 2+ \nb 3 \rangle$. \end{enumerate} Summarizing, we have the following distinct orbits: \begin{longtable}{ll} $\langle \alpha \nb 1+\beta \nb 2+ \nb 3 \rangle$ & $\langle \alpha \nb 1+\beta \nb 2+\gamma\nb 3+ \nb 4 \rangle$\\ $\langle \alpha \nb 1+\beta \nb 3+\gamma \nb 4+ \nb 5 \rangle$ & $\langle \alpha \nb 2+\beta \nb 3+\gamma \nb 4+ \delta \nb 5+ \nb 6 \rangle$\\ \multicolumn{2}{c}{$\langle \alpha \nb 1+ \beta \nb 2+\gamma \nb 4+\delta \nb 5+\varepsilon \nb 6+\nb 7 \rangle$}\\ \end{longtable} This gives the following new algebras: {\tiny \begin{longtable}{lllllllllllllll} ${\bf N}^4_{21}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=e_4$&$e_2e_1=e_3+\beta e_4$& $e_2 e_2=e_3$ \\ \hline ${\bf N}^4_{22}(\alpha,\beta,\gamma)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=\gamma e_4$&$e_2e_1=e_3+\beta e_4$& $e_2 e_2=e_3$&$e_3e_1=e_4$ \\ \hline ${\bf N}^4_{23}(\alpha,\beta,\gamma)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=\beta e_4$&$e_2e_1=e_3$\\ && $e_2 e_2=e_3$&$e_2e_3=e_4$&$e_3e_1=\gamma e_4$ \\ \hline ${\bf N}^4_{24}(\alpha,\beta,\gamma,\delta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=\beta e_4$&$e_2e_1=e_3+\alpha e_4$& $e_2 e_2=e_3$ \\ &&$e_2e_3=\delta e_4$&$e_3e_1=\gamma e_4$ & $e_3e_2=e_4$\\ \hline ${\bf N}^4_{25}(\alpha, \beta, \gamma, \delta, \varepsilon)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2= \alpha e_4$& $e_2e_1=e_3+\beta e_4$ & $e_2 e_2=e_3$\\ &&$e_2e_3=\delta e_4$ & $e_3e_1=\gamma e_4$ & $e_3e_2=\varepsilon e_4$ &$e_3e_3=e_4$\\ \hline \end{longtable} } \subsection{$1$-dimensional central extensions of $\cd 3{03}$} Here we collect all information about ${\mathfrak{CD}_{03}^{3}}$: \begin{longtable}{|l|l|l|} \hline Algebra & Multiplication & Cohomology \\ \hline ${\mathfrak{CD}}_{03}^{3}$ & $\begin{array}{l} e_1 e_1 = e_2\\ e_2 e_1=e_3 \end{array}$& $\begin{array}{lcl} {\rm H^2_{\mathfrak{CD}}}({\mathfrak{CD}}_{03}^{3})&=&\langle \Dl 12, \Dl 22, \Dl 13-2\Dl 31 \rangle \\ {\rm H^2}({\mathfrak{CD}}_{03}^{3})&=&{\rm H^2_{\mathfrak{CD}}}({\mathfrak{CD}_{03}^3}) \oplus \langle \Dl 31, \Dl 23, \Dl 32, \Dl 33 \rangle \end{array}$\\ \hline \end{longtable} Let us use the following notations \begin{longtable}{llll} $\nb 1 = \Dl 12$&$ \nb 2 = \Dl 22$& $\nb 3 = \Dl 13-2\Dl 31$ &$\nb 4=\Dl 31$\\ $\nb 5 = \Dl 23$&$ \nb 6 = \Dl 32$& $\nb 7 = \Dl 33$. \end{longtable} Take $\theta=\sum_{i=1}^7\alpha_i\nb i\in {\rm H}^2(\cd 3{03})$. If $$ \phi= \begin{pmatrix} x & 0 & 0\\ y & x^2 & 0\\ z & xy & x^3 \end{pmatrix}\in\aut{\cd 3{03}}, $$ then $$ \phi^T\begin{pmatrix} 0 & \alpha_1 & \alpha_3\\ 0 & \alpha_2 & \alpha_5\\ \alpha_4-2\alpha_3 & \alpha_6 & \alpha_7 \end{pmatrix} \phi= \begin{pmatrix} \alpha^* & \alpha^*_1 & \alpha^*_3\\ \alpha^{**} & \alpha^*_2 & \alpha^*_5\\ \alpha^*_4-2\alpha^*_3 & \alpha^*_6 & \alpha^*_7 \end{pmatrix}, $$ where \begin{longtable}{lclc} $\alpha^*_1$ &$=$&$ (\alpha_1x^2 + (\alpha_2+\alpha_3)xy + \alpha_5y^2 + \alpha_6xz + \alpha_7yz)x,$\\ $\alpha^*_2$ &$=$&$ (\alpha_2x^2 + (\alpha_5+\alpha_6)xy + \alpha_7y^2)x^2,$\\ $\alpha^*_3$ &$=$&$ (\alpha_3x + \alpha_5y + \alpha_7z)x^3,$\\ $\alpha^*_4$ &$=$&$ (\alpha_4x + (2\alpha_5+\alpha_6)y + 3\alpha_7z)x^3,$\\ $\alpha^*_5$ &$=$&$ (\alpha_5x + \alpha_7y)x^4,$\\ $\alpha^*_6$ &$=$&$ (\alpha_6x + \alpha_7y)x^4,$\\ $\alpha^*_7$ &$=$&$ \alpha_7x^6.$ \end{longtable} Hence, $\phi\langle\theta\rangle=\langle\theta^*\rangle$, where $\theta^*=\sum\limits_{i=1}^7 \alpha_i^* \nb i.$ We are only interested in $\theta$ with $(\alpha_4, \alpha_5, \alpha_6, \alpha_7)\neq (0,0,0,0)$. \begin{enumerate} \item $\alpha_7\ne 0$. Then put $y=-\frac{\alpha_6x}{\alpha_7}$ and $z=\frac x{\alpha_7^2}(\alpha_5\alpha_6 - \alpha_3\alpha_7)$ to make $\alpha^*_3=\alpha^*_6=0$. \begin{enumerate} \item $\alpha_5\ne \alpha_6$. Then choosing $x=\frac {\alpha_5 - \alpha_6}{\alpha_7}$, we obtain the family of representatives of distinct orbits $\langle\alpha\nb 1+\beta\nb 2+\gamma\nb 4+\nb 5+\nb 7\rangle$. \item $\alpha_5=\alpha_6$ and $\alpha_4\ne 3\alpha_3$. Then choosing $x=\sqrt{\frac {\alpha_4-3\alpha_3}{3\alpha_7}}$, we obtain the family of representatives $\langle\alpha\nb 1+\beta\nb 2+\nb 4+\nb 7\rangle$, where two distinct pairs $(\alpha,\beta)$ and $(\alpha',\beta')$ determine the same orbit if and only if $(\alpha,\beta)=(-\alpha',\beta')$. \item $\alpha_5=\alpha_6$, $\alpha_4=3\alpha_3$ and $\alpha_6^2\ne \alpha_2\alpha_7$. Then choosing $x=\frac 1{\alpha_7}\sqrt{\alpha_2\alpha_7-\alpha_6^2}$, we obtain the family of representatives $\langle\alpha\nb 1+\nb 2+\nb 7\rangle$, where two distinct parameters $\alpha$ and $\alpha'$ determine the same orbit if and only if $\alpha=-\alpha'$. \item $\alpha_5=\alpha_6$, $\alpha_4=3\alpha_3$ and $\alpha_6^2=\alpha_2\alpha_7$. Then we obtain 2 representatives $\langle\nb 7\rangle$ and $\langle\nb 1+\nb 7\rangle$ depending on whether $\alpha_3\alpha_6=\alpha_1\alpha_7$ or not. \end{enumerate} \item $\alpha_7=0$, $\alpha_5\ne 0$ and $\alpha_6\ne 0$. Then $\alpha^*_7=0$, and we put $y=-\frac{\alpha_3x}{\alpha_5}$ and $z=\frac x{\alpha_5\alpha_6}(\alpha_2\alpha_3 - \alpha_1\alpha_5)$ to make $\alpha^*_1=\alpha^*_3=0$. \begin{enumerate} \item $(\alpha_4-2\alpha_3)\alpha_5 \ne \alpha_3\alpha_6$. Then choosing $x=\frac 1{\alpha_5\alpha_6}((\alpha_4-2\alpha_3)\alpha_5 - \alpha_3\alpha_6)$, we obtain the family of representatives of distinct orbits $\langle\alpha\nb 2+\nb 4+\beta\nb 5+\nb 6\rangle_{\beta\ne 0}$. \item $(\alpha_4-2\alpha_3)\alpha_5 = \alpha_3\alpha_6$ and $(\alpha_2 - \alpha_3)\alpha_5 \ne \alpha_3\alpha_6$. Then choosing $x=\frac 1{\alpha_5\alpha_6}((\alpha_2 - \alpha_3)\alpha_5 - \alpha_3\alpha_6)$, we obtain the family of representatives of distinct orbits $\langle\nb 2+\alpha\nb 5+\nb 6\rangle_{\alpha\ne 0}$. \item $(\alpha_4-2\alpha_3)\alpha_5 = \alpha_3\alpha_6=(\alpha_2 - \alpha_3)\alpha_5$. Then we obtain the family of representatives of distinct orbits $\langle\alpha\nb 5+\nb 6\rangle_{\alpha\ne 0}$. \end{enumerate} \item $\alpha_7=\alpha_6=0$ and $\alpha_5\ne 0$. Then $\alpha^*_6=\alpha^*_7=0$, and we put $y=-\frac{\alpha_3x}{\alpha_5}$ to make $\alpha^*_3=0$. \begin{enumerate} \item $\alpha_4\ne 2\alpha_3$. Then choosing $x=\frac {\alpha_4-2\alpha_3}{\alpha_5}$, we obtain the family of representatives of distinct orbits $\langle\alpha\nb 1+\beta\nb 2+\nb 4+\nb 5\rangle$. \item $\alpha_4=2\alpha_3$ and $\alpha_2\ne\alpha_3$. Then choosing $x=\frac {\alpha_2-\alpha_3}{\alpha_5}$, we obtain the family of representatives of distinct orbits $\langle\alpha\nb 1+\nb 2+\nb 5\rangle$. \item $\alpha_4=2\alpha_3$ and $\alpha_2=\alpha_3$. Then we obtain 2 representatives $\langle\nb 5\rangle$ and $\langle\nb 1+\nb 5\rangle$ depending on whether $\alpha_3^2 = \alpha_1\alpha_5$ or not. \end{enumerate} \item $\alpha_7=\alpha_5=0$ and $\alpha_6\ne 0$. Then $\alpha^*_5=\alpha^*_7=0$, and we put $y=-\frac{\alpha_4x}{\alpha_5}$ and $z=\frac x{\alpha_6^2}((\alpha_2 + \alpha_3)\alpha_4 - \alpha_1\alpha_6)$ to make $\alpha^*_1=\alpha^*_4=0$. \begin{enumerate} \item $\alpha_2\ne\alpha_4$. Then choosing $x=\frac {\alpha_2-\alpha_4}{\alpha_6}$, we obtain the family of representatives of distinct orbits $\langle\nb 2+\alpha\nb 3+\nb 6\rangle$. \item $\alpha_2=\alpha_4$. Then we obtain 2 representatives $\langle\nb 6\rangle$ and $\langle\nb 3+\nb 6\rangle$ depending on whether $\alpha_3=0$ or not. We will join $\langle\nb 6\rangle$ with the family $\langle\alpha\nb 5+\nb 6\rangle_{\alpha\ne 0}$ found above. \end{enumerate} \item $\alpha_7=\alpha_6=\alpha_5=0$ and $\alpha_4\ne 0$. Then $\alpha^*_5=\alpha^*_6=\alpha^*_7=0$. \begin{enumerate} \item $\alpha_2\ne-\alpha_3$. Then choosing $y = -\frac{\alpha_1x}{\alpha_2 + \alpha_3}$, we obtain the family of representatives of distinct orbits $\langle\alpha\nb 2+\beta\nb 3+\nb 4\rangle_{\beta\ne-\alpha}$. \item $\alpha_2=-\alpha_3$. Then we obtain two families of representatives of distinct orbits $\langle\alpha\nb 2-\alpha\nb 3+\nb 4\rangle$ and $\langle\nb 1+\alpha\nb 2-\alpha\nb 3+\nb 4\rangle$ depending on whether $\alpha_1=0$ or not. The family $\langle\alpha\nb 2-\alpha\nb 3+\nb 4\rangle$ will be joined with the family $\langle\alpha\nb 2+\beta\nb 3+\nb 4\rangle_{\beta\ne-\alpha}$ from the previous item. \end{enumerate} \end{enumerate} Summarizing, we have the following distinct orbits: \begin{longtable}{lll} $\langle\nb 1+\alpha\nb 2-\alpha\nb 3+\nb 4\rangle$ & $\langle\alpha\nb 1+\beta\nb 2+\nb 4+\nb 5\rangle$ & $\langle\alpha\nb 1+\beta\nb 2+\gamma\nb 4+\nb 5+\nb 7\rangle$ \\ \multicolumn{2}{l}{$\langle\alpha\nb 1+\beta\nb 2+\nb 4+\nb 7\rangle^{O(\alpha,\beta)=O(-\alpha,\beta)}$}& $\langle\alpha\nb 1+\nb 2+\nb 5\rangle$ \\ $\langle\alpha\nb 1+\nb 2+\nb 7\rangle^{O(\alpha)=O(-\alpha)}$ & $\langle\nb 1+\nb 5\rangle$ & $\langle\nb 1+\nb 7\rangle$ \\ $\langle\alpha\nb 2+\beta\nb 3+\nb 4\rangle$ & $\langle\nb 2+\alpha\nb 3+\nb 6\rangle$& $\langle\alpha\nb 2+\nb 4+\beta\nb 5+\nb 6\rangle_{\beta\ne 0}$ \\ $\langle\nb 2+\alpha\nb 5+\nb 6\rangle_{\alpha\ne 0}$ & $\langle\nb 3+\nb 6\rangle$ & $\langle\nb 5\rangle$ \\ $\langle\alpha\nb 5+\nb 6\rangle$& $\langle\nb 7\rangle$ \end{longtable} They correspond to the following new algebras: {\tiny \begin{longtable}{lllllllllllllll} ${\bf N}^4_{26}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$&$e_1e_3=-\alpha e_4$ &$e_2e_1=e_3$ &$e_2e_2=\alpha e_4$&$ e_3e_1=(1+2 \alpha) e_4$ \\ \hline ${\bf N}^4_{27}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_1=e_3$ &$e_2e_2=\beta e_4$&$e_2e_3=e_4$&$e_3e_1=e_4$& \\ \hline ${\bf N}^4_{28}(\alpha,\beta,\gamma)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_1=e_3$&$e_2e_2=\beta e_4$\\&&$e_2e_3=e_4$&$e_3e_1=\gamma e_4$&$e_3e_3=e_4$& \\ \hline ${\bf N}^4_{29}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_2=\beta e_4$ &$e_2e_1=e_3$&$e_3e_1=e_4$&$e_3e_3=e_4$& \\ \hline ${\bf N}^4_{30}(\alpha) $ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_1=e_3$&$e_2e_2=e_4$&$e_2e_3=e_4$ \\ \hline ${\bf N}^4_{31}(\alpha) $ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_1=e_3$&$e_2e_2=e_4$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{32}$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$&$e_2e_1=e_3$&$e_2e_3=e_4$ \\ \hline ${\bf N}^4_{33}$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$&$e_2e_1=e_3$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{34}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=\beta e_4$&$e_2e_1=e_3$&$e_2e_2=\alpha e_4$& \multicolumn{2}{l}{$e_3e_1=(1-2\beta) e_4$} \\ \hline ${\bf N}^4_{35}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=\alpha e_4$&$e_2e_1=e_3$ &$e_2e_2=e_4$&$e_3e_1=-2\alpha e_4$&$e_3e_2=e_4$& \\ \hline ${\bf N}^4_{36}(\alpha,\beta)_{\beta\neq 0}$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_3$&$e_2e_2=\alpha e_4$ &$e_2e_3=\beta e_4$&$e_3e_1=e_4$&$e_3e_2=e_4$& \\ \hline ${\bf N}^4_{37}(\alpha)_{\alpha\neq0}$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_3$&$e_2e_2=e_4$&$e_2e_3=\alpha e_4$&$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{38}$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=e_4$&$e_2e_1=e_3$&$e_3e_1=-2e_4$&$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{39}$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_3$&$e_2e_3=e_4$& \\ \hline ${\bf N}^4_{40}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_3$&$e_2e_3=\alpha e_4$&$e_3e_2=e_4$& \\ \hline ${\bf N}^4_{41}$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_3$&$e_3e_3=e_4$& \\ \hline \end{longtable} } \begin{longtable}{ll} ${\bf N}^4_{29}(\alpha,\beta) \cong {\bf N}^4_{29}(-\alpha,\beta)$ & ${\bf N}^4_{31}(\alpha) \cong {\bf N}^4_{31}(-\alpha)$ \end{longtable} \subsection{$1$-dimensional central extensions of $\cd 3{04}$}\label{sec-CD^3_04} Here we collect all information about ${\mathfrak{CD}_{04}^{3}}:$ \begin{longtable}{|l|l|l|} \hline Algebra & Multiplication & Cohomology \\ \hline ${\mathfrak{CD}}_{04}^{3}$ & $\begin{array}{l} e_1 e_1 = e_2\\ e_1 e_2=e_3\\ e_2 e_1=\lambda e_3 \end{array}$& $\begin{array}{lcl} {\rm H^2_{\mathfrak{CD}}}({\mathfrak{CD}}_{04}^{3})&=&\langle (\lambda-2)\Dt 13-(2\lambda-1)\Dt 31, \Dt 21, \Dt 22 \rangle \\ {\rm H^2}({\mathfrak{CD}}_{04}^{3})&=&{\rm H^2_{\mathfrak{CD}}}({\mathfrak{CD}_{04}^3}) \oplus \langle \Dl 13-2 \Dl 31, \Dl 23, \Dl 32, \Dl 33 \rangle \end{array}$\\ \hline \end{longtable} Let us use the following notations \begin{longtable}{llll} $\nb 1 = (\lambda-2)\Dl 13-(2\lambda-1)\Dl 31$ &$\nb 2 = \Dl 21$ & $\nb 3 = \Dl 22$\\ $\nb 4= \Dl 13-2 \Dl 31$ & $\nabla_5=\Dl 23$ & $\nabla_6=\Dl 32$ & $\nabla_7=\Dl 33$ \end{longtable} Take $\theta=\sum_{i=1}^3\alpha_i\nb i\in {\rm H}^2_{\mathfrak{CD}}(\cd 3{04})$. If $$ \phi= \begin{pmatrix} x & 0 & 0\\ y & x^2 & 0\\ z & (\lambda+1)xy & x^3 \end{pmatrix}\in\aut{\cd 3{04}}, $$ then $$ \phi^T\begin{pmatrix} 0 & 0 & (\lambda-2)\alpha_1+\alpha_4\\ \alpha_2 & \alpha_3 & \alpha_5\\ -(2\lambda-1)\alpha_1-2 \alpha_4 & \alpha_6 & \alpha_7 \end{pmatrix} \phi= \begin{pmatrix} \alpha^* & \alpha^{**} & (\lambda-2)\alpha_1^*+\alpha^*_4\\ \alpha_2^*+\lambda\alpha^{**} & \alpha_3^* & \alpha^*_5\\ -(2\lambda-1)\alpha_1^*-2 \alpha^*_4 & \alpha^*_6 & \alpha^*_7 \end{pmatrix}, $$ where \begin{longtable}{lll} $\alpha^*_1$& $=$& $\frac{x^3}{3} (3 x \alpha_1-y (2 \alpha_5+\alpha_6)-3 z \alpha_7)$\\ $\alpha^*_2$& $=$& $x (x^2 \alpha_2+y (1+\lambda ) (z \alpha_7 (1-\lambda )+y (\alpha_6-\alpha_5 \lambda ))+x (z (\alpha_5- \alpha_6 \lambda) -y (\alpha_3 (\lambda-1 )+$\\ &&$\alpha_1 (\lambda-1 ) (\lambda+1 )^2+\alpha_4 (2+3 \lambda +\lambda ^2))))$\\ $\alpha_3^*$ &$=$&$x^2 (x^2 \alpha_3+x y (\alpha_5+\alpha_6) (1+\lambda )+y^2 \alpha_7 (\lambda+1 )^2)$\\ $\alpha_4^*$ &$=$&$\frac{ x^3}{3} (3 x \alpha_4+y \alpha_6 (\lambda-2 )+3 z \alpha_7 (\lambda-1 )+y \alpha_5 (2 \lambda-1 ))$\\ $\alpha_5^*$ &$=$&$x^4 (x \alpha_5+y \alpha_7 (\lambda+1 ))$\\ $\alpha_6^*$ &$=$&$x^4 (x \alpha_6+y \alpha_7 (\lambda+1 ))$\\ $\alpha_7^*$ &$=$&$x^6 \alpha_7$ \end{longtable} Hence, $\phi\langle\theta\rangle=\langle\theta^*\rangle$, where $\theta^*=\sum\limits_{i=1}^7 \alpha_i^* \nb i.$ We are only interested in elements with $(\alpha_4, \alpha_5, \alpha_6,\alpha_7)\neq (0,0,0,0).$ \begin{enumerate} \item $\alpha_7\ne 0$ and $\lambda\neq -1$. We may assume $\alpha_7=1$. We put $z = \alpha_1x - \frac 13(2\alpha_5 + \alpha_6)y$ to make $\alpha^*_1=0$. \begin{enumerate} \item $\alpha_6\neq \alpha_5$. Then choosing $y=-\frac{3((\lambda - 1)\alpha_1 + \alpha_4)x}{\alpha_5 - \alpha_6}$ we have $\alpha^*_4=0$. Now, if $\alpha_6\ne\frac{3(\lambda+1)((\lambda-1)\alpha_1 + \alpha_4)}{\alpha_5 - \alpha_6}$, then choosing $x=\alpha_6-\frac{3(\lambda+1)((\lambda-1)\alpha_1 + \alpha_4)}{\alpha_5 - \alpha_6}$ we have the family of representatives of distinct orbits $\langle \alpha \nb 2+\beta \nb 3+\gamma \nb 5+ \nb 6+ \nb 7\rangle_{\lambda\ne-1,\gamma\ne 1}$. Otherwise, choosing $x=\alpha_5-\alpha_6$ we have the family of representatives of distinct orbits $\langle \alpha \nb 2+\beta \nb 3+\nb 5+ \nb 7\rangle_{\lambda\ne-1}$. \item $\alpha_6= \alpha_5$ and $\alpha_4 \neq (1 - \lambda )\alpha_1.$ Then choosing $y=-\frac{\alpha_5x}{\lambda + 1}$ and $x=\sqrt{\alpha_4+(\lambda-1)\alpha_1}$, we have the family of representatives $\langle \alpha \nb 2+ \beta \nb 3+ \nb 4+ \nb 7\rangle_{\lambda\ne-1}$, where $(\alpha,\beta)$ and $(\alpha',\beta')$ determine the same orbit if and only if $(\alpha,\beta)=(\pm\alpha',\beta')$. \item $\alpha_6= \alpha_5,$ $\alpha_4 = (1 - \lambda )\alpha_1$ and $\alpha_3\ne \alpha_5^2.$ Then choosing $y=-\frac{\alpha_5x}{\lambda + 1}$ and $x=\sqrt{ \alpha_3-\alpha_5^2}$, we have the family of representatives $\langle \alpha \nb 2+ \nb 3+ \nb 7\rangle_{\lambda\ne-1}$, where $\alpha$ and $\alpha'$ determine the same orbit if and only if $\alpha=\pm\alpha'$. \item $\alpha_6= \alpha_5,$ $\alpha_4=(1-\lambda)\alpha_1$ and $\alpha_3=\alpha_5^2$. Then have two representatives $\langle \nb 7\rangle_{\lambda\ne-1}$ and $\langle \nb 2+ \nb 7\rangle_{\lambda\ne-1}$ depending on whether $\alpha_2=(\lambda -1)\alpha_1 \alpha_5$ or not. \end{enumerate} \item $\alpha_7\ne 0$ and $\lambda=-1$. We may assume $\alpha_7=1$. \begin{enumerate} \item $\alpha_6\neq0$ and $\alpha_6\neq \alpha_5.$ Then choosing $x=\alpha_6,$ $y=\frac{3(2 \alpha_1-\alpha_4) x}{\alpha_5-\alpha_6}$ and $z=\frac{(\alpha_4 (2 \alpha_5+\alpha_6)-3 \alpha_1 (\alpha_5+\alpha_6))x}{\alpha_5-\alpha_6},$ we have the family of representatives of distinct orbits $\langle \alpha \nb 2+\beta \nb 3+\gamma \nb 5+\nb 6+ \nb 7\rangle_{\lambda=-1, \gamma \neq 1}$. \item $\alpha_6\neq0, \alpha_6= \alpha_5$ and $\alpha_3\neq \alpha_5^2$. Then choosing $x=\alpha_5,$ $y=\frac{(\alpha_2+2 \alpha_1 \alpha_5)x}{2 (\alpha_5^2-\alpha_3)}$ and $z=\frac{(\alpha_2 \alpha_5+2 \alpha_1 \alpha_3)x}{2 (\alpha_3-\alpha_5^2)},$ we have the family of representatives of distinct orbits $\langle \alpha \nb 3+ \beta \nb 4+ \nb 5+\nb 6+ \nb 7\rangle_{\lambda=-1,\alpha\ne 1}$. \item $\alpha_6\neq0, \alpha_6= \alpha_5$ and $\alpha_3= \alpha_5^2.$ Then choosing $x=\alpha_6,$ $y=\alpha_1$ and $z=0,$ we have the family of representatives of distinct orbits $\langle \alpha \nb 2+ \nb 3+ \beta \nb 4+ \nb 5+\nb 6+ \nb 7\rangle_{\lambda=-1}$. \item $\alpha_6=0$ and $\alpha_5\neq 0.$ Then choosing $x=\alpha_5,$ $y=3(2\alpha_1-\alpha_4)$ and $z=(2 \alpha_4-3 \alpha_1)\alpha_5,$ we have the family of representatives of distinct orbits $\langle \alpha \nb 2+ \beta \nb 3+ \nb 5 + \nb 7\rangle_{\lambda=-1}$. \item $\alpha_6=0,$ $\alpha_5= 0,$ $\alpha_4\neq 2 \alpha_1$ and $\alpha_3\neq 0.$ Then choosing $x=\sqrt{\alpha_4-2 \alpha_1},$ $y=-\frac{\alpha_2x}{2\alpha_3}$ and $z=\alpha_1x,$ we have the family of representatives of distinct orbits $\langle \alpha \nb 3+ \nb 4+ \nb 7\rangle_{\lambda=-1, \alpha\ne 0}$. \item $\alpha_6=0,$ $\alpha_5= 0,$ $\alpha_4= 2 \alpha_1$ and $\alpha_3\neq 0.$ Then choosing $x=\sqrt{\alpha_3},$ $y=-\frac{\alpha_2x}{2 \alpha_3}$ and $z=\alpha_1x,$ we have the representative $\langle \nb 3+\nb 7\rangle_{\lambda=-1}$. \item $\alpha_6=0,$ $\alpha_5= 0,$ $\alpha_4= 2 \alpha_1$ and $\alpha_3= 0.$ Then choosing $z=\alpha_1x$, we have two representatives $\langle \nb 7\rangle_{\lambda=-1}$ and $\langle \nb 2+ \nb 7\rangle_{\lambda=-1}$, depending on whether $\alpha_2=0$ or not. \end{enumerate} \item $\alpha_7=0$ and $\alpha_6\ne 0.$ We may assume $\alpha_6=1$. \begin{enumerate} \item $ \alpha_5\not\in\{-\frac 12,\lambda\}$. Then the system $\alpha^*_1=\alpha^*_2=0$ has a unique solution in $y$ and $z$. Now, if $\alpha_4\neq -\frac{((2 \lambda-1)\alpha_5+\lambda-2)\alpha_1}{2\alpha_5 + 1}$, then choosing the suitable value of $x$ we have the family of representatives distinct orbits $\langle \alpha \nb 3+ \nb 4 +\beta \nabla_5+ \nb 6\rangle_{\beta\not\in\{-\frac 12,\lambda\}}$. Otherwise, we have two families of representatives distinct orbits $\langle \alpha \nabla_5+ \nb 6\rangle_{\alpha\not\in\{-\frac 12,\lambda\}}$ and $\langle \nb 3+ \alpha \nabla_5+\nb 6\rangle_{\alpha\not\in\{-\frac 12,\lambda\}}$ depending on whether $\alpha_3=-\frac{3(\lambda + 1)(\alpha_5 + 1)\alpha_1}{2\alpha_5 + 1}$ or not. \item $\alpha_5= \lambda\ne-\frac 12$ and $\alpha_4 \neq \frac{2 (1-\lambda ^2)\alpha_1}{2\lambda+1}$. Then choosing $x=\alpha_4+\frac{2 (\lambda^2-1)\alpha_1}{2 \lambda+1}$ and $y=\frac{3 \alpha_1 x}{2 \lambda+1}$ we have the family of representatives of distinct orbits $\langle \alpha \nb 2+\beta \nb 3+ \nb 4+\lambda \nabla_5+\nb 6\rangle_{\lambda\ne -\frac 12}$. \item $\alpha_5= \lambda\ne-\frac 12$ and $\alpha_4 = \frac{2 (1-\lambda ^2)\alpha_1}{2\lambda+1}$. We put $y=\frac{3 \alpha_1 x}{2 \lambda+1}$ to make $\alpha^*_1=0$. Now, if $\alpha_3\neq -\frac{3 (\lambda+1)^2\alpha_1}{2 \lambda + 1}$, then choosing $x=\alpha_3+\frac{3(\lambda+1)^2\alpha_1}{2 \lambda+1}$, we have the family of representatives of distinct orbits $\langle \alpha \nb 2+ \nb 3 +\lambda \nabla_5+ \nb 6\rangle_{\lambda\ne -\frac 12}$. Otherwise, we have two representatives $\langle \lambda \nabla_5+\nb 6\rangle_{\lambda\ne -\frac 12}$ and $\langle \nb 2+ \lambda \nabla_5+\nb 6\rangle_{\lambda\ne -\frac 12}$, depending on whether $\alpha_2=-\frac{9(\lambda - 1)(\lambda + 1)^2\alpha_1^2}{(2\lambda + 1)^2}$ or not. \item $\alpha_5=-\frac 12 \ne\lambda$. Then put $y=2 \alpha_4 x$ and $z=\frac{2(\alpha_2-2 \alpha_4 (\lambda-1) (\alpha_3+\alpha_1 (\lambda+1)^2)x}{2 \lambda+1}$ to make $\alpha^*_2=\alpha^*_4=0$. Now, if $\alpha_3\ne -(\lambda+1)\alpha_4$, then choosing $x=\alpha_3+(\lambda+1)\alpha_4,$ we have the family of representatives of distinct orbits $\langle \alpha \nb 1+ \nb 3 -\frac 12\nb 5+ \nabla_6\rangle_{\lambda\ne -\frac 12}$. Otherwise, we have two representatives $\langle -\frac 12\nb 5+ \nabla_6\rangle_{\lambda\ne -\frac 12}$ and $\langle \nb 1 -\frac 12\nb 5+ \nabla_6\rangle_{\lambda\ne -\frac 12}$, depending on whether $\alpha_1=0$ or not. \item $\alpha_5=-\frac 12=\lambda$ and $\alpha_4\ne-2 \alpha_3$. Then choosing $x=\alpha_3+\frac{\alpha_4}2$ and $y=2\alpha_4x$ we have the family of representatives of distinct orbits $\langle \alpha\nabla_1 +\beta \nabla_2+\nabla_3 -\frac 12\nb 5+ \nabla_6\rangle_{\lambda= -\frac 12}$. \item $\alpha_5=-\frac 12=\lambda$ and $\alpha_4=-2 \alpha_3$. Then we put $y=-4\alpha_3 x$ to make $\alpha^*_4=0$. Now, if $\alpha_2\neq\frac {3\alpha_3}2 (\alpha_1+4 \alpha_3)$, then choosing $x=\sqrt{\alpha_2-\frac{3\alpha_3}2 (\alpha_1+4 \alpha_3)}$, we have the family of representatives of distinct orbits $\langle \alpha\nabla_1 + \nabla_2 -\frac 12\nb 5+ \nabla_6\rangle_{\lambda= -\frac 12}$. Otherwise, we have two representatives $\langle \nabla_1 -\frac 12\nb 5+ \nabla_6\rangle_{\lambda= -\frac 12}$ and $\langle -\frac 12\nb 5+ \nabla_6\rangle_{\lambda= -\frac 12}$, depending on whether $\alpha_1=0$ or not. \end{enumerate} \item $\alpha_7=\alpha_6=0$ and $\alpha_5\ne 0$. We may assume $\alpha_5=1$. Then the system $\alpha^*_1=\alpha^*_2=0$ has a unique solution in $y$ and $z$. Now, if $2 \alpha_4+(2 \lambda-1)\alpha_1\neq 0$, then choosing $x=\alpha_4+(\lambda-\frac 12)\alpha_1$ we have the family of representatives of distinct orbits $\langle \alpha \nabla_3+ \nabla_4+\nb 5\rangle$. Otherwise, we have two representatives $\langle \nb 5\rangle$ and $\langle \nabla_3+ \nb 5\rangle$ depending on whether $2 \alpha_3+3(\lambda+1) \alpha_1=0$ or not. \item $\alpha_7=\alpha_6=\alpha_5=0$ and $\alpha_4\ne 0$. We may assume $\alpha_4=1$. \begin{enumerate} \item $(\lambda-1) ((\lambda+1)^2\alpha_1+\alpha_3)+(\lambda + 1)(\lambda + 2)\neq 0.$ Then choosing $y=\frac{\alpha_2x}{(\lambda-1) ((\lambda+1)^2\alpha_1+\alpha_3)+(\lambda + 1)(\lambda + 2)}$ we have the family of representatives of distinct orbits $\langle \alpha \nabla_1+ \beta \nabla_3+\nb 4\rangle$, where $(\lambda-1) ((\lambda+1)^2\alpha+\beta)+(\lambda + 1)(\lambda + 2)\neq 0$. \item $(\lambda-1) ((\lambda+1)^2\alpha_1+\alpha_3)+(\lambda + 1)(\lambda + 2)= 0$. Then $\lambda\neq 1$ and we have two families of representatives of distinct orbits $\Big\langle \alpha \nabla_1+ \frac{(\lambda+1) ((\lambda^2-1)\alpha + \lambda + 2)}{1-\lambda} \nabla_3+\nb 4\Big\rangle_{\lambda\ne 1}$ and $\Big\langle \alpha \nabla_1+\nabla_2+ \frac{(\lambda+1) ((\lambda^2-1)\alpha + \lambda + 2)}{1-\lambda} \nabla_3+\nb 4\Big\rangle_{\lambda\ne 1}$, depending on whether $\alpha_2=0$ or not. The first family will be joined with the family from the previous item. \end{enumerate} \end{enumerate} Summarizing, we have the following representatives of distinct orbits: \begin{longtable}{lllllll} $\Big\langle \alpha \nabla_1+\nabla_2+ \frac{(\lambda+1) ((\lambda^2-1)\alpha + \lambda + 2)}{1-\lambda} \nabla_3+\nb 4\Big\rangle_{\lambda\ne 1}$ & $\langle \alpha\nabla_1 +\beta \nabla_2+\nabla_3 -\frac 12\nb 5+ \nabla_6\rangle_{\lambda= -\frac 12}$\\ $\langle \alpha\nabla_1 + \nabla_2 -\frac 12\nb 5+ \nabla_6\rangle_{\lambda= -\frac 12}$& $\langle \alpha \nabla_1+ \beta \nabla_3+\nb 4\rangle$\\ $\langle \alpha \nb 1+ \nb 3 -\frac 12\nb 5+ \nabla_6\rangle_{\lambda\ne -\frac 12}$& $\langle \nabla_1 -\frac 12\nb 5+ \nabla_6\rangle_{\lambda= -\frac 12}$\\ $\langle \alpha \nb 2+\beta \nb 3+ \nb 4+\lambda \nabla_5+\nb 6\rangle_{\lambda\ne -\frac 12}$& $\langle \alpha \nb 2+ \nb 3+ \beta \nb 4+ \nb 5+\nb 6+ \nb 7\rangle_{\lambda=-1}$\\ $\langle \alpha \nb 2+ \beta \nb 3+ \nb 4+ \nb 7\rangle_{\lambda\ne-1}^{O(\alpha,\beta)=O(-\alpha,\beta)}$& $\langle \alpha \nb 2+ \nb 3 +\lambda \nabla_5+ \nb 6\rangle_{\lambda\ne -\frac 12}$\\ $\langle \alpha \nb 2+\beta \nb 3+\gamma \nb 5+ \nb 6+ \nb 7\rangle_{\gamma\ne 1}$& $\langle \alpha \nb 2+\beta \nb 3+\nb 5+ \nb 7\rangle$\\ $\langle \alpha \nb 2+ \nb 3+ \nb 7\rangle_{\lambda\ne-1}^{O(\alpha)=O(-\alpha)}$& $\langle \nb 2+ \lambda \nabla_5+\nb 6\rangle_{\lambda\ne -\frac 12}$\\ $\langle \nb 2+ \nb 7\rangle$& $\langle \alpha \nabla_3+ \nabla_4+\nb 5\rangle$\\ $\langle \alpha \nb 3+ \nb 4 +\beta \nabla_5+ \nb 6\rangle_{\beta\not\in\{-\frac 12,\lambda\}}$& $\langle \alpha \nb 3+ \beta \nb 4+ \nb 5+\nb 6+ \nb 7\rangle_{\lambda=-1,\alpha\ne 1}$\\ $\langle \alpha \nb 3+ \nb 4+ \nb 7\rangle_{\lambda=-1, \alpha\ne 0}$& $\langle \nabla_3+ \nb 5\rangle$\\ $\langle \nb 3+ \alpha \nabla_5+\nb 6\rangle_{\alpha\not\in\{-\frac 12,\lambda\}}$& $\langle \nb 3+\nb 7\rangle_{\lambda=-1}$\\ $\langle \nb 5\rangle$& $\langle \alpha \nabla_5+ \nb 6\rangle$\\ $\langle \nb 7\rangle$ & \end{longtable} The corresponding algebras are: {\tiny \begin{longtable}{lllllllllllllll} ${\bf N}^4_{42}(\lambda,\alpha)_{\lambda\ne 1}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ && $e_1e_3=(\alpha(\lambda-2)+1)e_4$\\ && $e_2 e_1=\lambda e_3+e_4$& \multicolumn{2}{l}{$e_2e_2=\frac{(\lambda+1) ((\lambda^2-1)\alpha + \lambda + 2)}{1-\lambda}e_4$} & $e_3e_1=(\alpha(1-2\lambda)-2)e_4$ \\ \hline ${\bf N}^4_{43}(\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=-\frac {5\alpha}2e_4$ & $e_2 e_1=-\frac 12e_3+\beta e_4$\\ && $e_2e_2=e_4$ & $e_2 e_3=-\frac 12 e_4$ & $e_3e_1=2\alpha e_4$ & $e_3e_2 = e_4$ \\ \hline ${\bf N}^4_{44}(\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=-\frac {5\alpha}2e_4$ & $e_2 e_1=-\frac 12e_3+e_4$\\ && $e_2 e_3=-\frac 12 e_4$ & $e_3e_1=2\alpha e_4$ & $e_3e_2 = e_4$ \\ \hline ${\bf N}^4_{45}(\lambda,\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & \multicolumn{2}{l}{$e_1e_3=(\alpha(\lambda-2)+1)e_4$}\\ && $e_2 e_1=\lambda e_3$& $e_2e_2=\beta e_4$ & \multicolumn{2}{l}{$e_3e_1=(\alpha(1-2\lambda)-2)e_4$} \\ \hline ${\bf N}^4_{46}(\lambda,\alpha)_{\lambda\ne-\frac 12}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=\alpha(\lambda-2)e_4$ & $e_2 e_1=\lambda e_3$\\ && $e_2e_2=e_4$ & $e_2 e_3=-\frac 12 e_4$ & $e_3e_1=\alpha(1-2\lambda)e_4$ & $e_3e_2 = e_4$ \\ \hline ${\bf N}^4_{47}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=-\frac 52e_4$ & $e_2 e_1=-\frac 12e_3$\\ && $e_2 e_3=-\frac 12 e_4$ & $e_3e_1=2 e_4$ & $e_3e_2 = e_4$ \\ \hline ${\bf N}^4_{48}(\lambda,\alpha,\beta)_{\lambda\ne -\frac 12}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=e_4$ & $e_2 e_1=\lambda e_3+\alpha e_4$\\ && $e_2e_2=\beta e_4$ & $e_2 e_3=\lambda e_4$ & $e_3e_1=-2e_4$ & $e_3e_2 = e_4$ \\ \hline ${\bf N}^4_{49}(\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=\beta e_4$\\ && $e_2 e_1=-e_3+\alpha e_4$ & $e_2e_2=e_4$ & $e_2 e_3=e_4$\\ && $e_3e_1=-2\beta e_4$ & $e_3e_2 = e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{50}(\lambda,\alpha,\beta)_{\lambda\ne-1}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=e_4$ & $e_2 e_1=\lambda e_3+\alpha e_4$\\ && $e_2e_2=\beta e_4$ & $e_3e_1=-2e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{51}(\lambda,\alpha)_{\lambda\ne -\frac 12}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+\alpha e_4$\\ && $e_2e_2=e_4$ & $e_2e_3=\lambda e_4$ & $e_3e_2=e_4$ \\ \hline ${\bf N}^4_{52}(\lambda,\alpha,\beta,\gamma)_{\gamma\ne 1}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+\alpha e_4$ & $e_2e_2=\beta e_4$\\ && $e_2e_3=\gamma e_4$ & $e_3e_2=e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{53}(\lambda,\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+\alpha e_4$\\ && $e_2e_2=\beta e_4$ & $e_2e_3=e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{54}(\lambda,\alpha)_{\lambda\ne -1}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+\alpha e_4$\\ && $e_2e_2=e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{55}(\lambda)_{\lambda\ne -\frac 12}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+e_4$\\ && $e_2e_3=\lambda e_4$ & $e_3e_2=e_4$ \\ \hline ${\bf N}^4_{56}(\lambda)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{57}(\lambda,\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=e_4$ & $e_2 e_1=\lambda e_3$\\ && $e_2e_2=\alpha e_4$ & $e_2e_3=e_4$ & $e_3e_1=-2e_4$ \\ \hline ${\bf N}^4_{58}(\lambda,\alpha,\beta)_{\beta\not\in\{-\frac 12,\lambda\}}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=e_4$ & $e_2 e_1=\lambda e_3$\\ && $e_2e_2=\alpha e_4$ & $e_2e_3=\beta e_4$ & $e_3e_1=-2e_4$ & $e_3e_2=e_4$ \\ \hline ${\bf N}^4_{59}(\alpha,\beta)_{\alpha\ne 1}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=\beta e_4$\\ && $e_2 e_1=-e_3$ & $e_2e_2=\alpha e_4$ & $e_2e_3=e_4$\\ && $e_3e_1=-2\beta e_4$ & $e_3e_2=e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{60}(\alpha)_{\alpha\ne 0}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=e_4$ & $e_2 e_1=-e_3$\\ && $e_2e_2=\alpha e_4$ & $e_3e_1=-2e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{61}(\lambda)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$\\ && $e_2e_2=e_4$ & $e_2e_3=e_4$ \\ \hline ${\bf N}^4_{62}(\lambda,\alpha)_{\alpha\not\in\{-\frac 12,\lambda\}}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$\\ && $e_2e_2=e_4$ & $e_2e_3=\alpha e_4$ & $e_3e_2=e_4$ \\ \hline ${\bf N}^4_{63}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=-e_3$\\ && $e_2e_2=e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{64}(\lambda)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$ & $e_2e_3=e_4$ \\ \hline ${\bf N}^4_{65}(\lambda,\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$\\ && $e_2e_3=\alpha e_4$ & $e_3e_2=e_4$ \\ \hline ${\bf N}^4_{66}(\lambda)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$ & $e_3e_3=e_4$\\ \hline \end{longtable}}\begin{longtable}{ll} ${\bf N}^4_{50}(\lambda,\alpha,\beta) \cong {\bf N}^4_{50}(\lambda,-\alpha,\beta)$ & ${\bf N}^4_{54}(\lambda,\alpha) \cong {\bf N}^4_{54}(\lambda,-\alpha)$ \end{longtable} \section{The classification theorem}\label{S:class} \begin{theorem} \label{teo-alg} Let ${\bf N}$ be a complex $4$-dimensional nilpotent algebra. Then ${\bf N}$ is isomorphic to an algebra from the following list: {\tiny \begin{longtable}{lllll llll} \hline$\cd {4*}{01}$&$:$& $e_1 e_1 = e_2$\\ \hline$\cd {4*}{02}$&$:$& $e_1 e_1 = e_3$ &$ e_2 e_2=e_3$ \\ \hline$\cd {4*}{03}$&$:$& $e_1 e_2=e_3$ & $e_2 e_1=-e_3$ \\ \hline$\cd {4*}{04}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_2 e_1=e_3$ & $e_2 e_2=e_3$\\ \hline$\cd {4*}{05}$ & $:$ & $e_1e_1 = e_3$&$ e_2e_2=e_4$ \\ \hline$\cd {4*}{06}$ & $:$ & $e_1e_2 = e_4$&$ e_3e_1 = e_4$ \\ \hline$\cd {4*}{07}$ & $:$ & $e_1e_2 = e_3$&$ e_2e_1 = e_4$&$ e_2e_2 = -e_3$\\ \hline$\cd {4*}{08}(\alpha)$ & $:$&$ e_1e_1 = e_3$& $e_1e_2 = e_4$& $e_2e_1 = -\alpha e_3$&$e_2e_2 = -e_4$ \\ \hline$\cd {4*}{09}(\alpha)$&$:$&$e_1e_1 = e_4$&$ e_1e_2 = \alpha e_4$&$ e_2e_1 = -\alpha e_4$\\&&$ e_2e_2 = e_4$&$ e_3e_3 = e_4$\\ \hline$\cd {4*}{10}$&$:$&$ e_1e_2 = e_4$&$ e_1e_3 = e_4$&$ e_2e_1 = -e_4$\\&&$ e_2e_2 = e_4$&$ e_3e_1 = e_4$\\ \hline$\cd {4*}{11}$&$:$&$ e_1e_1 = e_4$&$ e_1e_2 = e_4$&$ e_2e_1 = -e_4$&$ e_3e_3 = e_4$\\ \hline$\cd {4*}{12}$&$:$&$ e_1e_2 = e_3$&$ e_2e_1 = e_4$ \\ \hline$\cd {4*}{13}$&$:$&$ e_1e_1 = e_4$&$ e_1e_2 = e_3$&$ e_2e_1 = -e_3$&$ e_2e_2=2e_3+e_4$\\ \hline$\cd {4*}{14}(\alpha)$&$:$&$ e_1e_2 = e_4$&$ e_2e_1 =\alpha e_4$&$ e_2e_2 = e_3$\\ \hline$\cd {4*}{15}$&$:$&$e_1e_2 = e_4$&$ e_2e_1 = -e_4$&$ e_3e_3 = e_4$\\ \hline $\D{4}{00} $&$:$& $e_1e_1=e_4$ & $e_1e_2=e_4$& $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$ & $e_2e_3=e_4$ \\\hline $\D{4}{01}(\lambda,\alpha,\beta)$&$:$& $e_1 e_1 = \lambda e_3 + e_4$ & $e_1 e_3 = \alpha e_4$ & $e_2 e_1=e_3$ \\ &&$ e_2 e_2 = e_3$ & $e_2 e_3 = \beta e_4$ & $e_3e_1 = e_4$ \\ \hline $\D{4}{02}(\lambda,\alpha,\beta)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1 e_3 = \alpha e_4$ & $e_2 e_1=e_3$ \\ && $e_2 e_2 = e_3$ & $e_2 e_3 = \beta e_4$ & $e_3e_1 = e_4$ \\ \hline $\D{4}{03}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1 e_2 = e_4$ & $e_2 e_1=e_3$ \\ && $e_2 e_2 = e_3$ & $e_2 e_3 = \alpha e_4$ & $e_3e_1 = e_4$ \\ \hline $\D{4}{04}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_2 e_1=e_3$ & $e_2 e_2 = e_3$ \\ && $e_2 e_3 = \alpha e_4$ & $e_3e_1 = e_4$ \\ \hline $\D{4}{05}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1 e_2 = \lambda e_4$ & $e_2 e_1=e_3 + \lambda\alpha e_4$ \\ && $e_2 e_2 = e_3$ & $e_2 e_3 = \Theta e_4$ & $e_3e_1 = \lambda e_4$ \\ \hline $\D{4}{06}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1 e_2 = e_4$ & $e_2 e_1=e_3 + \alpha e_4$ \\ && $e_2 e_2 = e_3$ & $e_2 e_3 = \Theta^{-1} e_4$ & $e_3e_1 = e_4$ \\ \hline $\D{4}{07}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_2 e_1=e_3 + \lambda e_4$ & $e_2 e_2 = e_3$ \\&& $e_2 e_3 = \Theta e_4$ & $e_3e_1 = \lambda e_4$ \\ \hline $\D{4}{08}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_2 e_1=e_3 + e_4$ & $e_2 e_2 = e_3$ \\&& $e_2 e_3 = \Theta^{-1} e_4$ & $e_3e_1 = e_4$ \\ \hline $\D{4}{09}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1 e_2 = e_4$& $e_1 e_3 = \alpha e_4$ \\ && $e_2 e_1=e_3$ & $e_2 e_2 = e_3 $& $e_3e_1 = e_4$ \\ \hline $\D{4}{10}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$& $e_1 e_3 = \alpha e_4$ & $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$ & $e_3e_1 = e_4$ \\ \hline $\D{4}{11}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3 + e_4$& $e_1e_2 = \alpha e_4$ & $e_1 e_3 = -e_4$ \\ && $e_2 e_1=e_3$ & $e_2 e_2 = e_3$ & $e_3e_1 = e_4$ \\ \hline $\D{4}{12}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3 + e_4$& $e_1e_3 = \alpha e_4$ & $e_2 e_1=e_3$ \\ && $e_2 e_2 = e_3$ & $e_3e_1 = \Theta e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{13}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3 + e_4$& $e_1e_3 = \alpha e_4$ & $e_2 e_1=e_3$ \\ && $e_2 e_2 = e_3$ & $e_3e_1 = (1-\Theta)e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{14}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$& $e_1e_3 = \alpha e_4 $ & $e_2 e_1=e_3$ \\ && $e_2 e_2 = e_3$ & $e_3e_1 = \Theta e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{15}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$& $e_1e_3 = \alpha e_4$ & $e_2 e_1=e_3$ \\ && $e_2 e_2 = e_3$ & $e_3e_1 = (1-\Theta)e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{16}(\alpha)$&$:$& $e_1e_3 = \alpha e_4$ & $e_2 e_1=e_3 + e_4$ & $e_2 e_2 = e_3$ \\&& $e_3e_1 = e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{17}(\alpha)$&$:$& $e_1 e_1 = e_4$ & $e_1e_3 = -e_4$ & $e_2 e_1=e_3 + \alpha e_4$ \\ && $e_2 e_2 = e_3$ & $e_3e_1 = e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{18}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3 + e_4$& $e_1e_3 = \Theta\alpha e_4$ & $e_2 e_1=e_3$ & $e_2 e_2 = e_3$ \\ && $e_2 e_3 = \alpha e_4$ & $e_3e_1 = \Theta e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{19}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3 + e_4$& $e_1e_3 = (1-\Theta)\alpha e_4$ & $e_2 e_1=e_3$ & $e_2 e_2 = e_3$ \\&& $e_2 e_3 = \alpha e_4$ & $e_3e_1 = (1-\Theta)e_4$& $e_3e_2 = e_4$ \\ \hline $\D{4}{20}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$& $e_1e_3 = \Theta\alpha e_4$ & $e_2 e_1=e_3$ & $e_2 e_2 = e_3$ \\ && $e_2 e_3 = \alpha e_4$ & $e_3e_1 = \Theta e_4$& $e_3e_2 = e_4$ \\ \hline $\D{4}{21}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$& $e_1e_3 = (1-\Theta)\alpha e_4$ & $e_2 e_1=e_3$ & $e_2 e_2 = e_3$ \\&& $e_2 e_3 = \alpha e_4$ & $e_3e_1 = (1-\Theta)e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{22}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + (1-2\lambda)e_4$& $e_1 e_2 = e_4$& $e_1e_3 = (\Theta - 1) e_4$ & $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$ & $e_2 e_3 = (1-\Theta^{-1}) e_4$& $e_3e_1 = \Theta e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{23}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + \lambda(1-2\lambda) e_4$& $e_1 e_2 = \lambda e_4$& $e_1e_3 = -\lambda\Theta e_4 $& $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$ & $e_2 e_3 = -\Theta^2 e_4$ & $e_3e_1 = \lambda(1-\Theta)e_4$ & $e_3e_2 = \lambda e_4$ \\ \hline $\D{4}{24}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + \Theta e_4$& $e_1 e_2 = e_4$& $e_1e_3 = (\Theta - 1) e_4$ & $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$ & $e_2 e_3 = (1-\Theta^{-1}) e_4$ & $e_3e_1 = \Theta e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{25}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + \lambda(1-\Theta)e_4$& $e_1 e_2 = \lambda e_4$& $e_1e_3 = -\lambda\Theta e_4$ & $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$ & $e_2 e_3 = -\Theta^2 e_4$ & $e_3e_1 = \lambda (1-\Theta)e_4$ & $e_3e_2 = \lambda e_4$ \\ \hline $\D{4}{26}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + \Theta e_4$& $e_1 e_2 = e_4$& $e_1e_3 = -\Theta e_4$ & $e_2 e_1=e_3$\\ & & $e_2 e_2 = e_3$ & $e_2 e_3 = -e_4$ & $e_3e_1 = \Theta e_4$ & $e_3e_2 = e_4 $\\ \hline $\D{4}{27}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + (1-\Theta)e_4$& $e_1 e_2 = e_4$& $e_1e_3 = (\Theta-1) e_4$ & $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$ & $e_2 e_3 = -e_4$ & $e_3e_1 = (1-\Theta)e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{28}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + (1-\Theta)e_4$& $e_1 e_2 = e_4$& $e_1e_3 = -\Theta e_4$ & $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$ & $e_2 e_3 = -e_4$ & $e_3e_1 = \Theta e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{29}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + \Theta e_4$& $e_1 e_2 = e_4$& $e_1e_3 = (\Theta-1) e_4$ & $e_2 e_1=e_3$ \\ && $e_2 e_2 = e_3$ & $e_2 e_3 = -e_4$ & $e_3e_1 = (1-\Theta)e_4$ & $e_3e_2 = e_4$ \\ $\D{4}{30}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + e_4$ & $e_1e_3 = (\Theta-1) e_4$ & $e_2 e_1=e_3$ & $e_2 e_2 = e_3$ \\ & & $e_2 e_3 = -e_4$ & $e_3e_1 = \Theta e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{31}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + e_4$ & $e_1e_3 = -\Theta e_4$ & $e_2 e_1=e_3$ & $e_2 e_2 = e_3$ \\ && $e_2 e_3 = -e_4$ & $e_3e_1 = (1-\Theta)e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{32}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_3 = (\Theta-1) e_4$ & $e_2 e_1=e_3$ & $e_2 e_2 = e_3$ \\ && $e_2 e_3 = -e_4$ & $e_3e_1 = \Theta e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{33}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_3 = -\Theta e_4$ & $ e_2 e_1=e_3$ & $e_2 e_2 = e_3$ \\ && $e_2 e_3 = -e_4$ & $e_3e_1 = (1-\Theta)e_4$ & $e_3e_2 = e_4$ \\ \hline $\D{4}{34}$&$:$& $e_1 e_1 = e_4$ & $e_2 e_1=e_3 + e_4$ & $e_2 e_2 = e_3$\\& & $e_3 e_1 = e_4$& $e_3 e_2 = e_4$ \\ \hline $\D{4}{35}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_3 = e_4$ & $e_2 e_1=e_3 + e_4$ & $e_2 e_2 = e_3$ \\ \hline $\D{4}{36}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_3 = e_4$ & $e_2 e_1=e_3$ & $e_2 e_2 = e_3$ \\ \hline $\D{4}{37}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + e_4$ & $e_1e_3 = \Theta e_4$ & $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$& $e_2 e_3 = e_4$ \\ \hline $\D{4}{38}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3 + e_4$ & $e_1e_3 = (1-\Theta)e_4$ & $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$& $e_2 e_3 = e_4$ \\ \hline $\D{4}{39}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_3 = \Theta e_4$ & $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$& $e_2 e_3 = e_4$ \\ \hline $\D{4}{40}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_3 = (1-\Theta)e_4$ & $e_2 e_1=e_3$ \\&& $e_2 e_2 = e_3$& $e_2 e_3 = e_4$\\ \hline $\cd 4{01}$&$:$& $e_1 e_1 = e_2$ & $e_2 e_2=e_3$ \\ \hline $\cd 4{02}$&$:$& $e_1 e_1 = e_2$ & $e_2 e_1= e_3$ & $e_2 e_2=e_3$ \\ \hline$\cd 4{03}$&$:$& $e_1 e_1 = e_2$ & $e_2 e_1=e_3$ \\ \hline$\cd 4{04}(\lambda)$&$:$& $ e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$ \\ \hline $\cd 4{05}$ & $:$ & $e_1 e_1 = e_2$ & $e_2 e_1=e_4$ & $e_2 e_2=e_3$ \\ \hline$\cd 4{06}$ & $:$ & $ e_1 e_1 = e_2$ & $e_1 e_2=e_4$ & $e_2 e_1=e_3$ \\ \hline$\cd 4{07}(\lambda)$& $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_4$ & $e_2 e_1=\lambda e_4$ & $e_2 e_2=e_3$ \\\hline $\cd 4{08}(\alpha)$ &$:$& $e_1 e_1 = e_2$ & $e_1e_3=e_4$ & $e_2 e_1=e_3$ \\&& $e_2e_2=\alpha e_4$ & $e_3e_1=-2e_4$ \\ \hline$\cd 4{09}$ &$:$& $e_1 e_1 = e_2$ & $e_1e_2=e_4$ & $e_1e_3=e_4$ \\ && $e_2 e_1=e_3$ & $e_2e_2=- e_4$ & $e_3e_1=-2e_4$\\ \hline $\cd 4{10}(\alpha)$ &$:$& $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=-e_4$ \\ && $e_2 e_1= e_3 +e_4$ & $e_2e_2=\alpha e_4$ & $e_3e_1=- e_4$ \\ \hline$\cd 4{11}(\lambda)$ &$:$& $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=(\lambda-2)e_4$ \\ $\lambda \neq1$&& $e_2 e_1= \lambda e_3+e_4$ & $e_2e_2=-(\lambda+1)^2 e_4$ &$e_3e_1= (1-2\lambda) e_4$ \\ \hline$\cd 4{12}(\alpha, \lambda)$ &$:$& $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=(\lambda-2)e_4$ \\&& $e_2 e_1=\lambda e_3$ & $e_2e_2=\alpha e_4$ & $e_3e_1=(1-2\lambda) e_4$ \\ \hline $\cd {4}{13}(\alpha)$&$:$& $e_1 e_1 = e_2$ & $e_1e_2=e_4$ & $e_1e_3=e_4$& $e_2e_1=e_4$ \\ $\alpha\neq \frac{1}{2}$&& $e_2e_3=\alpha e_4$& $e_3e_1=e_4$& \multicolumn{2}{l}{$e_3e_2=(\alpha +1)e_4$}\\ \hline $\cd {4}{14}(\alpha, \beta)$&$:$& $e_1 e_1 = e_2$ & $e_1 e_2 = e_4$& $e_1 e_3 = \alpha e_4$& $e_2 e_1 = e_4$\\ && $e_2 e_2 = e_4$& $e_3 e_1 = \alpha e_4$& $e_3 e_2 = e_4$& $e_3 e_3 =\beta e_4$& \\ \hline $\cd {4}{15}(\alpha)$&$:$& $e_1 e_1 = e_2$ & $e_1 e_2 = \alpha e_4$& $e_1 e_3 = e_4$\\&& $e_2 e_1 =(\alpha+1) e_4$ & $e_3 e_1 = e_4$\\ \hline $\cd {4}{16}$&$:$& $e_1 e_1 = e_2$ & $e_1e_2=e_4$ & $e_2 e_1 = e_4$\\ && $e_2 e_3 = -\frac{1}{2} e_4$& $e_3 e_2 =\frac{1}{2} e_4$& $e_3 e_3 = e_4$&\\ \hline $\cd {4}{17}(\alpha)$&$:$& $e_1 e_1 = e_2$ & $e_1 e_2 = e_4$& $e_2 e_1 = e_4$\\&& $e_2 e_3 =\alpha e_4$& $e_3 e_2 = (\alpha+1) e_4$&\\ \hline $\cd {4}{18}(\alpha)$&$:$& $e_1 e_1 = e_2$ & $e_1 e_2 =\alpha e_4$& \multicolumn{1}{l}{$e_2 e_1 =(\alpha+1) e_4$}& $e_3 e_3 = e_4$&\\ \hline $\cd {4}{19}$&$:$& $e_1 e_1 = e_2$ & $e_1 e_2 = e_4$& $e_2 e_1 = e_4$\\&& $e_3 e_1 = e_4$& $e_3 e_3 = e_4$&\\ \hline $\cd {4}{20}$&$:$& $e_1 e_1 = e_2$ & $e_1 e_2 = e_4$& $e_2 e_1 = e_4$& $e_3 e_1 = e_4$&\\ \hline $\cd {4}{21}(\alpha)$&$:$& $e_1 e_1 = e_2$ & $e_1 e_3 =\alpha e_4$& $e_2 e_1 = e_4$& $e_2 e_2 = e_4$&\\ && $e_2 e_3 = e_4$& $e_3 e_1 =\alpha e_4$& $e_3 e_2 = e_4$& $e_3 e_3 = e_4$\\ \hline $\cd {4}{22}$&$:$& $e_1 e_1 = e_2$ & $e_1 e_3 = e_4$& $e_2 e_3 = -\frac{1}{2} e_4$\\&& $e_3 e_1= e_4$& $e_3 e_2 =\frac{1}{2} e_4$& $e_3 e_3 = e_4$&\\ \hline $\cd {4}{23}(\alpha)$&$:$& $e_1 e_1 = e_2$& $e_1 e_3 = e_4$& $e_2 e_3 = \alpha e_4$\\&& $e_3 e_1 = e_4$& $e_3 e_2 =(\alpha +1)e_4$&\\ \hline $\cd {4}{24}(\alpha)$&$:$& $e_1 e_1 = e_2$& $e_1 e_3 = e_4$& $e_2 e_2 = e_4$\\ && $e_3 e_1 = e_4$& $e_3 e_2 = e_4$& $e_3 e_3 =\alpha e_4$&\\ \hline $\cd {4}{25} $&$:$& $e_1 e_1 = e_2$& $e_1 e_3 = e_4$& $e_2 e_1 = e_4$\\&& $e_3 e_1 = e_4$& $e_2 e_2 = e_4$&\\ \hline $\cd {4}{26}(\alpha)$&$:$& $e_1 e_1 = e_2$& $e_1 e_3 = \alpha e_4$& $e_2 e_2 = e_4$& \multicolumn{2}{l}{$e_3 e_1 =(\alpha+1) e_4$}&\\ \hline $\cd {4}{27}$&$:$& $e_1 e_1 = e_2$& $e_2 e_1 = e_4$& $e_2 e_3 = e_4$\\&& $e_3 e_2 = e_4$& $e_3 e_3 = e_4$&\\ \hline $\cd {4}{28}(\alpha)$&$:$& $e_1 e_1 = e_2$& $e_2 e_1 = e_4$& $e_2 e_2 = e_4$\\ $\alpha\ne 1$&& $e_2 e_3 = e_4$& $e_3 e_2 = e_4$& $e_3 e_3 =\alpha e_4$&\\ \hline $\cd {4}{29}$&$:$& $e_1 e_1 = e_2$& $e_2 e_3 =-\frac{1}{2} e_4$& $e_3 e_2 =\frac{1}{2} e_4$& $e_3 e_3 = e_4$&\\ \hline $\cd {4}{30}$&$:$& $e_1 e_1 = e_2$& $e_2 e_1 = e_4$& $e_2 e_3 = e_4$& $e_3 e_2 = e_4$&\\ \hline $\cd {4}{31}$&$:$& $e_1 e_1 = e_2$& $e_2 e_3 = e_4$& $e_3 e_1 = e_4$& $e_3 e_2 = e_4$&\\ \hline $\cd {4}{32}(\alpha)$&$:$& $e_1 e_1 = e_2$& $e_2 e_3 = \alpha e_4$& \multicolumn{2}{l}{$e_3 e_2 = (\alpha+1) e_4$}& \\ \hline $\cd {4}{33}$&$:$& $e_1 e_1 = e_2$& $e_2 e_1 = e_4$& $e_2 e_2 = e_4$& $e_3 e_3 = e_4$& \\ \hline $\cd {4}{34}$&$:$& $e_1 e_1 = e_2$& $e_2 e_2 = e_4$& $e_3 e_3 = e_4$& \\ \hline $\cd {4}{35}$&$:$& $e_1 e_1 = e_2$& $e_2 e_2 = e_4$& $e_3 e_1 = e_4$& $e_3 e_3 = e_4$& \\ \hline $\cd {4}{36}(\alpha)$&$:$& $e_1 e_1 = e_2$& $e_2 e_2 = e_4$& $e_3 e_2= e_4$& $e_3 e_3 =\alpha e_4$&\\ \hline $\cd {4}{37}$&$:$& $e_1 e_1= e_2$& $e_1 e_2= e_4$& $e_2 e_1= e_4$& $e_3 e_3= e_4$\\ \hline $\cd {4}{38}$&$:$& $e_1 e_1= e_2$& $e_2 e_3= e_4$& $e_3 e_2= e_4$&\\ \hline $\cd 4{39}$ & $:$ & $e_1 e_1 = e_3+e_4$ & $e_1e_2=\frac i2 e_4$ & $e_1e_3=e_4$ & $e_2e_1=\frac i2 e_4$\\ && $ e_2 e_2=e_3$ & $e_2e_3=-2ie_4$ & $e_3e_1=2e_4$ & $e_3e_2=-ie_4$\\ \hline $\cd 4{40}$ & $:$ & $e_1 e_1 = e_3+e_4$ & $e_1e_2=\frac i2 e_4$ & $e_1e_3=-\frac 12e_4$ & $e_2e_1=\frac i2 e_4$\\ && $ e_2 e_2=e_3$ & $e_2e_3=-\frac i2e_4$ & $e_3e_1=\frac 12e_4$ & $e_3e_2=\frac i2e_4$\\ \hline $\cd 4{41}$ & $:$ & $e_1 e_1 = e_3+e_4$ & $e_1e_2=-\frac i2 e_4$ & $e_1e_3=e_4$ & $e_2e_1=-\frac i2 e_4$\\ && $ e_2 e_2=e_3$ & $e_2e_3=-2ie_4$ & $e_3e_1=2e_4$ & $e_3e_2=-ie_4$\\ \hline $\cd 4{42}$ & $:$ & $e_1 e_1 = e_3+e_4$ & $e_1e_2=-\frac i2 e_4$ & $e_1e_3=-\frac 12e_4$ & $e_2e_1=-\frac i2 e_4$\\ && $ e_2 e_2=e_3$ & $e_2e_3=-\frac i2e_4$ & $e_3e_1=\frac 12e_4$ & $e_3e_2=\frac i2e_4$\\ \hline $\cd 4{43}(\alpha)$ &$:$& $e_1 e_1 = e_3 + e_4$ & $e_1 e_2 = \alpha e_4$ & $e_1 e_3 = -\frac 12 e_4$ \\ && $e_2 e_1 = \alpha e_4$ & $e_2 e_2 = e_3$ & $e_3 e_1 = \frac 12 e_4$\\ \hline $\cd 4{44}(\alpha,\beta,\gamma)$ & $:$ & $e_1 e_1 = e_3 + \alpha e_4$ & $e_1 e_2 = \beta e_4$ & $e_2 e_1 = (\beta+\gamma) e_4$ \\ && $e_2 e_2 = e_3$ & $e_3 e_1 = e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{45}$ & $:$ & $e_1 e_1 = e_3 + 2i e_4$ & $e_1 e_2 = e_4$ & $e_2 e_1 = e_4$& $e_2 e_2 = e_3$\\ && $e_3 e_1 = e_4$ & $e_3 e_2 = i e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{46}(\alpha)$ & $:$ & $e_1 e_1 = e_3 - 2i\alpha e_4$ & $e_1 e_2 = \alpha e_4$ & $e_2 e_1 = \alpha e_4$ & $e_2 e_2 = e_3$ \\ $\alpha\ne 0$& & $e_3 e_1 = e_4$ & $e_3 e_2 = i e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{47}(\alpha,\beta)$ & $:$ & $e_1 e_1 = e_3 + e_4$ & $e_1 e_3 = \alpha e_4$ & $e_2 e_2 = e_3$\\ $\beta\ne 0$ && $e_2 e_3 = \beta e_4$ & $e_3 e_1 =(\alpha+1) e_4$ & $e_3 e_2 = \beta e_4$ \\ \hline $\cd 4{48}(\alpha)$ & $:$ & $e_1 e_1 = e_3 + \alpha e_4$ & $e_2 e_1 = i\alpha e_4$ & $e_2 e_2 = e_3$ \\ $\alpha\ne 0$&& $e_3 e_1 = e_4$ & $e_3 e_2 = i e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{49}(\alpha)$ & $:$ & $e_1 e_1 = e_3 + \alpha e_4$ & $e_2 e_1 = -i\alpha e_4$ & $e_2 e_2 = e_3$\\ $\alpha\ne 0$& & $e_3 e_1 = e_4$& $e_3 e_2 = i e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{50}(\alpha)$ & $:$ & $e_1 e_1 = e_3 + \alpha e_4$ & $e_2 e_1 = e_4$ & $e_2 e_2 = e_3$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{51}(\alpha)$ & $:$ & $e_1 e_1 = e_3 + \alpha e_4$ & $e_2 e_2 = e_3$ & $e_3 e_1 = e_4$\\ && $e_3 e_2 = i e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{52}$ & $:$ & $e_1 e_1 = e_3 + e_4$ & $e_2 e_2 = e_3$ & $e_3 e_3 = e_4$\\ \hline $ \cd 4{53}$ & $:$ & $e_1 e_1 = e_3$ & $e_1e_2=-\frac 12e_4$ & $e_1e_3=e_4$ & $e_2e_1=\frac 12e_4$\\ && $e_2 e_2 = e_3$ & $e_2e_3=ie_4$ & $e_3e_1=e_4$ & $e_3e_2=ie_4$ \\ \hline $\cd 4{54}(\alpha)$ & $:$ & $e_1 e_1 = e_3$ & $e_1e_2=e_4$ & $e_1e_3=\alpha e_4$ & $e_2e_1=e_4$\\ && $ e_2 e_2=e_3$ & $e_2e_3=-i(\alpha+1)e_4$ & $e_3e_1=(\alpha+1) e_4$ & $e_3e_2=-i\alpha e_4$\\ \hline $\cd 4{55}(\alpha)$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_2 = e_4$ & $e_1 e_3 = \alpha e_4$ \\ & & $e_2 e_1 = e_4$ & $e_2 e_2 = e_3$ & $e_3 e_1 = (\alpha+1) e_4$ \\ \hline $\cd 4{56}$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_2 = e_4$ & $e_2 e_1 = -e_4$ & $e_2 e_2 = e_3$ \\ & & $e_3 e_1 = e_4$ & $e_3 e_2 = i e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{57}(\alpha,\beta)$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_2 = \alpha e_4$ & $e_2 e_1 = (\alpha+\beta)e_4$ & $e_2 e_2 = e_3$ \\ $\beta\not\in\{0,-2\alpha\}$& & $e_3 e_1 = e_4$ & $e_3 e_2 = i e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{58}$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_3 = e_4$ & $e_2 e_1 = e_4$ & $e_2 e_2 = e_3$\\ && $e_2 e_3 = i e_4$ & $e_3 e_1 = e_4$ & $e_3 e_2 = i e_4$ \\ \hline $\cd 4{59}(\alpha,\beta)$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_3 = \alpha e_4$ & $e_2 e_2 = e_3$\\ $\beta\ne 0$ && $e_2 e_3 = \beta e_4$ & $e_3 e_1 = (\alpha+1) e_4$ & $e_3 e_2 = \beta e_4$ \\ \hline $\cd 4{60}$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_3 = i e_4$ & $e_2 e_2 = e_3$\\ && $e_2 e_3 = e_4$ & $e_3 e_1 = (i+1) e_4$ & $e_3 e_2 = (i+1) e_4$ \\ \hline $\cd 4{61}(\alpha)$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_3 = -i\alpha e_4$ & $e_2 e_2 = e_3$\\ && $e_2 e_3 = \alpha e_4$ & $e_3 e_1 = (1-i\alpha) e_4$ & $e_3 e_2 = (\alpha+i) e_4$ \\ \hline $\cd 4{62}$ & $:$ & $e_1 e_1 = e_3$ & $e_1e_3=e_4$ & $e_2 e_2=e_3$\\ && $e_2e_3=-2ie_4$ & $e_3e_1=2e_4$ & $e_3e_2=-ie_4$\\ \hline $\cd 4{63}$ & $:$ & $e_1 e_1 = e_3$ & $e_1e_3=-\frac 12e_4$ & $ e_2 e_2=e_3$\\ && $e_2e_3=-\frac i2e_4$ & $e_3e_1=\frac 12e_4$ & $e_3e_2=\frac i2e_4$\\ \hline $\cd 4{64}(\alpha)$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_3 = \alpha e_4$ & $e_2 e_2 = e_3$ & $e_3 e_1 = (\alpha+1) e_4$ \\ \hline $\cd 4{65}(\alpha)$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_3 = \alpha e_4$ & $e_2 e_2 = e_3$ \\ $\alpha\ne 0$&& $e_3 e_1 = (\alpha+1) e_4$ & $e_3 e_2 = i e_4$ \\ \hline $\cd 4{66}$ & $:$ & $e_1 e_1 = e_3$& $e_2 e_1 = e_4$ & $e_2 e_2 = e_3$ \\ && $e_2 e_3 = e_4$ & $e_3 e_2 = e_4$ \\ \hline $\cd 4{67}$ & $:$ & $e_1 e_1 = e_3$ & $e_2 e_2 = e_3$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{68}$ & $:$ & $e_1 e_1 = e_3+e_4$ & $e_1 e_3 = i e_4$ &$e_2 e_2 = e_3$ &\\ &&$e_2 e_3 = e_4$&$e_3 e_1 = ie_4$ & $e_3 e_2 = e_4$ \\ \hline $\cd 4{69}$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_3 = ie_4$ &$e_2 e_2 = e_3$ &\\ &&$e_2 e_3 = e_4$&$e_3 e_1 = ie_4$ & $e_3 e_2 = e_4$ \\ \hline $\cd 4{70}$ & $:$ & $e_1 e_1 = e_3$ & $e_1 e_3 = e_4$ &$e_2 e_2 = e_3$ & $e_3 e_1 = e_4$ & \\ \hline $\cd 4{71}$ & $:$ & $e_1 e_1 = e_4$ & $e_1 e_2 = e_3$ & $e_1 e_3 = e_4$ \\&& $e_2 e_1 = -e_3$ & $e_3 e_1 = -e_4$\\ \hline $\cd 4{72}$ & $:$ & $e_1 e_1 = e_4$ & $e_1 e_2 = e_3$ & $e_1 e_3 = e_4$\\ && $e_2 e_1 = -e_3$ & $e_2 e_2 = e_4$ & $e_3 e_1 = -e_4$\\ \hline $\cd 4{73}$ & $:$ & $e_1 e_2 = e_3 + e_4$ & $e_1 e_3 = e_4$ & $e_2 e_1 = -e_3$ & $e_3 e_1 = -e_4$\\ \hline $\cd 4{74}(\alpha)$ & $:$ & $e_1 e_2 = e_3$ & $e_1 e_3 = (\alpha+1)e_4$ & $e_2 e_1 = -e_3$ & $e_3 e_1 = -\alpha e_4$\\ \hline $\cd 4{75}(\alpha)$ & $:$ & $e_1 e_2 = e_3$ & $e_1 e_3 = (\alpha+1)e_4$ & $e_2 e_1 = -e_3$ \\&& $e_2 e_2 = e_4$ & $e_3 e_1 = -\alpha e_4$\\ \hline $\cd 4{76}$ & $:$ & $e_1 e_2 = e_3$ & $e_1 e_3 = e_4$ & $e_2 e_1 = -e_3$ \\&& $e_2 e_2 = e_4$ & $e_3 e_1 = -e_4$\\ \hline $\cd 4{77}$ & $:$ & $e_1 e_2 = e_3$ & $e_1 e_3 = e_4$ & $e_2 e_1 = -e_3$ \\&& $e_2 e_3 = e_4$ & $e_3 e_2 = -e_4$\\ \hline $\cd 4{78}$ & $:$ & $e_1 e_2 = e_3$ & $e_1 e_3 = e_4$ & $e_2 e_1 = -e_3$\\ && $e_2 e_2 = e_4$ & $e_2 e_3 = e_4$ & $e_3 e_2 = -e_4$\\ \hline $\cd 4{79}(\alpha)$ & $:$ & $e_1 e_2 = e_3+\alpha e_4$ & $e_1 e_3 = e_4$ & $e_2 e_1 = -e_3$ \\&& $e_2 e_3 = e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{80}$ & $:$ & $e_1 e_2 = e_3+e_4$ & $e_1 e_3 = e_4$ & $e_2 e_1 = -e_3$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{81}$ & $:$ & $e_1 e_2 = e_3+e_4$ & $e_2 e_1 = -e_3$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{82}$ & $:$ & $e_1 e_2 = e_3$ & $e_1 e_3 = e_4$ & $e_2 e_1 = -e_3$ \\&& $e_2 e_2 = e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{83}( \alpha)$ & $:$ & $e_1 e_2 = e_3$ & $e_2 e_1 = -e_3$ & $e_2 e_2 = \alpha e_4$ \\ $\alpha\ne 0$&& $e_2 e_3 = e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{84}$ & $:$ & $e_1 e_2 = e_3$ & $e_2 e_1 = -e_3$ & $e_2 e_2 = e_4$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{85}$ & $:$ & $e_1 e_2 = e_3$ & $e_2 e_1 = -e_3$ & $e_3 e_3 = e_4$\\ \hline $\cd 4{86}$ & $:$ & $e_1 e_2 = e_3$ & $e_2 e_1 = -e_3$ & $e_2 e_3 = e_4$ & $e_3 e_2 = -e_4$\\ \hline $\cd {4}{87}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3+(2 \Theta-1)e_4$& $e_1 e_2=e_4$ & $e_1e_3=e_4$& \multicolumn{2}{l}{$e_2 e_1=e_3-(1- \Theta)^2 \lambda^{-1}e_4$}\\ $\lambda \neq 0, \frac{1}{4}$&& $e_2 e_2=e_3$& $e_2e_3=\Theta\lambda^{-1}e_4$ &$e_3e_3=e_4$\\ \hline$\cd {4}{88}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3+(1-2 \Theta)e_4$& $e_1 e_2=e_4$ & $e_1e_3=e_4$& \multicolumn{2}{l}{$e_2 e_1=e_3- \Theta^2 \lambda^{-1}e_4$}\\ $\lambda \neq 0, \frac{1}{4}$&& $e_2 e_2=e_3$& $e_2e_3=\Theta\lambda^{-1}e_4$ &$e_3e_3=e_4$\\ \hline$\cd {4}{89}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3+(2 \Theta-1)e_4$& $e_1 e_2=e_4$ & $e_1e_3=e_4$& \multicolumn{2}{l}{$e_2 e_1=e_3-(1- \Theta)^2 \lambda^{-1}e_4$}\\ $\lambda \neq 0, \frac{1}{4}$&& $e_2 e_2=e_3$& $e_2e_3=(1-\Theta)\lambda^{-1}e_4$ &$e_3e_3=e_4$\\ \hline$\cd {4}{90}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3+(1-2 \Theta)e_4$& $e_1 e_2=e_4$ & $e_1e_3=e_4$& \multicolumn{2}{l}{$e_2 e_1=e_3- \Theta^2 \lambda^{-1}e_4$}\\ $\lambda \neq 0, \frac{1}{4}$&& $e_2 e_2=e_3$& $e_2e_3=(1-\Theta)\lambda^{-1}e_4$ &$e_3e_3=e_4$\\ \hline$\cd {4}{91}(\lambda, \alpha)$&$:$& $e_1 e_1 = \lambda e_3+(2 \Theta-1)e_4$& $e_1 e_2=e_4$ & $e_1e_3=\alpha e_4$& \\ $\lambda \neq 0, \frac{1}{4}$&& $e_2 e_1=e_3- (1-\Theta)^2 \lambda^{-1}e_4$& $e_2 e_2=e_3$& $e_3e_3=e_4$\\ \hline$\cd {4}{92}(\lambda, \alpha)$&$:$& $e_1 e_1 = \lambda e_3+(1-2 \Theta)e_4$& $e_1 e_2=e_4$ & $e_1e_3=\alpha e_4$& \\ $\lambda \neq 0, \frac{1}{4}$&& $e_2 e_1=e_3- \Theta^2 \lambda^{-1}e_4$& $e_2 e_2=e_3$& $e_3e_3=e_4$\\ \hline$\cd {4}{93}( \alpha)$&$:$& $e_1 e_1 = e_4$ & $e_1 e_2=e_4$ &$e_1e_3=\alpha e_4$ & $e_2e_1=e_3+e_4$\\ &&$ e_2 e_2=e_3$ &$e_2e_3=\alpha e_4$ & $e_3e_3=e_4$\\ \hline$\cd {4}{94}(\alpha, \beta)$&$:$& $e_1 e_1 = e_4$ & $e_1e_2=e_4$& $e_1e_3=\alpha e_4$&\\ $\alpha\neq0$&&$e_2 e_1=e_3+\beta e_4$ & $e_2 e_2=e_3$ &$e_3e_3=e_4$\\ \hline$\cd {4}{95}(\alpha)$&$:$& $e_1 e_1 = e_4$ & $e_1e_2=e_4$ &$e_1e_3=\alpha e_4$ & $e_2 e_1=e_3$\\ && $e_2 e_2=e_3$ & $e_2e_3=\alpha e_4$&$e_3e_3=e_4$\\ \hline$\cd {4}{96}(\alpha)$&$:$& $e_1 e_1 = e_4$ & $e_1e_2=e_4$ & $e_2 e_1=e_3+\alpha e_4$\\ && $e_2 e_2=e_3$ & $e_2e_3= e_4$&$e_3e_3=e_4$\\ \hline$\cd {4}{97}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_2=e_4$& $e_1e_3=\Theta e_4 $&$e_2 e_1=e_3-e_4$ \\ && $e_2 e_2=e_3$ &$e_2e_3=e_4$ &$e_3e_3=e_4$\\ \hline$\cd {4}{98}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_2=e_4$& $e_1e_3=(1-\Theta) e_4 $&$e_2 e_1=e_3-e_4$ \\ $\lambda \neq \frac{1}{4}$ && $e_2 e_2=e_3$ &$e_2e_3=e_4$ &$e_3e_3=e_4$\\ \hline$\cd {4}{99}(\alpha)$&$:$& $e_1e_2=e_4$& $e_1e_3=e_4$& $e_2 e_1=e_3-e_4$\\ $\alpha\neq1$ && $e_2 e_2=e_3$ &$e_2e_3=\alpha e_4$&$e_3e_3=e_4$\\ \hline$\cd {4}{100}(\alpha)$&$:$& $e_1 e_1 = \frac{1}{4} e_3$ & $e_1e_2=e_4$& $e_1e_3=\alpha e_4$& $e_2 e_1=e_3-e_4$ \\ $\alpha\notin\{0, \frac{1}{2}\}$&& $e_2 e_2=e_3$ &$e_2e_3=2 \alpha e_4$ &$e_3e_3=e_4$\\ \hline$\cd {4}{101}(\alpha, \beta)$&$:$& $e_1e_2=e_4$& $e_1e_3=\alpha e_4$& $e_2 e_1=e_3$ \\ && $e_2 e_2=e_3$&$e_2e_3=\beta e_4$ &$e_3e_3=e_4$\\ \hline$\cd {4}{102}(\lambda, \alpha)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_2=e_4$& $e_2 e_1=e_3-e_4$ \\ $\lambda\neq0$&& $e_2 e_2=e_3$&$ e_2e_3=\alpha e_4$ &$e_3e_3=e_4$\\ \hline$\cd {4}{103}$&$:$& $e_1 e_2 = e_4$ & $e_2 e_1=e_3-e_4$ & $e_2 e_2=e_3$ &$e_3e_3=e_4$\\ \hline$\cd {4}{104}$&$:$& $e_1 e_3 = e_4$ & $e_2 e_1=e_3+e_4$ & $e_2 e_2=e_3$ \\&&$e_2e_3=e_4$&$e_3e_3=e_4$\\ \hline$\cd {4}{105}(\lambda, \alpha,\beta)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_3=e_4$&$e_2 e_1=e_3+\alpha e_4$ \\ $ \lambda\ne 0, \alpha\ne 0$&& $e_2 e_2=e_3$ & $e_2e_3=\beta e_4$&$e_3e_3=e_4$\\ \hline$\cd {4}{106}(\alpha)$&$:$& $e_1 e_3 = e_4$ & $e_2 e_1=e_3+\alpha e_4$ & $e_2 e_2=e_3$ &$e_3e_3=e_4$\\ \hline$\cd {4}{107}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_3=\Theta e_4$& $e_2 e_1=e_3$ \\ && $e_2 e_2=e_3$ & $e_2e_3=e_4$&$e_3e_3=e_4$\\ \hline$\cd {4}{108}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_1e_3=(1-\Theta) e_4$& $e_2 e_1=e_3$ \\ $\lambda \not\in \{0, \frac{1}{4}\}$&& $e_2 e_2=e_3$ & $e_2e_3=e_4$&$e_3e_3=e_4$\\ \hline$\cd {4}{109}(\lambda,\alpha)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_2 e_1=e_3+e_4$ \\&& $e_2 e_2=e_3$ &$e_2e_3=\alpha e_4$&$e_3e_3=e_4$\\ \hline$\cd {4}{110}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_2 e_1=e_3$ & $e_2 e_2=e_3$ \\&&$e_2e_3= e_4$&$e_3e_3=e_4$\\ \hline$\cd {4}{111}(\lambda)$&$:$& $e_1 e_1 = \lambda e_3$ & $e_2 e_1=e_3$ & $e_2 e_2=e_3$ &$e_3e_3=e_4$\\ \hline$\cd {4}{112}(\lambda, \alpha, \beta, \gamma)$&$:$& $e_1 e_1 = \lambda e_3+e_4$ & $e_1e_3=\alpha e_4$ & $e_2 e_1=e_3+\beta e_4$ \\ &&$e_2 e_2=e_3$& $e_2e_3=\gamma e_4$&$e_3e_3=e_4$\\ \hline ${\bf N}^4_{01}(\alpha)$ &$:$& $e_1 e_1 = e_2$& $e_1e_2=\alpha e_4$& $e_1e_3=e_4$ \\& & $e_2e_1=e_4$& $e_2 e_2=e_3$ \\ \hline ${\bf N}^4_{02}(\alpha,\beta,\gamma,\delta)$ &$:$& $e_1 e_1 = e_2$& $e_1e_2=\alpha e_4$& $e_1e_3=\gamma e_4$& $e_2e_1=\beta e_4$\\&& $e_2 e_2=e_3$& $e_2e_3=\delta e_4$& $e_3e_2=e_4$& $e_3e_3=e_4$\\ \hline ${\bf N}^4_{03}(\alpha)$ &$:$& $e_1 e_1 = e_2$& $e_1e_2=\alpha e_4$& $e_1e_3=e_4$\\&& $e_2 e_2=e_3$ & $e_2e_3=e_4$\\ \hline ${\bf N}^4_{04}(\alpha,\beta,\gamma)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=\gamma e_4$&$e_2e_1=\beta e_4$\\ && $e_2 e_2=e_3$&$e_2e_3=e_4$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{05}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=e_4$ \\&&$e_2e_1=\beta e_4$& $e_2 e_2=e_3$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{06}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_1=e_4$\\&& $e_2 e_2=e_3$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{07}$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$&$e_1e_3=e_4$& $e_2 e_2=e_3$ \\ \hline ${\bf N}^4_{08}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$&$e_1e_3=\beta e_4$\\&&$e_2e_1=\alpha e_4$& $e_2 e_2=e_3$ &$e_3e_1=e_4$\\ \hline ${\bf N}^4_{09}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=\beta e_4$ \\&& $e_2 e_2=e_3$ &$e_2e_3=e_4$&$e_3e_1=e_4$ \\ \hline ${\bf N}^4_{10}$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$& $e_2 e_2=e_3$ &$e_2e_3=e_4$ \\ \hline ${\bf N}^4_{11}$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$& $e_2 e_2=e_3$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{12}(\alpha)$ &$:$& $e_1 e_1 = e_2$& $e_1e_3=\alpha e_4$&$e_2e_1=e_4$\\& & $e_2 e_2=e_3$ &$e_3e_1=e_4$ \\ \hline ${\bf N}^4_{13}(\alpha,\beta,\gamma)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=\beta e_4$&$e_2e_1=\alpha e_4$& $e_2 e_2=e_3$\\ &&$e_2e_3=\gamma e_4$&$e_3e_1=e_4$&$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{14}(\alpha, \beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3 =e_4$&$e_2e_1= \alpha e_4$ \\&& $e_2 e_2=e_3$ &$e_2e_3=\beta e_4$&$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{15}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_4$& $e_2 e_2=e_3$ \\&&$e_2e_3=\alpha e_4$ &$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{16}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=\alpha e_4$ & $e_2 e_2=e_3$ &$e_3e_1=e_4$\\ \hline ${\bf N}^4_{17}$ &$:$& $e_1 e_1 = e_2$& $e_1e_3=e_4$ & $e_2 e_2=e_3$\\ \hline ${\bf N}^4_{18}$ &$:$& $e_1 e_1 = e_2$& $e_2 e_2=e_3$ &$e_2e_3=e_4$\\ \hline ${\bf N}^4_{19}(\alpha)$ &$:$& $e_1 e_1 = e_2$& $e_2 e_2=e_3$ &$e_2e_3=\alpha e_4$ &$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{20}$ &$:$& $e_1 e_1 = e_2$& $e_2 e_2=e_3$ &$e_3e_3=e_4$\\ \hline ${\bf N}^4_{21}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=e_4$ \\&&$e_2e_1=e_3+\beta e_4$& $e_2 e_2=e_3$ \\ \hline ${\bf N}^4_{22}(\alpha,\beta,\gamma)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=\gamma e_4$ \\&&$e_2e_1=e_3+\beta e_4$& $e_2 e_2=e_3$&$e_3e_1=e_4$ \\ \hline ${\bf N}^4_{23}(\alpha,\beta,\gamma)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_1e_3=\beta e_4$&$e_2e_1=e_3$\\ && $e_2 e_2=e_3$&$e_2e_3=e_4$&$e_3e_1=\gamma e_4$ \\ \hline ${\bf N}^4_{24}(\alpha,\beta,\gamma,\delta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=\beta e_4$&$e_2e_1=e_3+\alpha e_4$& $e_2 e_2=e_3$ \\ &&$e_2e_3=\delta e_4$&$e_3e_1=\gamma e_4$ & $e_3e_2=e_4$\\ \hline ${\bf N}^4_{25}(\alpha, \beta, \gamma, \delta, \varepsilon)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2= \alpha e_4$& $e_2e_1=e_3+\beta e_4$ & $e_2 e_2=e_3$\\ &&$e_2e_3=\delta e_4$ & $e_3e_1=\gamma e_4$ & $e_3e_2=\varepsilon e_4$ &$e_3e_3=e_4$\\ \hline ${\bf N}^4_{26}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$&$e_1e_3=-\alpha e_4$ \\&&$e_2e_1=e_3$ &$e_2e_2=\alpha e_4$&$ e_3e_1=(1+2 \alpha) e_4$ \\ \hline ${\bf N}^4_{27}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_1=e_3$\\& &$e_2e_2=\beta e_4$&$e_2e_3=e_4$&$e_3e_1=e_4$& \\ \hline ${\bf N}^4_{28}(\alpha,\beta,\gamma)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_1=e_3$&$e_2e_2=\beta e_4$\\&&$e_2e_3=e_4$&$e_3e_1=\gamma e_4$&$e_3e_3=e_4$& \\ \hline ${\bf N}^4_{29}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_2=\beta e_4$\\& &$e_2e_1=e_3$&$e_3e_1=e_4$&$e_3e_3=e_4$& \\ \hline ${\bf N}^4_{30}(\alpha) $ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_1=e_3$\\&&$e_2e_2=e_4$&$e_2e_3=e_4$ \\ \hline ${\bf N}^4_{31}(\alpha) $ &$:$& $e_1 e_1 = e_2$&$e_1e_2=\alpha e_4$&$e_2e_1=e_3$\\&&$e_2e_2=e_4$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{32}$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$&$e_2e_1=e_3$&$e_2e_3=e_4$ \\ \hline ${\bf N}^4_{33}$ &$:$& $e_1 e_1 = e_2$&$e_1e_2=e_4$&$e_2e_1=e_3$&$e_3e_3=e_4$ \\ \hline ${\bf N}^4_{34}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=\beta e_4$&$e_2e_1=e_3$\\&&$e_2e_2=\alpha e_4$& \multicolumn{2}{l}{$e_3e_1=(1-2\beta) e_4$} \\ \hline ${\bf N}^4_{35}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=\alpha e_4$&$e_2e_1=e_3$\\& &$e_2e_2=e_4$&$e_3e_1=-2\alpha e_4$&$e_3e_2=e_4$& \\ \hline ${\bf N}^4_{36}(\alpha,\beta)$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_3$&$e_2e_2=\alpha e_4$\\ $\beta\neq 0$& &$e_2e_3=\beta e_4$&$e_3e_1=e_4$&$e_3e_2=e_4$& \\ \hline ${\bf N}^4_{37}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_3$&$e_2e_2=e_4$ \\ $\alpha\neq0$&&$e_2e_3=\alpha e_4$&$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{38}$ &$:$& $e_1 e_1 = e_2$&$e_1e_3=e_4$&$e_2e_1=e_3$\\&&$e_3e_1=-2e_4$&$e_3e_2=e_4$ \\ \hline ${\bf N}^4_{39}$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_3$&$e_2e_3=e_4$& \\ \hline ${\bf N}^4_{40}(\alpha)$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_3$&$e_2e_3=\alpha e_4$&$e_3e_2=e_4$& \\ \hline ${\bf N}^4_{41}$ &$:$& $e_1 e_1 = e_2$&$e_2e_1=e_3$&$e_3e_3=e_4$& \\ \hline ${\bf N}^4_{42}(\lambda,\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & \multicolumn{2}{l}{$e_1e_3=(\alpha(\lambda-2)+1)e_4$}\\ $\lambda\ne 1$&& $e_2 e_1=\lambda e_3+e_4$ & \multicolumn{2}{l}{$e_2e_2=\frac{(\lambda+1) ((\lambda^2-1)\alpha + \lambda + 2)}{1-\lambda}e_4$} & $e_3e_1=(\alpha(1-2\lambda)-2)e_4$ \\ \hline ${\bf N}^4_{43}(\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=-\frac {5\alpha}2e_4$ & $e_2 e_1=-\frac 12e_3+\beta e_4$\\ && $e_2e_2=e_4$ & $e_2 e_3=-\frac 12 e_4$ & $e_3e_1=2\alpha e_4$ & $e_3e_2 = e_4$ \\ \hline ${\bf N}^4_{44}(\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=-\frac {5\alpha}2e_4$ & $e_2 e_1=-\frac 12e_3+e_4$\\ && $e_2 e_3=-\frac 12 e_4$ & $e_3e_1=2\alpha e_4$ & $e_3e_2 = e_4$ \\ \hline ${\bf N}^4_{45}(\lambda,\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & \multicolumn{2}{l}{$e_1e_3=(\alpha(\lambda-2)+1)e_4$}\\ && $e_2 e_1=\lambda e_3$& $e_2e_2=\beta e_4$ & \multicolumn{2}{l}{$e_3e_1=(\alpha(1-2\lambda)-2)e_4$} \\ \hline ${\bf N}^4_{46}(\lambda,\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=\alpha(\lambda-2)e_4$ & $e_2 e_1=\lambda e_3$\\ $\lambda\ne-\frac 12$&& $e_2e_2=e_4$ & $e_2 e_3=-\frac 12 e_4$ & $e_3e_1=\alpha(1-2\lambda)e_4$ & $e_3e_2 = e_4$ \\ \hline ${\bf N}^4_{47}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=-\frac 52e_4$ & $e_2 e_1=-\frac 12e_3$\\ && $e_2 e_3=-\frac 12 e_4$ & $e_3e_1=2 e_4$ & $e_3e_2 = e_4$ \\ \hline ${\bf N}^4_{48}(\lambda,\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=e_4$ & $e_2 e_1=\lambda e_3+\alpha e_4$\\ $\lambda\ne -\frac 12$&& $e_2e_2=\beta e_4$ & $e_2 e_3=\lambda e_4$ & $e_3e_1=-2e_4$ & $e_3e_2 = e_4$ \\ \hline ${\bf N}^4_{49}(\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=\beta e_4$\\ && $e_2 e_1=-e_3+\alpha e_4$ & $e_2e_2=e_4$ & $e_2 e_3=e_4$\\ && $e_3e_1=-2\beta e_4$ & $e_3e_2 = e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{50}(\lambda,\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=e_4$ & $e_2 e_1=\lambda e_3+\alpha e_4$\\ $\lambda\ne-1$&& $e_2e_2=\beta e_4$ & $e_3e_1=-2e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{51}(\lambda,\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+\alpha e_4$\\ $\lambda\ne -\frac 12$&& $e_2e_2=e_4$ & $e_2e_3=\lambda e_4$ & $e_3e_2=e_4$ \\ \hline ${\bf N}^4_{52}(\lambda,\alpha,\beta,\gamma)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+\alpha e_4$ & $e_2e_2=\beta e_4$\\ $\gamma\ne 1$&& $e_2e_3=\gamma e_4$ & $e_3e_2=e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{53}(\lambda,\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+\alpha e_4$\\ && $e_2e_2=\beta e_4$ & $e_2e_3=e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{54}(\lambda,\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+\alpha e_4$\\ $\lambda\ne -1$&& $e_2e_2=e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{55}(\lambda)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+e_4$\\ $\lambda\ne -\frac 12$ && $e_2e_3=\lambda e_4$ & $e_3e_2=e_4$ \\ \hline ${\bf N}^4_{56}(\lambda)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3+e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{57}(\lambda,\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=e_4$ & $e_2 e_1=\lambda e_3$\\ && $e_2e_2=\alpha e_4$ & $e_2e_3=e_4$ & $e_3e_1=-2e_4$ \\ \hline ${\bf N}^4_{58}(\lambda,\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=e_4$ & $e_2 e_1=\lambda e_3$\\ $\beta\not\in\{-\frac 12,\lambda\}$&& $e_2e_2=\alpha e_4$ & $e_2e_3=\beta e_4$ & $e_3e_1=-2e_4$ & $e_3e_2=e_4$ \\ \hline ${\bf N}^4_{59}(\alpha,\beta)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=\beta e_4$\\ $\alpha\ne 1$&& $e_2 e_1=-e_3$ & $e_2e_2=\alpha e_4$ & $e_2e_3=e_4$\\ && $e_3e_1=-2\beta e_4$ & $e_3e_2=e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{60}(\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_1e_3=e_4$ & $e_2 e_1=-e_3$\\ $\alpha\ne 0$&& $e_2e_2=\alpha e_4$ & $e_3e_1=-2e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{61}(\lambda)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$\\ && $e_2e_2=e_4$ & $e_2e_3=e_4$ \\ \hline ${\bf N}^4_{62}(\lambda,\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$\\ $\alpha\not\in\{-\frac 12,\lambda\}$&& $e_2e_2=e_4$ & $e_2e_3=\alpha e_4$ & $e_3e_2=e_4$ \\ \hline ${\bf N}^4_{63}$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=-e_3$\\ && $e_2e_2=e_4$ & $e_3e_3=e_4$ \\ \hline ${\bf N}^4_{64}(\lambda)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$ & $e_2e_3=e_4$ \\ \hline ${\bf N}^4_{65}(\lambda,\alpha)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$\\ && $e_2e_3=\alpha e_4$ & $e_3e_2=e_4$ \\ \hline ${\bf N}^4_{66}(\lambda)$ & $:$ & $e_1 e_1 = e_2$ & $e_1 e_2=e_3$ & $e_2 e_1=\lambda e_3$ & $e_3e_3=e_4$\\ \end{longtable}} All of these algebras are pairwise non-isomorphic, except for the following: {\tiny \begin{longtable}{c} $\D{4}{01}(\lambda,0,\beta) \cong \D{4}{02}(\lambda,0,\beta) \cong \D{4}{04}(\lambda,\beta),\quad \D{4}{01}(\lambda,\alpha,0)_{\alpha \neq -1} \cong \D{4}{02}(\lambda,\alpha,0) \cong \D{4}{10}(\lambda,\alpha),\quad \D{4}{01}(\lambda,-1,0) \cong \D{4}{11}(\lambda,0),$\\ $\D{4}{03}(\lambda,0) \cong \D{4}{09}(\lambda,0),\quad \D{4}{03}\left(\lambda,(1-\Theta)^{-1}\right)_{\lambda \neq 0} \cong \D{4}{05}(\lambda,0)_{\lambda \neq 0}, \D{4}{03}\left(\lambda,\Theta^{-1}\right)\cong \D{4}{06}(\lambda,0),\quad \D{4}{04}(\lambda,0) \cong \D{4}{10}(\lambda,0),$\\ $\D{4}{05}(1/4,\alpha) \cong \D{4}{06}(1/4,\alpha),\quad \D{4}{07}(1/4) \cong \D{4}{08}(1/4), \quad \D{4}{05}(0,\alpha) \cong \D{4}{07}(0) \cong \D{4}{23}(0) \cong \D{4}{25}(0) \cong \D{4}{40}(0),$\\ $\D{4}{12}(\lambda,0) \cong \D{4}{18}(\lambda,0),\quad \D{4}{12}(1/4,\alpha) \cong \D{4}{13}(1/4,\alpha),\quad \D{4}{12}(0,\alpha)_{\alpha \neq -1} \cong \D{4}{14}(0,\alpha),\quad \D{4}{12}(0,-1) \cong \D{4}{17}(0),$\\ $\D{4}{13}(\lambda,0) \cong \D{4}{19}(\lambda,0),\quad \D{4}{14}(\lambda,0) \cong \D{4}{20}(\lambda,0),\quad \D{4}{14}(1/4,\alpha) \cong \D{4}{15}(1/4,\alpha),\quad \D{4}{15}(\lambda,0) \cong \D{4}{21}(\lambda,0),$\\ $\D{4}{18}(1/4,\alpha) \cong \D{4}{19}(1/4,\alpha),\quad \D{4}{18}(0,0) \cong \D{4}{22}(0) \cong \D{4}{24}(0),\quad \D{4}{18}(1/4,-1) \cong \D{4}{19}(1/4,-1) \cong \D{4}{30}(1/4) \cong \D{4}{31}(1/4),$\\ $\D{4}{20}(1/4,\alpha) \cong \D{4}{21}(1/4,\alpha),\quad \D{4}{20}(1/4,-1) \cong \D{4}{21}(1/4,-1) \cong \D{4}{32}(1/4) \cong \D{4}{33}(1/4),$\\ $\D{4}{22}(1/4) \cong \D{4}{23}(1/4) \cong \D{4}{24}(1/4) \cong \D{4}{25}(1/4) \cong \D{4}{26}(1/4) \cong \D{4}{27}(1/4) \cong \D{4}{28}(1/4) \cong \D{4}{29}(1/4),$\\ $ \D{4}{37}(1/4) \cong \D{4}{38}(1/4),\quad \D{4}{39}(1/4) \cong \D{4}{40}(1/4).$ \end{longtable} \begin{longtable}{lll} $\cd 4{43}(\alpha)\cong\cd 4{43}(-\alpha)$ & $\cd 4{44}(\alpha,\beta,\gamma)\cong\cd 4{44}(\alpha,-\beta,-\gamma)$& $\cd 4{47}(\alpha,\beta)\cong \cd 4{47}(\alpha,-\beta)$\\ $\cd 4{50}(\alpha)=\cd 4{50}(-\alpha)$& $\cd 4{54}(\alpha)\cong\cd 4{54}(-\alpha-1)$& $\cd 4{57}(\alpha,\beta)\cong \cd 4{57}(\alpha+\beta,-\beta)$\\ $\cd 4{59}(\alpha,\beta)\cong\cd 4{59}(\alpha,-\beta)$& $\cd {4}{91}(\lambda, \alpha)\cong\cd {4}{91}(\lambda, -\alpha)$ & $\cd {4}{92}(\lambda, \alpha)\cong\cd {4}{92}(\lambda, -\alpha)$ \\ $\cd {4}{93}(\alpha)\cong\cd {4}{93}(-\alpha)$ & $\cd {4}{94}(\alpha,\beta)\cong\cd {4}{94}(-\alpha,\beta)$ & $\cd {4}{95}(\alpha)\cong\cd {4}{95}(-\alpha)$ \\ $\cd {4}{100}(\alpha)\cong\cd {4}{100}(-\alpha)$ & $\cd {4}{101}(\alpha,\beta)\cong\cd {4}{101}(-\alpha,-\beta)$ & $\cd {4}{109}(\lambda,\alpha)\cong\cd {4}{109}(\lambda,-\alpha)$ \\ \multicolumn{3}{c}{$\cd {4}{112}(\lambda,\alpha,\beta,\gamma)\cong\cd {4}{112}(\lambda,-\alpha,\beta,-\gamma)$} \\ \multicolumn{3}{c}{$\cd {4}{112}(\lambda,\alpha,\beta,\gamma)\cong \cd {4}{112}\left(\lambda,(\gamma-\alpha\beta)\sqrt{\frac{-\lambda}{1-\beta+\lambda\beta^2}},\frac 1\lambda-\beta,(\frac\gamma\lambda-\frac\alpha\lambda-\beta\gamma)\sqrt{\frac{-\lambda}{1-\beta+\lambda\beta^2}}\right)$, if $\lambda\ne 0$, $\beta\ne\frac{1\pm\sqrt{1-4\lambda}}{2\lambda}$}\\ ${\bf N}^4_{02}(\alpha,\beta,\gamma,\delta) \cong {\bf N}^4_{02}(-(\alpha,\beta,\gamma),\delta)$ & ${\bf N}^4_{04}(\alpha,\beta,\gamma) \cong {\bf N}^4_{04}(-(\alpha,\beta,\gamma))$ & ${\bf N}^4_{05}(\alpha,\beta) \cong {\bf N}^4_{05}(\sqrt[3]{1}(\alpha,\beta))$ \\ ${\bf N}^4_{29}(\alpha,\beta) \cong {\bf N}^4_{29}(-\alpha,\beta)$ & ${\bf N}^4_{31}(\alpha) \cong {\bf N}^4_{31}(-\alpha)$ & ${\bf N}^4_{50}(\lambda,\alpha,\beta) \cong {\bf N}^4_{50}(\lambda,-\alpha,\beta)$ \\ & ${\bf N}^4_{54}(\lambda,\alpha) \cong {\bf N}^4_{54}(\lambda,-\alpha)$ \end{longtable} } \end{theorem} \section{Applications}\label{S:apps} \subsection{The algebraic classification of $4$-dimensional nilpotent Lie-admissible algebras} The variety of Lie-admissible algebras is defined by the following identity $$[[x,y],z]+[[y,z],x]+[[z,x],y]=0,$$ where $[x,y]=xy-yx.$ Lie-admissible algebras satisfy the following fundamental property: {\it under the commutator multiplication each Lie-admissible algebra is a Lie algebra}. \begin{corollary} Let $\bf A$ be a complex $4$-dimensional nilpotent Lie-admissible algebra. Then $\bf A$ is isomorphic to an algebra from the list given in Theorem \ref{teo-alg}. \end{corollary} \begin{Proof} Thanks to \cite{kkl19} each $4$-dimensional nilpotent anticommutative algebra is Lie. Hence, each $4$-dimensional nilpotent algebra is a Lie-admissible algebra. \end{Proof} \subsection{The algebraic classification of $4$-dimensional nilpotent Alia type algebras} Let $\bf A$ be an algebra with a certain bilinear multiplication $(x,y)\mapsto xy$, which we call standard. Let us define \begin{center}$\textsf{T}(x,y,z, {\textsf P})={\textsf P}([x,y],z)+{\textsf P}([y,z],x)+{\textsf P}([z,x],y)$ \end{center} where $[x,y]=xy-yx$ and ${\textsf P}$ is a bilinear multiplication on the same vector space. Now we are ready to introduce Alia type algebras, which appeared in \cite{dzhuma1, dzhuma2}: \begin{enumerate} \item the variety of $0$-Alia (also known as $0$-anti-Lie-admissible) algebras is defined by the identity $\textsf{T}(x,y,z, {\textsf P})=0,$ where ${\textsf P}$ is the standard multiplication; \item the variety of $1$-Alia (also known as $1$-anti-Lie-admissible) algebras is defined by the identity $\textsf{T}(x,y,z, {\textsf P})=0,$ where ${\textsf P}$ is the Jordan commutator multiplication ${\textsf P}(x,y)=xy+yx;$ \item the variety of two-sided Alia (also known as anti-Lie-admissible) algebras is defined by the identities $\textsf{T}(x,y,z, {\textsf P}_1)=0$ and $\textsf{T}(x,y,z, {\textsf P}_2)=0,$ where ${\textsf P}_1$ is the standard multiplication and ${\textsf P}_2$ is the opposite multiplication. \end{enumerate} \begin{corollary} Let $\bf A$ be a complex $4$-dimensional nilpotent $0$-Alia ($1$-Alia, or two sided Alia) algebra. Then $\bf A$ is isomorphic to an algebra from the list given in Theorem \ref{teo-alg}, except \begin{longtable}{c} $\cd {4}{79} - \cd {4}{85}, \cd {4}{87} - \cd {4}{112}, {\bf N}^4_{25}, \ {\bf N}^4_{28}, \ {\bf N}^4_{29}, \ {\bf N}^4_{31}, \ {\bf N}^4_{33}, \ {\bf N}^4_{41}, \ {\bf N}^4_{49}, \ {\bf N}^4_{50}(\lambda\neq 1), $\\ $\ {\bf N}^4_{52}(\lambda\neq 1),\ {\bf N}^4_{53}(\lambda\neq 1), \ {\bf N}^4_{54}(\lambda\neq 1),\ {\bf N}^4_{56}(\lambda\neq 1),{\bf N}^4_{59}, {\bf N}^4_{60}, {\bf N}^4_{63}, {\bf N}^4_{66}(\lambda\neq 1).$ \end{longtable} \end{corollary} \begin{Proof} After verification, we have that all $3$-dimensional nilpotent algebras are $0$-Alia, $1$-Alia, and two-sided Alia. Analyzing the cocycles on $3$-dimensional algebras we see that the subspaces of $0$-Alia cocycles, $1$-Alia cocycles, and two-sided Alia cocycles coincide and are listed below. \begin{longtable}{|l|l|} \hline Algebra & Cocycles \\ \hline ${\mathfrak{CD}}_{01}^{3}, {\mathfrak{CD}}_{04}^{3}(1), {\mathfrak{CD}}_{01}^{3*},{\mathfrak{CD}}_{02}^{3*}$ & $\langle \Delta_{ij} \rangle_{1\leq i,j\leq 3} $ \\ \hline ${\mathfrak{CD}}_{02}^{3}, {\mathfrak{CD}}_{03}^{3}, {\mathfrak{CD}}_{04}^{3}(\lambda \neq 1), {\mathfrak{CD}}_{03}^{3*},{\mathfrak{CD}}_{04}^{3*}$ & $\langle \Delta_{ij} \rangle_{1\leq i,j\leq 3; (i,j) \neq (3,3) } $ \\ \hline \end{longtable} It remains to choose the algebras from Theorem \ref{teo-alg} that are determined by Alia cocycles. \end{Proof}
{ "timestamp": "2020-12-02T02:23:37", "yymm": "2012", "arxiv_id": "2012.00525", "language": "en", "url": "https://arxiv.org/abs/2012.00525", "abstract": "We give the complete algebraic classification of all complex 4-dimensional nilpotent algebras. The final list has 234 (parametric families of) isomorphism classes of algebras, 66 of which are new in the literature.", "subjects": "Rings and Algebras (math.RA)", "title": "The algebraic classification of nilpotent algebras", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.977022632097108, "lm_q2_score": 0.8198933381139645, "lm_q1q2_score": 0.8010543472429897 }
https://arxiv.org/abs/1611.03919
Additive Collatz Trajectories
Collatz Conjecture (also known as Ulam's conjecture and 3x+1 problem) concerns the behavior of the iterates of a particular function on natural numbers. A number of generalizations of the conjecture have been subjected to extensive study. This paper explores Additive Collatz Trajectories, a particular case of a generalization of Collatz conjecture and puts forward a sufficient and necessary condition for looping of Additive Collatz Trajectories, along with two minor results. An algorithm to compute the number of equivalence classes when natural numbers are quotiented by the limiting behavior of their corresponding trajectories is also proposed.
\section{Introduction} \subsection{The Collatz Conjecture} The $3x + 1$ problem is most simply stated in terms of the Collatz function $C(x)$ defined on integers as: \[ C(x) = \begin{cases} 3x + 1 & \text{if } x \equiv 1 \mod 2\\ \frac{x}{2} &\text{if } x \equiv 0 \mod 2\\ \end{cases} \] The Collatz Conjecture states that every for every $m \in \mathbb{N}$, there exists $k \in \mathbb{N}$ such that iterate $T^{(k)}(m) = 1$. \cite{lagarias}. One natural generalization of Collatz function would be to consider an arbitrary affine linear map of $x$ instead of $3x + 1$. \subsection{Generalization} We introduce the concept of a generalized Collatz function $C_{a,d,m}'(x)$ as: \[ C'_{a,d,m}(x) = \begin{cases} mx + a & \text{if } x \not\equiv 0 \mod d\\ \frac{x}{d} &\text{if } x \equiv 0 \mod d\\ \end{cases} \] where $x,a,d,m \in \mathbb{N}$. For an arbitrary choice of $(x, a, d, m)$ and for sufficient large values of $k \in \mathbb{N}$, we would like to explore the nature of $C_{a,d,m}'^{(k)}(x)$ \cite{generalized}. In this paper, we look at a particular case of the generalized Collatz function, which we term as the Additive Collatz function. An additive Collatz function $T_{a,d}(x)$ is defined as $C_{a,d,1}'(x)$. \subsection{Terminology} We use the following definitions: \begin{definition} An additive Collatz trajectory $O_{a,d}$ of an integer $x$ is the infinite tuple $$O_{a,d}(x) = (x, T_{a,d}(x), T_{a,d}^{(2)}(x), ... )$$ \end{definition} \begin{definition} A trajectory $O = (o_0, o_1, o_2 ... )$ is said to loop if $\exists k, N \in \mathbb{N}$ such that $\forall n > N$, $o_n = o_{n+k}$. \end{definition} \begin{definition} Given $a, d \in \mathbb{N}$, two natural numbers $x_1$ and $x_2$ are said to be equivalent under the orbit equivalence relation if $\exists n_1, n_2, N \in \mathbb{N}$ such that $\forall k > N$: $$ T_{a,d}^{(k+n_1)}(x_1) = T_{a,d}^{(k+n_2)}(x_2) $$ \end{definition} \begin{definition} Given $a, d \in \mathbb{N}$, an orbit is an element of partition of $\mathbb{N}$ under the orbit equivalence relation. \end{definition} \section{Analysis of Additive Collatz Trajectories} The limiting behavior of an additive Collatz trajectory is identified based on the eventual formation of loops. It is a straightforward observation that if $a$ and $d$ are not co-prime, then the trajectory does not necessarily loop. \begin{proposition} For non-coprime $a,d \in \mathbb{N}$, there exists $x \in \mathbb{N}$ such that $O_{a,d}(x)$ does not loop. \end{proposition} \begin{proof} Consider $ r \not\equiv 0 \mod \gcd(a,d) $. Then, $O_{a,d}(r)$ does not loop. This can be proved by induction on natural numbers. \\ Claim: $\forall k \in \mathbb{N}, T_{a,d}^{(k)}(r) = r + ak$ \\ Base Case: Trivially true for $k = 0$\\ Induction Hypothesis: $\exists k \in \mathbb{N} \ni T_{a,d}^{(k)}(r) = r + ak$ \\ Induction Step: Let $\delta= \gcd(a,d)$. $$T_{a,d}^{(k)}(r) = r + ak \equiv r \mod \delta \implies T_{a,d}^{(k)}(r) \not\equiv 0 \mod d$$ $$\implies T_{a,d}^{(k+1)}(r) = T_{a,d}^{(k)}(r) + a = r + ka + a = r + (k+1)a$$ Hence, the trajectory is an increasing progression, and does not loop. \\ \end{proof} It can be observed that if $r$ is not a multiple of $\gcd(a,d)$, then $O_{a,d}(r)$ does not loop. We would now like to analyze the converse of this statement. \begin{lemma} Given $a,d \in \mathbb{N}$, if $r \equiv 0 \mod \delta$, $$T_{a,d}^{(k)}(r) = \delta T_{\frac{a}{\delta},\frac{d}{\delta}}^{(k)}\Big( \frac{r}{\delta} \Big)$$ where $\delta = \gcd(a,d)$ \end{lemma} If $r$ is a multiple of $\gcd(a,d)$, then Lemma 1 permits us to reduce our analysis to a case where $a$ and $d$ are co-prime. We now only consider the cases where $a$ and $d$ are co-prime and prove that for all natural numbers $x$, the additive Collatz trajectory $O_{a,d}(x)$ loops. This analysis is divided into three propositions. \begin{proposition} For co-prime $a,d \in \mathbb{N}$ given $x \in \mathbb{N}$, there exists $N(x) \in \mathbb{N}$ such that $T_{a,d}^{N(x)}(x) \leq a$ \end{proposition} \begin{proof} Consider $(n_i)$ as a sub-trajectory of $O_{a,d}(x)$ such that $$n_0 = x$$ $$n_i = T_{a,d}^{(z_i)}(x)$$ where $T_{a,d}^{(z_i - 1)}(x) = dT_{a,d}^{(z_i)}(x)$. As $a$ and $d$ are co-prime, B\'{e}zout's Lemma forces the existence of such a sub-trajectory. Let $y_i = z_{i+1} - z_i - 1$. Then, \begin{equation} \label{1} n_{i+1} = \frac{n_i + ay_i}{d} \end{equation} On solving the recursion, $$n_{k+1} = \frac{n_0 + a\sum_{i = 0}^k y_id^i}{d^{k+1}}$$ By definition, $y_i$ is the least non-negative integer such that $n_i + ay_i$ is divisible by $d$, and hence $y_i$ is strictly less than $d$. Hence, $$n_{k+1} \leq \frac{n_0 + a\sum_{i = 0}^k (d-1)d^i}{d^{k+1}}$$ Which simplifies to $$n_{k+1} \leq \frac{n_0 - a}{d^{k+1}} + a$$ For sufficiently large $k$, $d^{k+1} > n_0 - a$. Hence, there is an element in the trajectory which is less than or equal to $a$. \\ \end{proof} \begin{proposition} For co-prime $a,d \in \mathbb{N}$, $O_{a,d}(x)$ loops for all $x \leq a$. \end{proposition} \begin{proof} Equation (1) and the fact that $y_i \leq (d-1) $ implies that if $ n_i \leq a$, then $ n_{i+1} \leq \dfrac{a + a(d-1)}{d} = a $ . Thus if $x = n_0 \leq a $ then $ \forall j \in \mathbb{N} $ $n_j \leq a $. As there are only finitely many natural numbers less than or equal to $a$, trajectory loops. \end{proof} \begin{proposition} Given $a,d \in \mathbb{N}$, $O_{a,d}(r)$ loops for all $r \equiv 0 \mod \gcd(a,d)$. \end{proposition} \begin{proof} By Lemma 1, $O_{a,d}(r)$ loops if and only if $O_{\frac{a}{\delta},\frac{d}{\delta}}(\frac{r}{\delta})$ loops where $\delta = \gcd(a,d)$. Let $a' = a/\delta$, $d' = d/\delta$ and $r' = r/\delta$. By Proposition 2, there exists $N(r')$ such that $T_{a,d}^{(N(r'))}(r') \leq a$. Let $T_{a,d}^{(N(r'))}(r') = k$. As the each term of the trajectory depends only on the previous term, $O_{a',d'}(r')$ is equal to $O_{a',d'}(k)$. By Proposition 3, $O_{a',d'}(k)$ loops, hence $O_{a',d'}(r')$ loops. This implies that $O_{a,d}(r)$ loops. \\ \end{proof} A straightforward implication of Proposition 4 is that if $a$ and $d$ are co-prime, then $\forall x \in \mathbb{N}$, $O_{a,d}(x)$ loops. \section{Orbit Counting} For co-prime $a$ and $d$, we have shown that $O_{a,d}(x)$ loops. We would now like to characterize the number of unique loops possible. For this, the concept of orbit equivalence relation is proposed. Two natural numbers are said to be equivalent, if their trajectories are the same, eventually. Formally speaking, given $a, d \in \mathbb{N}$, two natural numbers $x_1$ and $x_2$ are said to be equivalent under the orbit equivalence relation if $\exists n_1, n_2, N \in \mathbb{N}$ such that $\forall k > N$: $$ T_{a,d}^{(k+n_1)}(x_1) = T_{a,d}^{(k+n_2)}(x_2) $$ Under the orbit equivalence relation, the set of natural numbers can be partitioned into equivalence classes. The number of equivalence classes is the same as the number of unique loops of trajectories possible. Proposition 3 and Proposition 4 imply that the number of unique loops formed by $O_{a,d}(x)$ where $x \in \mathbb{N}$ is the same as the number of unique loops formed by $O_{a,d}(x)$ where $x \in \mathbb{Z}/a\mathbb{Z}$. In order to count the equivalence classes, we identify them with orbits of a group action and use P\'{o}lya's Enumeration Theorem to count them. \subsection{Group Action} Consider the sub-trajectory $(n_i)$ as defined in the proof of Proposition 2. We observe that given a trajectory $O_{a,d}(x)$, one can construct the sub-trajectory $(n_i)$ and vice-versa. Hence, the number loops of the sub-trajectories have a one-to-one correspondence with the loops of the trajectories. By definition of the sub-trajectory $(n_i)$ $$n_{i+1} = d^{-1}n_i \mod a$$ For some $k$, all elements of the sub-trajectory are eventually smaller than $a$. This allows us to identify the limiting behavior of the sub-trajectory with $$(n_k,\: d^{-1}n_k \mod a,\: d^{-2}n_k \mod a,\: d^{-3}n_k \mod a,\: ...)$$ where $n_k \leq a$. Each element of the sub-trajectory is a power of $d^{-1}$ multiplied by $n_k$ and hence, the sub-trajectory can be identified with the action of the group of negative powers of $d$ on $n_k$ under multiplication modulo $a$. Under the binary operation of multiplication modulo a natural number $a$, the numbers co-prime to $a$ (modulo $a$) form an Abelian group denoted by $(\mathbb{Z}/a\mathbb{Z})^*$. For some $d$ in $(\mathbb{Z}/a\mathbb{Z})^*$, let $H$ be the subgroup generated by $d$. $$H = \{ d^i : i \in \mathbb{Z} \}$$ Then, the number of orbits formed by quotienting natural numbers by the nature of limiting behavior of the additive Collatz trajectory $O_{a,d}(x)$ is given by the number of orbits under the action of $H$ on $\mathbb{Z}/a\mathbb{Z}$ under the binary operation multiplication modulo $a$. \subsection{Computation} Let $\xi(a,d)$ denote the number of orbits of $(\mathbb{Z}/n\mathbb{Z})$ when $H$ acts on it. By P\'{o}lya's Enumeration Theorem, we have: $$\xi(a,d) = \frac{1}{|H|} \sum_{x \in S} |H_x|$$ where $H_x = \{g \in H: gx \equiv x \mod a \}$. $|H_x|$ can be computed by finding the number of solutions for $d^t \in H$ in the equation: \begin{equation} d^tx \equiv x \mod a \end{equation} Let $m_x = \gcd(x,a)$, $p_x = \frac{a}{m_x}$ and $q_x = \frac{x}{m_x}$. On substitution in equation (2), we have: \begin{equation} d^tm_xq_x \equiv m_xq_x \mod (m_xp_x) \end{equation} As $\gcd(p_x,q_x) = 1$, the number of solutions to equation (3) is same as the number of solutions to : $$d^t \equiv 1 \mod p_x$$ Let the smallest solution to equation be termed as $\alpha_{p_x} (d)$. Therefore the total number of solutions to equation (2) are: $\frac{|H|}{\alpha_{p_x}(d)}$ Hence, we have: $$\xi(a,d) = \frac{1}{|H|}\sum_{x \in S}\frac{|H|}{\alpha_{p_x}(d)} = \sum_{x \in S} \frac{1}{\alpha_{p_x}(d)}$$ For each $p_x$, $q_x$ takes values co prime to $p_x$, hence, by counting repetitions, we get: $$\xi(a,d) = \sum_{f | a} \frac{\phi(f)}{\alpha_{f}(d)}$$ where $\phi$ is the Euler-totient function. \subsection{Upper and Lower Bounds} We can set a lower bound on $\xi(a,d)$ by considering the Carmichael function $\lambda$. $$\lambda(m) = \max \{ \alpha_m(d) : d \in (Z/mZ) \}$$ Hence, $$\xi(a,d) = \sum_{f | a} \frac{\phi(f)}{\alpha_{f}(d)} \geq \sum_{f | a} \frac{\phi(f)}{\lambda(f)} = \xi_{inf}(a)$$ One can further claim that $\xi_{inf}(a)$ is a strong lower bound for $\xi(a,d)$ as for every $a$, there exists $d$ such that for all factors $f$ of $a$, $\alpha_{f}(d) = \lambda(f)$. The proof of this claim relies on the decomposition of $(\mathbb{Z}/n\mathbb{Z})^*$ into cyclic groups. The strong upper bound for $\xi(a,d)$ is $a$, which is attained when $d$ is $1$. \subsection{Applications} The computation of $\xi(a,d)$ employs factorization as well as the discrete logarithm function (in computation of $\alpha_{f}(d)$). There are no known efficient algorithms to compute either of them, making computation of $\xi(a,d)$ difficult. This difficulty can be employed for public-key cryptography. Knowing the prime factorization of $a$, can make the computation of $\xi(a,d)$ easier. Consider two primes $p$ and $q$. Then, for some $d$ co-prime to $pq$, $$\xi(pq,d) = 1 + \frac{\phi(p)}{\alpha_{p}(d)} + \frac{\phi(q)}{\alpha_{q}(d)} + \frac{\phi(pq)}{\alpha_{pq}(d)}$$ On simplification, we have: $$\xi(pq,d) = 1 + \frac{p-1}{\alpha_{p}(d)} + \frac{q-1}{\alpha_{q}(d)} + \frac{(p-1)(q-1)}{\alpha_{p}(d)\alpha_{q}(d)}\gcd(\alpha_{p}(d),\alpha_{q}(d))$$ Computing $\xi(pq,d)$ would be cumbersome without using equation (9), however, it is much simpler using the prime factorization. This provides much hope for the possibility of design of a public key cryptography algorithm or key exchange system using additive Collatz trajectories. \section{Results} This paper puts forward the concept of Additive Collatz Trajectories and provides an analysis of their limiting behavior. A necessary and sufficient condition is provided for eventual looping of the Additive Collatz trajectories, along with a formula to compute the number of unique trajectories possible up to the orbit equivalence relation. \section{Further Scope} \subsection{Generalized Collatz Trajectories} The spirit and strategy of this paper can be used to deal with generalized Collatz trajectories. One immediate result is that if the equation: \begin{equation} m^{r+1} x + a ( m^r + m^{r-1} + ... +m +1) \equiv 0 \mod d \end{equation} does not have a solution for any $r$, then for all $x \in \mathbb{N}$, the trajectory formed by iteration of $C_{a,d,m}'(x)$ would not loop. Also, if $ m \equiv 1 \mod d $, then the equation will always have a solution. We can then define a sub-trajectory similar to that done in Proposition 2 as: $$n_0 = x$$ $$n_{i} = C_{a,d,m}'^{(z_i)}(x)$$ where $C_{a,d,m}'^{(z_i - 1)}(x) = dC_{a,d,m}'^{(z_i)}(x)$. We will be able to show that: If $ d \not | n_i $ $$ n_{i+1} = \dfrac{m^{r_i}n_i + a( m^{r_i-1} + ... +m +1 )}{d}$$ where $ r \equiv -a^{-1}n_i \mod d$ and $ 0 < r_i \leq (d-1) $ \\ else,$$ n_{i+1} = \dfrac{n_i}{d} $$ \subsection{Public-key Cryptography} As mentioned in subsection 3.4, there is a hope for developing a public-key cryptography system that relies on Additive Collatz Trajectories, particularly on counting the number of equivalence classes formed under the orbit equivalence relation. Much effort and study is required to design an implementable cryptography design, as there are a number of challenges. Firstly, there is no trivial characteristic that is common among the elements of an orbit equivalence class. Secondly, the formula for counting the number of orbit equivalence classes employs a number of functions, hence there is no natural way to compute the inverse for decryption. Lastly, the formula uses the discrete logarithm function which cannot practically be computed for larger cases. One must deal with these challenges in the process of an encryption algorithm design using the results proved in this paper.
{ "timestamp": "2016-11-15T02:01:51", "yymm": "1611", "arxiv_id": "1611.03919", "language": "en", "url": "https://arxiv.org/abs/1611.03919", "abstract": "Collatz Conjecture (also known as Ulam's conjecture and 3x+1 problem) concerns the behavior of the iterates of a particular function on natural numbers. A number of generalizations of the conjecture have been subjected to extensive study. This paper explores Additive Collatz Trajectories, a particular case of a generalization of Collatz conjecture and puts forward a sufficient and necessary condition for looping of Additive Collatz Trajectories, along with two minor results. An algorithm to compute the number of equivalence classes when natural numbers are quotiented by the limiting behavior of their corresponding trajectories is also proposed.", "subjects": "Number Theory (math.NT)", "title": "Additive Collatz Trajectories", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9912886165717258, "lm_q2_score": 0.8080672135527632, "lm_q1q2_score": 0.801027830219688 }
https://arxiv.org/abs/1711.07288
When Fourth Moments Are Enough
This note concerns a somewhat innocent question motivated by an observation concerning the use of Chebyshev bounds on sample estimates of $p$ in the binomial distribution with parameters $n,p$. Namely, what moment order produces the best Chebyshev estimate of $p$? If $S_n(p)$ has a binomial distribution with parameters $n,p$, there it is readily observed that ${\rm argmax}_{0\le p\le 1}{\mathbb E}S_n^2(p) = {\rm argmax}_{0\le p\le 1}np(1-p) = \frac12,$ and ${\mathbb E}S_n^2(\frac12) = \frac{n}{4}$. Rabi Bhattacharya observed that while the second moment Chebyshev sample size for a $95\%$ confidence estimate within $\pm 5$ percentage points is $n = 2000$, the fourth moment yields the substantially reduced polling requirement of $n = 775$. Why stop at fourth moment? Is the argmax achieved at $p = \frac12$ for higher order moments and, if so, does it help, and compute $\mathbb{E}S_n^{2m}(\frac12)$? As captured by the title of this note, answers to these questions lead to a simple rule of thumb for best choice of moments in terms of an effective sample size for Chebyshev concentration inequalities.
\section{Introduction} This note concerns a somewhat innocent question motivated by an observation concerning the use of Chebyshev bounds on sample estimates of $p$ in the binomial distribution with parameters $n,p$. Namely, what moment order produces the best Chebyshev estimate of $p$? Chebyshev is arguably the most basic concentration inequality to occur in risk probability estimates and the use of second moments is a textbook example in elementary probability and statistics. Consider i.i.d. Bernoulli $0-1$ random variables $X_1,X_2,\dots,X_n$ with parameter $p\in[0,1]$, and let $S_n(p) = \sum_{j=1}^n(X_j-p)$. There it is readily observed that ${\rm argmax}_{0\le p\le 1}{\mathbb E}S_n^2(p) = {\rm argmax}_{0\le p\le 1}np(1-p) = \frac12$. It is also a well-known probability exercise to check that 4-th moment Chebyshev bounds improve the rate of convergence that can more generally be used for a proof of the strong law of large numbers, e.g. see (Bhattacharya and Waymire, 2016; p.100). Somewhat relatedly, Rabi Bhattacharya (personal communication) recently noticed, after a mildly tedious calculation to check ${\rm argmax}_{0\le p\le 1}{\mathbb E}S_n^4(p) = \frac12$, that the second moment Chebyshev bound is rather significantly improved by consideration of fourth moments as well. In particular, while the second moment Chebyshev sample size for a $95\%$ confidence estimate within $\pm 5$ percentage points is $n = 2000$, the fourth moment yields the substantially reduced polling requirement of $n = 775$. While the Chebyshev inequality is one among several inequalities used to obtain sample estimates, it is no doubt the simplest; see (Bhattacharya and Waymire, 2016) for comparison of fourth order Chebyshev to other concentration inequality bounds, and (Skinner, 2017) for numerical comparisons to higher order Chebyshev bounds. So why stop at fourth moments? Is ${\rm argmax}_{0\le p\le 1}\mathbb{E}S_n^{2m}(p) = \frac12$ for all $m,n$ and, if so, does it improve the estimate? Somewhat surprisingly we were not able to find a resolution of such basic questions in the published literature. In any case, with the argmax question resolved in part $(a)$ of the theorem below, part $(b)$ provides a direct computation of $\mathbb{E}S_n^{2m}(\frac12)$. Part $(c)$ then provides a more readily computable version. \begin{theorem}\label{mainthm} \begin{enumerate}[label=$(\alph*)$] \item For all $m\ge 1$ and all $n$ sufficiently large, ${\rm argmax}_{0\le p\le 1}\mathbb{E}S^{2m}_n(p) = \frac12$. \item For all positive $m$ and $n$, $\mathbb{E}S^{2m}_n(\frac12) = 4^{-m}\sum_{\mu\in\pi(m), |\mu|\le m\wedge n}\binom{2m}{2\mu_1,\dots,2\mu_{|\mu|}} \binom{n}{|\mu|}$, \item For all positive $m$ and $n$, ${\mathbb E}S^{2m}_n(\frac12) = 2^{-2m-n}\sum_{k=0}^n \binom{n}{k} (2k-n)^{2m}$. \end{enumerate}\noindent Here $\pi(m)$ is the set of ordered integer partitions of $m$, also referred to as integer compositions, and $|\mu|$ denotes the number of parts of $\mu\in\pi(m)$. We refer to $|\mu|$ as the {\it size} of the partition $\mu$. \end{theorem} The equivalent calculus challenge is to show for fixed $m$ that for all sufficiently large $n$, \begin{equation}\label{eqMoment} \text{argmax}_{0\le p\le 1}\frac{d^{2m}}{dt^{2m}}(pe^{qt} + qe^{-pt})^n|_{t=0} = \frac12. \end{equation} The example below illustrates the challenge to locating absolute maxima for such polynomials (in $p$), especially to proofs by mathematical induction. The proof given here is based on explicit combinatorial computation of ${\mathbb E}S_n^{2m}(p)$ in terms of ordered partitions of $2m$, after introducing a few preliminary lemmas. The lemmas are relatively simple to check using the statistical independence and identical distributions of the terms $X_i-p$ and $X_j-p$, $i\neq j$, and make good exercises in calculus, probability, and number theory. However let us first observe that part $(a)$ of the theorem does not hold for $m > n$. \ \noindent {\bf Counter example to Theorem \ref{mainthm}$(a)$ for (small) $n< m$:} Observe for $n = 1$ and $m=2$, the function $${\mathbb E}S_1^4(p) = p - 4p^2 + 6p^3 - 3p^4, \quad 0\le p\le 1,$$ has a {\it minimum} at $p=\frac12$, with two local maxima at $\frac12 \pm \frac{\sqrt{2}}{4}$. In particular, $${\rm argmax}_{0\le p\le 1}{\mathbb E}S_1^4(p) = \frac12 \pm \frac{\sqrt{2}}{4}.$$ In particular, the polynomial is generally {\it not} unimodal. So the restriction to sufficiently large $n$ is necessary for part $(a)$ of Theorem \ref{mainthm}. There is also the question of how large is sufficiently large. We do not address this here, but computations suggest a bound along the lines of $m\le c\cdot n^{\varepsilon}$, with $\varepsilon$ a little less than $\frac12$. We let $m_n$ denote the largest value of $m$, dependent on $n$, such that Theorem \ref{mainthm}$(a)$ holds for all $m\le m_n$. We leave it as an open problem to determine an exact formula for $m_n$ and to determine a formula for ${\rm argmax}_{0\le p\le1}\mathbb{E}_n^{2m}(p)$ for $m>m_n$. \section{Proofs and Remarks} Let $\pi(2m)$ denote the set of ordered partitions of $2m$. We we will use $|\mu| = k$ to denote the number of parts of $\mu$. Finally, for $\mu\in\pi(2m)$, let \begin{equation*} f_i(\mu,p) = pq^{\mu_i} + q(-p)^{\mu_i}, \quad 0\le p\le 1, q=(1-p), 1\le i\le |\mu|. \end{equation*} \begin{lemma} \label{basiclemma} Let $0\le p\le 1$ and $q = 1-p$. The following hold, \begin{enumerate}[label=$(\alph*)$] \item $S_n(p) = -^{\text{dist}} S_n(q)$, \item $\mathbb{E}S^{2m}_n(p) = \mathbb{E}S^{2m}_n(q)$, \item $\mathbb{E}S^{2m}_n(p) = \sum_{\mu\in\pi(2m)}\binom{n}{|\mu|} \binom{2m}{\mu_1,\dots,\mu_{|\mu|}}\prod_{i=1}^{|\mu|}f_i(\mu,p)$, \item $\frac{d}{dp}\mathbb{E}S^{2m}_n(p) = \sum_{\mu\in\pi(2m)}\binom{n}{|\mu|} \binom{2m}{\mu_1,\dots,\mu_{|\mu|}}\sum_{i=1}^{|\mu|} f_i^\prime(\mu,p)\prod_{j\ne i}^{|\mu|}f_j(\mu,p)$. \end{enumerate} \end{lemma} \begin{lemma} \label{derivlem} Let $\mu\in\pi(2m)$ and $1\le i\le |\mu|$. Then, $$\frac{d}{dp}f_i(\mu,p) = q^{\mu_i}\left(1-\frac{p}{q}\mu_i\right) + (-1)^{\mu_i+1} p^{\mu_i}\left(1-\frac{q}{p}\mu_i\right).$$ \end{lemma} It now follows easily that \begin{equation} \label{fi} f_i\left(\mu,\frac12 \right) = \begin{cases} 2^{-\mu_i} & \text{for even}\ \mu_i,\\ 0 & \text{for odd}\ \mu_i, \end{cases} \end{equation} \begin{equation} \label{fiprime} f^\prime_i\left(\mu,\frac12 \right) = \begin{cases} 0 & \text{for even}\ \mu_i,\\ -2(\mu_i-1)2^{-\mu_i} & \text{for odd}\ \mu_i. \end{cases} \end{equation} The keys to the following proof of Theorem \ref{mainthm} reside in (1) the parity conflicts between \eqref{fi} and \eqref{fiprime} and (2) the expansion $(d)$ in Lemma \ref{basiclemma}, viewed as a polynomial in $n$. \begin{proof}[Proof (of theorem)] That $p=\frac12$ is a critical point follows from $(d)$ of Lemma \ref{basiclemma} together with \eqref{fi} and \eqref{fiprime} by examining the terms $f_i^\prime(\mu,\frac12)\prod_{j\ne i}^{|\mu|}f_j(\mu,\frac12)$. In particular, for partitions of $2m$, if $\mu_i$ is odd then there must be a $j\neq i$ such that $\mu_j$ is odd as well. To see that $p=\frac12$ is an absolute maximum, the trick is to observe that for $0\le p < \frac12 < q$, the leading coefficient of $\frac{d}{dp}\mathbb{E}S_n^{2m}(p)$, viewed as a polynomial in $n$, is obtained at the $m$-part composition $\mu = (2,2,\dots,2)$ of $2m$. Namely, it is obtained from $\binom{n}{m}\binom{2m}{2,2,\dots,2}m(q^2-p^2)(pq)^{m-1}$, and therefore is positive for all $p<\frac12$. Thus, for sufficiently large $n$, $$\frac{d}{dp}\mathbb{E}S_n^{2m}(p) > 0, \quad\mbox{for } 0\le p < 1/2.$$ In view of the symmetry expressed in $(b)$ of Lemma \ref{basiclemma}, it follows that $p =\frac12$ is the unique global maximum. For part $(b)$ of the theorem one simply computes from independence, writing $\tilde{X}_i = X_i-\frac12, i =1,2,\dots, n$. In particular, $\tilde{X}_i = \pm\frac12$ with equal probabilities. So, for $m\ge 1$, \begin{align*} \mathbb{E}S_n^{2m}\left(\frac12\right) &= \sum_{1\le j_1,\dots,j_{2m}\le n}{\mathbb E}\prod_{i=1}^{2m}\tilde{X}_{j_i} \\ &= \sum_{2m_1+\cdots +2m_n = 2m}\prod_{i=1}^{n}{\mathbb E}\tilde{X}_{i}^{2m_i} \\ &= \sum_{k=1}^{m\wedge n}\sum_{2m_1+\cdots +2m_n = 2m, \#\{j:m_j\ge 1\}=k}\prod_{i=1}^{n}4^{-m_i} \\ &= \sum_{k=1}^{m\wedge n}\binom{n}{k}\sum_{\mu=(\mu_1,\dots,\mu_k)\in\pi(m)} \binom{2m}{2\mu_1,\dots,2\mu_k}4^{-m}. \end{align*} Here one adopts the convention that a sum over an empty set is zero so that if there are no partitions $\mu$ of $m$ with $|\mu|=k$ then the indicated sum is zero for this choice of $k$. So nonzero contributions to the sum are provided by ordered partitions $\mu$ of size $|\mu|\le m\wedge n$. To simplify the computation in terms of ordered partitions $(b)$ one may proceed as follows to obtain the formula in $(c)$. We instead compute $\mathbb{E}S_n^{2m}(\frac12)$ as the $2m$-th moment of $S_n(\frac12)$ as given in \eqref{eqMoment}. By the binomial theorem, we have that \begin{align*} \mathbb{E}S_n^{2m}\left(\frac12\right) &= \frac{d^{2m}}{dt^{2m}} \left[\left( \frac{e^{\frac{t}{2}}}{2} + \frac{e^{-\frac{t}{2}}}{2} \right)^n\right]_{t=0} = \frac{d^{2m}}{dt^{2m}} \left[ 2^{-n} \sum_{k=0}^n \binom{n}{k} e^{\frac{t}{2}(2k-n)} \right]_{t=0} \\&= 2^{-n-2m}\sum_{k=0}^n \binom{n}{k} (2k-n)^{2m}. \end{align*} \end{proof} \begin{remark} A linear recurrence in $m$ is possible to aid the pre-asymptotic (in $n$) computation of $\mathbb{E}S_n^{2m}(\frac12)$. Namely, \begin{align} \label{recur} \mathbb{E}S_n^{2m+2\ell+2}\left(\frac12\right) &= \sum_{j=0}^\ell c_j 2^{2j-2\ell-2}\mathbb{E}S_n^{2m+2j}\left(\frac12\right), \end{align} where $\ell=\left\lfloor\frac{n-1}{2}\right\rfloor$, $a_k = (2k-n)^2$, and $(c_0,c_1,\dotsc,c_\ell)$ is the unique solution to \begin{align*} \begin{pmatrix} a_0^0&a_0^1&\dots&a_0^\ell\\ a_1^0&a_1^1&\dots&a_1^\ell\\ \vdots \\ a_\ell^0&a_\ell^1&\dots&a_\ell^\ell\\ \end{pmatrix} \begin{pmatrix} c_0\\c_1\\ \vdots\\ c_\ell \end{pmatrix} &= \begin{pmatrix} a_0^{\ell+1}\\ a_1^{\ell+1}\\ \vdots\\ a_\ell^{\ell+1} \end{pmatrix}. \end{align*} To see this, write \begin{align*} \mathbb{E}S_n^{2m}\left(\frac12\right) &= 2^{-2m-n+1}\sum_{k=0}^{\ell} \binom{n}{k} (2k-n)^{2m} . \end{align*} Then \eqref{recur} follows since \begin{align*} &\mathbb{E}S_n^{2m+2\ell+2}\left(\frac12\right) - \sum_{j=0}^\ell c_j 2^{2j-2\ell-2}\mathbb{E}S_n^{2m+2j}\left(\frac12\right) \\ &= 2^{-2m-2\ell-n-1}\sum_{k=0}^{\ell}\binom{n}{k}a_k^{m+\ell+1} - \sum_{j=0}^{\ell}c_j 2^{-2m-2\ell-n-1} \sum_{k=0}^{\ell}\binom{n}{k}a_k^{m+j} \\ &= 2^{-2m-2\ell-n-1}\sum_{k=0}^{\ell}\binom{n}{k}a_k^{m} \left(a_k^{\ell+1}-\sum_{j=0}^{\ell}c_ja_k^j \right) =0. \end{align*} \end{remark} \ \ For the application to statistical estimation one may combine Theorem \ref{mainthm} with Chebyshev's inequality to obtain, \begin{corollary} For $\epsilon > 0$, we have that $ P\left(|\frac{1}{n}S_n(p)| > \epsilon\right) \le \min_{1\le m\le m_n} \left( \frac{\sqrt[2m]{{\mathbb E}S^{2m}_n(\frac12)}} {n\epsilon}\right)^{2m}. $ \end{corollary} Noting the scaling invariance $ \text{argmax}_{0\le p \le 1}\mathbb{E}S_n^{2m}(p) = \text{argmax}_{0\le p\le 1}\mathbb{E}\frac{S_n^{2m}(p)}{n^m}, $ and ${\mathbb E}Z^{2m} = 2^{-m}\frac{(2m)!}{m!}$ for the standard normal random variable $Z$, in the limit ``$n \to\infty, \epsilon \to 0, n\epsilon^2\to \tilde{n}$'' one has \begin{equation*} B_m := \mathbb{E}\frac{S^{2m}_n(\frac12)}{n^{2m}\epsilon^{2m}} = \mathbb{E} \frac{\left(\frac{S_n(\frac12)}{\sqrt{n/4}}\right)^{2m}}{n^{2m}\epsilon^{2m}} \left(\frac{n}{4}\right)^m \to 2^{-2m}\tilde{n}^{-m}{\mathbb E}Z^{2m} = 2^{-3m}\frac{(2m)!}{m!}\tilde{n}^{-m}. \end{equation*} In particular, one may ask for the best choice of $m$ for large $n$, i.e, in the above limit as $n\to\infty,\epsilon\downarrow 0, n\epsilon^2\to \tilde{n}$. The quantity $\tilde{n} = n\epsilon^2$ denotes an {\it effective sample size} in the sense of the risk assessment defined by $P(|S_n(p)| > n\epsilon) < \epsilon$; see (Duchi et al. 2013) for an introduction of this artful terminology in a much broader context. Observe that in the limit of large $n$ \begin{equation*} \lim_{n\to\infty, \epsilon\downarrow 0, n\epsilon^2 = \tilde{n}} \frac{B_{m+1}}{B_m} = \frac{2m+1}{4\tilde{n}} \begin{cases} \le 1\\ = 1\\ \ge 1 \end{cases} \end{equation*} if and only if \begin{equation*} m \begin{cases} \le 2\tilde{n} -\frac12\\ = 2\tilde{n} -\frac12\\ \ge 2\tilde{n} -\frac12. \end{cases} \end{equation*} \ \ The take-away is perhaps best summarized in terms of the following informally interpreted optimal estimation principle. \ \ \noindent {\bf Approximate Rule of Thumb:} {\it For large $n$ the optimal moment order $2m$ for the Chebyshev bound is quadruple the effective sample size. In particular, the fourth moment is optimal for a one unit effective sample size!} \section{References} \noindent Bhattacharya, R.N., and Waymire, E.C. (2016), ``A Basic Course in Probability Theory'', 2nd ed., Universitext, Springer, NY. \ \noindent Duchi, J. and Wainwright, M. J. and Jordan, M. I. (2013), ``Local privacy and minimax bounds: Sharp rates for probability estimation'', in Advances in Neural Information Processing Systems, 1529-1537. \ \noindent Skinner, Dane, 2017, ``Concentration of Measure Inequalities'', Master of Science, Oregon State University \ \end{document} \begin{bibdiv} \begin{biblist} \bibselect{DSEW_bib} \end{biblist} \end{bibdiv} \end{document}
{ "timestamp": "2017-11-21T02:17:22", "yymm": "1711", "arxiv_id": "1711.07288", "language": "en", "url": "https://arxiv.org/abs/1711.07288", "abstract": "This note concerns a somewhat innocent question motivated by an observation concerning the use of Chebyshev bounds on sample estimates of $p$ in the binomial distribution with parameters $n,p$. Namely, what moment order produces the best Chebyshev estimate of $p$? If $S_n(p)$ has a binomial distribution with parameters $n,p$, there it is readily observed that ${\\rm argmax}_{0\\le p\\le 1}{\\mathbb E}S_n^2(p) = {\\rm argmax}_{0\\le p\\le 1}np(1-p) = \\frac12,$ and ${\\mathbb E}S_n^2(\\frac12) = \\frac{n}{4}$. Rabi Bhattacharya observed that while the second moment Chebyshev sample size for a $95\\%$ confidence estimate within $\\pm 5$ percentage points is $n = 2000$, the fourth moment yields the substantially reduced polling requirement of $n = 775$. Why stop at fourth moment? Is the argmax achieved at $p = \\frac12$ for higher order moments and, if so, does it help, and compute $\\mathbb{E}S_n^{2m}(\\frac12)$? As captured by the title of this note, answers to these questions lead to a simple rule of thumb for best choice of moments in terms of an effective sample size for Chebyshev concentration inequalities.", "subjects": "Probability (math.PR); Statistics Theory (math.ST)", "title": "When Fourth Moments Are Enough", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9825575183283514, "lm_q2_score": 0.8152324871074608, "lm_q1q2_score": 0.8010128093929564 }
https://arxiv.org/abs/2205.10968
Eigenvalue bounds of the Kirchhoff Laplacian
We prove that each eigenvalue l(k) of the Kirchhoff Laplacian K of a graph or quiver is bounded above by d(k)+d(k-1) for all k in {1,...,n}. Here l(1),...,l(n) is a non-decreasing list of the eigenvalues of K and d(1),..,d(n) is a non-decreasing list of vertex degrees with the additional assumption d(0)=0. We also prove that in general the weak Brouwer-Haemers lower bound d(k) + (n-k) holds for all eigenvalues l(k) of the Kirchhoff matrix of a quiver.
\section{The theorem} \paragraph{} Let $G=(V,E)$ be a {\bf finite simple graph} with $n$ vertices. Denote by $\lambda_1 \leq \lambda_2 \leq \cdots \leq \lambda_n$ the ordered list of eigenvalues of the {\bf Kirchhoff matrix} $K=B-A$, where $B$ is the diagonal vertex degree matrix with ordered vertex degrees $d_1 \leq d_2 \leq \cdots \leq d_n$ and where $A$ is the adjacency matrix of $G$. \paragraph{} We assume $d_0=0$ so that $d_1+d_0=d_1$ and prove \begin{thm} $\lambda_k \leq d_k+d_{k-1}$, for all $1 \leq k \leq n$. \label{1} \end{thm} The case $d=n$ is the {\bf spectral radius} estimate $\lambda_n \leq d_n+d_{n-1}$ of Anderson and Morley \cite{AndersonMorley1985} which we will use in the proof. The case $k=1$ is obvious because $\lambda_1=0$ and $k=2$ is a special case of the Schur-Horn inequality \cite{Brouwer} as $\lambda_2 = \lambda_1+\lambda_2 \leq d_1+d_2$. Already the case $\lambda_3 \leq d_3+d_2$ appears to be new as it goes beyond the Schur-Horn inequality. Note that any estimate on the spectral radius $\lambda_n$ provides upper bounds on $\lambda_k$ and there are many improvements since the ground breaking Anderson-Morley paper \cite{GroneMerrisSunder1,GroneMerrisSunder2,Zhang2004,Das2004, Guo2005,Shi2007,ShiuChan2009, LiShiuChan2010,ZhouXu}. They would give better upper bounds also for $\lambda_k$ but look less elegant. \paragraph{} Already the corollary $\lambda_k \leq 2d_k$ is stronger than what the {\bf Gershgorin circle theorem} \cite{Gershgorin,GershgorinAndHisCircles} gives in this case: the circle theorem provides in every interval $[0,2d_k]$ at least one eigenvalue $\lambda_l$ of $K$. It does not need to be the k'th one. In the Kirchhoff case, where the Gershgorin circles are nested, $0$ is always in the spectrum. Theorem~(\ref{1}) gives more information. The spectral data $\lambda_1=0,\lambda_2=10,\lambda_3=10$ for example are Gershgorin compatible to $d_1=1, d_2=3, d_3=7$ because there is an eigenvalue in each closed ball $[0,2], [0,6]$ and $[0,14]$. But these data contradict Theorem~(\ref{1}) as $\lambda_2 = 2 d_2+4$. Theorem~\ref{1} keeps the eigenvalues more aligned with the degree sequence, similarly as the Schur-Horn theorem does. \paragraph{} An application is that the pseudo determinant ${\rm Det}(K)= \prod_{k=2}^n \lambda_k$ which by the Kirchhoff-Cayley {\bf matrix tree theorem} count the number of rooted spanning trees has an upper bound $2^n \prod_k d_k$ and that ${\rm det}(1+K)$ by the Chebotarev-Shamis matrix forest theorem count the number of rooted spanning forests has an upper bound $2^n \prod_k (1+d_k)$. These determinant inequalities do not follow from the Gershgorin nor from the Schur-Horn inequalities. We like to rephrase the determinant bounds as bound on the spectral potential $U(z)=(1/n) \log \det(K-z) \leq 2 + (1/n) \log\det(M-z)$, where $M={\rm Diag}(d_1,\dots,d_n)$ is the diagonal matrix and $z \leq 0$ is real. We made use of Theorem~(\ref{1}) to show that $U(z)$ has a Barycentric limit. It bounds the potential $U$ of the interacting system with the potential $(1/n) \log\det(M-z)$ of the non-interacting system where $M$ can be thought of as the Kirchhoff matrix of a non-interacting system where we have $d_k$ self-loops at vertices $k$, even-so the tree and forest interpretations do not make sense any more. \section{The proof} \paragraph{} If we make he statement slightly stronger, also the induction assumption becomes more powerful. Let $\mathcal{K}$ the class of symmetric matrices which are obtained as principal submatrices of a Kirchhoff matrices of a finite simple graph obtained by deleting the rows and columns to a maximal diagonal entry. In other words, we close the class of Kirchhoff matrices under the operation of taking principal submatrices. The stronger statement is that for all $A \in \mathcal{K}$, the eigenvalues $\lambda_k$ and diagonal entries $d_k$, when ordered in an ascending order, satisfy $\lambda_k \leq d_k+d_{k-1}$. \paragraph{} The class $\mathcal{K}$ is invariant under the operation $\phi$ which removes corresponding rows and columns. This allows induction. The induction foundation $n=1$ is obvious because the result holds for any $1 \times 1$ matrix with non-negative entries. For $2 \times 2$ matrices $K=\left[ \begin{array}{cc} d_1 & -t \\ -t & d_2 \end{array} \right]$ we have with the trace $T=d_1+d_2$ the eigenvalue $\lambda_2= (T+\sqrt{T^2+4t^2-4d_1 d_2})/2$ which is $\leq T=d_1+d_2$ if $t \leq \sqrt{d_1 d_2}$ and especially if $t \leq d_1$. \paragraph{} The induction step uses the {\bf Cauchy interlace theorem} or {\bf Separation theorem} \cite{HornJohnson2012}: if $\mu_k$ are the eigenvalues of $\phi(K)$, then $\lambda_k \leq \mu_k \leq \lambda_{k+1}$. (The interlace theorem follows from the {\bf Hermite interlace theorem} for real polynomials \cite{Hwang2004,Fisk2005}. If $f$ is a monic polynomial of degree $n$ with real roots $\lambda_1 \leq \cdots \leq \lambda_n$ and $g$ is a monic polynomial of degree $n-1$ with real roots $\mu_1 \leq \cdots \leq \mu_{n-1}$ then {\bf $g$ interlaces $f$} if $\lambda_1 \leq \mu_1 \leq \lambda_2 \leq \cdots \leq \mu_{n-1} \leq \lambda_n$. Equivalent for $f$ and $g$ to interlace is that the interpolation $t f+(1-t)g$ has real roots for all $t \in [0,1]$.) \paragraph{} The interlace theorem does not catch the largest eigenvalue $\lambda_n$. This requires an upper bound for the {\bf spectral radius}. This is where Anderson and Morley \cite{AndersonMorley1985} come in. They realized that $K=F^* F$ is essentially isospectral to $F F^*$. The later is the adjacency matrix of the line graph but with the modification that all diagonal entries are $2$. Since the row or column sum of $F F^*$ is bound by $d_k(a) + d_k(b)$ for any edge $(a,b)$, the upper bound holds. But we need to extend this estimate to $\mathcal{K}$. \begin{lemma}[Upgrade of Anderson and Morley] For any matrix in $\mathcal{K}$, the spectral radius satisfies $\lambda_n \leq d_n + d_{n-1}$. \end{lemma} \paragraph{} We can assume that after a coordinate change the diagonal entries of $K \in \mathcal{K}$ are ordered. This simply requires to order the vertices compatible with the order of the $d_j$. \cite{AndersonMorley1985} estimate $\lambda_n \leq d_n+d_{n-1}$ for Kirchhoff matrices and even show that $\lambda_n \leq {\rm max}_{(a,b) \in E(G)} d(a)+d(b)$. We now verify that this extends to $\mathcal{K}$. \paragraph{} We have already seen the induction assumption for $n=1$ and $n=2$. Let $K$ be a $n \times n$ matrix in $\mathcal{K}$ and assume we know the answer already for all $(n-1) \times (n-1)$ matrices in $\mathcal{K}$. When taking a principal submatrix of $K$ by deleting the first row and first column, we do not change $d_n$ nor $d_{n-1}$. The reason is that decoupling the lowest degree vertex does not change the largest two diagonal elements. The upper bound $d_n+d_{n-1}$ is therefore not affected. \paragraph{} Let $\mu_1 \leq \mu_2 \leq \cdots \leq \mu_{n-1}$ be the eigenvalues of the principal submatrix where the first row and column are deleted. The Cauchy interlacing result gives $\lambda_{n-1} \leq \mu_{n-1} \leq \lambda_n$ The largest eigenvalue $\mu_{n-1}$ of the deformed matrix is smaller or equal than than the largest eigenvalue $\lambda_n$ and so smaller or equal than $d_n + d_{n-1}$. The inequality $\mu_{n-1} \leq d_n + d_{n-1}$ is what we wanted to show. \paragraph{} One could also explicitly track how the maximal eigenvalue decreases if all the non-diagonal entries in the first row and column is multiplied by $t$. If $v$ is the normalized eigenvector to the largest eigenvalue $\lambda_n$, then since the matrices are symmetric, $v$ is perpendicular to $[1,1,\dots, 1,1]$ the eigenvector to $\lambda_1=0$. This implies $\sum_{k=1}^{n-1} v_k = -v_1$. Let now $E$ be the matrix in which the first row and first column except $E_{11}$ contains only 1 and everything else is zero. We need to understand how $K - t E$ changes the largest eigenvalue $\lambda$. The {\bf first Hadamard deformation formula} gives $\lambda(t)' = v^T E v = v_1^2>0$. The {\bf second Hadamard deformation formula} would even show $\lambda(t)''>0$, illustrating {\bf eigenvalue repulsion}. In any case, the largest eigenvalue increases under the deformation. Since at $t=1$, the end of the deformation, we have $\lambda_n \leq d_n+d_{n-1}$ this is also the case for $t=0$, where the connections to the weak vertex link have all been capped. \begin{comment} s = RandomGraph[{20, 50}]; a = RandomChoice[VertexList[s]]; S = VertexList[NeighborhoodGraph[s, a]]; A = Normal[KirchhoffMatrix[s]]; Q[t_]:=Module[{U = A},Do[U[[a, S[[k]]]] = -t; U[[S[[k]], a]] = -t, {k, Length[S]}]; U] F[t_]:=Module[{}, B = Q[t]; l = Max[Sort[N[Eigenvalues[B]]]]; d = Sort[Table[B[[k, k]], {k, Length[B]}]]; d1 = Prepend[Delete[d, Length[d]], 0]; l] Plot[F[t], {t, 0, 1}] \end{comment} \section{Remarks} \paragraph{} Similarly than the Schur-Horn inequality $\sum_{j=1}^k \lambda_j \leq \sum_{j=1}^k d_j$ which is true for any symmetric matrix with diagonal entries $d_j$, Theorem~(\ref{1}) controls how close the ordered eigenvalue sequence is to the ordered vertex degree sequence. But it is different. For example, if the eigenvalues increase exponentially like $\lambda_k = 6^k$, then the inequality $\lambda_k \leq 2 d_k$ implies on a logarithmic scale $\log(\lambda_k) \leq \log(2) + \log(d_k)$. Schur-Horn does not provide that. As for the {\bf Gershgorin circle theorem}, it would only establish $\log(\lambda) \leq \log(2) + \log(d_n)$, where $d_n$ is the largest entry because the theorem assures only that in each Gershgorin circle, there is at least one eigenvalue. \paragraph{} Anderson and Morley \cite{AndersonMorley1985} have the better bound $\max_{(x,y) \in E} d(x) + d(y)$ and Theorem~(\ref{1}) could be improved in that $d_k+d_{k-1}$ can be replaced by ${\rm max}_{(x,y) \in E} d_k+d_j$. Any general better upper bound of the spectral radius leads to better results. The Anderson-Morley estimate as an early use of a {\bf McKean-Singer super symmetry} \cite{McKeanSinger} (see \cite{knillmckeansinger} for graphs). in a simple case where one has only $0$-forms and $1$-forms. There, it reduces to the statment that $K=F^* F$ is essentially isospectral to $F F^*$ which is true for all matrices. It uses that the Laplacian $K$ is of the form $F^* F = {\rm div} {\rm grad}$, where $F={\rm grad}$ is the {\rm incidence matrix} $F f( (a,b) ) = f(b)-f(a)$ for functions $f$ on vertices (0-forms) leading to functions on oriented edges (1-form). \paragraph{} There is much work on the spectral radius of the Kirchhoff Laplacian. It is bounded above by the spectral radius of the {\bf signless Kirchhoff Laplacian} $|K|$ in which one takes the absolute values for each entry. This matrix is a non-negative matrix in the connected case has a power $|K|^n$ with all positive entries so that by the Perron-Frobenius theorem, the maximal eigenvalue is unique. (The Kirchhoff matrix itself of course can have multiple maximal eigenvalues like for the case of the complete graph). Also, unlike $K$ which is never invertible $|K|$ is invertible if $G$ is not bipartite. If we treat a graph as a one-dimensional simplicial complex (ignoring 2 and higher dimensional simplices in the graph), and denote $d$ the exterior derivative of this skeleton complex, then $(d+d^*)^2 = K_0 + K_1$ where $K=K_0=d^* d$ is the Kirchhoff matrix and $K_1=d d^*$ is the one-form matrix with the same spectral radius. This leads to \cite{AndersonMorley1985}. Much work has gone in improving this \cite{BrualdiHoffmann,Stanley1987,LiZhang1997,Zhang2004,FengLiZhang,Shi2007,LiShiuChan2010}. We have used that an identity coming from connection matrices \cite{Hydrogen}. \paragraph{} We stumbled on the theorem when looking for bounds on the {\bf tail distribution} $\mu([x,\infty))$ of the {\bf limiting density of states} $\mu$ of the Barycentric limit $\lim_{n \to \infty} G_n$ of a finite simple graph $G=G_0$, where $G_n$ is the Barycentric refinement of $G_{n-1}$ in which the complete subgraphs of $G_{n-1}$ are the vertices and two are connected if one is contained in the other \cite{KnillBarycentric,KnillBarycentric2}. We wanted a {\bf potential} $U(z) = \int_0^{\infty} \log(z-w) \; d\mu(w)$ because $U(-1)$ measures the exponential growth of rooted spanning trees while $U(0)$ measures the exponential growth of the rooted spanning forests. The connection is that in general, the {\bf pseudo determinant} \cite{cauchybinet} ${\rm Det}(K)=\prod_{\lambda \neq 0} \lambda$ is the number of {\bf rooted spanning trees} in $G$ by the {\bf Kirchhoff matrix tree theorem} and ${\rm det}(1+K)$ is the number of rooted spanning forests in $G$ by the {\bf Chebotarev-Schamis matrix forest theorem} \cite{ChebotarevShamis1,ChebotarevShamis2,Knillforest}. All these relations follow directly from the {\bf generalized Cauchy-Binet theorem} that states that for any $n \times m$ matrices $F,G$, one has the pseudo determinant version ${\rm Det}(F^T G) = \sum_{|P|=k} \det(F_P) \det(G_P)$ with $k$ depends on $F,G$ and $\det(1+x F^T G) = \sum_P x^{|P|} \det(F_P) \det(G_P)$. {\bf Pythagorean identities} like ${\rm Det}(F^T F) = \sum_{|P|=k} \det^2(F_P)$ and $\det(1+F^T F) = \sum_P \det^2(F_P)$ follow for an arbitrary $n \times m$ matrix $F$. Applied to the {\bf incidence matrix} $F$ of a connected graph, where $k(A)=n-1$ is the rank of $K=F^T F$, the first identity counts on the right spanning trees and the second identity counts on the right the number of spanning forests \paragraph{} Having noticed that the {\bf tree-forest ratio} $\tau(G)={\rm Det}(1+K)/{\rm Det}(K)=\prod_{\lambda \neq 0} (1+1/\lambda)$ has a Barycentric limit $\lim_{n \to \infty} (1/|G_n|) \log(\tau(G_n))$, we interpreted this as $U(-1)-U(0)$ requiring the normalized potential to exist. By the way, for complete graphs $G=K_n$ the tree-forest ratio is $(1+1/n)^{n-1}$ and converges to the {\bf Euler number} $e$. For triangle-free graphs, $\log(\tau(G_n))/|V(G_n)|$ converges to $\log(\phi^2)$, where $\phi$ is the {\bf golden ratio}. For example, for $G=C_n$, where ${\rm Det}(K)=n^2$ and the number of rooted spanning forests is the alternate {\bf Lucas number} recursively given by $L(n+1) = 3 L(n)-L(n-1)+2, L(0)=0,L(1)=1$. We proved in general that $\log(\tau(G_n))/|V(G_n))|$ converges under {\bf Barycentric refinements} $G_0 \to G_1 \to G_2 \dots$ for arbitrary graphs to a universal constant that only depends on the maximal dimension of $G=G_0$. \paragraph{} Theorem~(\ref{1}) needs to be placed in the context of the spectral graph literature like \cite{Biggs,VerdiereGraphSpectra,Godsil,Brouwer,Chung97,DvetkovicDoobSachs,Spielman2009} or articles like \cite{Guo2005,Stephen,GroneMerrisSunder}. Most research work in this area has focused on small eigenvalues like $\lambda_2$ or large eigenvalues like the spectral radius $\lambda_n$. For $\lambda_2$, there is a {\bf Cheeger estimate} $h^2/(2d_n) \leq \lambda_2 \leq h$ for the eigenvalue $\lambda_2$ and Cheeger constant $h=h(G)$ \cite{Cheeger1970} first defined for Riemannian manifolds, meaning in the graph case that one needs remove $h |H|$ edges to remove a subgraph $H$ from $G$. For the largest eigenvalue $\lambda_n$ the Anderson-Morley bound has produced an industry of results. \paragraph{} Let us look at some example. Figure~\ref{(1)} shows more visually what happens in some examples of graphs with $n=10$ vertices. \\ a) For the cyclic graph $C_4$, the Kirchhoff eigenvalues are $\lambda_1=0,\lambda_2=2,\lambda_3=2,\lambda_4=4$ and the edge degrees are $d_1=d_2=d_3=d_4=2$. \\ b) For the star graph with $n-1$ spikes, the eigenvalues are $\lambda_1=0,\lambda_2= \cdots \lambda_{n-1}=1, \lambda_n=n$ while the degree sequence is $d_1=\cdots = d_{n-1}=1, d_n=n$. \\ c) For a complete bipartite graph $K_{n,m}$ with $m \leq m$, we have $0 \leq m \leq \cdots \leq m \leq n \cdots n$ with $m-1$ eigenvalues $m$ and $n-1$ eigenvalues $n$. The degree sequence has $m \leq m \leq \cdots \leq m \leq n \leq \cdots \leq n$ with $m$ entries $m$ and $n$ entries $n$. \paragraph{} From all the $38$ isomorphism classes of connected graphs with $4$ vertices, there are only $3$, for which equality holds in Theorem~(\ref{1}) and $7$ for which $\lambda_k = 2 d_k$. From the $728$ isomorphism classes of connected graphs with $5$ vertices, there are none for Theorem~(\ref{1}) and $5$ for $\lambda_k = 2 d_k$. From the $26704$ isomorphism classes of connected graphs with $6$ vertices, there are $70$ for which Theorem~(\ref{1}) has equality and $76$ for which $\lambda_k = 2 d_k$. It is always only the largest eigenvalue, where we have seen equality $\lambda_n=2d_n$ to hold. \paragraph{} The theorem implies $\sum_{j=1}^k \lambda_j \leq 2\sum_{j=1}^k d_j$ which is weaker than the Schur-Horn inequality $\sum_{j=1}^k \lambda_j \leq \sum_{j=1}^k d_j$. The later result is of wide interest. It can be seen in the context of partial traces \cite{TaoSchurHorn} $\sum_{j=1}^k \lambda_j = {\rm inf}_{dim(V)=k} {\rm tr}(A|V)$ and is special case of the {\bf Atiyah-Guillemin-Sternberg convexity theorem} \cite{Atiyah1982,GuilleminSternberg1982}. The Schur inequality has been sharpened a bit for Kirchhoff matrices to $\sum_{j=1}^k \lambda_j \leq \sum_{j=1}^k d_j-1$ \cite{Brouwer} Proposition 3.10.1. \paragraph{} There is a general lower bound $\lambda_k \geq d_k-(n-k)+1$ which had been conjectured of Guo \cite{Guo2007} and is now a theorem \cite{BrouwerHaemers2008} (see also \cite{Brouwer} Proposition 3.10.2.). This Brouwer-Haemers estimate from 2008 generalizes $\lambda_n \geq d_n+1$ \cite{GroneMerrisSunder2} $\lambda_{n-1} \geq d_{n-1}$ (\cite{LiPan1999} and $\lambda_{n-2} \geq d_{n-2}-1$ \cite{Guo2007}. It follows also from $K=-A+B$ with $B={\rm Diag}(d_1,\cdots d_n)$ positive semi definite that $\lambda_k(K) \geq \lambda_k(-A)$ \cite{HornJohson2012} (Corollary 4.3.12). \paragraph{} Unlike the Schur-Horn inequality, Theorem~(\ref{1}) does not extend to general symmetric matrices. Already for $A=\left[ \begin{array}{cc} 1 & 3 \\ 3 & 1 \end{array} \right]$ which has eigenvalues $\lambda_1=-2, \lambda_2=4$ and with $d_1=1,d_2=1$, the inequality $\lambda_2 \leq 2 d_2$ fails. It also does not extend to symmetric matrices with non-negative eigenvalues, but also this does not work as $A=\left[ \begin{array}{ccc} 1 & 1 & 1 \\ 1 & 1 & 1 \\ 1 & 1 & 1 \end{array} \right]$ shows, as this matrix has eigenvalues $0,0,3$ and diagonal entries $1,1,1$. The case of $n \times n$ matrices with constant entries $1$ is an example with eigenvalues $0$ and $n$ showing that no estimate $\lambda_n \leq C d_n$ is in general possible for symmetric matrices even when asking the diagonal entry to dominate the other entries. \section{Open ends} \paragraph{} Theorem~(1) also would follow from the statement \begin{equation} \sum_{j=1}^k d_j - \sum_{j=1}^k \lambda_j \leq d_{k} \end{equation} which would be of independent interest as it estimates the {\bf Schur-Horn error}. Indeed, the Schur-Horn gives together with such a hypothetical error bound $0 \leq \sum_{j=1}^k d_j - \sum_{j=1}^k \lambda_j + d_{k+1}-\lambda_{k+1} \leq d_k+ d_{k+1} - \lambda_{k+1}$, which is $\lambda_k \leq d_k+d_{k+1}$. Can we prove the above Schur-Horn error estimate (1)? We do not know yet but our experiments indicate: \\ {\bf Conjecture A:} [Schur-Horn error] Estimate (1) holds for all finite simple graphs \paragraph{} We have mentioned the Brouwer-Haemers bound $\lambda_k \geq d_k-(n-k)+1$ which is very good for large $k$ but far from optimal for smaller $k$. (Note that part of the graph theory literature labels the eigenvalues in decreasing order. We use an ordering more familiar in the manifold case, where one has no largest eigenvalue and which also appears in the earlier literature like \cite{HornJohnson2012}.) The guess $d_{k-1}/2 \leq \lambda_k$ assuming $d_{-1}=0$ is a {\bf rule of thumb}, as it fails only in rare cases. The next thing to try is $d_{k-2}/2 \leq \lambda_k$ and this is still wrong in general but there are even less counter examples. We can try $d_{k-2}/3$ for which we have not found a counter example yet but it might just need to look for larger networks to find a counter example. We still believe that there is an affine lower bound. This is based only on limited experiments like for $A=1/3$ where it already looks good. {\bf Conjecture B:} [Affine Brouwer-Haemers bound] There exist constants $0<A<1$ and $B$ such that for all graphs and $1 \leq k \leq n$ we have $A d_k - B \leq \lambda_k$. \paragraph{} We see for many Erd\"os-R\'enyi graphs that an upper bound $\lambda_k \leq C d_k$ holds for most graphs for any $C>1$, if the graphs are large. A possible conjecture is that \\ {\bf Conjecture C:} [Linear bound for Erdoes-Renyi] For all $C>1$ and all $p \in [0,1]$, the probability of the set of graphs in the Erd\"os-R\'enyi probability space $E(n,p)$ with $\lambda_k \leq C d_k$ for all $1 \leq k \leq n$ goes to $1$ for $n \to \infty$. \section{Illustration} \paragraph{} Here are a few examples of spectra with known upper and lower bounds: \begin{figure}[!htpb] \scalebox{0.6}{\includegraphics{figures/stargraph.pdf}} \scalebox{0.6}{\includegraphics{figures/cyclegraph.pdf}} \scalebox{0.6}{\includegraphics{figures/lineargraph.pdf}} \scalebox{0.6}{\includegraphics{figures/wheelgraph.pdf}} \scalebox{0.6}{\includegraphics{figures/completegraph.pdf}} \scalebox{0.6}{\includegraphics{figures/bipartite.pdf}} \scalebox{0.6}{\includegraphics{figures/petersen.pdf}} \scalebox{0.6}{\includegraphics{figures/gridgraph.pdf}} \scalebox{0.6}{\includegraphics{figures/randomgraph.pdf}} \label{Spectra} \caption{ This figure shows examples of spectra and compares them with known upper and lower bounds. We see first the {\bf Star, Cycle and Path graph}, then the {\bf Wheel, Complete, Bipartite graph}, and finally the {\bf Petersen, Grid and Random graph}, all with $10$ vertices. The eigenvalues are outlined thick (in red). Above we see the upper bound (in blue) as proven in the present paper. Then there are the Brouwer-Haemers and Horn-Johnson lower bounds which, as the examples show, do not always compete but complement each other. } \end{figure} \bibliographystyle{plain}
{ "timestamp": "2022-05-27T02:04:52", "yymm": "2205", "arxiv_id": "2205.10968", "language": "en", "url": "https://arxiv.org/abs/2205.10968", "abstract": "We prove that each eigenvalue l(k) of the Kirchhoff Laplacian K of a graph or quiver is bounded above by d(k)+d(k-1) for all k in {1,...,n}. Here l(1),...,l(n) is a non-decreasing list of the eigenvalues of K and d(1),..,d(n) is a non-decreasing list of vertex degrees with the additional assumption d(0)=0. We also prove that in general the weak Brouwer-Haemers lower bound d(k) + (n-k) holds for all eigenvalues l(k) of the Kirchhoff matrix of a quiver.", "subjects": "Combinatorics (math.CO); Discrete Mathematics (cs.DM)", "title": "Eigenvalue bounds of the Kirchhoff Laplacian", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9825575132207566, "lm_q2_score": 0.8152324848629214, "lm_q1q2_score": 0.8010128030236902 }
https://arxiv.org/abs/1610.07836
Classification of crescent configurations
Let $n$ points be in crescent configurations in $\mathbb{R}^d$ if they lie in general position in $\mathbb{R}^d$ and determine $n-1$ distinct distances, such that for every $1 \leq i \leq n-1$ there is a distance that occurs exactly $i$ times. Since Erdős' conjecture in 1989 on the existence of $N$ sufficiently large such that no crescent configurations exist on $N$ or more points, he, Pomerance, and Palásti have given constructions for $n$ up to $8$ but nothing is yet known for $n \geq 9$. Most recently, Burt et. al. had proven that a crescent configuration on $n$ points exists in $\mathbb{R}^{n-2}$ for $n \geq 3$. In this paper, we study the classification of these configurations on $4$ and $5$ points through graph isomorphism and rigidity. Our techniques, which can be generalized to higher dimensions, offer a new viewpoint on the problem through the lens of distance geometry and provide a systematic way to construct crescent configurations.
\section{Introduction} Erd\H{o}s once wrote, ``my most striking contribution to geometry is, no doubt, my problem on the number of distinct distances,'' \cite{Erd96}. The referred question, which asks what is the minimum number of distinct distances determined by $n$ points, was first asked in 1946 \cite{Erd46} and marked the beginning of a chain of variants. See \cite{She} and \cite{SM15} for a survey on these. Although one would expect all distances between $n$ points to be different if they were to be placed in the plane at random, if the distances are regularly placed, such as on a lattice, then many distances may repeat. Erd\H{o}s' conjectured lower bound, $\Omega(n/\sqrt{\log{n}})$, attained by a $\sqrt{n}\times \sqrt{n}$ integer lattice, was essentially proven up to a $\sqrt{\log{n}}$ factor by Guth and Katz in 2010. \cite{GK} The variant we study in this paper is one where the distances have prescribed multiplicities. One says $n$ points are in crescent configuration in $\mathbb{R}^d$ if they are in general position and determine $n-1$ distinct distances such that for every $1 \leq i \leq n-1$, there is a distance that occurs exactly $i$ times. Erd\H{o}s conjectured that there exists a sufficiently large $N$ such that no crescent configuration exists on $N$ or more points \cite{Erd89}. Though constructions have been provided for $n=5,6,7,8$ by Erd\H{o}s, I. Pal\'asti and C. Pomerance \cite{Pal87, Pal89, Erd89}, little progress has been made towards a construction for $n \geq 9$. One problem often encountered in the search for these configurations is the lack of understanding of their properties and the difficulty in exhibiting the configurations' information combinatorially. As such, we take a new approach to studying these crescent configurations, borrowing techniques from distance geometry and graph theory. Our main theorems are the results of two algorithms that search for and classify all crescent configurations on any $n \geq 4$ up to graph isomorphism and find geometric realizations for each of these isomorphism classes in the plane. \begin{theorem} \label{thm:3on4} Given a set of three distinct distances $\{d_1,d_2,d_3\}$ on four points, there are only three allowable crescent configurations up to graph isomorphism. In Figure \ref{fig:mcr} we provide graph realizations for each type. \end{theorem} \begin{figure}[h] \includegraphics[width=\linewidth]{MCRMain.png} \caption{Types M, C, and R.} \label{fig:mcr} \end{figure} \begin{theorem}\label{thm:27on5} Given a set of four distinct distances $\{d_1,d_2,d_3,d_4\}$ on five points, there are 27 allowable crescent configurations up to graph isomorphism. In Figure \ref{fig:allFIVE} we provide graph realizations for each type. \end{theorem} \begin{figure}[h] \includegraphics[width=\textwidth]{configson52.png} \caption{Representatives for all possible crescent configurations on five points.} \label{fig:allFIVE} \end{figure} The advantage of this algorithmic method is that it can be applied in higher dimensions, though we hope that the running time ($\mathcal{O}(n^{n})$) can still be vastly improved. In Appendix \ref{app: 5points} we include distance sets and realizable distances for each crescent configuration on four and five points. \par In Section 2, we introduce our distance geometry approach and prove a classification of crescent configurations for a general $n$. We follow this in Section 3 with an outline of the first half of the algorithm used to achieve Theorems \ref{thm:3on4} and \ref{thm:27on5}. We then move to Section 4 where we discuss how distance geometry methods may be applied to determine whether a distance set is realizable. In Section 5, we outline the second half of the algorithm, completing the proofs for Theorems \ref{thm:3on4} and \ref{thm:27on5}. After the theorems have been established, we discuss another way to verify the uniqueness of our isomorphism classes and also a different method to characterize these configurations through analysis of rigidity in Section 6. Lastly, we discuss potential future work based on our approach using graph theory, rigidity, and distance geometry. \begin{rem} The authors are happy to provide copies of any code referenced in the course of this paper. Please email \href{mailto:Steven.Miller.MC.96@aya.yale.edu}{Steven.Miller.MC.96@aya.yale.edu}. \end{rem} \section{Classification of Crescent Configurations}\label{sec:classification} In this section we provide the key definitions and theorems that we use to classify crescent configurations. \begin{defi}[General Position \cite{SM15}]\label{def:genpos} We say that n points are in general position in $\mathbb{R}^d$ if no d+1 points lie on the same hyperplane and no d+2 lie on the same hypersphere. \end{defi} \begin{defi}[Crescent Configuration \cite{SM15}] We say $n$ points are in crescent configuration (in $\mathbb{R}^{d}$) if they lie in general position in $\mathbb{R}^{d}$ and determine $n - 1$ distinct distances, such that for every $1 \leq i \leq n - 1$ there is a distance that occurs exactly $i$ times. \end{defi} The notion of general position is very important in the construction of crescent configuration. Without this notion, the problem of placing $n$ points in $\mathbb{R}^d$ to determine $n-1$ distinct distances satisfying the prescribed multiplicities becomes trivial. By simply placing $n$ points on a line in an arithmetic progression, we solve the problem in any dimension. \vspace{0.2cm} \begin{defi}[Distance Coordinate]The \emph{distance coordinate}, $D_{A}$, of a point $A$ is the multiset of all distances, counting multiplicity, between $A$ and the other points in a set $\mathcal{P}$. Order does not matter. \end{defi} \begin{defi}[Distance Set] The \emph{distance set}, $\mathcal{D}$ corresponding to a set of points, $\mathcal{P}$, is the multiset of the distance coordinates for each point in the $\mathcal{P}$. \end{defi} \begin{defi} [Isomorphism for Labeled Graph \cite{Gervasi}]\label{def:isomorphism} Graph A is isomorphic to graph B if and only if there exists a bijective function $f: V(A) \mapsto V(B)$, (where V(A) and V(B) are the vertex spaces) such that: \begin{enumerate} \item $\forall a_{i} \in A, \ l_{A}(a_{i}) = l_{B}(f(a_{i}))$, \item $\forall a_{i},a_{j} \in V, \hspace{2mm} \{a_{i},a_{j}\}\in E_{A}\leftrightarrow \{f(a_{i}),f(a_{j})\}\in E_{B}$, and \item $\forall \{a_{i},a_{j}\}\in E_{A}, w_{A}(\{a_{i},a_{j}\})=w_{B}(f(\{a_{i},a_{j}\}))$, \end{enumerate} where $\{l_{A}, l_{B}\}$ and $\{w_{A}, w_{B}\}$ are functions that define the labels of the vertices and edges of A and B respectively. \end{defi} \vspace{0.2cm} We note that a crescent configuration on $n$ points can be considered a weighted complete graph with $n-1$ distinct weights associated to the edges in a certain manner, so that the configuration can be realized in $\mathbb{R}^d$. The adjacency matrix, thus, is a natural way to store information about the configuration. Should we rearrange the weights that each vertex is incident to, we would most likely have to draw a configuration differently. Due to this insight, we have the following theorem. \begin{theorem}\label{thm: graphisom} Let $A$ and $B$ be two crescent configurations on the same number of points $n$. If $A$ and $B$ have the same distance sets, then there exists a graph isomorphism from $A$ to $B$. \end{theorem} \begin{proof} Consider two crescent configurations, $A$ and $B$, each on $n$ points, $\{a_{1},...,a_{n}\}$ and $\{b_{1},...,b_{n}\}$, such that $A$ and $B$ have the same distance sets (up to order of the entries of the coordinates). We note that we do not yet care about the specific distances and, instead, designate each distance by the points that define them, as will be shown below. \par First we show that the ordering of the elements of the distance coordinates does not matter when comparing the distance sets. Then we prove that $A$ and $B$ are isomorphic when we view them as labeled graphs. \par Consider $A$. First, we number each point of $A$ from $1$ to $n$. For each point $a_{i}$, $1\leq i \leq n$, in the configuration, we re-order its distance coordinate so that the distance between $a_{i}$ and $a_{1}$, $d_{a_{i,1}}$, is in the first slot, the distance between $a_{i}$ and $a_{2}$, $d_{a_{i,2}}$, is in the second slot, and so on, inserting 0 into the $i$th slot, representing the distance between $a_{i}$ and $a_{i}$. We will call these \emph{augmented distance coordinates}. All of our augmented distance coordinates uniquely determine the points that they represent. Note that the arrangement of these augmented coordinates depends on how we choose to index the points of $A$. The nonzero entries, however, only encode information about the set of distances between a point and all other points in a given configuration. They do not account for relative position. That is to say, until assign a number to each of the points of $A$ (i.e., $a_1$, $a_2$, and so on), the order of the entries of the distance coordinates does not matter. The distance coordinate $\{d_1,d_2,d_3\}$ is the same as the coordinate $\{d_3,d_1,d_2\}$.\\ \par Next we do the same thing for $B$. Since the distance sets of $A$ and $B$ are the same, we know that for every augmented coordinate in $A$, there is an augmented coordinate in $B$ with the same set of non-zero entries. Thus, let $a_{i}$ and $b_{j}$ have the same non-zero entries in their augmented coordinates. We define each point by its augmented coordinate, $$D_{a_{i}}=(d_{a_{i,1}},...,d_{a_{i,i-1}},0,d_{a_{i,i+1}},...,d_{a_{i,n}})$$ $$D_{b_{j}}=(d_{b_{j,1}},...,d_{b_{j,j-1}},0,d_{b_{j,j+1}},...,d_{b_{j,n}}).$$ Since the set of non-zero entries in $D_{a_{i}}$ and $D_{b_{j}}$ are the same, we know that there exists a permutation, $f_{ij}$ such that $f_{ij}(D_{a_{i}})= D_{b_{j}}$. Since we know that every distance coordinate in $A$ has a match in $B$, there exists a permutation function of this sort that maps $D_{a_{i}}$ to some $D_{b_{j}}$ for all $1\leq i \leq n$. \par Furthermore, we actually only need one permutation for all of the reordered coordinates of $A$. This can be quickly seen by recognizing that the ordering of a single augmented distance coordinate of $A$ determines the ordering of all augmented distance coordinates if we wish to retain all information. To prove that it this is indeed the case, we point out that the reordered distance coordinates of $A$ are, in fact, the rows and columns of a weighted adjacency matrix. The adjacency matrix below models a configuration on $3$ points. The element in row $1$, column $2$ represents the distance between points $a_1$ and $a_2$. $$\begin{bmatrix} 0&d_{1,2}&d_{1,3}\\ d_{1,2}&0&d_{2,3}\\ d_{1,3}&d_{2,3}&0 \end{bmatrix}$$ From this, a proof by induction will show that if we permute the entries of row $1$, the other rows (and columns) must undergo the same permutation in order to preserve the symmetry of the matrix. \par We now identify each point of $A$ and $B$ by these augmented and rematched distance coordinates to create two labelled and weighted graphs where the vertices are the points of $A$ and $B$, and each edge $\{a_i,a_j\}$ is weighted by the distance between the two points, $d_{a_{i,j}}$. In these graphs, the vertex labels are the augmented distance coordinates of each point, where the coordinates of $A$ are further transformed by $f_{ij}$, and the edge labels are the distinct distances, $d_1,d_2,....,d_{n-1}$ that define our crescent configurations. Given this information, it is very straightforward to prove that these graphs are isomorphic by showing that they satisfy the three conditions in Definition \ref{def:isomorphism}. \par Let $F: V(A_{D}) \mapsto V(B_{D})$ be such that $F(a_{i})=b_j$, where $b_j$ is chosen so that the distance coordinates of $a_i$ and $b_j$ have the same nonzero values. Since we augmented and transformed our distance coordinates by the permutation function $f_{ij}$, we know that $a_{i}$ and $b_{j}$ have the same vertex labels. Thus, we know $F$ is one-to-one because the distance sets of $A$ and $B$ are the same. Furthermore, we know that $F$ is onto because $A$ and $B$ both have $n$ points. Therefore, $\forall b_{j}\in B$, $\exists \ a_{i} \in A$ such that $F(D_{a_{i}})=D_{b_{j}}$. Therefore, F is a bijection. \par \textit{Condition 1:} Let $l_A$ and $l_B$ be defined as in Definition \ref{def:isomorphism}. This menas that $l_{A}(a_i)=f_{ij}(D_{a_{i}})=D_{b_{j}}$ and $l_{B}(b_{j})=D_{b_{j}}$. Thus, given $F(a_{i})=b_{j}$, we have that $\forall a_{i} \in A, \ l_{A}(a_i)=l_{B}(F(a_{i}))$. \par \textit{Condition 2:} This condition holds because crescent configurations are complete graphs, and $F$ is a bijection. Therefore $\{a_{i},a_{j}\}\in E_{A} \leftrightarrow \{F(a_{i}), F(a_{j})\} \in E_{B}$, where $E_A$ and $E_B$ are the edge sets of $A$ and $B$ respectively. \par \textit{Condition 3:} To prove this condition, we recall that applying $f_{ij}$ to $a_i$ is equivalent to re-indexing the points of $A$. Therefore, if we let $w_{A}$ and $w_{B}$ be functions that return the edge labels of $A$ and $B$, respectively, as defined as in Definition \ref{def:isomorphism}, then we know that $$w_{A}(\{a_i,a_j\})=d_{a_{i,j}}=d_{b_{l,k}}=w_{B}(\{b_l,b_k\})=w_{B}(\{F(a_{i}),F(a_{j})\}),$$ satisfying this condition. Thus A and B are isomorphic. \end{proof} \begin{remark} Since the algorithm can only distinguish the similar permutations according to the distance labels, the resulting distance sets may define crescent configurations that are not geometrically realizable. We address this concern later in the paper (see Section \ref{sec:geomREAL} on geometric realizability). \end{remark} \section{Method for Counting Isomorphism Classes}\label{sec: counting} As a direct result of Theorem \ref{thm: graphisom}, we now have a method for classifying all crescent configurations on $n$ points into isomorphism classes. In this section we provide a sketch of the algorithm used to find these isomorphism classes. A pseudocode is provided in Appendix \ref{app: algorithms}. \par Consider a set of distances, $\{d_1,d_2,d_2,...,d_{n-1}\}$, associated to a crescent configuration on $n$ points. This set of distances may be threaded through an $n \times n$ adjacency matrix like the one shown in the proof of Theorem \ref{thm: graphisom}. \begin{comment} Consider a set of distances associated to a crescent configuration on $n$ points: $\{d_1,d_2,d_2,...,d_{n-1}\}$. This set of distances may be threaded through the rows and columns of an adjacency matrix to define a complete graph on $n$ vertices as shown in \ref{eq:threadEx} \begin{equation}\label{eq:threadEx} \{d_1,d_2,d_2,...,d_{n-1}\} \ \to \ \begin{pmatrix} 0&d_1&d_2&d_2...&{}&...\\ d_1&0&...&{}&.&...\\ d_2&...&0&...&{}&...\\ d_2&{}&...&0&{}&...\\ .&.&{}&{}&.&...\\ .&{}&...&d_{n-1}&d_{n-1}&0\\ \end{pmatrix} \end{equation} \end{comment} From this point on, we refer to these matrices as \textit{distance matrices}; however, they are the equivalent of weighted adjacency matrices in graph theory. \begin{defi}[Distance Matrix] Let $A$ be an $n \times n$ matrix. We say $A$ is a \textbf{distance matrix} if the following conditions are satisfied \begin{enumerate} \item $A$ is symmetric, \item the entries along the main diagonal of $A$ are all 0, and \item $a_{ij}$ (the entry found in the $ith$ row and $jth$ column of $A$) refers to the distance between points $i$ and $j$ of the point configuration (or graph) defined by $A$. \end{enumerate} \end{defi} As each configuration has a distance matrix associated to it, we can generate all possible configurations by threading all permutations of $\{d_1,d_2,d_2,...,d_{n-1}\}$ through distance matrices. Using standard combinatorial techniques, we can quickly see that this method will generate 60 configurations on four points, 12,600 on five points, and 37,837,800 on six points. We would now like to apply Theorem \ref{thm: graphisom} and group these configurations into isomorphism classes (via graph isomorphism). \begin{comment} To do so requires the following lemma. \begin{lemma}\label{lem: matrix2isom} Let $A$ and $B$ be two $n \times n$ distance matrices describing two crescent configurations on $n$ points. For each matrix, we take the set of non-zero elements of row $i$ and call these sets $A_{i}$ and $B_{i}$. If $A_{i} \ \subset \{B_{i}\}_{i=1}^{n}$ for all $1 \leq i \leq n$, Then A and B are isomorphic. \end{lemma} \begin{proof} Note that the set of non-zero elements of row $i$ in a given distance matrix represents the distances between point $i$ and all other points in the configuration. Therefore, it is equivalent to the \textbf{distance coordinate} of point $i$. Therefore, given a crescent configuration on $n$ points, the $n$ distance coordinates are the sets of non-zero elements of each row of its distance matrix, implying that the \textbf{distance set} of this configuration is entirely determined by the distance matrix. The remainder of the proof is a direct application of \ref{thm: graphisom}. \end{proof} \par Using this lemma, we may group all configurations on $n$ points into their isomorphism classes and conduct the remainder of our analysis on one representative from each class. For $n=4$, this reduces our initial $60$ to $4$ classes. For $n=5$, it reduces our initial $12,600$ to $85$ classes. \end{comment} \par A computer program may then be used to group together all distance matrices defining configurations with identical distance sets. These groups then represent our isomorphism classes, and we can conduct the remainder of our analysis on one representative from each class. For $n=4$, this reduces our initial $60$ to $4$ classes. For $n=5$, it reduces our initial $12,600$ to $85$ classes. \par Having finished this classification, we note that there are three degenerate cases that will force the configuration to \textit{always} violate general position. These will remain unchanged under graph isomorphism since vertex and edge-set pairings stay the same, allowing us to eliminate entire isomorphism classes if one representative is proven to be degenerate. These cases are as follows. \begin{enumerate} \item The configuration contains one point at the center of a circle of radius $d_{i}$ with four or more points on this circle as seen in Figure \ref{fig:star}. \item The configuration contains three (or more) isosceles triangles sharing the same base. \item The configurations contains four points arranged on the vertices of an isosceles trapezoid. \end{enumerate} \begin{figure}[h] \includegraphics[width=0.3\textwidth]{star.png} \caption{The central point of this configuration has distance coordinate $\{d,d,d,d,d,d\}$.} \label{fig:star} \end{figure} Although there exist other cases that will force a class to violate general position, these three cases may be accounted for by only considering the distance matrices. \par Case 1 is very simple to account for and is only possible for $n\geq 5$. In order to eliminate these cases, we remove configurations containing one or more distance coordinates in which a particular distance, $d_{i}$, occurs four or more times. In Algorithm \ref{alg:class}, this case and case 3 are accounted for in the procedure REMOVECYCLIC. \par As with case 1, case 2 is only possible for $n\geq 5$. If three or more isosceles triangles share the same base, then all of their apexes must reside on the line bisecting this base, forcing them to violate general position. \par In a distance matrix, an isosceles triangle is indicated by a matching pair of distances occurring in a row. Therefore, we remove all distance matrices in which three or more rows contain a matching pair occurring in the same slots in each row. This case is accounted for in Algorithm \ref{alg:class} by the procedure REMOVELINEARCASE. \par Case 3 requires us to remove all configurations that contain a subset or subsets of four points defining an isosceles trapezoid, since isosceles trapezoids are always cyclic quadrilaterals. In Algorithm \ref{alg:class}, the procedures SUBMATRICES and REMOVECYCLIC are included to account for these cases, which may be identified by their distance matrices using the following lemma. \begin{lemma}\label{cla:isosceles} A $4\times 4$ distance matrix will always define an isosceles trapezoid if and only if one of the following holds: \begin{enumerate} \item the matrix has only one distinct row such that \begin{enumerate} \item the matrix has only two distinct distances, or \item the matrix only has three distinct distances, \end{enumerate} \item the matrix has two distinct rows, both with multiplicity two, such that \begin{enumerate} \item the matrix has three distinct distances with multiplicity no greater than three (note that this means each distance occurs no more than six times in the distance matrix), or \item the matrix has four distinct distances with multiplicity no greater than two (each distance occurs no more than four times in the distance matrix). \end{enumerate} \end{enumerate} \end{lemma} \begin{proof} ($\Leftarrow$) According to Halsted \cite{Halsted}, a necessary and sufficient condition for a quadrilateral to be an isosceles trapezoid is that it has at least one pair of opposite sides with equal length and diagonals of equal length. It is not possible for these two lengths (sides and diagonals) to be equal because this would create two isosceles triangles that would have to be congruent. Therefore there are three cases for isosceles trapezoids: (1) four distinct distances, (2) three distinct distances, or (3) two distinct distances. Figure \ref{fig: 4dist} presents possible realizations for each of these cases. From here, it is straightforward to show that each of these quadrilaterals satisfies one of the conditions stated in Lemma \ref{cla:isosceles}, thus completing this direction of the proof. \begin{figure}[h] \includegraphics[width=\linewidth]{isosceles_trap.png} \caption{(1) Four distinct distances, (2) three distinct distances, (3) two distinct distances. } \label{fig: 4dist} \end{figure} \par ($\Rightarrow$) We now prove the other direction. \par We begin with condition (1a): one distinct row and two distinct distances. \par Assume each row represents the distance coordinate $(a,a,b)$ (order of distances may differ among rows). Since we require all rows of the distance matrix to have the same distance coordinate, distance $b$ must touch every point yet only show up twice. Therefore, it must represent both diagonals or two opposite sides. Both cases yield a quadrilateral with a set of congruent opposite sides and congruent diagonals, a necessary and sufficient condition for an isosceles trapezoid. \par For condition 1b - one distinct row and three distinct distances - all three distances must be common to all points, so they must describe either a set of opposite sides or the diagonals, in all cases yielding an isosceles trapezoid. \par The proofs in the other directions all following similar arguments. In each case there are two distances that must be common to all points and at least one distance that can only be common to two points, implying that one set of opposite sides cannot be congruent, therefore the distances common to all points must describe the diagonals and only one set of opposite sides, thus describing an isosceles trapezoid in each case. \end{proof} Once these cases have been eliminated, we are left with three isomorphism classes for four points and $51$ for five points. \begin{comment}\subsection{Usefulness of Classification} \par The usefulness of this classification system becomes evident when considering the question of geometric realizability (see \ref{def:geomreal}). Take, for instance, the two configurations shown in Figure \ref{fig:realisom}. \par If we do not assign values to $d_{1},d_{2},\ or \ d_{3}$, these two configurations are isomorphic, and thus fall in to the same isomorphism class. However, they clearly do not possess the same set of distances. And yet, in Section \ref{sec:geomREAL} that the distances are actually two different solutions to the same system of equations. By placing all crescent configurations into isomorphism classes, we set ourselves up nicely to solve for ALL realizable crescent configurations in a single class with one system of equations. \end{comment} \begin{figure}[h] \includegraphics[width=0.3\textwidth]{isomfigs.png} \caption{Two isomorphic configurations.} \label{fig:realisom} \end{figure} \begin{rem}\label{rem:runtime} It should be noted that the runtime of Algorithm \ref{alg:class} is $\mathcal{O}(n^{n})$, so its use is limited to crescent configurations on relatively few points. However, we believe that with enough processing power, upper bounds can be established using the algorithm for small $n$ such as $7$ and $8$. As such, at this time, this does not pose much of an issue to the progress of this research, as no crescent configurations on more than $8$ points is yet to be found. \end{rem} \section{Geometrically Realizability of Crescent Configurations} \label{sec:geomREAL} In the previous sections we developed a way to find every isomorphism class of distance sets that corresponded to a crescent configuration on $n$ points; however, it is not clear that given some $m$ and one of these distance sets, there exists a set of points in $\ensuremath{\mathbb{R}}^m$ that realizes this distance set. Thus, we formally define geometric realizability. \begin{defi}\label{def:geomreal} A crescent configuration on $n$ points is \emph{geometrically realizable} in $\ensuremath{\mathbb{R}}^m$ if there exists some distances $d_1, \dots,d_{n-1}$ for which there exist points $P_1, P_2, \dots, P_n$ in $\ensuremath{\mathbb{R}}^m$ that realize the corresponding distance set. \end{defi} We have thus far established upper bounds on the number of crescent configurations on $n$ points based on the combinatorial aspects of this problem. We now sharpen these bounds by considering criteria for geometric realizability. \begin{remark} Burt et. al. \cite{SM15} showed that given an $n$, there exists an $m$ such that an $n-$crescent configuration exists in $\ensuremath{\mathbb{R}}^m$. In this section, we fix $m$ and determine whether a given distance set is geometrically realizable in $\ensuremath{\mathbb{R}}^m$. \end{remark} In Section \ref{sec: counting}, we classify crescent configurations in terms of the distances between each pair of points. The problem of determining information about a set of points based on the distances between them is well studied, and is known as the distance geometry problem. Thus, we are now able to use techniques from distance geometry in order to sharpen the bounds on the number of geometrically realizable configurations on $n$ points. We begin by introducing some of these techniques. \begin{defi} The \emph{Cayley-Menger matrix} corresponding to a set of $n$ points $\{P_1, P_2, \dots, P_n\}$ is an $(n+1)\times (n+1)$ matrix of the following form: $$ \left( \begin{matrix} 0 & d_{1,2}^2 & \ldots & d_{1,n}^2 & 1\\ d_{2,1}^2 & 0 & \ldots & d_{2,n}^2 &1\\ \vdots & \vdots & \ddots & \vdots &\vdots\\ d_{n,1}^2 & d_{n,2}^2 & \dots &0&1\\ 1 & 1 &\ldots & 1 & 0 \end{matrix} \right) ,$$ where $d_{i,j}$ is the distance between $P_i$ and $P_j$. \end{defi} The Cayley-Menger matrix corresponding to a set of points can be used to determine whether those points lie in a $d$-dimensional Euclidian space. Note that $n+1$ points always lie in a hyperplane of $\ensuremath{\mathbb{R}}^n$; thus we only need to consider collections of points of size at least $n+2$. \begin{theorem}[Cayley-Menger Matrix]\label{thm: cayley} \cite{LL} Let $M$ be a $(n+3) \times (n+3)$ matrix of the form specified above. $M$ is the Cayley-Menger matrix of a set of $n+2$ points in $\mathbb{R}^n$ if and only if $\det{M} = 0$. \end{theorem} \begin{corollary}\label{cor:cayley} Let $M$ be a $n+1\times n+1$ matrix of the form specified above; $M$ is the Cayley-Menger matrix of a set of $n$ points, $\{P_1, \dots, P_n\}$ in $\mathbb{R}^2$, if and only if the Cayley-Menger matrix corresponding to every size $m+2$ subset has determinant zero. \end{corollary} \begin{proof} $\Rightarrow$ Suppose that $M$ is the Cayley-Menger matrix of $n$ points that can be realized in $\ensuremath{\mathbb{R}}_m$. Thus, every subset of $m+2$ points can also be realized in $\ensuremath{\mathbb{R}}^m$, and thus has a corresponding Cayley-Menger matrix with determinant zero. \\ $\Leftarrow$ Suppose that the Cayley-Menger matrix corresponding to every size $m+2$ subset has determinant zero. Note that $P_1, \dots, P_{m+1}$ must define an $m$-dimensional subspace of any Euclidean space. Note that by assumption, for every $i$, the Cayley-Menger matrix corresponding to $P_1, \dots, P_{m+1}, P_i$ has determinant $0$. Thus, as a consequence of Theorem \ref{thm: cayley}, $P_i$ is in the same $m$-dimensional subspace. Thus, our set of $n$ points must be realizable in $\ensuremath{\mathbb{R}}^m$. \end{proof} Submatrices of the Cayley-Menger matrix can also determine whether the points determined by a set of distances lie on the same hypersphere. \begin{defi} The \emph{Euclidian distance matrix} corresponding to a set of $n$ points $\{P_1, P_2, \dots P_n\}$ is an $n\times n$ matrix of the following form: $$ \left( \begin{matrix} 0 & d_{1,2}^2 & \ldots & d_{1,n}^2 \\ d_{2,1}^2 & 0 & \ldots & d_{2,n}^2\\ \vdots & \vdots & \ddots & \vdots &\vdots\\ d_{n,1}^2 & d_{n,2}^2 & \dots &0 \end{matrix} \right) .$$ \end{defi} Note that the Euclidean distance matrices defined above are slightly different than the general notion of distance matrices used in previous sections. \begin{theorem}\label{thm: circles} \cite{CS} Let $E$ be the Euclidean distance matrix corresponding to points $P_1, \dots P_n$ in Euclidean space. These points lie on a hypersphere in $\ensuremath{\mathbb{R}}^{n-2}$ if and only if $\det{E} = 0$. \end{theorem} We apply these techniques to the problem of counting crescent configurations in the following corollary. \begin{cor}\label{cor: realize} Let $M$ be the Cayley-Menger matrix corresponding to a distance set on n points, $\mathcal{D}$. Then $\mathcal{D}$ is geometrically realizable in general position in $\ensuremath{\mathbb{R}}^m$ if and only if the following conditions hold. \begin{enumerate} \item Let $S$ be a size $m+3$ subset of $\{1, 2, \dots, n, n+1\}$ that contains $n+1$. Let $M_S$ be the submatrix of $M$ with rows and columns indexed by $S$. For every choice of $S$, $\det{M_S} = 0$. \item Let $S$ be a size $m+2$ subset of $\{1, \dots, n, n+1\}$. For every such choice of $S$, $\det{M_S} \neq 0$. \end{enumerate} \end{cor} \begin{proof} From Corollary \ref{cor:cayley}, a collection of distances between $n$ points is geometrically realizable in $\ensuremath{\mathbb{R}}^m$ if and only if the Cayley-Menger matrix corresponding to each subset $m+2$ points has determinant 0. Each of these matrices is one of the submatrices specified by the first condition. We consider the submatrices specified by the second condition in two parts. First, consider the submatrices $M_S$ for which $n+1 \in S$. We see that these comprise the Cayley-Menger matrices for each subset of $m+1$ points. We now consider the submatrices $M_S$ for which $n+1 \not\in S$. We see that these comprise the Euclidean distance matrices for each subset of $m+2$ points. Thus, we can see that the submatrices specified by the second condition is the set of all Euclidean distance matrices for each size $m+2$ subset of the $n$ points together with the set of all Cayley-Menger matrices for each size $m+1$ subset of the $n$ points. By Theorem \ref{thm: circles}, a Euclidean distance matrix on $m+2$ points has determinant 0 if and only if these points lie on the same hypersphere in $\ensuremath{\mathbb{R}}^m$. By Theorem \ref{thm: cayley}, the Cayley-Menger matrix of $m+1$ points has determinant 0 if and only if these points lie on the same hyperplane in $\ensuremath{\mathbb{R}}^m$. Thus, this second condition holds if and only if the points lie in general position in $\ensuremath{\mathbb{R}}^m$. \end{proof} \par Our application of this corollary to the distance sets on $4$ and $5$ points have allowed us to determine the geometric realizability of each of the distance sets found using techniques from earlier sections. These geometrically realizable configurations are discussed in the following section. Thus far, most of our attentions have been focused on crescent configurations in the plane. However, these techniques can be applied to finding crescent configurations in higher dimensions, furthering the work of Burt et. al. \cite{SM15}. \section{Finding Geometric Realizations for Crescent Configurations} As stated in Section \ref{sec: counting}, Algorithm \ref{alg:class} yields three crescent configurations on four points and 51 on five points, unique up to isomorphism. However, these procedures do not guarantee that each isomorphism class contains geometrially realizable configurations. \begin{comment}\begin{theorem} \label{thm:3on4} Given a set of three distinct distances, $\{d1,d2,d3\}$, on four points in crescent configuration, there are only three allowable crescent configurations up to graph isomorphism. We refer to these configurations as M-type, C-type, and R-type, respectively. In Figure \ref{fig:mcr} we provide graph realizations, adjacency matrices, and a set of distances for each. \end{theorem} \begin{theorem}\label{thm:27on5} Given a set of four distinct distances, $\{d_1,d_2,d_3,d_4\}$, on five points in crescent configuration, there are 27 allowable crescent configurations up to graph isomorphism. \end{theorem} In Appendix \ref{app: 5points} we include distance sets and realizable distances for each crescent configuration on five points. \end{comment} \par To check which of these configurations are geometrically realizable, we run Algorithm \ref{alg: check4}. A pseudocode of this algorithm can be found in Appendix \ref{app: algorithms}. Note that we assume $d_{1}=1$ in order to simplify the procedure. \par Algorithm \ref{alg: check4} is an extended application of Theorem \ref{cor: realize}. The first step of this algorithm is to take the Cayley-Menger determinants of all $4$-point subsets of each configuration found by Algorithm \ref{alg:class} and set them equal to zero. Doing so yields a system of $\begin{pmatrix} n\\ 4 \end{pmatrix}$ equations with unknowns : $\{d_2,d_3,...,d_{n-1}\}$. If the configuration is realizable in the plane, solving this system of equations will give all possible solutions to these distances in $\mathbb{R}^{2}$. Note that the values must be positive and real-valued. \par For each of these solutions, we check the Cayley-Menger determinants of all $3$-point subsets of the configuration. If one or more of these determinants equals zero, that solution forces the configuration to place three points on the same line, violating general position, so we throw it away. If none of the determinants are zero, we keep the solution. \par For each remaining solution, we take the determinant of the Euclidean distance matrix of each $4$-point subset of the configuration. If any of these determinants equal zero, the solution forces four points onto the same circle, violating general position, and we throw it away. \par Any remaining solutions represent the distances of a geometrically realizable crescent configuration. Applying this algorithm to the configurations returned by Algorithm \ref{alg:class} completes the proofs of Theorems \ref{thm:3on4} and \ref{thm:27on5}, as we find that there are exactly three realizable crescent configurations on four points and 27 realizable crescent configurations on five points. \par In Appendix \ref{app: 5points}, we provide a set of distances for every configuration on five points that had at least one remaining solution after applying this algorithm. \begin{comment} \begin{figure}[h] \definecolor{qqwuqq}{rgb}{0.,0.39215686274509803,0.} \definecolor{qqqqff}{rgb}{0.,0.,1.} \definecolor{ffqqqq}{rgb}{1.,0.,0.} \begin{tikzpicture}[line cap=round,line join=round,>=triangle 45,x=1.0cm,y=1.0cm] \clip(-3.810089413056916,1.0780675462585876) rectangle (15.017641730323243,7.476163793212526); \draw [color=ffqqqq] (-2.,6.)-- (-3.,5.); \draw [color=ffqqqq] (-3.,5.)-- (-1.,5.); \draw [color=ffqqqq] (-2.,6.)-- (-1.,5.); \draw [color=qqqqff] (-3.,5.)-- (-2.,3.); \draw [color=qqqqff] (-1.,5.)-- (-2.,3.); \draw [color=qqwuqq] (-2.,6.)-- (-2.,3.); \draw [color=ffqqqq] (0.,5.)-- (2.,5.); \draw [color=ffqqqq] (0.,5.)-- (1.,4.); \draw [color=ffqqqq] (2.,5.)-- (1.,4.); \draw [color=qqqqff] (0.,5.)-- (1.,3.); \draw [color=qqqqff] (2.,5.)-- (1.,3.); \draw [color=qqwuqq] (1.,4.)-- (1.,3.); \draw [color=qqqqff] (5.,3.)-- (6.,6.); \draw [color=qqqqff] (6.,6.)-- (7.,3.); \draw [color=qqwuqq] (5.,3.)-- (7.,3.); \draw [color=ffqqqq] (5.995999144407565,4.297200458023963)-- (6.,6.); \draw [color=ffqqqq] (5.,3.)-- (5.995999144407565,4.297200458023963); \draw [color=ffqqqq] (7.,3.)-- (5.995999144407565,4.297200458023963); \draw [color=qqwuqq] (10.,3.)-- (10.,5.); \draw [color=qqqqff] (10.,5.)-- (13.,6.); \draw [color=qqqqff] (13.,6.)-- (14.,3.); \draw [color=ffqqqq] (10.,5.)-- (14.,3.); \draw [color=ffqqqq] (10.,3.)-- (13.,6.); \draw [color=ffqqqq] (10.,3.)-- (14.,3.); \draw (-0.7443349613914741,2.610944772091302) node[anchor=north west] {M}; \draw (5.787054957374033,2.5109745182326466) node[anchor=north west] {C}; \draw (11.718623352987604,2.544297936185532) node[anchor=north west] {R}; \end{tikzpicture} \[ \hspace{1cm} \{d_1,d_2,d_2\},\{d_1,d_3,d_3\}, \hspace{1.5cm} \{d_1,d_2,d_3\},\{d_1,d_2,d_3\}, \hspace{1.5cm} \{d_1,d_2,d_3\},\{d_1,d_3,d_3\}, \] \[ \hspace{1cm} \{d_2,d_3,d_3\},\{d_2,d_3,d_3\} \hspace{1.7cm} \{d_2,d_2,d_3\},\{d_3,d_3,d_3\} \hspace{1.7cm} \{d_2,d_3,d_2\},\{d_3,d_3,d_2\}\] \[ \{d_{1} \ = \ \sqrt{3} \ + \ \sqrt{8}, \ d_{2} \ = \ 3, \ d_{3} \ = \ 2\} \hspace{0.5cm} \{d_{1} \ = \ 1, \ d_{2} \ = \ \frac{\sqrt{17}}{2}, \ d_{3} \ = \ \frac{17}{16}\} \hspace{0.5cm} \{d_{1} \ = \ \sqrt{\frac{3}{2}}, \ d_{2} \ = \ 2, \ d_{3} \ = \ 1\} \] \caption{Types M, C, and R.} \label{fig:mcr} \end{figure} \end{comment} \begin{comment} \begin{figure}[h] \includegraphics[width=\textwidth]{configson52.png} \caption{Representatives for all possible distance sets on five points.} \label{fig:allFIVE} \end{figure} \end{comment} \section{Rigidity of Crescent Configurations}\label{sec: rigid} So far in this paper, we have shown that crescent configurations can be classified into a finite number of isomorphism classes for each positive integer $n$ and defined a method to geometrically realize these configurations. We now turn our attention to testing whether our crescent configurations are rigid, which can help answer questions, such as, whether one distance set could define two different realizations of crescent configurations belonging to the same isomorphism class. For this reason, we shall treat our configurations as graphs and adopt the following definition of graph rigidity. \begin{defi}[Asimow \& Roth]\cite{AsimowRoth} Let $G$ be a graph $(V,E)$ on $v$ vertices in $\ensuremath{\mathbb{R}}^n$ then $G(p)$ is $G$ together with the point $ p = (p_1,p_2, \dots, p_v) \in \ensuremath{\mathbb{R}}^n \times \ensuremath{\mathbb{R}}^n \times \ensuremath{\mathbb{R}}^n \dots \ensuremath{\mathbb{R}}^n = \ensuremath{\mathbb{R}}^{nv}$. Let $K$ be the complete graph on $v$ vertices. The graph $G(p)$ is \emph{rigid} in $\ensuremath{\mathbb{R}}^n$ if there exists a neighbordhood $\mathbf{U}$ of $p$ such that $$e_K^{-1}(e_K(p)) \cap \mathbf{U} = e_G^{-1}(e_G(p)) \cap \mathbf{U},$$ where $e_K$ and $e_G$ are the edge functions of $K$ and $G$, which return the distances of edges of the associated graphs. \end{defi} In other words, a rigid graph cannot have its vertices be continuously moved to noncongruent positions while preserving the distances. The rigidity testing of these configurations would not only serve as a verification for our classification under graph isomorphism, but would also give us another way to characterize these crescent configurations. \par One important note is that all crescent configurations are complete graphs. It is a rather direct result of the following theorem and its corollary that all complete graphs are rigid in $\ensuremath{\mathbb{R}}^2$. \begin{theorem}[Laman]\label{thm:rigidR2}\cite{Laman} The edges of a graph $G = (V,E)$ are independent in two dimensions if and only if no subgraph $G' = (V',E')$ has more than $2n-3$ edges. \end{theorem} \begin{cor} A graph with $2n-3$ edges is rigid in two dimensions if and only if no subgraph $G'$ has more than $2n'-3$ edges. \end{cor} We can easily check that the complete graph on $1$, $2$ and $3$ points ($K_1$, $K_2$ and $K_3$ respectively) satisfy the condition and thus are rigid. For $n \geq 3$, the complete graph on $n$ points (or $K_n$ by convention) is composed of triangles, which are $K_3$. Since each $K_3$ composing $K_n$ is rigid, it must also be the case that one cannot move the vertices of $K_n$ to noncongruent positions while preserving the distances because this would inevitably change the distances in each $K_3$ subgraph. Therefore, all crescent configurations are rigid. \par Nonetheless, this fact does not imply our work is over, as there is more than one kind of rigidity for graphs. Thus, we want to study whether all crescent configurations fall into just one category of rigidity or if there are certain crescent configurations that are more rigid than others. In addition, Theorem \ref{thm:rigidR2} can only be used to show the rigidity of crescent configurations in two dimensions. Furthermore, in the near future, we want to extend our definition of crescent configuration to $\epsilon$ crescent configurations, where two edges are considered equal if their lengths are within a sufficiently small $\epsilon$ from each other. Thus we need another method that could accommodate the exploration in higher dimensions and the assessment of stability of an extended family of crescent configurations.\\ \par This leads us to using \textbf{rigidity matrices . These are an extremely powerful and flexible tool that can be used in $\ensuremath{\mathbb{R}}^d$ for all $d \in \mathbb{N}$. We introduce them and other necessary terminologies in Subsection \ref{sub:prelim} and apply them to our analysis in Subsection \ref{sub:analysis}. \subsection{Preliminaries}\label{sub:prelim} \par The following definitions and theorems will be used through out the rest of the paper to analyze the rigidity of crescent configurations on $4$ and $5$ points in $\ensuremath{\mathbb{R}}^2$. The formulation of each of these definitions have been adopted from one paper by Bruce Hendrickson \cite{Hendrickson} due to their accessible nature. For other formulations and characterizations of rigidity, see also \cite{AsimowRoth, Connelly, Roth, Roth2}.\\ \begin{defi}[Realization]\cite{Hendrickson} Let $G=(V,E)$ be a graph with some pairwise associated distance measurements. A realization $f$ of $G$ is a function that maps the vertices of $G$ to coordinates in some Euclidean space such that the distance measurements are realized. \end{defi} \begin{defi}[Framework]\cite{Hendrickson} A framework is a combination of a graph $G = (V,E)$ and a realization of $G$ in some Euclidean space, denoted $f(G)$ \end{defi} In our case, we will mostly be concerned with frameworks in $\ensuremath{\mathbb{R}}^2$, though the techniques can be generalized to higher dimension and in any Euclidean space. \begin{defi}[Flexibility of Frameworks] \label{def:flex} \cite{Hendrickson} A framework is called flexible if and only if it can be continuously deformed while preserving the distance constraints; otherwise the framework is rigid. A framework is redundantly rigid if and only if one can remove any edge and the remaining framework is rigid. \end{defi} \begin{defi}[Infinitesimal motion]\cite{Hendrickson} An \textbf{infinitesimal motion} is an assignment of velocity to each vertex such that $(v_i - v_j)(f_i - f_j) = 0$ for all pairs $(i,j) \in E$. A framework $f(G)$ is infinitesimally rigid if and only if it does not have any infinitesimal motion. \end{defi} \begin{theorem}[Gluck 1975]\label{thm:gluck} \cite{Gluck} If a graph has a single infinitesimally rigid realization, then all its generic realizations are rigid. \end{theorem} The main tool we will use to study the rigidity of our crescent configurations, as mentioned, are \textbf{rigidity matrices}. Each framework has a rigidity matrix associated to it. This matrix is the set of all equations whose solutions are the infinitesimal motions of that framework. Its rows correspond to the edges and its $nd$ columns correspond to the components of the vertices in $\mathbb{R}^d$. There are $2d$ nonzero elements in each row, one for each coordinate of the vertices connected to the corresponding edge. The differences in the coordinate values for the two vertices are these nonzero values. \begin{theorem}[Hendrickson 1992]\label{thm:Hend} \cite{Hendrickson} A framework $f(G)$ is rigid if and only if its rigidity matrix has rank exactly equal to $S(n,d)$,the number of allowed motions, which is equal to $nd - d(d+1)/2$ for $n \geq d$ and $n(n-1)/2$ otherwise. \end{theorem} All the above theorems and objects are tied together in the following lemma, which provides the basis for our analysis in the next section. \begin{lemma}\label{lem:BJackson} \cite{Jackson} The following statements are equivalent: \begin{enumerate} \item $G$ is rigid in $\ensuremath{\mathbb{R}}^d$, \item some framework $f(G)$ in $\ensuremath{\mathbb{R}}^d$ is infinitesimally rigid, \item every generic framework of $G$ in $\ensuremath{\mathbb{R}}^d$ is rigid. \end{enumerate} \end{lemma} \subsection{Rigidity Analysis for 4-point Crescent Configurations} \label{sub:analysis} \vspace{0.2cm} The first thing we need to do before we are able to study the rigidity matrix is to construct a realization. By Theorem \ref{thm:gluck}, it suffices to construct a single realization for each type of crescent configurations on $n$ points and study the rigidity of that framework. Figure 6 is a realization of type C obtained by fixing $d_1 = 1$. We note that since the distances in these configurations must satisfy general position as well as geometric realizability, fixing one distance to calculate the rest of the distances in the distance set does not affect the rigidity characterization of the configuration. \begin{figure}[h] \includegraphics[width=0.5\textwidth]{TypeC.png} \caption{Realization of type C obtained by fixing $d_1 = 1$} \end{figure} The above realization yield the following rigidity matrix, which we denote $A_C$: \[A_C = \begin{bmatrix} \frac{1}{2} & y + \sqrt{\frac{1+4y^2}{4}}&-\frac{1}{2} & -y - \sqrt{\frac{1+4y^2}{4}}&0&0&0&0\\ -\frac{1}{2}&y + \sqrt{\frac{1+4y^2}{4}}& 0& 0& \frac{1}{2}& -y - \sqrt{\frac{1+4y^2}{4}}&0&0\\ 0& \sqrt{\frac{1+4y^2}{4}}& 0&0& 0& 0& 0&-\sqrt{\frac{1+4y^2}{4}}\\ 0& 0& -1& 0& 1& 0& 0& 0\\ 0& 0& -\frac{1}{2}& -y& 0& 0& \frac{1}{2}& y\\ 0& 0& 0& 0& \frac{1}{2}& -y& -\frac{1}{2}& y\\ \end{bmatrix} .\] By row reduction operations, we get that Rank($A_C$)$=5 = S(4,2)$. Thus by Theorem \ref{thm:Hend}, the framework is infinitesimally rigid. Since the row reduction operations yield the same result whether $y > 0$ or $y < 0$, we can conclude that type $C$ is rigid for all $y \neq 0$ by Lemma \ref{lem:BJackson}.\\ Similarly, we can carry out the same analysis on type M. There are two realizations of type $M$, which are included in Figure 7. \begin{figure}[h] \includegraphics[width=0.7\textwidth]{TypeM.png} \caption{Two realizations of type M: $M_1$ and $M_2$} \end{figure} Note that $d_1$ is the only distance that differs in these two realizations. If we remove $d_1$ and denote the remaining framework as $M'_1$ and $M'_2$ respectively, we could continuously deform one into the other. This fact also means that type M cannot be redundantly rigid by definition. Therefore, type M is another rigid graph. \begin{comment} \[A_{M_1} = \begin{bmatrix} -2 x & 0 & 2 x & 0 & 0 & 0 & 0 & 0\\ -x & -x \sqrt{3} & 0 & 0 & x & x \sqrt{3} & 0 & 0 \\ -x & -x \sqrt{3} - y & 0 & 0 & 0 & 0 & x& x\sqrt{3} + y\\ 0 & 0 & x & -x \sqrt{3} & -x & x\sqrt{3} & 0 & 0 \\ 0 & 0 & x & -x \sqrt{3} - y & 0 & 0 & -x & x\sqrt{3} + y \\ 0 & 0 & 0 & 0 & 0 & -y & 0 & y \\ \end{bmatrix} \] Again by row reduction operations, Rank($A_{M_1}$) $=5$ thus by \ref{thm:Hend}, the framework is infinitesimally rigid therefore type M is another rigid graph. \end{comment} Last but not least, type R can also be studied using the same method. We start with a realization that is included in Figure 8 and a rigidity matrix $A_R$ to follow: \begin{figure}[ht] \includegraphics[width=0.7\textwidth]{TypeR.png} \caption{Realization of type R obtained by fixing $d_1 = 1$} \end{figure} \[A_R = \begin{bmatrix} -x & 0 & x & 0 & 0 & 0 & 0 & 0\\ \frac{-x}{2} & \frac{-x}{y} & 0 & 0 & \frac{x}{2} & \frac{x}{y} & 0 & 0 \\ \frac{-1}{2x} & \frac{-y}{2x} & 0 & 0 & 0 & 0 & \frac{1}{2x} & \frac{y}{2x}\\ 0 & 0 & x - \frac{x}{2} & \frac{-x}{2y} & -x + \frac{x}{2} & \frac{x}{2y} & 0 & 0\\ 0 & 0 & x - \frac{1}{2x} & \frac{-y}{2x} & 0 & 0 & -x + \frac{1}{2x} & \frac{y}{2x}\\ 0 & 0 & 0 & 0 & \frac{x}{2} - \frac{1}{2x} & \frac{x}{2y} - \frac{y}{2x} & \frac{-x}{2} + \frac{1}{2x}& \frac{-x}{2y} + \frac{y}{2x}\\ \end{bmatrix} .\] By carrying out row reduction operations once more, we find that Rank($A_R$) $= 6 > S(4,2)$. However, since removing any row in the matrix is equivalent to removing any edge in the framework and the rank of any remaining matrix obtained this way is always $5$, we conclude that type R is redundantly rigid by Definition \ref{def:flex}. By \cite{Hendrickson}, the conditions for unique realization of a graph is rigidity, $(d+1)$-connectedness, and redundant rigidity. Since type R satisfies all these conditions, it immediately follows that for each value of $x$ that we choose, there is a \emph{unique realization} of type R. In conclusion we have found that not all crescent configurations on four points have the same rigidity. As stated above, only type R has a unique realization in $\ensuremath{\mathbb{R}}^2$. Similarly, this analysis may be done on all $27$ crescent configurations on $5$ points to determine which ones have a unique realization in $\ensuremath{\mathbb{R}}^2$. \\ As mentioned previously, rigidity matrices are very accommodating. Suppose we need to lift our crescent configurations into $\ensuremath{\mathbb{R}}^d$ for $d > 2$, which we discuss in Subsection \ref{sub:high}. Then, the number of rows and columns of the associated rigidity matrices as well as $S(n,d)$ would change but the method of evaluating rigidity would remain the same. Similarly, with the extension of our definition from crescent configuration to $\epsilon$ crescent configuration (\ref{sub:epsilon}), we can insert $\epsilon$ into the matrix by constructing a realization that involves $\epsilon$ and then solve for $\epsilon$ depending on what stability type we want the configuration to exhibit. \section{Future Work} \subsection{Further Explorations in the Plane}\label{sub:further} Thus far, we have used our techniques to classify crescent configurations in the plane for $n=4$ and $n=5$. Because of the complexity of our algorithm, we have not been able to apply our techniques to higher $n$. As mentioned above, the runtime of our current algorithm is on the order of $n^n$, which prevents us from carrying out this process for large $n$. However, thus far no configurations have been found for $n>8$, so even running a similar algorithm for $n=9$ would yield significant progress on this problem. Thus, we are interested in the possibility of modifying our algorithm or finding a new technique that would allow us to count crescent configurations on higher $n$. In this way, we could develop a sequence of $\{c_i\}$, where each $c_i$ gives the number of crescent configurations on $i$ points. If Erd\H{o}s' conjecture is correct, then $\{c_i\}$ only has a finite number of non-zero terms. It would be interesting to see Erd\H{o}s' conjecture realized as a sequence that goes to zero. Since our techniques yield every possible crescent configuration for a given $n$, we can use these to observe patterns. For example, one can see from Figure \ref{fig:allFIVE} that many of crescent configurations on 5 points contain crescent configurations on 4 points. We may be able to develop techniques using such patterns that generate some of the possible crescent configurations for larger $n$. \subsection{Extensions to Higher Dimensions}\label{sub:high} As mentioned earlier, the distance geometry techniques that we use naturally extend to higher dimensions. Thus, we are interested in using these techniques to find the number of crescent configurations on $n$ points in a given dimension. Our goal is to construct a sequence for each $d$ consisting of the number of crescent configurations on $i$ points in $\ensuremath{\mathbb{R}}^d$ for each $i$. Currently, constructions in $\ensuremath{\mathbb{R}}^3$ have been found for 3, 4, and 5 points. Thus, even finding a single 6 point configuration in 3D would give new information. We have attempted to use techniques from distance geometry to find a realization in $\ensuremath{\mathbb{R}}^3$ of a known distance set for $n=6$ in the plane. However, the resulting systems of equations exceeded our computational resources. Recently, Burt et. al \cite{SM15} found that given $d$ high enough, one can always construct a crescent configuration on $n$ points in $\ensuremath{\mathbb{R}}^d$. We can consider similar questions using the concept of distance coordinates. We are interested in determining whether given a distance set there always exists a dimension in which the set is geometrically realizable. \subsection{Properties of Crescent Configuration Types}\label{sub:epsilon} Now that we have developed a way of classifying crescent configurations, we can examine certain properties for each of the types of crescent configurations. We started to explore in this direction with our rigidity calculations. \\ One direction we are interested in is to develop a concept of stability for these configurations, as we noticed that moving the points of the M, R, and C- type configurations resulted in different amounts of change in the distances. Further, should we define two distances to be equal if they are $\epsilon$ apart, then our study of the stability of crescent configurations could have some powerful applications to the study of molecules. \begin{comment} Lastly, the techniques here can extend to problems more general than the problem of finding crescent configurations. Given some restraints on the distances between a set of points, we can use these techniques to find geometric realizations. \begin{comment}\section {Proof of the classifications of crescent configurations on 4 points} \label{sec: 4crescent} \par The remaining three steps of the program require the user to input specific values for $\{d1,d2,d3\}$. Once the user has done this, the second step of the program checks to see if any of the three general crescent configurations defined above are geometrically realizable given these distances. To do this, the program uses the \textit{Caley-Menger Determinant}. \par \par Once this check has been completed, the final step of the program, "circleCheck4[distances,dim]," employs Ptolemy's formula to remove any of the remaining configurations that define cyclic quadrilaterals. The output of this final step is the list of all geometrically realizable crescent configurations for the specified distances.\\ \section{Proof of the classifications of crescent configurations on 5 points} \begin{theorem}\label{thm:5points} Given a set of four distinct distances, $\{d1,d2,d3,d4\}$, there are at most 85 five-point crescent configurations that may be formed with these distances. \end{theorem} \begin{proof} As with \ref{thm: 4point}, this result was determined using a Mathematica program to carry out the combinatorics. The full code can be found in Appendix B \par Just as in \ref{thm: 4point}, this program has four steps, and we again only consider the first step, the function "crescentClassStep1[distances,dim]," to acheive the upper bound of 85. The first step is nearly identical to the first step in the proof of \ref{thm: 4point}, except that we can no longer eliminate cases of isosceles trapezoids because, with four distinct distances, a collection of four points with only two distinct distance coordinates may define a non-square rhombus and its diagonals. A non-square rhombus is not a cyclic quadrilateral \par Instead, the program tests for distance coordinates with only one distinct distance. This distance coordinate would define a point at the center of a circle surrounded by four points on said circle, violating general position. \par Once these cases have been eliminated, we are left with only 85 distance matrices. \end{proof} CAYLEY-MENGER DETERMINANTS \end{comment}
{ "timestamp": "2016-10-26T02:05:01", "yymm": "1610", "arxiv_id": "1610.07836", "language": "en", "url": "https://arxiv.org/abs/1610.07836", "abstract": "Let $n$ points be in crescent configurations in $\\mathbb{R}^d$ if they lie in general position in $\\mathbb{R}^d$ and determine $n-1$ distinct distances, such that for every $1 \\leq i \\leq n-1$ there is a distance that occurs exactly $i$ times. Since Erdős' conjecture in 1989 on the existence of $N$ sufficiently large such that no crescent configurations exist on $N$ or more points, he, Pomerance, and Palásti have given constructions for $n$ up to $8$ but nothing is yet known for $n \\geq 9$. Most recently, Burt et. al. had proven that a crescent configuration on $n$ points exists in $\\mathbb{R}^{n-2}$ for $n \\geq 3$. In this paper, we study the classification of these configurations on $4$ and $5$ points through graph isomorphism and rigidity. Our techniques, which can be generalized to higher dimensions, offer a new viewpoint on the problem through the lens of distance geometry and provide a systematic way to construct crescent configurations.", "subjects": "Combinatorics (math.CO)", "title": "Classification of crescent configurations", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9883127402831221, "lm_q2_score": 0.8104789109591831, "lm_q1q2_score": 0.8010066334317508 }
https://arxiv.org/abs/1806.09810
On Representer Theorems and Convex Regularization
We establish a general principle which states that regularizing an inverse problem with a convex function yields solutions which are convex combinations of a small number of atoms. These atoms are identified with the extreme points and elements of the extreme rays of the regularizer level sets. An extension to a broader class of quasi-convex regularizers is also discussed. As a side result, we characterize the minimizers of the total gradient variation, which was still an unresolved problem.
\section{Application to some popular regularizers} \label{sec:app} We now show that the extreme points and extreme rays of numerous convex regularizers can be described analytically, allowing to describe important analytical properties of the solutions of some popular problems. The list given below is far from being exhaustive, but it gives a taste of the diversity of applications targeted by our main results. \subsection{Finite-dimensional examples} We first consider examples where one has $\mathrm{dim} \vecgal<+\infty$. In that case, $\Phi$ is continuous, and since the considered regularizations~$\reg$ are lower semi-continuous, we deduce that \begin{itemize} \item[$\circ$] the level set $\minset$ is closed, \item[$\circ$] the solution set $\sol=\minset\cap \Phi^{-1}(\{y\})$ is closed, and locally compact (even compact in most cases), hence it admits extreme points provided it contains no line. \end{itemize} \subsubsection{Nonnegativity constraints} \label{sec:nonneg} In a large number of applications, the signals to recover are known to be nonnegative. In that case, one may be interested in solving nonnegatively constrained problems of the form: \begin{equation}\label{eq:nonnegative} \inf_{u\in \RR^n_+} f(\Phi u - y). \end{equation} An important instance of this class of problems is the nonnegative least squares \cite{lawson1995solving}, which finds it motivation in a large number of applications. Applying the results of \Cref{sec:abstract} to Problem \eqref{eq:nonnegative} yields the following result. \begin{proposition} If the solution set of \eqref{eq:nonnegative} is nonempty, then it contains a solution which is $m$-sparse. In addition if $f$ is convex and the solution set is compact, then its extreme points are $m$-sparse. \end{proposition} Choosing $\reg$ as the characteristic function of ${\RR^n_+}$, the result simply stems from the fact that the extreme rays of the positive orthant are the half lines $\{\alpha e_i, \alpha \geq 0\}$, where $(e_i)_{1\leq i \leq n}$ denote the elements of the canonical basis. We have to consider $m$ atoms and not $m-1$ since $t^\star=\inf \reg=0$, see \cref{rem:infR}. It may come as a surprise to some readers, since the usual way to promote sparsity consists in using $\ell^1$-norms. This type of result is one of the main ingredients of \cite{donoho2005sparse} which shows that the $\ell^1$-norm can sometimes be replaced by the indicator of the positive orthant when sparse \emph{positive} signals are looked after. \subsubsection{Linear programming} Let $\psi \in \mathbb{R}^n$ be a vector and $\Phi \in \RR^{m\times n}$ be a matrix and consider the following linear program in standard (or equational) form: \begin{equation}\label{eq:linearprogramming} \inf_{\substack{u \in \RR^n_+ \\ \Phi u = y}} \langle \psi,u\rangle \end{equation} Applying Theorem \eqref{thm:first} to the problem \eqref{eq:linearprogramming}, we get the following well-known result (see e.g. \cite[Thm. 4.2.3]{matousek2007understanding}): \begin{proposition}\label{prop:linprog} Assume that the solution set of \eqref{eq:linearprogramming} is nonempty and compact. Then, its extreme points are $m$-sparse, i.e. of the form: \begin{equation} u = \sum_{i=1}^m \alpha_i e_i, \alpha_i\geq0, \end{equation} where $e_i$ denotes the $i$-th element of the canonical basis. \end{proposition} In the linear programming literature, solutions of this kind are called \emph{basic solutions}. To prove the result, we can reformulate \eqref{eq:linearprogramming} as follows: \begin{equation}\label{eq:linearprogramming2} \inf_{\substack{(u,t) \in \RR^n_+\times \RR \\ \Phi u = y \\ \langle \psi,u\rangle=t}} t \end{equation} \begin{sloppypar} Letting $R(u,t) = t + \iota_{\RR_+^n}(u)$, we get $\inf R = -\infty$. Hence, if a solution exists, we only need to analyze the extreme points and extreme rays of ${\minset = \{(x,t)\in \RR^n\times \RR, R(x,t) \le t^\star\}=\RR_+^n\times ]-\infty,t^\star]}$, where $t^\star$ denotes the optimal value. The extreme rays of this set (a shifted nonnegative orthant) are of the form $\{\alpha e_i, \alpha > 0 \}\times \{t^\star\}$ or $\{0\}\times]-\infty,t^\star[$. In addition $\minset$ possesses only one extreme point $(0,t^\star)$. Applying \cref{thm:first}, we get the desired result. \end{sloppypar} \subsubsection{$\ell^1$ analysis priors} An important class of regularizers in the finite dimensional setting $\vecgal=\RR^n$ contains the functions of the form $R(u) = \| Lu \|_1$, where {$L$ is a linear operator from $\RR^{n}$ to $\RR^{p}$}. They are sometimes called \emph{analysis priors}, since the signal $u$ is ``analyzed'' through the operator $L$. Remarkable practical and theoretical results have been obtained using this prior in the fields of inverse problems and compressed sensing, even though many of its properties are -to the belief of the authors- still quite obscure. Since $R$ is one-homogeneous, it suffices to describe the extremality properties of the unit ball $C=\{u\in \RR^n, \| Lu \|_1\leq 1\}$ to use our theorems. The lineality space is simply equal to $\mathrm{lin}(C) = \mathrm{ker}(L)$. Let $K=\mathrm{ker}(L)$, $K^\perp$ denote the orthogonal complement of $K$ in $\RR^n$ and $L^+:\RR^n\to K^\perp$ denote the pseudo-inverse of $L$. We can decompose $C$ as $C= K + C_{K^\perp}$ with $C_{K^\perp}=C\cap K^\perp$. Our ability to characterize the extreme points of $C_{K^\perp}$ depends on whether $L$ is surjective or not. Indeed, we have \begin{equation}\label{eq:extL1} \ext(C_{K^\perp})=L^+\left( \ext\left( \mathrm{ran}(L)\cap B_1^p\right) \right), \end{equation} where $B_1^p$ is the unit $\ell^1$-ball defined as \begin{equation*} B_1^p=\{z\in \RR^p, \|z\|_1\leq 1\}. \end{equation*} Property \eqref{eq:extL1} simply stems from the fact that $C_{K^\perp}$ and $D=\mathrm{ran}(L)\cap B_1^p$ are in bijection through the operators $L$ and $L^+$. \paragraph{The case of a surjective operator $L$} When $L$ is surjective {(hence $p\leq n$)}, the problem becomes quite elementary. \begin{proposition}\label{prop:extanalysis} If $L$ is surjective, the extreme points $u$ of $C_{K^\perp}$ are $\ext(C_{K^\perp}) = (\pm L^+ e_i)_{1\leq i \leq p}$, where $e_i$ denotes the $i$-th element of the canonical basis. {Consider Problem \eqref{eq::data-fitting:convex} and assume that at least one solution exists. } Then Problem \eqref{eq::data-fitting:convex} has solutions of the form \begin{equation}\label{eq:extremeanalysisfinite} u^\star = \sum_{i\in I} \alpha_i L^+e_i + u_K, \end{equation} where $u_K\in \mathrm{ker}(L)$ and $I\subset \{1,\hdots, p\}$ is a set of cardinality $|I|\leq m - \mathrm{dim}(\Phi\mathrm{ker}(L))$. \end{proposition} {The proof of Proposition \ref{prop:extanalysis} follows from Corollary \ref{coro:lines} and Section \ref{subsec:nonconvex}, with $j=0$ and observing that $\projv$ is the orthogonal projection on $K^\perp$.} \paragraph{The case of an arbitrary operator $L$} When $L$ is not surjective the des\-crip\-tion of the extreme points $\ext\left(D\right)$ becomes untractable in general. A rough upper-bound on the number of extreme points can be obtained as follows. We assume that $L$ has full rank {$n$} and that $\mathrm{ran}(L)$ is in general position. The extreme points of $\mathrm{ran}(L)\cap B_1^p$ correspond to the intersections of some faces of the $\ell^1$-ball with a subspace of dimension {$n$}. In order for some $k$-face to intersect the subspace $\mathrm{ran}(L)$ on a singleton, $k$ should satisfy $n+k-p=0$, i.e. $k=p-n$. The $k$-faces of the $\ell^1$-ball contain $(k+1)$-sparse elements. The number of $(k+1)$-sparse supports in dimension $p$ is $\binom{p}{k+1}$. For a fixed support, the number of sign patterns is upper-bounded by $2^{k+1}$. Hence, the maximal number of extreme points satisfies $|\ext\left( \mathrm{ran}(L)\cap B_1^p\right)|\leq 2^{k+1}\binom{p}{k+1}$. This upper-bound is pessimistic since the subspace may not cross all extreme points, but it provides an idea of the combinatorial explosions that may happen in general. Notice that enumerating the number of faces of a polytopes is usually a hard problem. For instance, the Motzkin conjecture \cite{motzkin1957comonotone} which upper bounds the number of $k$ faces of a $d$ polytope with $z$ vertices was formulated in 1957 and solved by Mc Mullen \cite{mcmullen1970maximum} in 1970 only. \subsubsection{Matrix examples} \label{sec:matrix:examples} In several applications, one deals with optimization problems in matrix spaces. The following regularizations/convex sets are commonly used. \paragraph{Semi-definite matrix constraint} Similarly to \Cref{sec:nonneg}, one may consider in $\RR^{n\times n}$ the following constrained problem \begin{equation}\label{eq:sdpcone} \inf_{M\succeq 0} f(\Phi M - y), \end{equation} where $M\succeq 0$ means that $M$ must be symmetric positive semi-definite (p.s.d.). The extreme rays of the positive semi-definite cone $\minset$ are the p.s.d. matrices of rank $1$ (see for instance~\cite[Sec. 2.9.2.7]{dattorro_convex_2005}). Hence, arguing as in \Cref{sec:nonneg}, we may deduce that if there exists a solution to~\eqref{eq:sdpcone}, there is also a solution which has rank (at most) $m$. However, that conclusion is not optimal, as in that case a theorem by Barvinok~\cite[Th. 2.2]{barvinok_problems_1995} ensures that there exists a solution $M$ with \begin{equation}\label{eq:ranksdp} \rank(M) \leq \frac{1}{2}\left(\sqrt{8m+1}-1\right). \end{equation} To understand the gap with Barvinok's result, let us note that the p.s.d.\ cone has a very special structure which makes the Minkowski-Carath\'eodory theorem (or its extension by Klee) too pessimistic. By~\cite[Sec. 2.9.2.3]{dattorro_convex_2005}, given $M\succeq 0$, the smallest face of the p.s.d. cone which contains $M$ (\ie{} the set of p.s.d matrices which have the same kernel) has dimension \begin{equation}\label{eq:ranksdp2} d= \frac{1}{2}\rank(M)(\rank(M)+1). \end{equation} \begin{sloppypar} Equivalently, if the smallest face which contains $M$ has dimension $d$, then $\rank(M)=\frac{1}{2}\left(\sqrt{8d+1}-1\right)$, hence $M$ is a convex combination of ${\frac{1}{2}\left(\sqrt{8d+1}-1\right)}$ points in extreme rays, a value which is less than the value $d$ predicted by Klee's extension of Carath\'eodory's theorem. As a result, we recover Barvinok's result by noting that, as ensured by the first claim\footnote{or, more precisely, by its variant when $t^\star=\inf \reg$, see \cref{rem:infR}.} of \cref{thm:first}, any extreme point $M$ of the solution set belongs to a face of dimension $m$. Then, taking into account~\eqref{eq:ranksdp2} improves upon the second claim of \cref{thm:first}, and we immediately obtain~\eqref{eq:ranksdp}. \end{sloppypar} \paragraph{Semi-definite programming} Semi-definite programs are problems of the form: \begin{equation}\label{eq:semidefiniteprogramming} \inf_{\substack{M\succeq 0\\ \Phi(M) = y}} \langle A,M\rangle, \end{equation} where $A\in \RR^{n\times n}$ is a matrix and $\langle A,M\rangle\eqdef \mathrm{Tr}(A M)$. Arguing as in \cref{prop:linprog}, if the solution set of \eqref{eq:semidefiniteprogramming} is nonempty, our main result allows to state that its extreme points are matrices of rank $m$. In view of the above discussion, it is possible to refine this statement and show that~\eqref{eq:ranksdp} holds. \paragraph{The nuclear norm} The nuclear norm of a matrix $M\in \RR^{p\times n}$ is often denoted $\|M\|_*$ and defined as the sum of the singular values of $M$. It gained a considerable attention lately as a regularizer thanks to its applications in matrix completion \cite{candes2009exact} or blind inverse problems \cite{ahmed2014blind}. The geometry of the unit ball $\{M\in \RR^{p\times n}, \|M\|_*\leq 1\}$ is well studied due to its central role in the field of semi-definite programming \cite{pataki2000geometry}. Its extreme points are the rank one matrices $M= uv^T$, with $\|u\|_2=\|v\|_2=1$. Combining \cref{thm:first} with this result explains why regularizing pro\-blems over the space of matrices with the nuclear norm allows recovering \emph{rank-m solutions}. \paragraph{The rank-sparsity ball} The \emph{rank-sparsity} ball is the set $\{M\in \RR^{m\times n}, \|M\|_* + \|M\|_1 \leq 1\}$, where $\|M\|_1$ is the $\ell^1$-norm of the entries of $M$. The corresponding regularization is sometimes used in order to favor sparse and low-rank matrices. The authors of \cite{drusvyatskiy2015extreme} have described the extreme points of this unit ball. They have proved that \emph{the extreme points $M$ of the rank sparsity ball satisfy $\frac{r(r+1)}{2}-|I|\leq 1$}, where $|I|$ denotes the number of non-zero entries in $M$ and $r$ denotes its rank. This result partly explains why using the rank-sparsity gauge promotes sparse and low rank solution. Let us outline that this effect might be better obtained using different strategies \cite{Richard2013Intersecting,richard2014tight}. \paragraph{Bi-stochastic matrices} A doubly stochastic matrix is a matrix with nonnegative rows and columns summing to one. The set of such matrices is called the Birkhoff polytope. The Birkhoff-von Neumann theorem states that its extreme points are the permutation matrices. We refer the interested reader to \cite{Fogel2015Convex} for an use of such matrices in DNA sequencing. \subsection{Examples in infinite dimension} In this section, we provide results in infinite dimensional spaces, which echoe the ones described in finite dimension. \subsubsection{Problems formulated in Hilbert or Banach sequence spaces} The case of Hilbert spaces (or countable sequences) can be treated within our formalism and all the examples presented previously have their na\-tu\-ral counterpart in this setting. {In the same vein, one can also treat Banach sequence spaces $\ell^p$ for $1\leq p\leq\infty$}. We do not reproduce the results here for space limitations. Let us however mention that two works treat this specific case with $\ell^1$ regularizers \cite{unser2016representer,adcock2016generalized}. \subsubsection{Linear programming and the moment problem}\label{sec:momprob} ~ Let $\Omega$ be a {compact} metric space, $\mathcal{M}(\Omega)$ be the set of Radon measures on $\Omega$ and let $\mathcal{M}_+(\Omega)\subseteq \mathcal{M}(\Omega)$ be the cone of nonnegative measures on $\Omega$. Let $\psi$ and $(\phi_i)_{1\leq i \leq m}$ denote a collection of continuous functions on $\Omega$. Now, let $\Phi:\mathcal{M}(\Omega)\to \RR^m$ be defined by $(\Phi \mu)_i = \langle \phi_i, \mu\rangle$, where $ \langle \phi_i, \mu\rangle\eqdef\int_\Omega \phi_i \d \mu$, and consider the following linear program in standard form: \begin{equation}\label{eq:linearprogramminginf} \inf_{\substack{\mu \in \mathcal{M}_+(\Omega) \\ \Phi \mu = y}} \langle \psi,u\rangle. \end{equation} Applying Theorem \eqref{thm:first} to the problem \eqref{eq:linearprogramminginf}, we get \cref{prop:linearprogramminginf} below. We do not provide a proof here, since it mimics very closely the one given for linear programming in finite dimension. The extreme rays of $\mathcal{M}_+(\Omega)$ can be described, arguing as in~\cite[Th.~15.9]{aliprantis_infinite_2006}, as the rays directed by the Dirac masses. \begin{proposition}\label{prop:linearprogramminginf} Assume that the solution set \eqref{eq:linearprogramminginf} is {nonempty}. Then, its extreme points are $m$-sparse, i.e. of the form: \begin{equation} u = \sum_{i=1}^m \alpha_i \delta_{x_i}, x_i\in \Omega, \alpha_i\geq0. \end{equation} \end{proposition} { To make sure that the above proposition is non-trivial, one may wish to ensure that the solution set $\sol$ has indeed extreme points, using arguments from Section~\ref{sec:existextreme}. It is straightforward that $\sol$ is convex and does not contain any line. Now, let us endow $\mathcal{M}(\Omega)$ with the weak-* topology (\ie{} the coarsest topology for which $\mu\mapsto \int_{\Omega}\eta\d \mu$ is continuous for every $\eta\in \Cder{}(\Omega)$). By lower semi-continuity, $\sol$ is closed. Moreover, $\sol$ is locally compact since the closed convex cone $\mathcal{M}_+$ is itself locally compact (take any $\mu\in \mathcal{M}_+(\Omega)$, its neighborhood $\enscond{\nu\in \mathcal{M}_+(\Omega)}{\nu(\Omega)\leq \mu(\Omega)+1}$ is compact in the weak-* topology). } {Proposition~\ref{prop:linearprogramminginf}} is well known, see e.g. \cite{shapiro2001duality}. Note that if we optimize the linear form $\langle \psi,u \rangle$ over the set of probability measures instead of the set of nonnegative measures, we get the so-called moment problem \cite{shohat1943problem} for which we can obtain a similar result. \subsubsection{The total variation ball} Let $\Omega$ denote an open subset of $\RR^d$ and $\mathcal{M}(\Omega)$ denote the set of Radon measures on $\Omega$. The total variation ball $B_\mathcal{M} = \{u \in \mathcal{M}(\Omega), \|u\|_{\mathcal{M}(\Omega)}\leq 1\}$ plays a critical role for problems such as super-resolution \cite{candes2014towards,Tang2013Compressed,Duval2015Exact}. {It is compact for the weak-* topology and }its extreme points are the Dirac masses: $\ext(B_\mathcal{M})=\{\pm \delta_x, x\in \Omega\}$. Hence total variation regularized problems of the form: \begin{equation*} \inf_{u \in \mathcal{M}} f(\Phi u) + \|u\|_\mathcal{M}, \end{equation*} yield $m$-sparse solutions (under an existence assumption). A few variations around this central result were provided in \cite{fernandez2016super}. \paragraph{Demixing of Sines and Spikes} In \cite[Page 262]{fernandez2016super}, the author presents a regularization of the type \[ \|\mu\|_{\mathcal{M}}+\eta\|v\|_1 \] where $\eta>0$ is a tuning parameter, $\mu$ a complex measure and $v\in\mathbb C^n$ a sparse vector. Define $E$ as the set of $(\mu,v)$ where $\mu$ is a complex Radon measure on a domain $\Omega$ and $v\in\mathbb C^n$. Consider the unit ball \[ B\eqdef\{(\mu,v)\in E\ :\ \|\mu\|_{\mathcal{M}}+\eta \|v\|_1\leq 1\}. \] Its extreme points are the points \begin{itemize} \item $(a\delta_t,0)$ for all $t\in\Omega$ (and $\delta_t$ denotes the Dirac mass at point $t$) and all $a\in\mathbb C$ such that $|a|=1$, \item $(0,ae_k)$ for all $k=1,\ldots,n$ and all $a\in\mathbb C$ such that $|a|=1/\eta$ and $e_k$ denotes the vector with $1$ at entry $k$ and $0$ otherwise. \end{itemize} \paragraph{Group Total Variation: Point sources with a common support} In \cite[Page 266]{fernandez2016super}, the author presents a regularization of the type \[ \|\mu\|_{\mathcal{M}^n}:=\sup_{F:\Omega\to\mathbb C^n,\ \| F(t) \|_2\leq 1,\ t\in\Omega}\int_\Omega \langle F(t), \nu(t)\rangle \mathrm d|\mu|(t) \] were $F$ is continuous and vanishing at infinity, and $\mu$ is a vectorial Radon measure on~$\Omega$ such that $|\mu|$-a.e. $\mu=\nu \cdot |\mu|$ with~$\nu$ a measurable function from~$\Omega$ onto $\mathbb S^{n-1}$ the $n$-sphere and $|\mu|$ a positive finite measure on $\Omega$. Consider the unit ball \[ B\eqdef\{\mu, \|\mu\|_{\mathcal{M}^n}\leq 1\}. \] Its extreme points are $a\delta_t$ for all $t\in\Omega$ (and $\delta_t$ denotes the Dirac mass at point $t$) and all $a\in\mathbb C^n$ such that $\|a\|_2=1$. \subsubsection{Analysis priors in Banach spaces} \label{sec:prior} The analysis of extreme points of analysis priors in an infinite dimensional setting is more technical. Fisher and Jerome~\cite{fisher_spline_1975} proposed an inte\-res\-ting result, which can be seen as an extension of \eqref{eq:extremeanalysisfinite}. This result was recently revisited in \cite{unser2017splines} and \cite{flinth2017exact}. Below, we follow the presentation in \cite{flinth2017exact}. { Let $\Omega$ denote an open set in $\RR^d$. Let $\mathcal{D}'(\Omega)$ denote the set of distributions on~$\Omega$ and let $L:\mathcal{D}'(\Omega) \to \mathcal{D}'(\Omega)$ denote a linear operator with kernel $K=\mathrm{ker}(L)$. We let $E=\{u\in \mathcal{D}'(\Omega), Lu\in \mathcal{M}(\Omega)\}$ and let $\|\cdot\|_K$ denote a semi-norm on $E$, which restricted to $K$ is a norm. We define a function space $\mathcal{B}(\Omega)$ as follows: $$\mathcal{B}(\Omega) = \{u\in E, \|Lu\|_{\mathcal{M}(\Omega)}+\|u\|_K<+\infty\}$$ and equip it with the norm $\|u\|_{\mathcal{B}(\Omega)} = \|Lu\|_{\mathcal{M}(\Omega)} + \|u\|_K$. We assume that $L$ is surjective, i.e. $\mathcal{M}(\Omega) = L(\mathcal{B}(\Omega))$, and that $K$ has a topological complement (with respect to $\mathcal{B}(\Omega)$), which we denote by $K^\perp$. This setting encompasses all surjective Fredholm operators for instance. Under the stated assumptions, we can define a pseudo-inverse $L^+$ of $L$ relative to $K^\perp$ \cite{beutler} The representer theorems in \cite{fisher_spline_1975,unser2017splines,flinth2017exact} can be obtained using \cref{thm:first} as exemplified below. \begin{proposition} Let $B=\{u \in \mathcal{B}(\Omega), \|Lu\|_{\mathcal{M}(\Omega)}\leq 1\}$. Then the extreme points of the set $C_{K^\perp}=B\cap K^\perp$ are of the form $\pm L^+\delta_{x}$, for $x\in \Omega$. Let $f:\mathbb{R}^m\to \mathbb{R}\cup \{+\infty\}$ denote a convex function and define \begin{equation*} \sol = \argmin_{u \in \mathcal{B}(\Omega)} f(\Phi u) + \|Lu\|_{\mathcal{M}(\Omega)}. \end{equation*} Assume that $\sol$ is nonempty and does not contain $0$. Then the extreme points (if they exist) of $\pi_K(\sol)$ are of the form $u = \sum_{i=1}^m \alpha_i L^+ \delta_{x_i}$. \end{proposition} \begin{proof} The proof mimics the finite dimensional case \eqref{eq:extremeanalysisfinite}. First notice that $B=L^{-1}(B_\mathcal{M})$, where $L^{-1}(\{\mu\})$ is the pre-image of $\mu$ by $L$ and $B_\mathcal{M}$ is the unit total variation ball. We have $L^{-1}(B_\mathcal{M})=L^+(B_\mathcal{M})+K$ and we can identify $C_{K^\perp}$ with $L^+(B_\mathcal{M})$. Since $L^+$ is bijective from $\mathcal{M}(\Omega)$ to $K^\perp$, the extreme points of $C_{K^\perp}$ are the image by $L^+$ of the Dirac masses. The end of the proposition follows from Corollary \ref{cor:convex_fit} and from the fact that the lineality space of $\{u \in \mathcal{B}(\Omega), \|Lu\|_{\mathcal{M}(\Omega)}\leq 1\}$ is equal to $K$. \end{proof} Let us mention that, although the description of the extreme points follows directly from the results of Section 3, proving the existence of minimizers and the existence of extreme points is a considerably more difficult problem which needs a careful choice of topologies. The paper \cite{unser2017splines} provides a systematic way to construct Banach spaces and pseudo-inverse $L^+$ for ``spline admissible operators'' $L$ such as the fractional Laplacian. In addition, they prove existence of solutions by adding weak-* continuity assumptions on the sensing operator $\Phi$.} \subsubsection{The total gradient variation} Since its introduction in the field of image processing \cite{rudin1992nonlinear}, the total gradient variation proved to be an extremely valuable regularizer in diverse fields of data science and engineering. It is defined, for any locally integrable function $u$ as \begin{equation*} TV(u) \eqdef \sup\left(\int u\mathrm{div}(\phi) \, dx, \phi \in C^1_c(\RR^d)^d, \sup_{x\in \RR^d} \|\phi(x)\|_2\leq 1\right). \end{equation*} If the above quantity is finite, we say that $u$ has bounded variation and its gradient $D u$ is a Radon measure, with \begin{equation*} TV(u) = \int_{\RR^d}|D u| = \|Du \|_{(\Mm(\RR^d))^d}. \end{equation*} Working in $\vecgal=L^{d/(d-1)}(\RR^d)$, one is led to consider the convex set $\cvx=\{ u\in \vecgal, TV(u)\leq 1\}$, referred to as the TV unit ball. The generalized gradient operator is not a surjective operator. Hence, the analysis of \Cref{sec:prior} cannot help finding the extreme points of the TV ball. Still, those have been described in the fifties by Fleming in \cite{fleming1957functions} and refined analyses have been proposed more recently by Ambrosio, Caselles, Masnou and Morel in \cite{ambrosio2001connected}. \begin{theorem}[Extreme points of the TV ball \cite{fleming1957functions,ambrosio2001connected}] \label{thm:extremeBV} The extreme points of the unit TV unit ball are the indicators of simple sets normalized by their perimeter, i.e. functions of the form $u=\pm \frac{\mathbbm{1}_F}{TV(\mathbbm{1}_F)}$, where $F$ is an indecomposable and saturated subset of $\RR^d$. \end{theorem} Informally, the simple sets of $\RR^d$ are the simply connected sets with no hole. We refer the reader to \cite{ambrosio2001connected} for more details. Using \cref{thm:extremeBV} in conjunction with our results tell us that functions minimizing the total variation subject to a finite number of linear constraints can be expressed as a sum of a small number of indicators of simple sets, see for instance \cref{fig:extremeTV}, which is yet another theoretical result explaining the common observation that total variation tends to produce stair-casing \cite{nikolova2000local}. \begin{figure} \begin{center} \begin{tabular}{cc} \includegraphics[height=4cm]{meas_support_3.pdf} & \includegraphics[height=4cm]{blobs.pdf} \\ { (a)} & { (b)} \end{tabular} \caption{\label{fig:extremeTV}Illustration for the total gradient variation problem $\min \left\{ TV(u) : \Phi(u) =y \right\}$. Here, $\Phi$ is a linear mapping giving access to $3$ measurements, $y\in \mathbb{R}^3$, by performing the mean of an image $u$ of size $200 \times 200$ on $3$ different disks represented in (a). The TV problem is solved using a primal-dual algorithm, also known as the Chambolle-Pock algorithm \cite{Chambolle2011}. The recovered image is displayed in (b): it can be represented as the sum of $3$ indicator functions of simple sets. } \end{center} \end{figure} \section*{Acknowledgments} This work was initially started by two different groups composed of A. Chambolle, F. de Gournay and P. Weiss on one side and C. Boyer, Y. De Castro and V. Duval on the other side. The authors realized that they were working on a similar topic when a few of them met at the Cambridge semester in mathematical imaging during the Isaac Newton Institute (Cambridge) semester ``Variational methods and effective algorithms for imaging and vision'', supported by EPSRC Grant N.~EP/K032208/1. They therefore decided to write a joint paper and wish to thank the organizers for giving them this opportunity. This work initially started with the help of T. Pock through a few numerical experiments and with discussions with J. Fadili and C. Poon. The work of A.C. was partially supported by a grant of the Simons Foundation. \bibliographystyle{abbrv} \section{Conclusion} In this paper we have developed representer theorems for convex regularized inverse problems \eqref{eq::mainproblem}, based on fundamental properties of the geometry of convex sets: the solution set can be entirely described using convex combinations of a small number of extreme points and extreme rays of the regularizer level set. Obviously, the conclusion of Theorem~\ref{thm:first} is only nontrivial when $\minset$ has a ``sufficiently flat boundary'', in the sense that two or more faces of $\minset$ have dimension larger than $m$. For instance, if $\minset$ is strictly convex (\ie{} has only $0$-dimensional faces, except its interior\footnote{In this example, to simplify the discussion, we assume that $\vecgal$ has finite dimension.}), then the solution set $\sol$ is always reduced to a single extreme point of $\minset$! Nevertheless, several regularizers which are commonly used in the literature (notably sparsity-promoting ones) have that flatness property, and Theorem~\ref{thm:first} then provides interesting information on the structure of the solution set, as illustrated in Section~\ref{sec:app}. To conclude, the structure theorem presented in this paper highlights the importance of describing the extreme points and extreme rays of the regularizer: this yields a fine description of the set of solutions of variational problems of the form~\eqref{eq::mainproblem}. Our theorem also suggests a principled way to design a regularizer. If a particular family of solutions is expected, then one may construct a suitable regularizer by taking the convex hull of this family. Finally, representer theorems have had a lot of success in the fields of approximation theory and machine learning \cite{scholkopf2001generalized}, in the frame of reproducible kernel Hilbert spaces. A reason of this success is that they allow to design efficient numerical procedures that yield \emph{infinite dimensional} solutions by solving \emph{finite dimensional} li\-near systems. Such numerical procedures have recently been extended to the case of Banach spaces for some simple instances of the problems described in this paper \cite{fernandez2016super,de2017exact,flinth2017exact}. The price to pay when going from a Hilbert space to a Banach space is that semi-infinite convex programs have to be solved instead of simpler linear systems. We foresee that the results in this paper may help designing new efficient numerical procedures, since they allow to parameterize the solutions using extreme points and extreme rays only. \section{Introduction} Let $\vecgal$ denote a {real} vector space. Let $\Phi : \vecgal\to \RR^m$ be a linear mapping called \emph{sensing operator} and $u\in \vecgal$ denote a signal. The main results in this paper describe the structural properties of certain solutions of the following problem: \begin{equation}\label{eq::mainproblem} \inf_{u \in \vecgal} f(\Phi u) + \reg(u), \end{equation} where $R:\vecgal\to \mathbb{R}\cup\{+\infty\}$ is a convex function called \emph{regularizer} and $f$ is an arbitrary convex or non-convex function called \emph{data fitting term}. { In many applications, one looks for ``sparse solutions'' that are linear sums of a few atoms. This article investigates the theoretical legitimacy of this usage. } \paragraph{Representer theorems and Tikhonov regularization} The name \emph{representer theorem} comes from the field of machine learning \cite{Scholkopf2002Learning}. To provide a first concrete example\footnote{Here, we follow the presentation of \cite{gupta2018continuous}.}, assume that $\Phi \in \RR^{m\times n}$ is a finite dimensional measurement operator and $L\in \mathbb{R}^{p\times n}$ is a linear transform. Solving an inverse problem using Tikhonov regularization amounts to finding the mini\-mi\-zers of \begin{equation}\label{eq:tikhonov} \min_{u\in \RR^m} \frac{1}{2}\|\Phi u - y\|_2^2 + \frac{1}{2}\|Lu\|_2^2. \end{equation} Provided that $\mathrm{ker}\Phi\cap \mathrm{ker} L=\{0\}$, it is possible to show that, whatever the data $y$ is, solutions are always of the form \begin{equation}\label{eq:firstrepresentertheorem} u^\star = \sum_{i=1}^m \alpha_i \psi_i + u_K, \end{equation} where $u_K\in \mathrm{ker}(L)$ and $\psi_i=(\Phi^T\Phi + L^TL)^{-1}(\phi_i)$, where $\phi_i^T\in \RR^n$ is the $i$-th row of~$\Phi$. This result characterizes structural properties of the minimi\-zers without actually needing to solve the problem. In addition, when $\vecgal$ is an infinite dimensional Hilbert space, Equation \eqref{eq:firstrepresentertheorem} sometimes allows to compute exact solutions, by simply solving a finite dimensional linear system. This is a critical observation that explains the practical success of kernel methods and radial basis functions \cite{Wendland2005Scattered}. \paragraph{Representer theorems and convex regularization} The Tikhonov regularization \eqref{eq:tikhonov} is a powerful tool when the number $m$ of observations is large and the operator $\Phi$ is not too ill-conditioned. However, recent results in the fields of compressed sensing \cite{Donoho2006Compressed}, matrix completion \cite{candes2009exact} or super-resolution \cite{Tang2013Compressed,candes2014towards} - to name a few - suggest that much better results may be obtained in general, by using convex regularizers, with level sets containing singularities. Popular examples of regularizers in the finite dimensional setting include the indicator of the nonnegative orthant \cite{donoho2005sparse}, the $\ell^1$-norm~\cite{Donoho2006Compressed} or its composition with a linear operator \cite{rudin1992nonlinear} and the nuclear norm \cite{candes2009exact}. Those results were nicely unified in \cite{chandrasekaran2012convex} and one of the critical arguments behind all these techniques is a representer theorem of type \eqref{eq:firstrepresentertheorem}. In most situations however, this argument is only implicit. The main objective of this paper is to state a generalization of \eqref{eq:firstrepresentertheorem} to arbitrary convex functions $R$. It covers all the aforementioned problems, but also new ones for problems formulated over the space of measures. To the best of our knowledge, the name ``{\it representer theorem}'' is new in the field of convex regularization and its first mention is due to Unser, Fageot and Ward in~\cite{unser2017splines}. Describing the solutions of~\eqref{eq::mainproblem} is however an old problem which has been studied since at least the 1940's in the case of Radon measure recovery. \paragraph{Total variation regularization of Radon measures} A typical example of inverse problem in the space of measures i \begin{equation}\label{eq::beurling} \min_{\mu\in \Mm(\Omega)} |\mu|(\Omega) \quad \mbox{s.t.}\quad \Phi \mu=y \end{equation} where $\Omega\subseteq \RR^N$, $\Mm(\Omega)$ denotes the space of Radon measures, $|\mu|(\Omega)$ is the total variation of the measure $\mu$ (see Section~\ref{sec:app}) and $\Phi \mu$ is a vector of \textit{generalized moments}, \ie{} $\Phi \mu=\left(\int_\Omega \varphi_i(x)\d\mu(x)\right)_{1\leq i\leq m}$ where $\{\varphi_i\}_{1\leq i\leq m}$ is a family of continuous functions (which ``vanish at infinity'' if $\Omega$ is not compact). Problems of the form~\eqref{eq::beurling} have received considerable attention since the pioneering works of Beurling~\cite{Beurling1938} and Krein \cite{Krein1938}, sometimes under the name \textit{L-moment problem} (see the monograph~\cite{krein_markov_1977}). To the best of our knowledge, the first ``representer theorem'' for problems of the form~\eqref{eq::mainproblem} is given for~\eqref{eq::beurling} by \zuho{}~\cite{Zuhovickii1948} (see~\cite[Th. 3]{Zuhovickii1962} for an English version). It essentially states that \begin{equation}\label{statement1} \mbox{\textit{There exists a solution to~\eqref{eq::beurling} of the form $\displaystyle\sum_{i=1}^{r}a_i \delta_{x_i}$, with $r\leq m$.}} \end{equation} {A more precise result was given by Fisher and Jerome in~\cite{fisher_spline_1975}. When considering the problem \eqref{eq::beurling}, and for a bounded domain $\Omega$, the result reads as follows: \begin{align}\label{statement2} \begin{split} &\mbox{\textit{The extreme points of the solution set to~\eqref{eq::beurling} are of the form}} \\ &\qquad\qquad \qquad\qquad\qquad \displaystyle\sum_{i=1}^{r}a_i \delta_{x_i}, \mbox{ \textit{with} } r\leq m. \end{split} \end{align} Incidentally, the Fisher-Jerome theorem considers more general problems of the form: \begin{equation}\label{eq::spline} \min_{u\in \vecgal} |Lu|(\Omega) \quad \mbox{s.t.}\quad Lu\in \Mm(\Omega) \qandq \Phi u =y, \end{equation} where $\vecgal\subseteq \mathcal{D}'(\Omega)$ is a suitably defined Banach space of distributions, $L:\mathcal{D}'(\Omega)\rightarrow \mathcal{D}'(\Omega)$ maps $\vecgal$ onto $\Mm(\Omega)$ and $\Phi:\vecgal\to \RR^m$ is a continuous linear operator. We refer to Section~\ref{sec:app} for precise assumptions. Let us mention that the initial results by Fisher-Jerome were extended to a significantly more general setting in \cite{unser2017splines}.} It is important to note that the Fisher-Jerome theorem~\cite{fisher_spline_1975} provides a much finer description of the solution set than \zuho's result \cite{Zuhovickii1948}. Indeed, the well-known Krein-Milman theorem states that, if $\vecgal$ is endowed with the topology of a locally convex Hausdorff vector space and $\cvx\subset \vecgal$ is compact convex, then $\cvx$ is the closed convex hull of its extreme points, \begin{equation} \label{eq:krein} \cl \mathrm{conv} \left(\ext(\cvx)\right)=\cvx. \end{equation} In other words, the solutions described by the Fisher-Jerome theorem are sufficient to recover \emph{the whole set of solutions}. Let us mention that the Krein-Milman theorem was extended by Klee~\cite{klee_extremal_1957} to unbounded sets: if $C$ is locally compact, closed, convex, and contains no line, then \begin{equation} \label{eq:klee} \cl \mathrm{conv} \left(\ext(\cvx)\cup\rext(\cvx)\right)=\cvx, \end{equation} where $\rext(\cvx)$ denotes the union of the extreme rays of $\cvx$ (see Section~\ref{sec:notations} below). \paragraph{``Representer theorems'' for convex sets} As the Dirac masses are the extreme points of the total variation unit ball, each of the above-mentioned ``representer theorems'' for inverse problems actually reflect some phenomenon in the geometry of convex sets. In that regard, the celebrated Minkowski-Carath\'eodory theorem~\cite[Th.~III.2.3.4]{hiriart-urruty_convex_1993} is fundamental: any point of a compact convex set in an $m$-dimensional space is a convex combination of (at most) $m+1$ of its extreme points. In~\cite[Th.~(3)]{klee_theorem_1963}, Klee removed the boundedness assumption and obtained the following extension: any point of a closed convex set in an $m$-dimensional space is a convex combination of (at most) $m+1$ extreme points, or $m$ points, each an extreme point or a point in an extreme ray. One purpose of the present paper is to point out the connection between the Fisher-Jerome theorem and a lesser known theorem by Dubins~\cite{dubins1962extreme} (see also~\cite[Exercise II.7.3.f]{bourbaki_espaces_2007}): \medskip \begin{center} \noindent\emph{The extreme points of the intersection of $\cvx$ with an affine space of codimension $m$ are convex combination of $($\!at most$)$\footnote{In the rest of the paper, we omit the mention ``at most'', with the convention that some points may be chosen identical.} $m+1$ extreme points of~$\cvx$}, \end{center} \medskip \noindent provided $\cvx$ is linearly bounded and linearly closed (see Section~\ref{sec:notations}). That theorem was extended by Klee~\cite{klee_theorem_1963} to deal with the unbounded case. Although the connection with the Fisher-Jerome is striking, Dubins' theorem actually provides one extreme point too many. In the case of~\eqref{eq::beurling}, it would yield two Dirac masses for one linear measurement. We provide in this paper a refined analysis of the case of variational problems, which ensures at most $m$ extreme points. \paragraph{Contributions} The main results of this paper yield a description of some solutions to \eqref{eq::mainproblem} of the following form: \begin{equation*} u^\star = \sum_{i=1}^r \alpha_i \psi_i + u_K, \end{equation*} where $r\leq m$, the atoms $\psi_i$ are identified with some extreme points (or points in extreme rays) of the regularizer level sets, and $u_K$ is an element of the so-called constancy space of $R$, \ie{} the set of directions along which $\reg$ is invariant. The results take the form \eqref{statement1}, when $f$ is an arbitrary function and the form \eqref{statement2} when it is convex. We provide tight bounds on the number of atoms $r$ that depend on the geometry of the level sets and on the link between the constancy space of $R$ and the measurement operator $\Phi$. Our general theorems then allow us to revisit many results of the literature (linear programming, semi-definite programming, nonnegative constraints, nuclear norm, analysis priors), yielding simple and accurate descriptions of the minimizers. Our analysis also allows us to characterize the solutions of a resisting problem: we provide a representation theorem for the minimizers of the total gradient variation \cite{rudin1992nonlinear} as sums of indicators of simple sets. This provides a simple explanation to the staircaising effect when only a few measurements are used. {Let us mention that, shortly after this work was posted on arXiv, similar results appeared, with somewhat different proofs, in a paper by Bredies and Carioni~\cite{bredies_sparsity_2018}.} \section{Abstract representer theorems} \label{sec:abstract} \subsection{Main result} Our main result describes the facial structure of the solution set to \begin{equation}\label{eq:thminreg} \min_{u\in\vecgal} \reg(u) \quad \mbox{s.t.}\quad \Phi u=y, \tag{$\Pp$} \end{equation} where $y\in \RR^m$, $\Phi:\vecgal \rightarrow \RR^m$ is a linear operator, and $m\leq \mathrm{dim} \vecgal$, $m<+\infty$. In the following, let $\minval$ denote the \emph{optimal value} of~\eqref{eq:thminreg}, $\sol$ denote its \emph{solution set}, and~$\minset$ denote the \emph{corresponding level set} of $\reg$, \begin{align} \label{eq:defminset} \minset\eqdef \enscond{u\in \vecgal}{\reg(u)\leq \minval}. \end{align} \begin{theorem} \label{thm:first} Let $\reg:\vecgal \to \RR\cup\{+\infty\}$ be a convex function. Assume that $\inf_\vecgal \reg< t^\star< +\infty$, that $\sol$ is nonempty and that the convex set $\minset$ is linearly closed and contains no line. Let $p\in\sol$ and let $j$ be the dimension of the face $\face{p}{\sol}$. Then $p$ belongs to a face of $\minset$ with dimension at most $m+j-1$. In particular, $p$ can be written as a convex combination of: \begin{itemize} \item[$\circ$] $m+j$ extreme points of $\minset$, \item[$\circ$] or $m+j-1$ points of $\minset$, each an extreme point of $\minset$ or in an extreme ray of $\minset$. \end{itemize} Moreover, $\rec{\sol}=\rec{\minset}\cap \mathrm{ker}(\Phi)$ and therefore $\mathrm{lin}(\sol)=\mathrm{lin}(\minset)\cap \mathrm{ker}(\Phi)$. \end{theorem} \begin{figure} \centering \tdplotsetmaincoords{63}{55} \input{fig-th} \caption{An illustration of \cref{thm:first} for $m=2$. The solution set $\sol=\minset\cap \Phi^{-1}(\{y\})$ is made of an extreme point and an extreme ray. The extreme point is a convex combination of~$\{e_0,e_1\}$. Depending on their position, the points in the ray are a convex combination of~$\{e_0,e_1,e_2\}$ or a pair of points, one in $\rho_1$ and the other in $\rho_2$.} \label{fig:thm} \vspace{0.5cm} \end{figure} The proof of \cref{thm:first} is given in Section~\ref{sec:prooffirst}. Before extending this theorem to a wider setting, let us formulate some remarks. \begin{remark}[Extreme points and extreme rays of $\sol$] In particular ($j=0$), \emph{each extreme point} of $\sol$ is a convex combination of $m$ extreme points of~$\minset$, or a convex combination of $m-1$ points of~$\minset$, each an extreme point of~$\minset$ or in an extreme ray. Similarly $(j=1)$, \emph{each point on an extreme ray} of~$\sol$ is a convex combination of $m+1$ extreme points of~$\minset$, or a convex combination of $m$ points of~$\minset$, each an extreme point of $\minset$ or in an extreme ray. {Hence, provided the assumptions of Klee's theorem (see \eqref{eq:klee}) hold, Theorem \ref{thm:first} completely charaterizes the solution set.} An illustration is provided in \cref{fig:thm}. \end{remark} \begin{remark}[The hypothesis $\inf_\vecgal \reg< t^\star$\label{rem:infR}] We have focused on the case $\minval>\inf_\vecgal \reg$ in the theorem since the case $\minval=\min \reg$ is easier. In that case, $M=\Phi^{-1}(\{y\})$ can be in arbitrary position (\ie{} not necessarily tangent) w.r.t.\ $\minset=\argmin \reg$, and one can only use the general Dubins-Klee theorem~\cite{dubins1962extreme,klee_theorem_1963} to describe their intersection. As a result the conclusions of \cref{thm:first} are slightly weakened, $p$ belongs to a face of $\minset$ with dimension $m+j$, and one must add one more point in the convex combination (\eg{}, for $j=0$, each extreme point of $\sol$ is a convex combination of $m+1$ extreme points of $\minset$, or $m$ points\dots). \end{remark} \begin{remark}[Gauge functions or semi-norms] A common practice in inverse problems is to consider positively homogeneous regularizers $\reg$, such as (semi)-norms or gauge functions of convex sets. In that case the extreme points of $\minset$ correspond, up to a rescaling, to the extreme points of $\{u\in\vecgal : \reg(u)\leq 1\}$. In several cases of interest, the extreme points of such convex sets are well understood, see Section~\ref{sec:app} for examples in Banach spaces or, for instance, the paper \cite[Sec. 2.2]{chandrasekaran2012convex} for examples in finite dimensional spaces. \end{remark} \begin{remark}[Extension to semi-strictly quasi convex functions] \Cref{thm:first} can be extended to the case where $R$ is a semi-strictly quasi-convex function. A function $R$ is said to be {\em semi-strictly quasi-convex} \cite{daniilidis2007some} if it is quasi-convex and if \[ R(x)< R(y) \Longrightarrow R(\lambda x + (1-\lambda) y) < R(y) \quad \forall \lambda \in \oi{0}{1}. \] In words, semi-strictly quasi-convex functions are functions that, when restricted to a line are successively decreasing, constant and increasing on their domain. In comparison, strictly quasi-convex functions are successively decreasing and increasing while quasi-convex functions successively non-increasing and non-decreasing. The set of semi-strictly quasi-convex functions is a subset of quasi-convex functions and it contains all convex and strictly quasi-convex functions. In the proof of \cref{thm:first}, only the semi-strictly quasi-convex property is required to ensure that \eqref{eq:for_cvx_like} holds. \end{remark} \begin{remark}[Topological properties\label{rem:lsc}] The assumption that $\minset$ is linearly closed is fulfilled in most practical cases, since $\vecgal$ is usually endowed with the topology of a Banach (or locally convex) vector space and $\reg$ is assumed to be lower semi-continuous (so as to guarantee the existence of a solution to~\eqref{eq:thminreg}). {Note also that if $\reg$ is lower semi-continuous on any line (for the natural topology of the line), the set $\minset$ is linearly closed. } \end{remark} \subsection{The case of level sets containing lines} The reader might be intrigued by the assumption of \cref{thm:first} that $\minset$ contains no line, since in several applications the regularizer $\reg$ is invariant by the addition of, \eg, constant functions or low-degree polynomials (see Section~\ref{sec:app}). In that case, one is generally interested in the non-constant or non-polynomial part, and it is natural to consider a quotient problem for which the theorem applies. We describe below (see \cref{coro:lines}) how our result extends to the case where $\minset$ contains lines. \begin{figure}[htbp] \centering \input{fig-coro1b} \caption{Taking the quotient by $\vecrec{}=\mathrm{lin}(\minset)$ yields a level set $\tilde\minset$ with no line. In this figure, to simplify the notation, we have omitted the isomorphism $\qiso$ (\ie{} in this figure $\minset$ shoud be replaced with $\qiso(\minset)$, and similarly for $\sol$ and $\Phi^{-1}(\{y\})$).} \label{fig:coro1b} \end{figure} If $\minset$ is linearly closed and contains some line, it is translation-invariant in the corresponding direction. The collection of all such directions is the lineality space of $\minset$ (see \Cref{sec:notations}), we denote it by $\vecrec{}\eqdef \mathrm{lin}(\minset)$ (typically, if $\reg$ is the composition of a linear operator and a norm, $\vecrec{}$ is the \emph{kernel} of that linear operator). Let $\projv:\vecgal \rightarrow \vecgal/\vecrec{}$ be the canonical projection map. We recall that there exists a linear isomorphism $\qiso: \vecgal\rightarrow (\vecgal/\vecrec{})\times \vecrec{}$ such that the first component of $\qiso(p)$ is $\projv(p)$ for all $p\in \vecgal$. We may now describe the equivalence classes (modulo $\vecrec{}$) of the solutions. \begin{corollary} \label{coro:lines} Let $\reg:\vecgal \to [-\infty,+\infty]$ be a convex function. Assume that $\inf_\vecgal \reg< t^\star< +\infty$, that $\sol$ is nonempty and that the convex set $\minset$ is linearly closed. Let $\vecrec{}\eqdef\mathrm{lin}(\minset)$ be the lineality space of $\minset$ and $d\eqdef \mathrm{dim} \Phi(\vecrec{})$. Let $p\in\sol$, let $\projv(p)$ denote its equivalence class, and let $j$ be the dimension of the face $\face{\projv(p)}{\projv(\sol)}$. Then, $\projv(p)$ belongs to a face of $\projv(\minset)$ with dimension at most $m+j-d-1$. In particular, \begin{itemize} \item[$\circ$] $\projv(p)$ is a convex combination of $m+j-d$ extreme points of $\projv(\minset)$, \item[$\circ$] or $\projv(p)$ is a convex combination of $m+j-d-1$ points of $\projv(\minset)$, each an extreme point of $\projv(\minset)$ or in an extreme ray of $\projv(\minset)$. \end{itemize} As a result, letting $\qe_1,\ldots, \qe_r$ denote those extreme points (or points in extreme rays), \begin{equation}\label{eq:convcomb} p= \sum_{i=1}^r \theta_i \qiso^{-1}(\qe_i,0) + u_{\vecrec{}}, \qwhereq \theta_i\geq 0,\ \sum_{i=1}^r \theta_i =1, \qandq u_{\vecrec{}}\in \vecrec{}. \end{equation} \end{corollary} The proof of \cref{coro:lines} is given in \Cref{sec:proofconvex}. One can have an explicit representation with elements of $E$ of a solution $p\in \sol $. Indeed, let $W$ be some linear complement to $K=\mathrm{lin}(\minset)$. One may decompose $\minset = \tilde \minset + K$, where $\tilde{\minset}=\cvx\cap W$, and observe that $\projv (\minset) $ and $\tilde \minset$ are isomorphic. In this case, \cref{coro:lines} implies that $p$ can be written as the sum of one point in $\mathrm{lin}(\minset)$ and of a convex combination of: \begin{itemize} \item[$\circ$] $m+j-d$ extreme points of $\tilde \minset$, \item[$\circ$] or $m+j-1-d$ points of $\tilde \minset$, each an extreme point of $\tilde \minset$ or in an extreme ray of $\tilde \minset$. \end{itemize} \subsection{Extensions to data fitting functions} In this section, we discuss the extension of the above results to more general problems of the form \begin{equation} \label{eq::data-fitting:convex} \inf_{u \in \vecgal} f(\Phi u)+\reg(u), \tag{$\Pp_f$} \end{equation} where $f:\RR^m\rightarrow \RR\cup\{+\infty\}$ is an arbitrary fidelity term. \subsubsection{Convex data fitting term} When $f$ is a \emph{convex} data fitting function $f$, we get the following result. \begin{corollary} \label{cor:convex_fit} Assume that $f$ is convex and that the solution set $\sol_{f}$ of \eqref{eq::data-fitting:convex} is nonempty. Let $p\in \sol_{f}$ such that $\minset\eqdef \{u\in \vecgal, \reg(u)\leq \reg(p)\}$ is linearly closed, $\vecrec{}\eqdef\mathrm{lin}{\minset}$ and let $j$ be the dimension of the face $\face{\projv(p)}{\projv(\sol_{f})}$. If $\inf_\vecgal \reg < \reg(p)$, then the conclusions of \cref{coro:lines} (or \cref{thm:first} if $\vecrec{}=\{0\}$) hold. If $\inf_\vecgal \reg = \reg(p)$, they hold with $1$ more dimension (see \cref{rem:infR}). \end{corollary} Let us recall that, in view of \cref{rem:lsc}, if $\vecgal$ is a topological vector space and $R$ is lower semi-continuous, $\minset$ is closed regardless of the choice $p$. \begin{proof} Let $y=\Phi p$ and consider the following problem: \begin{align} \label{eq:singleton} \tag{$\Pp_{\{y\}}$} \min_{u\in\vecgal } \reg(u) &\quad \mbox{s.t.}\quad \Phi u = y. \end{align} Let $\sol_{\{y\}}$ denote its solution set. It is a convex subset of $\sol_{f}$, with $p\in \sol_{\{y\}}$. Additionally, if $j$ is the dimension of $\face{p}{\projv(\sol_f)}$ (resp. $\face{p}{\sol_f}$ if $\minset$ contains no line), then the face $\face{p}{\projv(\sol_{\{y\}})}$ (resp. $\face{p}{\sol_{\{y\}}}$) has dimension at most $j$, since $\sol_{\{y\}}\subseteq \sol_{f}$. It suffices to apply \cref{coro:lines} (resp. \cref{thm:first}) to obtain the result. \end{proof} \begin{remark}[The case of a strictly-convex function] In the case when $f$ is strictly convex, it is known that $\Phi \sol_f$ is a singleton, which means that $\sol_{\{y\}}=\sol_f$. \end{remark} \begin{remark}[The case of quasi-convex functions] The result actually holds whenever the solution set $\sol_f$ is convex. In particular, this property holds when $f$ is quasi-convex and $R$ is convex. \end{remark} \subsubsection{Non-convex function} \label{subsec:nonconvex} In the general case, \ie{} when $f:\RR^m\rightarrow \RR\cup\{+\infty\}$ is an arbitrary function, it is difficult to describe the structure of the solution set. However, one may choose a solution $p_0$ (as before, provided it exists) and observe that it is also a solution to~\eqref{eq:thminreg} for $y\eqdef \Phi p_0$. Then, one may apply \cref{coro:lines}, but the difficult part is that the dimensions $j$ to consider are with respect to the solution set $\sol$ of the \emph{convex} problem~\eqref{eq:thminreg}. Nevertheless, if one is able to assert that the solution set $\sol$ has at least an extreme point $p$, then \cref{coro:lines} ensures that $p$ can be written in the form~\eqref{eq:convcomb}, where $r\leq m$ and the $\qe_i$'s are extreme points (or points in extreme rays) of $\minset$. Since $p$ must also be a solution to~\eqref{eq::data-fitting:convex}, one obtains that there exists a solution to~\eqref{eq::data-fitting:convex} of the form~\eqref{eq:convcomb}. \subsection{Ensuring the existence of extreme points}\label{sec:existextreme} It is important to note that, in \cref{thm:first}, the existence a face of $\sol$ with dimension $j$ is not guaranteed (nor, for $j=0$, the existence of extreme points). The convex set $\sol$ might not even have any finite-dimensional face! For instance, let $\vecgal$ be the space of Lebesgue-integrable functions on $[0,1]$. If $\reg(u)=\int_0^1|u(x)|\d x$, $\Phi:\vecgal\rightarrow \RR$ is defined by $u\mapsto \int_0^1 u(x)\d x$, and $y=1$, then \begin{equation} \sol=\enscond{u\in\vecgal}{\int_0^1 |u(x)|\d x\leq 1 \qandq \int_0^1 u(x)\d x=1}. \end{equation} It is possible to prove that such a set $\sol$ does not have any extreme point. As a consequence $\sol$ does not have any finite-dimensional face (otherwise an extreme point of the closure of a face would be an extreme point of $\sol$). However, \cref{thm:first} (in fact the Dubins-Klee theorem~\cite{dubins1962extreme,klee_theorem_1963}) asserts that, \textit{if there is} a finite-dimensional face in $\sol$, then $\minset$ has indeed extreme points (and possibly extreme rays), and the convex combinations of such points generate the above-mentioned face. As a result, it is crucial to be able to assert a priori the existence of some finite-dimensional face for $\sol$, and this is where topological arguments come into play. If $\vecgal$ is endowed with the topology of a locally convex (Hausdorff) vector space, the theorems~\cite[3.3 and 3.4]{klee_extremal_1957} which generalize the celebrated Krein-Milman theorem, state that \emph{$\sol$ has an extreme point provided} \begin{itemize} \item[$\circ$] $\sol$ is nonempty, convex, \item[$\circ$] $\sol$ contains no line, \item[$\circ$] and $\sol$ is closed, locally compact. \end{itemize} The last two conditions hold in particular if $\sol$ is compact. Moreover, as in \cref{coro:lines}, the second condition can be ensured by considering a suitable quotient map, provided it preserves the other topological properties (\eg{} if $\mathrm{lin}(\cvx)$ has a topological complement). { \begin{remark} Whereas local compactness is a very strong property for topological vector spaces (implying their finite-dimensionality, see~\cite[Th.~3,Ch.~1]{bourbaki_espaces_2007}), it is not so difficult to ensure the local compactness of $\sol$ in practice. Indeed, very often, even the existence of solutions is usually ensured using compactness arguments for a suitable weak or weak-* topology. The unbounded cases require more specific arguments, but let mention that there are examples of cones which are locally compact without being contained in any finite-dimensional vector space. In Section~\ref{sec:momprob} below, we discuss the example of the cone $\mathcal{M}^+(\Omega)$ of non-negative measures over a compact set for the weak-* topology. Another example of locally compact convex cone is \[ \mathcal C=\Big\{ x\in\RR^\NN \text{ such that } x_n \ge 0 \text{ and } \sum_{n\in\mathbb{N} } x_n \omega_n \le \sum_{n\in\mathbb{N} } x_n <+\infty \Big\}\subseteq \ell^1(\NN)\,, \] for some non-decreasing positive sequence $(\omega_n)_n$ converging to $+\infty$ (note that $\omega_0<{1}$ for the cone to be non-empty). The cone $\mathcal{C}$ is locally compact for the strong topology. Indeed, consider the intersection $K$ of the cone $\mathcal C$ and the strong unit ball, namely $K = \{x=(x_n)_n\ : \ x_n\geq0,\ \sum_n \omega_n x_n \le \sum_{n } x_n\leq 1\}$ and consider a sequence of elements of $K$ denoted~$(x^k)_k \subset K$. Using a diagonal argument, each $(x_n^k)_k$ converges to some $\bar{x}_n \geq0$. Furthermore, using that $\{n\,:\, w_n<1\}$ is finite and Fatou's lemma, it holds that \begin{align*} \sum_{n } \bar x_n&\leq \liminf_k \sum_{n } x^k_n\leq1\quad \mathrm{and}\\%\quad\text{and}\quad \sum_{n\,:\, w_n\geq1 } (w_n-1)\bar x_n&\leq \liminf_k \sum_{n\,:\, w_n\geq1 } (w_n-1) x^k_n\\ &\leq \liminf_k \sum_{n\,:\, w_n<1 } (1-w_n) x^k_n=\sum_{n\,:\, w_n<1 } (1-w_n)\bar x_n \end{align*} and we deduce that $\bar x= (\bar{x}_n)_n\in K$. Furthermore, one has \[ ||x^k-\bar{x}||_1 = \sum_{n=0}^M |x_n^k-\bar{x}_n| + \sum_{n>M} |x_n^k-\bar{x}_n| \] and $M$ will be chosen later. Finally, \begin{align*} \sum_{n>M} |x_n^k-\bar{x}_n| &\leq \sum_{n>M} (x_n^k+ \bar{x}_n) \leq (1/\omega_M) \sum_{n>M} \omega_n (x_n^k+\bar{x}_n) \leq 2/\omega_M \end{align*} since $\omega_n/\omega_M \geq 1$ for $n>M$. Therefore choosing $M$ large enough ensures that the second term $\sum_{n>M} |x_n^k-\bar{x}_n|$ is less than some $\varepsilon>0$. Choosing $k$ large enough leads to $||x^k-\bar{x}||_1\leq 2\varepsilon$. \end{remark} } \section{Notation and Preliminaries}\label{sec:notations} Throughout the paper, unless otherwise specified, $\vecgal$ denotes a finite or infinite dimensional real vector space and $\cvx\subseteq \vecgal$ is a convex set. Given two distinct points $x$ and $y$ in $\vecgal$, we let $\oi{x}{y}=\enscond{tx+(1-t)y}{0<t<1}$ and $[x,y]=\{tx+(1-t)y : $ $0\leq t\leq 1\}$ denote the open and closed segments joining $x$ to $y$. We recall the following definitions, and we refer to~\cite{dubins1962extreme,klee_extremal_1957} for more details. \paragraph{Lines, rays, and linearly closed sets} A \textit{line} is an affine subspace of $\vecgal$ with dimension $1$. An open half-line, \ie{} a set of the form $\rho=\{ p+tv : $ $t>0 \}$, where $p,v\in \vecgal$, $v\neq 0$, is called a \emph{ray} (through $p$). We say that the set $\cvx$ is \emph{linearly closed} (resp. linearly bounded) if the intersection of $\cvx$ and a line of $\vecgal$ is closed (resp. bounded) for the natural topology of the line. If $\vecgal$ is a topological vector space and $\cvx$ is closed for the corresponding topology, then $\cvx$ is linearly closed. If $\cvx$ is linearly closed and contains some ray $\rho=p+\RR_+^*v$, it also contains the endpoint $p$ as well as the rays $q+\mathbb{R}_+v$ for all $q\in \cvx$. Therefore, if $\cvx$ contains a ray (resp. line), it recesses in the corresponding direction. \paragraph{Recession cone and lineality space} The set of all $v\in\vecgal$ such that $\cvx+\RR_+^*v\subseteq \cvx$ is a convex cone called the \emph{recession cone of $\cvx$}, which we denote by $\rec{\cvx}$. If $\cvx$ is linearly closed then so is $\rec{\cvx}$, and $\rec{\cvx}$ is the union of $0$ and all the vectors $v$ which direct the rays of $\cvx$. In particular, $\cvx$ contains a line if and only the vector space \begin{equation} \label{eq::lineality space} \mathrm{lin}(\cvx)\eqdef \rec{\cvx}\cap (-\rec{\cvx}) \end{equation} is non trivial. The vector space $\mathrm{lin}(\cvx)$ is called the \emph{lineality space} of $\cvx$. It corresponds to the largest vector space of invariant directions for $\cvx$. If $\vecgal$ is finite dimensional and $\cvx$ is closed, the recession cone coincides with the \emph{asymptotic cone}. \paragraph{Extreme points, extremal rays, faces} An \textit{extreme point} of $\cvx$ is a point $p\in \cvx$ such that $\cvx\setminus \{p\}$ is convex. An \textit{extremal ray} of $\cvx$ is a ray $\rho \in \cvx$ such that if $x,y\in \cvx$ and $\oi{x}{y}$ intersects $\rho$, then $\oi{x}{y}\subset \rho$. If $\cvx$ contains the endpoint~$p$ of $\rho$ (\eg{} if $\cvx$ is linearly closed), this is equivalent to $p$ being an extreme point of $\cvx$ and $\cvx\setminus \rho$ being convex. Following~\cite{dubins1962extreme,klee_theorem_1963}, if $p\in \cvx$, the smallest face of $\cvx$ which contains $p$ is the union of $\{p\}$ and all the open segments in $\cvx$ which have $p$ as an inner point. We denote it by $\face{p}{\cvx}$. The (co)dimension of $\face{p}{\cvx}$ is defined as the (co)dimension of its affine hull. The collection of all elementary faces, $\{\face{p}{\cvx}\}_{p\in\cvx}$, is a partition of $\cvx$. Extreme points correspond to the zero-dimensional faces of $\cvx$, while extreme rays are (generally a strict subcollection of the) one-dimensional faces. \paragraph{Quotient by lines} As noted above, if $\cvx$ is linearly closed, it contains a line if and only if the vector space $\mathrm{lin}(\cvx)$ defined in \eqref{eq::lineality space} is nontrivial. In that case, letting $W$ denote some linear supplement to $\mathrm{lin}(\cvx)$ , we may write \begin{equation} \label{eq::quotientbyline} \cvx= \qcvx+\mathrm{lin}(\cvx), \textrm{ with } \qcvx\eqdef \cvx\cap W \end{equation} and the corresponding decomposition is unique (\ie{} any element of $\cvx$ can be decomposed in a unique way as the sum of an element of $\qcvx$ and $\mathrm{lin}(\cvx)$). The convex set $\qcvx$ (isomorphic to the projection of $\cvx$ onto the quotient space $\vecgal/\mathrm{lin}(\cvx)$) is then linearly closed, and the decomposition of $\cvx$ in elementary faces is exactly given by the partition $\{\face{p}{\qcvx}+\mathrm{lin}(\cvx)\}_{p\in\qcvx}$, where $\face{p}{\qcvx}$ is the smallest face of $p$ in $\qcvx$. One may check that $\qcvx$ contains no line, as its recession cone $\rec{\qcvx}$, the projection of $\rec{\cvx}$ onto $W$ parallel to $\mathrm{lin}(\cvx)$, is a \emph{salient} convex cone. \section{Proofs of Section~\ref{sec:abstract}} \label{sec:proof} \subsection{Proof of Theorem~\ref{thm:first}} \label{sec:prooffirst} The set of solutions $\sol$ is precisely $\minset\cap \Phi^{-1}(\{y\})$, and the statement of the theorem amounts to describing its elementary faces. Since $\Phi^{-1}(\{y\})$ is an affine space with codimension at most $m$, the main theorem of~\cite{klee_theorem_1963} almost provides the desired conclusion, but, for our particular case, it yields one extreme point/ray too many. Here is how to obtain the correct number. Let $p$ be a point of $\sol$ such that $\face{p}{\sol}$ has dimension $j$. Up to a translation, it is not restrictive to assume that $p=0$, so that $\sol = \minset \cap \mathrm{ker} \Phi$. Let $\vecrest$ be the union of $\{0\}$ and all the lines $\ell$ such that $\minset\cap \ell$ contains an open interval which contains $0$. Note that $\vecrest$ is a linear space, the linear hull of $\face{0}{\minset}$. We claim that $\codim_{\vecrest}\left(\vecrest\cap\mathrm{ker} \Phi\right)\leq m-1$. By contradiction assume that there is a complement $Z$ to $\vecrest\cap\mathrm{ker}\Phi$ in $\vecrest$ with dimension $m$. Then $\restr{\Phi}{Z}$ has rank $m$ and is a bijection, hence we may define \begin{equation*} z\eqdef -\frac{\theta}{(1-\theta)}(\restr{\Phi}{Z})^{-1}(\Phi u_0) \in Z\subset \vecrest, \end{equation*} where $\theta\in ]0,1[$ and $u_0\in \minset$ is such that $\inf \reg \leq \reg(u_0)<\minval$. For $\theta$ small enough, $z\in \minset$, hence $\reg(z)\leq \minval$. Moreover, \begin{equation} \Phi\left((1-\theta)z+\theta u_0 \right)= (1-\theta)\Phi z + \theta \Phi u_0 = 0. \end{equation} so that $(1-\theta)z+\theta u_0$ lies in {$\minset\cap \mathrm{ker} \Phi$}. Since $\reg(u_0) < \reg(z)$, and $\reg$ is convex \begin{equation} \label{eq:for_cvx_like} \reg\left((1-\theta)z+\theta u_0 \right)< \reg\left(z\right)\leq\minval, \end{equation} we obtain a contradiction with the fact that $\minval$ is the minimal value of~\eqref{eq:thminreg}. As a result, $\codim_{\vecrest}\left(\vecrest\cap\mathrm{ker} \Phi\right)\leq m-1$. Observing that $\vecrest\cap\mathrm{ker} \Phi$ is the linear hull of $\face{0}{\sol}$, hence $j= \mathrm{dim}\left(\vecrest\cap\mathrm{ker} \Phi\right)$, we deduce that \begin{equation} \mathrm{dim} \face{0}{\minset}\eqdef\mathrm{dim} \vecrest = \codim_{\vecrest}\left(\vecrest\cap\mathrm{ker} \Phi\right) + \mathrm{dim}\left(\vecrest\cap\mathrm{ker} \Phi\right) \leq m-1+j, \end{equation} and the first claim of the theorem is proved. Now, applying the Carath\'eodory-Klee theorem (3) in~\cite{klee_theorem_1963}, $p$ is convex combination of at most $m+j$ (resp. $m+j-1$) extreme points (resp. extreme points or in an extreme ray) of $\face{0}{\minset}$. The conclusion stems from the fact that the extreme points (resp. rays) of $\face{0}{\minset}$ are extreme points (resp. rays) of $\minset$, see the proof of the main theorem in~\cite{klee_theorem_1963}. \qed \subsection{Proof of Corollary~\ref{coro:lines}} \label{sec:proofconvex} \begin{sloppypar} Now, assume that the vector space ${\vecrec{}\eqdef\rec{\minset}\cap (-\rec{\minset})}$ is non trivial (otherwise the conclusion follows from Theorem~\ref{thm:first}). We note that for any $u\in \minset$, the convex function $v\mapsto \reg(u+v)$ is upper bounded by $\minval$ on $\vecrec{}$, hence is constant. As a result, possibly replacing $\reg$ with $\reg+\chi_{\minset}$, it is not restrictive to assume that $\reg$ is invariant by translation along $\vecrec{}$. \end{sloppypar} Now, let $\projv$, $\projp$ be the canonical quotient maps and define $\tilde{\reg}$ and $\tilde{\Phi}$ by the commutative diagrams \[ \begin{tikzcd} \vecgal \arrow{r}{\reg} \arrow[swap]{d}{\projv} & \left[-\infty,\infty\right] \\% \vecgal/\mvecrec \arrow[swap]{ru}{\tilde{\reg}}& \end{tikzcd}\qquad \begin{tikzcd} \vecgal \arrow{r}{\Phi} \arrow[swap]{d}{\projv} & \RR^m \arrow{d}{\projp} \\% \vecgal/\mvecrec \arrow{r}{\tilde{\Phi}}& \RR^m/\Phi(\mvecrec) \end{tikzcd}\] Note that $\tilde{\reg}$ is a convex function and that $\tilde{\Phi}$ is a linear map with rank $m-d$, where $d\eqdef \mathrm{dim}\left(\Phi(\mvecrec)\right)$. It is then natural to consider the problem \begin{equation}\label{eq:qthminreg} \min_{\tilde{u}\in \vecgal/\mvecrec} \tilde{\reg}(\tilde{u}) \quad \mbox{s.t.}\quad \tilde{\Phi} \tilde{u}=\tilde{y},\tag{$\tilde{\Pp}$} \end{equation} where $\tilde{y}\eqdef\projp(y)$. In other words, one still wishes to minimize $\reg(u)$, but one is satisfied if the constraint $\Phi u=y$ merely holds up to an additional term $\Phi v$, where $v\in\mvecrec$. We observe that \eqref{eq:thminreg} and \eqref{eq:qthminreg} have the same value $\minval$, and the level set \begin{equation} \qminset\eqdef\enscond{\tilde{u}\in\vecgal/\mvecrec}{\tilde{\reg}(\tilde{u})\leq t^\star}=\projv(\minset) \end{equation} is convex linearly closed \emph{and contains no line}. Let $\qsol$ be the solution set to~\eqref{eq:qthminreg}. Theorem~\ref{thm:first} now describes the elements of the $j$-dimensional faces of $\qsol$ as convex combinations of $m-d+j$ (resp.\ $m-d+j-1$) extreme points (resp. extreme points or points in an extreme ray) of $\qminset$ that we denote by $\qe_1$,$\qe_2$,\ldots, $\qe_r$. To conclude, we have obtained $\projv(p)=\sum_i \theta_i \qe_i$, for some $\theta\in \RR_+^r$ with $\sum_i\theta_i=1$. Equivalently, since $\qiso^{-1}(\cdot,0)$ provides one element in the corresponding class, this means that $p\in \qiso^{-1}(\sum_i \theta_i \qe_i,0)+\vecrec{}$. We get the claimed result by linearity of $\qiso^{-1}$.\qed \begin{remark} Incidentally, we note that for $\cible=\{y\}$, the face $\face{p}{\sol_{\{y\}}}$ is isomorphic (through $\qiso$) to $\face{\projv(p)}{\projv(\sol_{\{y\}})}\times (\vecrec{}\cap \mathrm{ker}\Phi)$. \end{remark}
{ "timestamp": "2018-11-27T02:29:11", "yymm": "1806", "arxiv_id": "1806.09810", "language": "en", "url": "https://arxiv.org/abs/1806.09810", "abstract": "We establish a general principle which states that regularizing an inverse problem with a convex function yields solutions which are convex combinations of a small number of atoms. These atoms are identified with the extreme points and elements of the extreme rays of the regularizer level sets. An extension to a broader class of quasi-convex regularizers is also discussed. As a side result, we characterize the minimizers of the total gradient variation, which was still an unresolved problem.", "subjects": "Optimization and Control (math.OC); Information Theory (cs.IT)", "title": "On Representer Theorems and Convex Regularization", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9883127416600689, "lm_q2_score": 0.8104789018037399, "lm_q1q2_score": 0.8010066254992959 }
https://arxiv.org/abs/1909.08687
Unusual elementary axiomatizations for abelian groups
One of the most studied algebraic structures with one operation is the Abelian group, which is defined as a structure whose operation satisfies the associative and commutative properties, has identical element and every element has an inverse element. In this article, we characterize the Abelian groups with other properties and we even reduce it to two the list of properties to be fulfilled by the operation. For this, we make use of properties that, in general, are hardly ever mencioned.
\section*{Introduction} \noindent The axiomatic presentation of mathematical theories allows the selection of different sets of axioms for its development. This choice depends on criteria of economy, elegance, simplicity or pedagogy. The definition of an abelian group was initially formulated for finite groups (with the axioms of closure, associativity, commutativity and existence of inverses) by Kronecker in 1870 and by Weber for infinite groups in 1893 \cite{waerden}. In 1878 Cayley introduces the notion of abstract group and in 1882 Dick Van presents the first explicit definition of this notion \cite{wussing}. In 1938 Tarski \cite{tarski} defines an abelian group $(G, +)$ as an associative and commutative quasigroup and characterizes it in terms of subtraction using only two axioms, one that indicates that the subtraction is an operation in $G$ and the other which is a property that includes three variables. In 1952 Higman and Neumann \cite{higman} give an axiomatization for groups with one axiom in terms of division, using three variables. In 1981 Neumann \cite{neumann81} proposes another single law in terms of multiplication and inversion, in equational form with four variables. In 1993 McCune \cite{mccune} presents cha\-rac\-terizations of abelian groups with one axiom that has three or five variables, using computational tools, but in terms of operations such as \{addition and inverse\}, \{double subtraction\}, \{double subtraction, identity\}, \{subtraction, identity\}, \{subtraction, inverse\}, \{double subtraction, inverse\}. In all cases in which the groups or abelian groups are characterized in equational form with one axiom, it has an extensive expression and the proofs are intricate. How\-ever, in 1996 McCune and Sands \cite{mcsands} proposed a single law but in implicative form, which is simpler than the equational form, not only in appearance but in the proofs too. In the present work we give some characterizations of abelian groups with two elementary axioms whose expressions display an elegant simplicity. The same applies to the proofs which can help to understand this basic algebraic structure. Algebraic structures can be classified, giving them special names, according to the operations they involve. We have limited ourselves to consider algebraic structures with one operation $(G,+)$ with $G$ a nonempty set, called \textit{magma} or \textit{groupoid}. The best known are \textit{Semigroup}, a groupoid with one associative operation (A); \textit{Monoid}, a semigroup that has neutral element (NE); \textit{Group}, a monoid such that all elements have inverse elements (IN); and \textit{Abelian group}, a commutative, (C), group. However, there are other properties (see \cite{ilse}) such as : for all $a$, $b$, $c$, $d \in G$ \begin{itemize} \item CAI. \textit{Cyclic associativity I}: $a + (b + c) = c + (a + b)$. \item CAII. \textit{Cyclic associativity II}: $a + (b + c) = (c + a) + b$. \item AGI. \textit{Abel-Grassmann I}: $a + (b + c) = c + (b + a)$. \item AGII. \textit{Abel-Grassmann II}: $a + (b + c) = (b + a) + c$. \item R. \textit{Reduced product property}: $(a + b) + c = a + (c + b)$. \item H. \textit{Hilbert property}\footnote{This property was presented as part of an axiomatization for real numbers in \cite[p. 51-52]{hilbert}.}: the equations $x + a = b$ and $a + y = b$ have a unique solution. \end{itemize} Algebraic structures whose operations satisfy some of these properties have also received special names such as \textit{Quasigroup}\footnote{This concept was introduced by B. A. Hausmann and O. Ore in 1937 \cite[p. 22]{ilse}. An equivalent definition appears in \cite[p. 50]{warner}.}, a groupoid that satisfy H and \textit{Loop}, a quasigroup having a neutral element. \begin{theorem}\label{teor1} If $(G, +)$ is a commutative semigroup then it satisfies the properties CAI, CAII, AGI, AGII and R. \end{theorem} \begin{theorem}\label{teor2} If $(G, +)$ is an abelian group then it satisfies H. \end{theorem} Although classical structures such as the abelian group and the commutative semigroup satisfy the properties mentioned, this does not mean that these properties are not independent. \section*{Examples} \begin{enumerate} \item The natural numbers with usual addition and multiplication are commutative semigroups, have neutral elements 0 and 1 respectively, but are not quasigroups. \item Integers, rational, real and complex numbers with the usual sum are commutative semigroups and loops. \item A lattice with the meet ($\land$) and join ($\lor$) operations is a commutative semigroup, but not a quasigroup. \item The integers with subtraction $x \circ y = x - y$ is a quasigroup and satisfies AGI but not AGII. It neither satisfies ACI nor ACII, R, A or C and also does not have a neutral element. \item The integers with reciprocal subtraction $x \bullet y = y - x$ is a quasigroup which satisfies AGII, but not AGI. It neither satisfies ACI nor ACII, R, A or C and does not have a neutral element. \item A set A with the second projection operation defined by $x \ \pi_2 \ y = y$, is a non-commutative semigroup, which satisfies AGII but not AGI. It neither satisfies ACI nor ACII or R. It has a neutral element and it is not a quasigroup. \item A set A with the first projection operation defined by $x \ \pi_1 \ y = x$, is a non-commutative semigroup, which satisfies R but not AGI. It neither satisfies AGII nor ACI or ACII. It does not have a neutral element and it is not a quasigroup. \item In the real interval $[0, 1]$ the operation $p * q = 1 - pq$ is commutative but not associative. It does not have a neutral element. It is neither AGI nor AGII, ACI, ACII or R and it is not a quasigroup. This operation is used in probability theory to determine the probability for two independent events to not occur simultaneously when the probability of occurrence of one is $p$ and of the other is $q$. \item In the ordered set $\{0, 1/2, 1\}$ the operation defined by table \ref{tabla1}, is commutative, but not associative. It has a neutral element 1. It is not AGI, neither AGII nor ACI, ACII or R and it is not a quasigroup. This operation is the logical equivalence which is used in a trivalent Heyting algebra and it was used by Reichenbach in a formulation of quantum mechanics \cite[366-367]{jammer}. \begin{table}[h] \begin{center} \begin{tabular}{c|ccc} $\leftrightarrow$&0&1/2&1\\ \hline 0&1&0&0 \\ 1/2&0&1&1/2 \\ 1&0&1/2&1\\ \end{tabular} \end{center} \caption{Trivalent logical equivalence} \label{tabla1} \end{table} \end{enumerate} \section{Substituting associativity and commutativity} A strategy to search for new characterizations of abelian groups is to exchange or replace some of the properties that define them by others such that these new properties, when mixed with the remaining ones, give us a new definition of the abelian group. To this end, we now establish some relations between structures that satisfy some of the properties mentioned above. We can find structures characterized by some of the unusual properties mentioned above which together with the property NE will result into the properties A and C. \begin{theorem}\label{teor3} If a groupoid $(G, +)$ has a neutral element, $e$, and satisfies the property AGII, then it is a commutative semigroup. \end{theorem} \begin{proof} We first show that $+$ is commutative. Applying the properties NE and AGII we obtain \[a + b = a + (b + e) = (b + a) + e = b + a\] From properties AGII and C we deduce that $+$ is associative: \[a + (b + c) = (b + a) + c = (a + b) + c \qedhere\] \end{proof} The proof of theorem \ref{teor4} below is analogous to the one of theorem \ref{teor3}. \begin{theorem}\label{teor4} If a groupoid $(G, +)$ has a neutral element and satisfies one of the pro\-perties CAI, CAII, AGI or R, then it is a commutative semigroup. \end{theorem} Note that from theorems \ref{teor3} and \ref{teor4} we can replace the associative and commutative properties by any of the properties CAI, CAII, AGI, AGII or R, in the de\-fi\-ni\-tion of an abelian group. This way we obtain another characterization of the structure under consideration, only with three axioms. \begin{theorem} The following conditions are equivalent: \begin{enumerate} \item $(G, +)$ is an abelian group. \item $(G, +)$ is a groupoid that satisfies the properties NE, IN and CAI. \item $(G, +)$ is a groupoid that satisfies the properties NE, IN and CAII. \item $(G, +)$ is a groupoid that satisfies the properties NE, IN and AGI. \item $(G, +)$ is a groupoid that satisfies the properties NE, IN and AGII. \item $(G, +)$ is a groupoid that satisfies the properties NE, IN and R. \end{enumerate} \end{theorem} It should be noted that the properties CAI, CAII, AGII and AGI have been used \cite[p. 10]{pad} for axiomatizing the lattice theory. \section{Substituting inverse elements and neutral element} From theorems \ref{teor1} and \ref{teor2} we deduce that an abelian group satisfies the properties CAI, CAII, AGI, AGII, R and H. The next theorem indicates how the property H may be used to characterize the abelian groups, but without forgetting the commutative and associative properties. \begin{theorem}\label{teor6} If $(G, +)$ is an associative and commutative quasigroup then it is an abelian group. \end{theorem} \begin{proof} Since $(G, +)$ is a quasigroup, for all $a \in G$, the equation $a + x = a$ has a unique solution, say $e_a$, i.e. $a + e_a = a$. By C, $a + e_a = e_a + a = a$. Now, let $b \in G$ then by A, $a + b = (a + e_a) + b = a + (e_a + b)$ and as the equation $a + y = d$ with $d = a + b$, has a unique solution, we conclude that $b = e_a + b$. Therefore, $e_a + b = b = e_b + b$ and again by uniqueness of the solution of the equation $y + b = b$, it follows that $e_a = e_b$. Hence, $e_a$ is the neutral element of $G$ since the above argument is valid for all $b \in G$. The existence of an inverse element for each element $a$ of $G$ is guaranteed by the exis\-tence of the solution of the equation $a + x = e$ with $e$ the neutral element, and pro\-perty C. \end{proof} From the arguments presented in the proof of theorem \ref{teor6} we can conclude: \begin{theorem}\label{CA} If $(G, +)$ is a quasigroup then it satisfies the property of being cancelative (CA). CA is defined as follows: for all $a, b, c \in G$, \center{if \ $a + b = a + c$ \ then \ $a = c$ \ \ and \ \ if \ $b + a = c + a$ \ then \ $b = c$} \end{theorem} Combining the results of theorems \ref{teor2} and \ref{teor6} we obtain other characterizations of abelian groups with three axioms. \begin{theorem} The following conditions are equivalent: \begin{enumerate} \item $(G, +)$ is an abelian group. \item $(G, +)$ is a groupoid that satisfies the properties H, A and C. \end{enumerate} \end{theorem} \section{Substituting all properties} Below we present results in which we cha\-rac\-te\-ri\-ze the structure of an abelian group without using the usual properties. We focus on replacing the property NE, a key pro\-per\-ty that has been used in previous results, without having to resort to the properties A and C. \begin{theorem}\label{CAI} If $(G, +)$ is a quasigroup that satisfies the property CAI, then it is a loop. \end{theorem} \begin{proof} For all $a \in G$, let \begin{equation} a + e_a = a \end{equation} with $e_a$ the unique solution of the equation $a + x = a$. Combining (1) with the pro\-per\-ty CAI we obtain $e_a + a = e_a + (a + e_a) = e_a + (e_a + a)$. From theorem \ref{CA} we get \begin{equation} e_a + a = a. \end{equation} Given $b \in G$, from (2) and CAI we have \[(e_a + b) + a = (e_a + b) + (e_a + a) = e_a + (a + (e_a + b)) = e_a + (b + (a + e_a))\] and by (1) and CAI we obtain \[e_a + (b + (a + e_a)) = e_a + (b + a) = b + (a + e_a) = b +a.\] Then $(e_a + b) + a = b +a$ and by theorem \ref{CA} we conclude that $e_a + b = b$. Hence, $e_a + b = e_b + b$ and again by theorem \ref{CA}, $e_a = e_b$. As this argument is valid for all $b \in G$ we arrive at the conclusion that $e_a$ is the neutral element of $G$ which proves the theorem. \end{proof} \begin{theorem}\label{CAII} If $(G, +)$ is a quasigroup that satisfies the property CAII, then it is a loop. \end{theorem} We shall give two proofs for this theorem: one, similar to the previous proof, sho\-wing directly that there is a neutral element and the other, proving that under these assumptions the property CAI holds and so the assertion follows from theorem \ref{CAI}. \begin{proof}[Proof 1] As for all $a \in G$ the equation $a + x = a$ has a unique solution, say $e_a$, i.e. \begin{equation} a + e_a = a \end{equation} Applying (3) and the property CAII we have \[e_a + a = e_a + (a + e_a) = (e_a + e_a) + a\] From theorem \ref{CA} we get \begin{equation} e_a + e_a = e_a \end{equation} Given $b \in G$, from (4) and CAII it follows that \[a + e_a = a + (e_a + e_a) = (e_a + a) + e_a\] Again by theorem \ref{CA} we have $a = e_a + a$ for all $a \in G$. Now let $b \in G$, by (3) and CAII we obtain $b + a = b + (a + e_a) = (e_a + b) + a$. By theorem \ref{CA}, $b = e_a + b$. Thus, $e_a + b = e_b + b$ and so $e_a = e_b$. As this argument is valid for all $b \in G$ we again conclude that $e_a$ is the neutral element of $G$ and as a consequence $(G, +)$ is a loop. \end{proof} \begin{proof}[Proof 2] We first show that $+$ is associative. Let $a, b, c \in G$, as $G$ is a quasigroup there is $u \in G$ such that $b + u = c$. Hence, \begin{equation} a + (b + c) = a + (b + (b + u)) \end{equation} Applying the property CAII repeatedly we get \[a + (b + (b + u)) = ((b + u) + a) + b = (u + (a + b)) + b = (a + b) + (b + u)\] and replacing $b+u$ by $c$ we conclude that $a + (b + c) = (a + b) + c$. Now let us prove that $+$ satisfies CAI. Again applying the property CAII repeatedly to (5), we obtain \begin{align*} &a + (b + (b + u)) = a + ((u + b) + b) = (b + a) + (u + b) \\ &\hspace*{0.5cm}= (b + (b + a)) + u = ((a + b) + b) + u = b + (u + (a + b)) \end{align*} By the property A and replacing $b+u$ by $c$, we have $a + (b + c) = c + (a + b)$. Then from theorem \ref{CAI} it follows that $(G, +)$ is a loop. \end{proof} \begin{theorem}\label{AGII} If $(G, +)$ is a quasigroup that satisfies the property AGII, then it is a loop. \end{theorem} \begin{proof} For each $a \in G$ let $e_a$ the unique solution of the equation $a + x = a$, i.e. \begin{equation} a + e_a = a \end{equation} By (6) and property AGII we have \[e_a + a = e_a + (a + e_a) = (a + e_a) + e_a = a + e_a = a\] Therefore, for all $a \in G$, it holds that \begin{equation} e_a + a = a = a + e_a \end{equation} Now, let $b \in G$, then by (7) and the property AGII we obtain \[b + a = b + (e_a + a) = (e_a + b) + a\] Furthermore, by theorem \ref{CA} we conclude that $b = e_a + b$. Hence, $e_a + b = e_b + b$ and so $e_a = e_b$. Since this argument is valid for all $b \in G$, $e_a$ is the neutral element of $G$ which proves the theorem. \end{proof} \begin{theorem}\label{R} If $(G, +)$ is a quasigroup which satisfies the property R, then it is a loop. \end{theorem} \begin{proof} Since $G$ is a quasigroup, for each $a \in G$ there is $e_a \in G$ such that \begin{equation} a + e_a = a \end{equation} Therefore, given any $b \in G$, by (8) and the property R we get \[a + b = (a + e_a) + b = a + (b + e_a) \] By the theorem \ref{CA} we obtain $b = b + e_a$. As a consequence $b + e_a = b + e_b$ and so $e_a = e_b$. This argument holds for all $b \in G$. We therefore conclude that there is a unique right neutral element, which we denote by $e$. On the other hand, for all $a \in G$ the equation $y + a = a$ has a unique solution, say $\hat{e}_a$, i.e. \begin{equation} \hat{e}_a + a = a \end{equation} As $e$ is a right neutral element we have $(e + a) + e = e + a$. Applying (9) and the pro\-per\-ty R, we obtain $e + a = e + (\hat{e}_a + a) = (e + a) + \hat{e}_a$. Then $(e + a) + e = (e + a) + \hat{e}_a$ and by theorem \ref{CA}, $e = \hat{e}_a$ for all $a \in G$. Thus, $e$ is also left neutral element of $G$ which completes the proof. \end{proof} Finally, the results of theorems \ref{CAI}-\ref{R} together with the theorems \ref{teor1}-\ref{teor4} and \ref{teor6} can be condensed into the following final result. \begin{theorem} The following conditions are equivalent: \begin{enumerate} \item $(G, +)$ is an abelian group. \item $(G, +)$ is a groupoid that satisfies the properties H and CAI. \item $(G, +)$ is a groupoid that satisfies the properties H and CAII. \item $(G, +)$ is a groupoid that satisfies the properties H and AGII. \item $(G, +)$ is a groupoid that satisfies the properties H and R. \end{enumerate} \end{theorem} Examples 1 and 4 show the independence of each of the properties CAI, CAII, AGII and R with respect to property H, and examples 1 and 5 show the independence of AGI and H. One would think that with AGI we would have an analogous theorem to those presented in this section, however this combination of properties does not even provide us a loop. For example, integers with subtraction satisfy H and AGI but it is not an abelian group as it does not have a neutral element and hence no inverses.
{ "timestamp": "2019-09-20T02:02:16", "yymm": "1909", "arxiv_id": "1909.08687", "language": "en", "url": "https://arxiv.org/abs/1909.08687", "abstract": "One of the most studied algebraic structures with one operation is the Abelian group, which is defined as a structure whose operation satisfies the associative and commutative properties, has identical element and every element has an inverse element. In this article, we characterize the Abelian groups with other properties and we even reduce it to two the list of properties to be fulfilled by the operation. For this, we make use of properties that, in general, are hardly ever mencioned.", "subjects": "Group Theory (math.GR)", "title": "Unusual elementary axiomatizations for abelian groups", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9796676460637103, "lm_q2_score": 0.8175744850834649, "lm_q1q2_score": 0.800951271283468 }
https://arxiv.org/abs/2202.11875
Characterizing Spectral Properties of Bridge
The Bridge graph is a special type of graph which are constructed by connecting identical connected graphs with path graphs. We discuss different types of bridge graphs $B_{n\times l}^{m\times k}$ in this paper. In particular, we discuss the following: complete-type bridge graphs, star-type bridge graphs, and full binary tree bridge graphs. We also bound the second eigenvalues of the graph Laplacian of these graphs using methods from Spectral Graph Theory. In general, we prove that for general bridge graphs, $B_{n\times l}^2$, the second eigenvalue of the graph Laplacian should be between $0$ and $2$, inclusive. In the end, we talk about future work on infinite bridge graphs. We created definitions and found the related theorems to support our future work about infinite bridge graphs.
\section{Introduction} Spectral graph theory is the process of characterizing graphs by means of the eigenvalues and eigenvectors of the graph Laplacian. It connects graphs to matrices, and allows us to understand properties of graph using more analytic means. Recently, spectral graph theory has found application in machine learning and deep learning. In particular, there are many clustering algorithms based on spectral methods, like spectral clustering, that are more effective than tradition clustering methods like K-means. Many theoretical properties of these algorithms rely on bounding eigenvalues of the graph Laplacian. \newline Research has already been done in extracting bounds of eigenvalues of special graphs such as complete graphs, path graphs, the binary tree, and so on. In this paper, we will be focusing on some special types of graphs which are constructed by connecting some identical connected graphs by a path or multiple edges, which we'll call Bridge graphs. Bridge graphs are constructed by using path graphs, $P_m$ with $n\geq 2$, and putting some identical graphs on each end of the path. \newline We will then bound the second eigenvalues of the graph Laplacians of the graphs we discussed above. The second eigenvalues are the most important because the first eigenvalue of the graph Laplacian is always $0$. We'll use test vectors and Loewner partial ordering to approach this. At the end of this paper, we'll discuss constructing infinite bridge graphs, which are constructed by connecting a countably infinite number of identical connected graphs using path graphs. We'll also discuss the general idea for bounding the spectrum of the generalized Laplacian operator. \section{Basic Definitions} The following definitions are from Dan Spielman\cite{b1}.Assume we have a graph $G$ with vertex set $V$ and edge set $E$. Assume that the number of vertices is $|V|=n$. We can label vertices to be $\{1,2,3,\dots,n-1,n\}$. \begin{definition}The adjacency matrix $\bold{M}$ of a weighted graph $G=(V,E, w)$ is defined as the matrix with the following entries $$ \bold{M}(a,b) = \begin{dcases} w_{a,b} \quad (a,b)\in E\\ 0 \quad (a,b)\notin E \end{dcases} $$ When the graph is unweighted, $w(a) = 1$ for all $(a,b) \in E$. \end{definition} \begin{definition} The degree of a vertex $a$ is the number of edges attached to it. For a weighted graph, the degree $d(a)$ of the vertex $a$ is the sum of the weights of the edges attached to it. \end{definition} \begin{definition} The degree matrix $\bold{D}$ of a graph $G=(V,E)$ is a diagonal matrix whose entries are given by $$ \bold{D}(a,b) = \begin{dcases} d(a)\quad &a=b\\ 0 \quad &a\neq b \end{dcases} $$ \end{definition} \begin{definition} The graph laplacian $\bold{L}$ of a graph $G$ is defined to be $$ \bold{L} =\bold{M}- \bold{D}. $$ \end{definition} \begin{definition}(Loewner partial order) Let $G_1$ and $G_2$ be graphs each with $n$ vertices. Then for the graph Laplacians of $G_1$ and $G_2$, $L_{G_1}$ and $L_{G_2}$, we write $L_{G_1} \succcurlyeq L_{G_2}$ if and only if $\bold{v}^TA\bold{v}\geq \bold{v}^TB\bold{v}$ for all vectors $\bold{v} \in \mathbb{R}^n$. The relation $\succcurlyeq$ above is called Loewner partial order. In this case, the graphs $G_1$ and $G_2$ also have relation $G_1 \succcurlyeq G_2$. \end{definition} \section{Basic Theorems} The following theorems are from Dan Spielman\cite{b1} \begin{theorem} Assume we have a weighted graph $G=(V,E)$, for every edge $e = (a,b)$, let the weight be $w_{a,b}$. For a function $\bold{x}: V \to \mathbb{R}^n$, the quadratic form of the graph Laplacian is $$ \bold{x}^T\bold{L}\bold{x}=\sum_{(a,b)\in E} \bold{w}_{a,b}(x(a)-x(b))^2 $$ \end{theorem} \begin{theorem} For a $n\times n$ symmetric matrix $\bold{A}$ with ordered eigenvalues $\lambda_1\le \lambda_2\le \dots\le\lambda_n$ and corresponding eigenvectors $\phi_1,\phi_2,\dots,\phi_n$ we have $$\phi_i=\min\limits_{\substack{(\bold{x},\phi_k)=0,\\1\le k \le i-1} }{\frac{\bold{x}^T \bold{L}\bold{x}}{\bold{x}^T\bold{x}}} $$ with $$ \phi_i=\arg\min\limits_{\substack{(\bold{x},\phi_k)=0, \\1\le k \le i-1}} {\frac{\bold{x}^T \bold{L}\bold{x}}{\bold{x}^T\bold{x}}}. $$ \end{theorem} \begin{theorem} For a graph $G=(V,E)$, with graph Laplacian $L_G$, ordered eigenvalues $\lambda_1\le \lambda_2\le \dots\le\lambda_n$, and corresponding eigenvectors $\phi_1,\phi_2,\dots,\phi_n$, we have $\lambda_1=0$ and $\phi_1=\bold{1}$ where $\bold{1} = (1, \ldots, 1)^{T}$. \end{theorem} \begin{theorem} For a unweigted graph $G=(V,E)$ And $L_G$ is the graph Laplacian with ordered eigenvalues $0=\lambda_1\le \lambda_2\le \dots\le\lambda_n$. Then $G$ is connected if and only if $\lambda_2>0$ \end{theorem} \begin{theorem} Suppose $G_1$ and $G_2$ are graphs with the relation $G_1\succcurlyeq cG_2$. Then $\lambda_k(G_1)\succcurlyeq \lambda_k(G_2)$ \end{theorem} \begin{theorem} If $G_1$ is a subgraph of $G_2$ then $G_1\preccurlyeq cG_2$. \end{theorem} \section{$K_n$ Type Bridge Graphs} Now we will discuss dumbbell-like graphs $D_n^m$, which are formed by joining two complete graphs with $n$ vertices, $K_{n,1}$ and $K_{n,2}$, with a path graph $P_m$. For example, if we connect two $K_8$'s together with $P_3$, we have the following: \begin{center} \includegraphics[scale = 0.35]{K2.png} \end{center} Notice that this is a simple example of a bridge graph. \\ \begin{corollary}(The Path Inequality) A path graph $P_{a,b}$ is a path from $a$ to $b$, and $G_{a,b}$ is a graph with a single edge $(a,b)$, then the following path inequality holds: $$ |P_{a,b}|P_{a,b} \succcurlyeq G_{a,b}. $$ \end{corollary} \begin{theorem} For the dumbbell-like graph we mentioned above, $D_n^m$, we know that $|V_{D_n^m}|=2n+m-2$ , we have the following bound on the eigenvalues: $$ \frac{2}{(2n+m-3)(m+1)}\le\lambda_2(D_n^m)\le\frac{12}{6(m-1)(n-1)+m(m-1)} $$ \begin{proof} Let $K_{n,1}$ be the first complete graph and $K_{n,2}$ be the second complete graph. We label the vertices of $K_{n,1}$ as $\{1,2,3, \dots ,n-1,n\}$ and suppose that the vertex shared by $K_{n,1}$ and $P_m$ is labeled as $n$. Then the next vertex on $P_m$, which is attached to the vertex $n$ is labeled as $n+1$. Repeat the same process until we label the vertex which is on both $P_m$ and $K_{n,2}$ as $n+m-1$. Finally we label the vertices of $K_{n,2}$ to be $\{n+m-1,n+m,\dots,2n+m-2\}$ \newline To get the upper bound we construct test vector $\bold{x}$ to be $$ \bold{x}(i) = \begin{dcases} m-1\quad &1\le i < n\\ 2n+m-1-2i \quad & n\le i <n+m-1\\ 1-m \quad &n+m-1\le i \le 2n+m-2 \end{dcases}. $$ The vertices $n$ and $n+m-1$ are both on one of the complete graphs and the path graphs, so we need to check their value on both graphs to make sure our construction of test vector $\bold{x}$ is consistent. \newline When $i=n$ we plug into $x(i)=m-1$ gets ${x(i)=m-1}$. Also, if substitute $i = n$ into $x(i)=2n+n-2i$, we get $x(i)=2n+m-1-2n=m-1$. This matches with the other graph. When $i=n+m-1$ we substitute into $x(i)=2n+m-1-2i$, which yields $x(i)=2n+m-1-2(n+m-1)=-m+1$. Also if we substitute $i = n$ into $x(i)=1-m$, we get $x(i)=1-m$, This also matches with the other graph. Hence we have verified the consistency of our test vector.\\ Now we need to calculate the inner product of test vector $\bold{x}$ and the vector $\bold{1}=(1,\dots,1)^T$. We have \begin{align*} (\bold{x},\bold{1}) &= \sum_{i\in V}x(i)\\ &=\sum_{i=1}^{2n+m-2} x(i)\\ &=\sum_{i=1}^{n-1} x(i) + \sum_{i=n}^{n+m-1}x(i)+\sum_{i=n+m}^{2n+m-2} x(i)\\ &=\sum_{i=1}^{n-1} (m-1)+\sum_{i=n}^{n+m-1} (2n+m-1-2i) + \sum_{i=n+m}^{2n+m-2} (1-m). \end{align*} We now deal with the three terms separately. For the first term, $$\sum_{i=1}^{n-1} (m-1) = (m-1)(n-1).$$ For the second second term, we get $$\sum_{i=n}^{n+m-1} (2n+m-1-2i) = (2n+m-1)(n+m+1-n-1)-2\cdot\frac{n+m-1+n}{2}.$$ Lastly, $$\sum_{i=n+m}^{2n+m-2} (1-m) = (1-m)(n-1).$$ Adding the terms up, we get $$(\bold{x},\bold{1}) = 0.$$ \newline From the calculation above, we know that we can use $\bold{x}$ to get the upper bound of $\lambda_2(D_n^m)$. From Theorem 4 we have \begin{align*} \lambda_2(D_n^m) &\le \frac{\bold{x}^T\bold{L}\bold{x}}{\bold{x}^T\bold{x}}\\ &\le \frac{\sum_{(a,b)\in E} (x(a)-x(b))^2}{\sum_{i=1}^{2n+m-2}x(i)^2}\\ & = \frac{\sum_{1\le i,j<n} (x(i)-x(j))^2}{\sum_{1\le i<n} x(i)^2+\sum_{n\le i<n+m-1} x(i)^2 +\sum_{n+m-1\le i\le 2n+m-1} x(i)^2}\\ &+ \frac{\sum_{n\le i,j<n+m-1} (x(i)-x(j))^2}{\sum_{1\le i<n} x(i)^2+\sum_{n\le i<n+m-1} x(i)^2 +\sum_{n+m-1\le i\le 2n+m-1} x(i)^2}\\ &+ \frac{\sum_{n+m-1\le i,j\le 2n+m-1} (x(i)-x(j))^2}{\sum_{1\le i<n} x(i)^2+\sum_{n\le i<n+m-1} x(i)^2 +\sum_{n+m-1\le i\le 2n+m-1} x(i)^2}. \end{align*} The first and last term of the sum are zero, so this means that \begin{align*} \lambda_2(D_n^m)&=\frac{0+\sum_{i=n}^{n+m-2} (x(i)-x(i+1))^2+0}{(m-1)(n-1)+\frac{m(m-1)(m+1)}{3}+(m-1)(n-1)}\\ &=\frac{12}{6(m-1)(n-1)+m(m-1)}. \end{align*} To get the lower bound, we use the Loewner partial order. \newline For every pair of edge $(a,b)\in E_{D_n^m}$,let the path graph $P_{a,b}$ be a path from $a$ to $b$, and $G_{a,b}$ be a graph with a single edge $(a,b)$, then from the corollary we have $|P_{a,b}|P_{a,b} \succcurlyeq G_{a,b}$. We know that if $1\le a \le n$, then this means $a$ is a vertex of the $K_{n,1}$. Thus, $a$ is connected to vertex $n$. If $n+m-1\le a \le 2n+m-2$ then this means $a$ is a vertex of the $K_{n,2}$ and $a$ is connected to vertex $n+m-1$. If $a$ and $b$ are in the same complete graph then the length of $P_{a,b}$ is $1$; if $a$ and $b$ are in the different complete graphs, then the length of $P_{a,b}$ will be $1+m-1+1=m+1$. If either of $a$ or $b$ are in the path graph, then the length of $P_{a,b}$ shorter than the case when $a$ and $b$ are in different complete graphs. Hence, we conclude that $|P_{a,b}|\le m+1$. \newline It follows that $$G_{a,b}\preccurlyeq |P_{a,b}|P_{a,b} \preccurlyeq(m+1)P_{a,b}\preccurlyeq(m+1)D_n^m. $$ Also, we notice that complete graph $K_{2n+m-2}$ is constructed by connecting all edges together. It has $\binom{2n+m-2}{2}$ single edges.Thus $$K_{2n+m-2}\preccurlyeq \sum_{(a,b)\in E_{K_{2n+m-2}}}G_{a,b}\preccurlyeq \binom{2n+m-2}{2} G_{a,b} \preccurlyeq \binom{2n+m-2}{2}(m+1) D_m^n. $$ Thus, $$ 2n+m-2=\lambda(K_{2n+m-2}) \le \binom{2n+m-2}{2}(m+1) \lambda(D_m^n). $$ From the above, we get that $\lambda(D_m^n)\geq \frac{2}{(2n+m-1)(m+1)}$. \end{proof} \end{theorem} We notice that when $m=1$ then $D_n^1$ is a graph constructed by connecting two complete graphs with a single edges. For a bridge graph $D_n^{2\times k}$ with $k\le n$ which is constructed by two identical complete graphs $K_{n,1}$ and $K_{n,2}$ k different edges $e_1,\dots,e_k$ with the edge length of them are all 2, and for every edge $e_i=(v_{i,1},v_{i,2})$ where $v_{i,1}$ is in $K_{n,1}$ and $v_{i,2}$ is in $K_{n,2}$. A picture is given below in the case where $n = 8$ with $k=2$ and $e_1=(8,9),e_2=(7,16)$: \begin{center} \includegraphics[scale = 0.35]{K.png} \end{center} We can generalize our results from before to the following theorem. \begin{theorem} For the graph we mentioned above, $D_n^{2\times K}$, we know that $|V_{D_n^{2\times 2}}|=2n+1$. We also have the following bound on the second eigenvalue of the graph laplacian: $$ \frac{2}{3(2n-1)}\le\lambda_2(D_n^{2\times k})\le\frac{4}{n}. $$ \begin{proof} From $D_n^2$ is a subgraph of $D_n^{2\times k}$ we can get $D_n^2 \preccurlyeq D_n^{2\times k}$. Thus $\frac{2}{3(2n-1)}=\lambda_2(D_n^2) \le\lambda_2(D_n^{2\times k})$. For the other half of the inequality we can label vertices the same way as the graph $D_n^2$. We know vertices $v_{i,1}$ and $v_{i,2}$ are adjacent to each other for $1\le i \le k$ such that $v_{i,i} \in K_{n,1}$ and $v_{i,2} \in K_{n,2}$. Then we have $1\le v_{i,1}\le n$ and $n+1\le v_{i,2}\le 2n$. Also, we can use the same test vector $\bold{x}$ as the graph $D_n^m$ too. Let $$ \bold{x}(i) = \begin{dcases} 1\quad &1\le i \le n\\ -1 \quad &n+1\le i \le 2n \end{dcases}. $$ We can easily verify that $(\bold{x},\bold{1})=0$. \newline Now we can estimate the upper bound of $\lambda_2({D_n^{2\times2}})$: \begin{align*} \lambda_2(D_n^{2\times 2}) &\le \frac{\bold{x}^T\bold{L}\bold{x}}{\bold{x}^T\bold{x}}\\ &\le \frac{\sum_{(a,b)\in E_{D_n^{2\times 2}}} (x(a)-x(b))^2}{\sum_{i=1}^{2n}x(i)^2}\\ & = \frac{\sum_{1\le i,j<n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le 2n} x(i)^2}\\ &+ \frac{\sum_{i=1}^{k}(x(v_{i,1})-x(v_{i,2}))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le 2n} x(i)^2}\\ &+ \frac{\sum_{n+2\le i,j\le 2n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le 2n} x(i)^2}. \end{align*} The first and last term of the sum are zero, so this means that \begin{align*} \lambda_2(D_n^m)&\le \frac{0+4k+0}{n+n}\\ &=\frac{4k}{2n}\\ &=\frac{2k}{n}. \end{align*} So we have finished bounding $D_n^{2\times k}$. \end{proof} \end{theorem} Consider a general bridge graph $B_n^{2\times k}$ which is constructed by two arbitrary identical graphs $G_{n,1}$ and $G_{n,2}$ with $k$ different edges $e_1,\dots,e_k$ and $k\le n$. We also assume that there are $k$ distinct edges connecting the two graphs. An example is given in the figure: \begin{center} \includegraphics[scale = 0.35]{K3.png} \end{center} For every edge $e_i=(v_{i,1},v_{i,2})$, where $v_{i,1}$ is in $G_{n,1}$ and $v_{i,2}$ is in $G_{n,2}$, we have the following theorem. \begin{theorem} For a bridge graph $B_n^{2 \times k}$ we have $$ 0< \lambda_2(B_n^{2\times k}) \le \frac{2k}{n}. $$ \begin{proof} The proof of the lower bound follows from the fact that $B_m^{2\times k}$ is a connected graph. Thus the second eigenvalue should be positive . The upper bound is a straight forward consequence of Theorem 4.2. We see that $B_n^{2\times k}$ is a subgraph of $B_n^{2\times k}$; hence $B_n^{2\times k}\preccurlyeq D_n^{2\times k}$. It follows that $\lambda_2(B_n^{2\times k}) \le \lambda_2(D_n^{2\times k})\le \frac{2k}{n}$. \end{proof} \end{theorem} \section{$S_n$ Type Bridge Graphs} The star graph, $S_n$, is another graph we will consider. The star graph is special because it is a complete bipartite graph, $K_{1,n-1}$. Now we can construct star-type bridge graphs, $S_n^m$, by connecting two identical star graph $S_{n,1}$ and $S_{n,2}$ with a path graph $P_m$. \begin{theorem} For the star-like graphs $S_n^m$ we mentioned above, we have the following bound on the eigenvalues: $$ \frac{2}{(2n+m-3)(m+3)} \le \lambda_2(S_n^m) \le \frac{4n+2}{2n+m-4}. $$ \begin{proof} Since $S_{n,1}$ is also a bipartite graph ,we can separate it to different set $V_{(1,1)}$ and $V_{(1,2)}$ where $V_{(1,1)}$ only has one vertex which is internal vertex for the tree $S_{n,1}$. And the remaining vertices are all leaves of tree and they are in $V_{(1,2)}$. Notice that no edges has both vertices in the same sets, and every edges that connect vertices in different set is part of the graph. We label the only vertex in $V_{(1,1)}$ as $1$ and remaining as $2,\dots,n$. \newline We know that there is a vertex $v^1$ in $S_{n,1}$ is also on the graph $P_m$. Then we label the vertex which is attached to $v^1$ but not in graph $S_{n,1}$ as $n+1$, repeat the same process until we label the vertex $n+m-2$.We can also separate it to different set $V_{(2,1)}$ and $V_{(2,2)}$ where $V_{(2,2)}$ only has one vertex which is internal vertex for the tree $S_{n,2}$. And the remaining vertices are all leaves of tree and they are in $V_{(2,2)}$. Also notice that no edges has both vertices in the same sets, and every edges that connect vertices in different set is part of the graph. We label the only vertex in $V_{(1,1)}$ as $n+m-1$ and remaining as $n+m,\dots,2n$. We notice that there is a vertex $v^2$ in $S_{n,2}$ is also on the graph $P_m$. Thus $n+m-1\le v^2 \le 2n$. \newline Now we can set the test vector. Now we need to discuss different cases based on whether $n$ is odd or even and based on the value of $v^1$ and $v^2$. \newline \textbf{Case 1:} $n$ is odd, $v^1=1$ and $v^2=n+m-1$, the figure below is the case when $n=9,m=3$ \newline \begin{center} \includegraphics[scale = 0.20]{S1.jpg} \end{center} We choose the test vector $$ \bold{x}(i) = \begin{dcases} 1\quad &i=1,n+m-1\\ 0\quad &i=n \text{ or } i=2n \text{ or } n+1\le i \le n+m-2\\ 1 \quad &2\le i \le \frac{n-1}{2}\\ -1 \quad &\frac{n+1}{2}\le i \le n-1\\ 1 \quad &n+m\le i \le \frac{3n+2m-3}{2}\\ -1\quad &\frac{3n+2m-1}{2}\le i \le 2n-1 \\ \end{dcases}. $$ Notice that \begin{align*} (\bold{x},\bold{1})&=\sum_{i=1}^{2n+m-2} x(i)\\ &=x(1)+x(n+m-1)+\sum_{i=n+1}^{n+m-2}x(i)+\sum_{i=2}^{\frac{n-1}{2}}x(i)+\sum_{\frac{n+1}{2}}^{n-1} x(i)\\ &+\sum_{i=n+m}^{\frac{3n+m-3}{2}}x(i)+\sum_{\frac{3n+m-1}{2}}^{2n} x(i)\\ &=1+1+0+\left(\frac{n-1}{2}-2+1\right) -\left(n-1-\frac{n+1}{2}+1\right)\\ &+\left(\frac{3n+2m-3}{2}-(n+m)+1\right)-\left(2n-1-\frac{3n+2m-1}{2}+1\right)\\ &=0. \end{align*} Hence it's possible to use our test vector to get the upper bound. \begin{align*} &\lambda_2(T_n^m)\\ &\le \frac{\bold{x}^T\bold{L}\bold{x}}{\bold{x}^T\bold{x}}\\ &=\frac{\sum_{(a,b)\in E_{T_n^{m}}} (x(a)-x(b))^2}{\sum_{i=1}^{2n+m-2}x(i)^2}\\ & = \frac{\sum_{1\le i,j\le n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2} x(i)^2}\\ &+ \frac{\sum_{n+1\le i,j\le n+m-2} (x(i)-x(j))^2+(x(1)-x(n+1))^2+(x(n+m-2)-x(n+m-1))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2}x(i)^2}\\ &+ \frac{\sum_{n+m-1\le i,j\le 2n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}. \end{align*} Notice that the first term is equal to \begin{align*} &\frac{\sum_{2\le j\le n} (x(1)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2} x(i)^2}\\ =& \frac{\sum_{2\le j\le n} (x(1)-x(j))^2}{n-1+m-2+n-1}\\ =& \frac{\sum_{j=2}^{\frac{n-1}{2}}(x(1)-x(j))^2+\sum_{j= \frac{n+1}{2}}^{j=n-1}(x(1)-x(j))^2+(x(1)-x(n))^2}{2n+m-4}\\ =&\frac{4\frac{n-1}{2}+1}{2n+m-4}\\ =&\frac{2n-3}{2n+m-4}. \end{align*} The second term is equal to \begin{align*} \frac{0+1+1}{2n+m-4}=\frac{2}{2n+m-4}. \end{align*} From symmetry, the third term and the second term are the same, so the third term is $\frac{2n-3}{2n+m-4}$. Add all terms together, and we get $$ \lambda_2(T_n^m)\le \frac{4n-6}{2n+m-4}. $$\newline \textbf{Case 2:} $n$ is odd, $v^1=1$ and $v^2\neq n+m-1$ or $n$ is odd, $v^1\neq 1$ and $v^2= n+m-1$ \newline \begin{center} \includegraphics[scale = 0.4]{S2.jpg} \end{center} We will only discuss when $n$ is odd, $v^1=1$ and $v^2\neq n+m-1$, or $n$ is odd because the other case will get us the same result from symmetry. When we label our vertices, we can make $v^2=n+m$ now. Then we can still use the same test vector as case 1. So the text vector is well defined. The upper bound process will be the same as case 1. Thus we will get the same bound as case 1.\newline \textbf{Case 3:} $n$ is odd, $v^1\neq1$ and $v^2\neq n+m-1$ \newline \begin{center} \includegraphics[scale = 0.4]{S3.png} \end{center} When we label the vertices, we can make $v^1=2$ and $v^2=n+m$ now. Then we can still use the same test vector as case 1. So the text vector is well defined. The upper bound process will be the same as case 1 thus we will get the same bound as case 1.\newline \textbf{Case 4:} $n$ is even, $v^1=1$ and $v^2=n+m-1$ \newline \begin{center} \includegraphics[scale = 0.4]{S4.jpg} \end{center} We define the test vector as $$ \bold{x}(i) = \begin{dcases} 1\quad &i=1,n+m-1\\ 0\quad & n+1\le i \le n+m-2\\ 1 \quad &2\le i \le \frac{n-1}{2}\\ -1 \quad &\frac{n+1}{2}\le i \le n\\ 1 \quad &n+m-1\le i \le \frac{3n+2m-3}{2}\\ -1\quad &\frac{3n+2m-1}{2}\le i \le 2n \\ \end{dcases}. $$ Notice that \begin{align*} (\bold{x},\bold{1})=&=\sum_{i=1}^{2n+m-2} x(i)\\ &=x(1)+x(n+m-1)+\sum_{i=n+1}^{n+m-2}x(i)+\sum_{i=2}^{\frac{n-1}{2}}x(i)\\ &+\sum_{\frac{n+1}{2}}^{n-1} x(i)+\sum_{i=n+m}^{\frac{3n+m-3}{2}}x(i)+\sum_{\frac{3n+m-1}{2}}^{2n} x(i)\\ &=1+1+\left(\frac{n-1}{2}-2+1\right) \\ &-\left(n-1-\frac{n+1}{2}+1\right)+\left(\frac{3n+2m-3}{2}-(n+m)+1\right)\\ &-\left(2n-\frac{3n+2m-1}{2}+1\right)\\ &=0. \end{align*} Now we get \begin{align*} &\lambda_2(T_n^m)\\ &\le \frac{\bold{x}^T\bold{L}\bold{x}}{\bold{x}^T\bold{x}}\\ &=\frac{\sum_{(a,b)\in E_{T_n^{m}}} (x(a)-x(b))^2}{\sum_{i=1}^{2n+m-2}x(i)^2}\\ & = \frac{\sum_{1\le i,j\le n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2} x(i)^2}\\ &+ \frac{\sum_{n+1\le i,j\le n+m-2} (x(i)-x(j))^2+(x(1)-x(n+1))^2+(x(n+m-2)-x(n+m-1))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2}x(i)^2}.\\ &+ \frac{\sum_{n+m-1\le i,j\le 2n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}. \end{align*} Notice that the first term is equal to \begin{align*} &\frac{\sum_{2\le j\le n} (x(1)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2} x(i)^2}\\ =& \frac{\sum_{2\le j\le n} (x(1)-x(j))^2}{n-1+m-2+n-1}\\ =& \frac{\sum_{j=2}^{\frac{n-1}{2}}(x(1)-x(j))^2+\sum_{j= \frac{n+1}{2}}^{j=n-1}(x(1)-x(j))^2}{2n+m-4}\\ =&\frac{4\frac{n}{2}}{2n+m-4}\\ =&\frac{2n}{2n+m-4}. \end{align*} From case 1 we know that the second term is $\frac{2}{2n+m-4}$, and from the symmetry, the third term and the second term are the same, so the third term is $\frac{2n}{2n+m-4}$. Adding three terms together gets us $$ \lambda_2(T_n^m)\le \frac{4n+2}{2n+m-4}. $$ \textbf{Case 5:} $n$ is even, $v^1=1$ and $v^2\neq n+m-1$ or $n$ is even, $v^1\neq 1$ and $v^2= n+m-1$ \newline \begin{center} \includegraphics[scale = 0.4]{S5.png} \end{center} We will only discuss when $n$ is odd,$v^1=1$ and $v^2\neq n+m-1$ or $n$ is odd because the other case will get us the same result from symmetry. When we label it we can make $v^2=n+m$ now. Then we can still use the same test vector as case 4. So the text vector is well defined.And the upper bound process will be the same as case 1 thus we will get the same bound as case 1.\newline \textbf{Case 6:} $n$ is even, $v^1\neq1$ and $v^2\neq n+m-1$\newline \begin{center} \includegraphics[scale = 0.4]{S6.png} \end{center} When we label it we can make $v^1=2$ and $v^2=n+m$ now. Then we can still use the same test vector as case 4. So the text vector is well defined.And the upper bound process will be the same as case 1 thus we will get the same bound as case 1. \newline Hence we have finished upper bound since we have exhausted all possible cases. \newline For the lower bound, we can compare our graphs to complete graphs. For every pair of edge $(a,b)\in E_{S_n^m}$,let the path graph $P_{a,b}$ be a path from $a$ to $b$, and $G_{a,b}$ be a graph with a single edge $(a,b)$. From the lemma we have $|P_{a,b}|P_{a,b} \succcurlyeq G_{a,b}$. We know that if $a$ and $b$ are both in the same star graph, without loss of generality, we can assume they are both in $S_{n,1}$, Thus, the length from vertex $a$ to $b$ is at most $2$. If $a$ and $b$ are in different star graphs, without loss of generality, we suppose $a$ is in $S_{n,1}$ and $b$ is in $S_{n,2}$. Then the length of the path $P_{a,b}$ is at most $2+m-1+2=m+3$.Hence, the length of the path $P_{a,b}$ is at most $2+m-1+2=m+3$. It follows that \begin{align*} G_{a,b}\preccurlyeq |P_{a,b}|P_{a,b} & \preccurlyeq(m+3) \\ &\preccurlyeq(m+3)S_n^m. \end{align*} Also we know that complete graph $K_{2n+m-2}$ has $\binom{2n+m-2}{2}$ single edges. Thus \begin{align*} K_{2n+m-2}\preccurlyeq \sum_{(a,b)\in E_{K_{2n+m-2}}}G_{a,b}&\preccurlyeq \binom{2n+m-2}{2} G_{a,b} \\ &\preccurlyeq \binom{2n+m-2}{2}(m+3) T_m^n. \end{align*} Hence, $$ 2n+m-2=\lambda_2(K_{2n+m-2}) \le \binom{2n+m-2}{2}(m+3) \lambda_2(S_m^n). $$ Finally, we arrive at $$ \lambda_2(S_m^n)\geq \frac{2}{(2n+m-3)(m+3)}. $$ \end{proof} \end{theorem} \begin{section}{$T_n$ Type Bridge Graphs} Now we will discuss binary tree-like graphs $T_n^m$, which are formed by by joining two full binary trees with $n$ vertices, $T_{n,1}$ and $T_{n,2}$, with a path graph $P_m$. Notice that this is also a simple example of a bridge graph. \\ \begin{theorem} For the binary tree-like graphs $T_n^m$ we mentioned above we have the following bound on the eigenvalues: $$ \frac{2}{(2n+m-1)(2 \log_2(n+1)+m-3)}\le \lambda_2{(T_n^m)}\le \frac{5}{2(n-1)}. $$ \begin{proof} Now we need to label $T_n^m$. We label $T_{n,1}$ the following way. We label the vertex which is ancestor of $T_{n,1}$ all other vertices as $1$. Then $1$ has two children. We label them as $2$ and $3$,then we label children of $2$ as $4$ and $5$, the children of $3$ as $6$ and $7$ and so on until $n$. \newline Then we label the path $P_m$. We know one end of $P_m$ is $i$ where $1\le i\le n$, then we label the vertex which is on the path $P_m$ and attached to $0$ as $n+1$ repeat the process until the vertex $n+m-2$, then the next vertex which in $P_m$ and attach to $n+m-2$ is on the graph $T_{n,2}$. \newline Then we label the graph $T_{n,2}$. We label the vertex which is ancestor of all other vertices of $T_{n,2}$ as $n+m-1$. Then $n+m-1$ has two children. We label them as $n+m$ and $n+m+1$,then we label children of $n+m$ as$n+m+1$ and $n+m+2$, the children of $n+m+1$ as $n+m+3$ and $n+m+4$ and so on until $2n+m-2$. Now we need to break into 3 cases depends on where the ends of $P_m$ locate at. We know that one end is between $1$ and $n$ and the other is between $n+m-1$ and $2n+m-2$. \\ \textbf{Case 1:} One end of $P_m$ is $1$ and the other end of $P_m$ is $n+m-2$. Figure $2$ demonstrates the case of $T_{7}^3$: \begin{center} \includegraphics[scale = 0.5]{T.png} \end{center} We can set the test vector to be: $$ \bold{x}(i) = \begin{dcases} 0\quad &i=1,n+m-1\\ 0\quad &n+1\le i \le n+m-2\\ 1 \quad &i=2,n+m\\ 1 \quad &2<i\le n \text{ and } i \text{ is a descendant of } 2\\ 1 \quad &n+m<i\le 2n+m-2 \text{ and } i \text{ is a descendant of } n+m\\ -1\quad & \text{otherwise} \end{dcases}. $$ We notice that for elements of $T_{n,1}$ the number of $1$ is $\frac{n-1}{2}$ and the number of $-1$ is $\frac{n-1}{2}$. For elements of $T_{n,1}$ the number of $1$ is $\frac{n-1}{2}$ and the number of $-1$ is $\frac{n-1}{2}$. Hence we have \begin{align*} (\bold{x},\bold{1})&=\sum_{i=1}^{2n+m-2} x(i)\\ &=x(1)+x(n+m-1) + \sum_{i=2}^{n}x(i)+\sum_{i=n+m}^{2n+m-2} x(i)\\ &=0+0+\frac{n-1}{2}-\frac{n-1}{2}+\frac{n-1}{2}-\frac{n-1}{2}\\ &=0. \end{align*} We have finished verifying $(\bold{x},\bold{1})=0$. Now we estimate the upper bound of $\lambda_2(T_n^m)$: \begin{align*} \lambda_2(T_n^m)&\le \frac{\bold{x}^T\bold{L}\bold{x}}{\bold{x}^T\bold{x}}\\ &=\frac{\sum_{(a,b)\in E_{T_n^{m}}} (x(a)-x(b))^2}{\sum_{i=1}^{2n+m-2}x(i)^2}\\ & = \frac{\sum_{1\le i,j\le n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2} x(i)^2}\\ &+ \frac{\sum_{n+1\le i,j\le n+m-2} (x(i)-x(j))^2+(x(1)-x(n+m-1))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2}x(i)^2}.\\ &+ \frac{\sum_{n+m-1\le i,j\le 2n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}. \end{align*} Notice that the first term is \begin{align*} &=\frac{ (x(1)-x(2))^2+(x(2)-x(3))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-1}x(i)^2}\\ &=\frac{2}{n-1+0+n-1}\\ &=\frac{1}{n-1} \end{align*} The second term is $0$, and the third term is \begin{align*} &=\frac{ (x(n+m-1)-x(n+m))^2+(x(n+m-1)-x(n+m+1))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-1}x(i)^2}\\ &=\frac{2}{n-1+0+n-1}\\ &=\frac{1}{n-1}. \end{align*} Adding all the three terms, we get $$ \lambda_2({T_n^m})\le \frac{1}{n-1}+\frac{1}{n-1}=\frac{2}{n-1}. $$\newline \textbf{Case 2:} One end of $P_m$ is $1$, and the other end of $P_m$ is $J$ where $n+m\le J \le 2n$ or one end of $P_m$ is $n+m-1$ and the other end of $P_m$ is $K$ where $1\le K \le n$. Without loss of generality we only discuss when one end of $P_m$ is $1$ and the other end of $P_m$ is $J$ because the other case follows by an identical argument. Figure $3$ demonstrates $T_7^3$ in this case: \begin{center} \includegraphics[scale = 0.5]{T2.png} \end{center} We can set the test vector to be same as case 1. Since we already know from case 1 that $(\bold{x},\bold{1})=0$, \begin{align*} \lambda_2(T_n^m)&\le \frac{\bold{x}^T\bold{L}\bold{x}}{\bold{x}^T\bold{x}}\\ &=\frac{\sum_{(a,b)\in E_{T_n^{m}}} (x(a)-x(b))^2}{\sum_{i=1}^{2n+m-2}x(i)^2}\\ & = \frac{\sum_{1\le i,j\le n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2} x(i)^2}\\ &+ \frac{\sum_{n+1\le i,j\le n+m-2+(x(1)-x(J))^2} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2}x(i)^2}\\ &+ \frac{\sum_{n+m-1\le i,j\le 2n+m-2} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2}x(i)^2}. \end{align*} Notice that that the second term is $$ \frac{0+1}{n-1+n-1}=\frac{1}{2(n-1)}. $$ We have calculated the first term and the third term in case 1. Hence $$ \lambda_2({T_n^m})\le \frac{1}{n-1}+\frac{1}{2(n-1)}+\frac{1}{n-1}=\frac{5}{2(n-1)}. $$\newline \textbf{Case 3:} One end of $P_m$ is $J$ where $2\le J \le n$ and the other end of $P_m$ is $K$ where $n+m\le K \le 2n$. Figure $4$ demonstrates $T_7^3$ in this case: \begin{center} \includegraphics[scale = 0.5]{T3.png} \end{center} Similar to before, we can set the test vector to be $$ \bold{x}(i) = \begin{dcases} 0\quad &i=1,n+m-1\\ 0\quad &n+1\le i \le n+m-2\\ 1 \quad &i=J,K\\ 1 \quad &2<i\le n \text{ and } i \text{ is a descendant of } J \\ 1 \quad &2<i\le n \text{ and } i \text{ is a ancestor of } J \\ 1 \quad &i<n+m\le n \text{ and } i \text{ is a descendant of } K\\ 1 \quad &i<n+m\le n \text{ and } i \text{ is a ancestor of } K\\ -1\quad & \text{otherwise} \end{dcases}. $$ We notice that for elements of $T_{n,1}$, the number of $1$'s in $\bold{x}$ is $\frac{n-1}{2}$ and the number of $-1$'s in $\bold{x}$ is $\frac{n-1}{2}$. For elements of $T_{n,1}$ the number of $1$'s in $\bold{x}$ is $\frac{n-1}{2}$ and the number of $-1$'s in $\bold{x}$ is $\frac{n-1}{2}$. Hence we have \begin{align*} (\bold{x},\bold{1})&=\sum_{i=1}^{2n+m-2} x(i)\\ &=x(1)+x(n+m-1) + \sum_{i=2}^{n}x(i)+\sum_{i=n+m}^{2n+m-2} x(i)\\ &=0+0+\frac{n-1}{2}-\frac{n-1}{2}+\frac{n-1}{2}-\frac{n-1}{2}\\ &=0. \end{align*} Now we can estimate the upper bound of second eigenvalue \begin{align*} \lambda_2(T_n^m)&\le \frac{\bold{x}^T\bold{L}\bold{x}}{\bold{x}^T\bold{x}}\\ &=\frac{\sum_{(a,b)\in E_{T_n^{m}}} (x(a)-x(b))^2}{\sum_{i=1}^{2n+m-2}x(i)^2}\\ & = \frac{\sum_{1\le i,j\le n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2} x(i)^2}\\ &+ \frac{\sum_{n+1\le i,j\le n+m-2} (x(i)-x(j))^2+(x(J)-x(K)^2)}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2}x(i)^2}.\\ &+ \frac{\sum_{n+m-1\le i,j\le 2n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+m-2} x(i)^2+\sum_{n+m-1\le i\le 2n+m-2}x(i)^2}. \end{align*} Notice that the second term is $$ \frac{0}{n-1+n-1}=0. $$ The first term and the second term calculation is basically the same as case 1, and the result is the same too. It follows that $$ \lambda_2(T_n^m)\le \frac{1}{n-1}+0+\frac{1}{n-1}=\frac{2}{(n-1)}, $$ which gives us our upper bound estimation. \newline For the lower bound, we can compare our graphs to complete graphs. For every pair of edge $(a,b)\in E_{T_n^m}$,let the path graph $P_{a,b}$ be a path from $a$ to $b$, and $G_{a,b}$ be a graph with a single edge $(a,b)$, then from the lemma we have $|P_{a,b}|P_{a,b} \succcurlyeq G_{a,b}$. We know that if $a$ and $b$ are both in the same binary tree, without loss of generality we assume they are both in $P_{n,1}$,from the definition of full binary tree we know that the length from the vertex $1$ to the vertices which have no children is $\log_2{(n+1)}-1$, so the length from vertex $a$ to $b$ is at most $2log_2{(n+1)}-2$. If $a$ and $b$ are in different binary trees, without loss of generality, we suppose $a$ is in $T_{n,1}$ and $b$ is in $T_{n,2}$ then the length of the path $P_{a,b}$ is the longest when $a$ and $b$ are the vertices which have no descendants. Hence the length of the path $P_{a,b}$ is at most $2\log_2{(n+1)}+m-3$. That is, $|P_{a,b}|\le 2\log_2(n+1)+m-3$. It follows that \begin{align*} G_{a,b}\preccurlyeq |P_{a,b}|P_{a,b} & \preccurlyeq((2\log_2(n+1)+m-3)P_{a,b} \\ &\preccurlyeq(2\log_2(n+1)+m-3)T_n^m. \end{align*} Also we know that complete graph $K_{2n+m-2}$ t has $\binom{2n+m-2}{2}$ single edges. Thus \begin{align*} K_{2n+m-2}\preccurlyeq \sum_{(a,b)\in E_{K_{2n+m-2}}}G_{a,b}&\preccurlyeq \binom{2n+m-2}{2} G_{a,b} \\ &\preccurlyeq \binom{2n+m-2}{2}(2log_2(n+1)+m-3) T_m^n. \end{align*} Hence, $$ 2n+m-2=\lambda(K_{2n+m-2}) \le \binom{2n+m-2}{2}(2log_2(n+1)+m-3) \lambda(T_m^n). $$ From the above, we conclude $\lambda_2(T_m^n)\geq \frac{2}{(2n+m-1)(2log_2(n+1)+m-3}$. \end{proof} \end{theorem} We notice that when $m=2$ the graph $T_n^2$ is connected by a single edge. Now we are doing the same thing as we did for the complete graphs. When graphs $T_{n,1}$ and $T_{n,2}$ are connected by $k$ different single edges $e_1,e_2,\dots e_k$ we get the graph $T_n^{2\times k}$. An example for $k = 3$ is given below: \begin{center} \includegraphics[scale = 0.5]{T4.jpg} \end{center} We now have following theorem. \begin{theorem} For the graph $T_n^{2\times k}$ described above, we have the following bound: $$ \frac{2}{(2n+1)(2\log_2(n-1)+1)}\le \lambda_2(T_n^{2\times k})\le \frac{2m+2}{n-1}+\chi_{k\geq n-1}\frac{-3m+3n-3}{2(n-1)} $$ where $\chi$ denotes the characteristic function. \begin{proof} We know that $T_n^2$ is a subgraph of $T_n^{2\times k}$ so we have $T_n^2\preccurlyeq T_n^{2\times k}$.Hence $$\frac{2}{(2n+1)(2\log_2(n-1)+1)}\le \lambda_2({T_n^2})\le \lambda_2(T_n^{2\times k}) $$ Thus we finished the lower bound. \newline For the upper bound we still use test vector. Now we notice the graph $T_{n,1}$ contains three different sets of vertices. One set, $V_{1,1}$, only contain the vertex $1$, the second set, $V_{1,2}$, contains vertex $2$ and all of it's descendants. The third set, $V_{1,3}$, contains vertex $3$ and its descendants. The graph $T_{n,2}$ also contains three different vertices sets. One set, $V_{2,1}$, only contains the vertex $n+1$, the second set, $V_{2,2}$, contains vertex $n+2$ and all of it's descendants. The third set, $V_{2,3}$, contains vertex $n+3$ and all of its descendants. \newline We know that edges $e_1,e_2,\ldots e_k$ contain $k$ vertices in $T_{n,1}$ and $k$ different vertices in $T_{n,2}$. For those $k$ vertices in $T_{n,1}$ we know that they might be in vertex set $V_{1,1}$ or $V_{1,2}$ or $V_{1,2}$. We also know that there are at most one vertex in $V _{1,1}$. Hence when $k\ge 2$ there are one or more vertices in either $V_{1,2}$ or $V_{1,3}$. \newline Without loss of generality, assume there are more vertices in set $V_2$. And for graph $T_{n,2}$ we also assume there are more vertices in $V_{2,2}$ We set the test vector to be $$ \bold{x}(i) = \begin{dcases} 0\quad &i=1,n+1\\ 1 \quad &i\in V_{1,2} \text{ or } i\in V_{2,2}\\ -1\quad & \text{otherwise} \end{dcases}. $$ We notice that the test vector here is basically the same when we define the test vector for the graph $T_n^m$. For elements of $T_{n,1}$, the number of occurrences of $1$ is $\frac{n-1}{2}$ and the number of occurrences of $-1$ is $\frac{n-1}{2}$. For elements of $T_{n,1}$ the number of occurrences of $1$ is $\frac{n-1}{2}$ and the number of occurrences of $-1$ is $\frac{n-1}{2}$. Hence we have \begin{align*} (\bold{x},\bold{1})&=\sum_{i=1}^{2n} x(i)\\ &=x(1)+x(n+2) + \sum_{i\in V_{1,2}}x(i)+\sum_{i\in V_{2,2}}^ x(i)+\sum_{i\in V_{1,3}}x(i)+\sum_{i\in V_{2,3}}x(i)\\ &=0+0+\frac{n-1}{2}+\frac{n-1}{2}-\frac{n-1}{2}-\frac{n-1}{2}\\ &=0. \end{align*} We have finished verifying $(\bold{x},\bold{1})=0$. Now we can try to bound $\lambda_2(T_n^m)$: \begin{align*} \lambda_2(T_n^{2\times k})&\le \frac{\bold{x}^T\bold{L}\bold{x}}{\bold{x}^T\bold{x}}\\ &=\frac{\sum_{(a,b)\in E_{T_n^{2\times k}}} (x(a)-x(b))^2}{\sum_{i=1}^{2n}x(i)^2}\\ & = \frac{\sum_{1\le i,j\le n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n} x(i)^2}\\ &+ \frac{\sum_{(i,j)\in {\{e_1,e_2,\dots e_k\}}} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}\\ &+ \frac{\sum_{n+m-1\le i,j\le 2n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}. \end{align*} Notice that the first term is \begin{align*} &\frac{\sum_{1\le i,j\le n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n} x(i)^2}\\ =&\frac{ (x(1)-x(2))^2+(x(2)-x(3))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le n+1} x(i)^2+\sum_{n+1\le i\le 2n}x(i)^2}\\ =&\frac{2}{n-1+n-1}\\ =&\frac{1}{n-1}. \end{align*} The second term is \begin{align*} \frac{\sum_{(i,j)\in {\{e_1,e_2,\dots e_k\}}} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2} &=\frac{\sum_{i\in V_{1,1} \text{ or } j\in V_{2,1}}(x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}\\ &+\frac{\sum_{i\in V_{1,2}, j\in V_{2,2}(x(i)-x(j))^2}}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}\\ &+\frac{\sum_{i\in V_{1,2}, j\in V_{2,3}(x(i)-x(j))^2}}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}\\ &+\frac{\sum_{i\in V_{1,3}, j\in V_{2,2}(x(i)-x(j))^2}}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}\\ &+\frac{\sum_{i\in V_{1,3}, j\in V_{2,3}(x(i)-x(j))^2}}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}\\ \end{align*} We know that there are at most $\frac{n-1}{2}$ vertices which have nonzero value connected to vertex $1$ and at most $\frac{n-1}{2}$ vertices which have nonzero value connected to vertex $n+1$. Hence we have $$ \frac{\sum_{i\in V_{1,1} \text{ or } j\in V_{2,1}}(x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2} \le \frac{2(n-1)(0-1)^2}{2(n-1}=1. $$ The second term in the sum above is $0.$ We know that $i=j\le \frac{n-1}{2}$, where $i\in V_{1,2},j\in V_{2,3}$. Thus, the third term in the sum above obeys $$ \frac{\sum_{i\in V_{1,2}, j\in V_{2,3}}(x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2} < \frac{(n-1)(1+1)^2}{2(n-1)}=2. $$ We know that $i=j\le \frac{n-1}{2}$ where $i\in V_{1,3},j\in V_{2,2}$, Thus, the fourth term in the sum above satisfies $$ \frac{\sum_{i\in V_{1,3}, j\in V_{2,2}}(x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2} < \frac{(n-1)(-1-1)^2}{2(n-1)}=2. $$ Lastly, the fifth term in the sum above is $0$.\\ We can also see that \begin{align*} &\frac{\sum_{n+m-1\le i,j\le 2n} (x(i)-x(j))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+m-1\le i\le 2n}x(i)^2}\\ =&\frac{ (x(n+1)-x(n+2))^2+(x(n+1)-x(n+3))^2}{\sum_{1\le i\le n} x(i)^2+\sum_{n+1\le i\le 2n}x(i)^2}\\ =&\frac{2}{n-1+n-1}\\ =&\frac{1}{n-1}. \end{align*} But we don't need to add above terms when $k$ is relatively small. We can make the inequality tighter depending on the value of $k$. We notice that when $k\le n-1$ the $\lambda_2({T_n^{2\times k}})$ takes the greatest value when $i\in V_{1,3},j\in V_{2,2}$ or $i\in V_{1,2}, j\in V_{2,3}$. Hence, we actually have $$ \lambda_2({T_n^{2\times k}})\le \frac{m(-1-1)^2}{2(n-1)}+\frac{2}{n-1}=\frac{2m+2}{n-1} $$ when $k> n-1$. Also, $\lambda_{T_n^{2\times k}}$ takes the greatest value when $i\in V_{1,3},j\in V_{2,2}$ or $i\in V_{1,2}, j\in V_{2,3}$. It's follows that \begin{align*} \lambda_{T_2(n^{2\times k}})&\le 2+ \frac{(m-(n-1))(0+-1)^2}{2(n-1)}+\frac{2}{n-1} \\ &=\frac{m+3n+1}{2(n-1)} \\ &=\frac{2m+2}{n-1}+\frac{-3m+3n-3}{2(n-1)}. \end{align*} From the above we get that \begin{align*} \lambda_{T_n^{2\times k}}\le \frac{2m+2}{n-1}+\chi_{k\geq n-1}\frac{-3m+3n-3}{2(n-1)} \end{align*} \end{proof} \end{theorem} Now we can construct a graph $T_{n\times l}^2$ which is connected by $l$ identical full binary graphs $T_{n,1},\dots T_{n,l}$ using single edge. For every graph $T_{n,j}$ where $1\le j \le l-1$, there is a vertex $v^i$ which is ancestor of all other vertices in $T_{n,j}$, we connect that with $v^{j+1}$. As an example, we display $T_{7\times 3}^2$: \begin{center} \includegraphics[scale = 0.7]{T5.png} \end{center} We have following theorem. \begin{theorem} For the graphs $T_{n\times l}^2$ which we described above we have following bound of the second eigenvalues: $$ \frac{2}{(nl-1)(l\cdot \log_2(n-1)-1)}\le T_{n\times l}^2 \le \frac{l}{n-1}. $$ \begin{proof} For the vertex $v^i$ which is ancestor of all other vertices in $T_{n,i}$, we label it as $(j-1)n+1$. Then $(j-1)n+1$ has two children. We label them as $(j-1)n+2$ and $(j-1)n+3$. Then we label children of $(j-1)n+2$ as $(j-1)n+4$ and $(j-1)n+5$, the children of $(j-1)n+3$ as $(j-1)n+6$ and $(j-1)n+7$, and so on until $jn$. To get the upper bound we still need to use a test vector. We can set the test vector to be: $$ \bold{x}(i) = \begin{dcases} 0\quad &j\in\{1,\dots l\},(j-1)n+1\\ 1 \quad &j\in\{1,\dots l\},(j-1)n+2\\ 1 \quad &j\in\{1,\dots l\},(j-1)n+2<i\le jn \text{ and } i \text{ descendant of } (j-1)n+2\\ -1\quad & \text{otherwise} \end{dcases}. $$ We notice that for elements of $T_{n,j}$ where $j\in \{1,\dots l\}$, the number of occurrences of $1$'s in $\bold{x}$ is $\frac{n-1}{2}$ and the number of occurrences of $-1$'s in $\bold{x}$ is $\frac{n-1}{2}$. Hence we have \begin{align*} (\bold{x},\bold{1})&=\sum_{j=1}^{l}\sum_{i=(j-1)n+1}^{jn} x(i)\\ &=l(0)+l\frac{n-1}{2}-l\frac{n-1}{2}\\ &=0. \end{align*} We have finished verifying $(\bold{x},\bold{1})=0$. Now we estimate the upper bound of $\lambda_2(T_n^m)$ : \begin{align*} \lambda_2(T_{n\times l}^2)&\le \frac{\bold{x}^T\bold{L}\bold{x}}{\bold{x}^T\bold{x}}\\ &=\frac{\sum_{(a,b)\in E_{T_{n\times l}^{2}}} (x(a)-x(b))^2}{\sum_{i=1}^{nl}x(i)^2}.\\ \end{align*} We notice that only when $a=(j-1)n+1$ with $b=(j-1)n+2$ or $b=(j-1)n+3$, then the term $(x(a)-x(b))^2$ is not zero. There are $n-1$ vertices in each $T_{n,j}$ such that $x{a}^2=1$. Hence the above equation has the following form: \begin{align*} & = \frac{\sum_{j=1}^{l}\sum_{1\le a,b\le nj} (x(a)-x(b))^2}{\sum_{j=1}^{l}\sum_{1\le a\le nj} x(a)^2}\\ & = \frac{2l}{(n-1)l}\\ & = \frac{2}{(n-1)}. \end{align*} \newline Now we need to find the lower bound. We still compare our graphs to complete graphs. For every pair of edge $(a,b)\in E_{T_n^m}$, let the path graph $P_{a,b}$ be a path from $a$ to $b$, and $G_{a,b}$ be a graph with a single edge $(a,b)$; then from the lemma, we have $|P_{a,b}|P_{a,b} \succcurlyeq G_{a,b}$. We notice that length of the path $P_{a,b}$ is the longest when $a$ and $b$ where $a\in T_{n,1}$ and $b\in T_{n,l}$, and $a$ and $b$ have no descendants. Hence the length of the path $P_{a,b}$ is at most $l \cdot \log_2{(n-1)}-1$, which means that $|P_{a,b}|\le 2 \log_l(n-1)-1$. Now, \begin{align*} G_{a,b}&\preccurlyeq |P_{a,b}|P_{a,b} \\ &\preccurlyeq(l \cdot \log_2(n-1)-1)P_{a,b} \\ &\preccurlyeq(l \cdot \log_2(n-1)-1)T_{n\times l}^m. \end{align*} Also, we know that complete graph $K_{nl}$ has $\binom{nl}{2}$ edges, so $$K_{nl}\preccurlyeq \sum_{(a,b)\in E_{K_{nl}}}G_{a,b}\preccurlyeq \binom{nl}{2} G_{a,b} \preccurlyeq \binom{nl}{2}(llog_2(n-1)-1) T_m^n. $$ Hence, $$ nl=\lambda_2(K_{nl}) \le \binom{nl}{2}(l\cdot \log_2(n-1)-1) \lambda(T_{n\times l}^2). $$ From the above, we get that $\lambda_2(T_m^n)\geq \frac{2}{(nl-1)(l \log_2(n-1)-1)}$. \end{proof} \end{theorem} We have finished discussing the graph $T_{n\times l}^2$. Now we will discuss a more general case when a graph $B_{n\times l}^2$ which is connected by $l$ identical graphs $G_{n,1},\dots G_{n,l}$ using a single edge. Graphs $G_{n,1},\dots G_{n,l}$ are all identical, but they can be any arbitrary graph. We have following theorem. \begin{theorem} For graphs of the form $B_{n\times l}^2$, which we described above, we have following bound on the second eigenvalue: $$ 0<\lambda_2(B_{n\times l}^2)\le 2 $$ \end{theorem} \begin{proof} Now we need to label graph $B_{n\times l}^2$ first. For graph $G_{n,1}$, we know that there is a vertex which is attached to another vertex in graph $G_{n,2}$; we label this vertex as $n$. All other vertices in the graph $G_{n,1}$ can be labeled from $1$ to $n-1$ without repeating. For graph $G_{n,l}$, we know that there is a vertex which is attached to another vertex in graph $G_{n,l-1}$; we label this vertex as $n(l-1)+1$. All other vertices in the graph $G_{n,l}$ can be labeled from $n(l-1)+2$ to $nl$ without repeating. For every graph $G_{n,i}$ where $2\le i \le l-1$, we know that there is a vertex which is attached to another vertex in graph $G_{n,i-1}$. We label this vertex as $n(i-1)+1$, and there is a vertex which is attached to another vertex in graph $G_{n,i+1}$. We label that vertex as $ni$. All other vertices in the graph $G_{n,i}$ can be labeled from $n(i-1)+2$ to $ni-1$ without repeating.\\ We notice that $B_{n\times l}^2$ is connected. Hence we have $$ \lambda_2(B_{n\times l}^2)>0. $$ Also, we notice that if we take away an edge $(n,n+1)$ from graph $B_{n\times l}^2,$ then we get a new graph $\tilde{B}_{n\times l}^2=B_{n\times l}^2\setminus(n,n-1),$ which is not connected. Thus, we have $\lambda_2({\tilde{B}_{n\times l}^2})=0$\\ From Theorem 3.2, we know that $$ 0=\lambda_2(B_{n\times l}^2)=\min\limits_{\substack{(\bold{x},\bold{1})=0,x\in R }}{\frac{\bold{x}^T \bold{L_{B_{n\times l}^2}}\bold{x}}{\bold{x}^T\bold{x}}} $$ and $$ \lambda_2({\tilde{B}_{n\times l}^2})=\min\limits_{\substack{(\bold{x},\bold{1})=0,x\in R }}{\frac{\bold{x}^T \bold{L_{\tilde{B}_{n\times l}^2}}\bold{x}}{\bold{x}^T\bold{x}}}. $$ Notice that \begin{align*} \lambda_2(B_{n\times l}^2)&=\min\limits_{\substack{(\bold{x},\bold{1})=0,x\in R }}{\frac{\bold{x}^T \bold{L_{B_{n\times l}^2}}\bold{x}}{\bold{x}^T\bold{x}}}\\ &=\min\limits_{\substack{(\bold{x},\bold{1})=0,x\in R }}\, \frac{1}{\bold{x}^T\bold{x}}\sum\limits_{(a,b)\in B_{n\times l}^2}(x(a)-x(b))^2\\ &\le \min\limits_{\substack{(\bold{x},\bold{1})=0, x\in R }}\, \left[\sum\limits_{\substack{(a,b)\in B_{n\times l}^2, \\ (a,b)\neq (n,n+1)}}\frac{(x(a)-x(b))^2}{{\bold{x}^T\bold{x}}}+\frac{(x(n)-x(n+1))^2}{{\bold{x}^T\bold{x}}}\right]\\ &= \min\limits_{\substack{(\bold{x},\bold{1})=0,x\in R }}\left[\sum\limits_{(a,b)\in \tilde{B}_{n\times l}^2}\frac{(x(a)-x(b))^2}{{\bold{x}^T\bold{x}}}+\frac{(x(n)-x(n+1))^2}{{\bold{x}^T\bold{x}}}\right]\\ &= \min\limits_{\substack{(\bold{x},\bold{1})=0,x\in R }}\left[\sum\limits_{(a,b)\in \tilde{B}_{n\times l}^2}\frac{(x(a)-x(b))^2}{{\bold{x}^T\bold{x}}} +\frac{(x(n)-x(n+1))^2}{{\bold{x}^T\bold{x}}}\right] \\ &= \min\limits_{\substack{(\bold{x},\bold{1})=0,x\in R }} (I_1 + I_2) \end{align*} where $$I_1 = \sum\limits_{(a,b)\in \tilde{G}_{n\times l}^2}\frac{(x(a)-x(b))^2}{{\bold{x}^T\bold{x}}} \text{ and } I_2 = \frac{(x(n)-x(n+1))^2}{{\bold{x}^T\bold{x}}}.$$ Notice that $I_1$ is the same as $$ 0=\lambda_2(\tilde{B}_{n\times l}^2)=\min\limits_{\substack{(\bold{x},\bold{1})=0,x\in R }}{\frac{\bold{x}^T \bold{L_{\tilde{B}_{n\times l}^2}}\bold{x}}{\bold{x}^T\bold{x}}}= \frac{1}{\bold{x}^T\bold{x}}\sum\limits_{(a,b)\in \tilde{B}_{n\times l}^2}(x(a)-x(b))^2. $$ We have the following bound for the denominator of $I_2$: \begin{align*} \bold{x}^T\bold{x}&=\sum_{a=1}^{a=nl}(x(a))^2\\ &\geq (x(n))^2+(x(n+1))^2\\ &= \frac{1}{2}(x(n))^2+\frac{1}{2}(x(n+1))^2+\frac{1}{2}(x(n))^2+\frac{1}{2}(x(n+1))^2-x(n)x(n+1)+x(n)x(n+1)\\ &= \frac{1}{2}((x(n))^2+(x(n+1))^2+2x(n)x(n+1))+\frac{1}{2}((x(n))^2+(x(n+1))^2-2x(n)x(n+1))\\ &= \frac{1}{2}(x(n)+x(n+1))^2+\frac{1}{2}(x(n)-x(n+1))^2\\ & \geq \frac{1}{2}(x(n)-x(n+1))^2. \end{align*} We now use this to bound the second term. $$ {\frac{(x(n)-x(n+1))^2}{\bold{x}^T\bold{x}}}\le \frac{(x(n)-x(n+1))^2}{\frac{1}{2}(x(n)-x(n+1))^2}=2. $$ Adding two terms together gets us $$ \lambda_2(B_{n\times l}^2)\le 2. $$ Thus, the result is proven. \end{proof} \begin{remark} The upper bound in the inequality above cannot be improved. We notice that when $G_{1\times 2}^2=P_2,$ we know that the graph $P_2$ is constructed by connecting two identical graphs $G_{1,1}$ together, where $G_{1,1}$ is a single vertex. Then we have $\lambda_2(G_{1\times 2}^2)=2$, so equality is achieved. \end{remark} \end{section} \begin{section}{Conclusion and future work} Our work from section 4 to section 6 went through various different graphs. We noticed that for the $K_n$ type of graphs $D_n^m$, we have the approximate bound $\lambda_2(D_n^m)\sim\frac{1}{n}$. We also observe that when $n$ and $m$ increase the second eigenvalues decrease. Similarly, for $D_n^{2\times k}$, we have $\lambda_2(D_n^{2\times k})\sim\frac{1}{n}$, and when $n$ increases the second eigenvalues decrease too. \newline We noticed that for $S_n$ type graphs, $S_n^m$, we still have $\lambda_2(S_n^m)\sim\frac{1}{n}$. This is expected because star graph is a type of complete bipartite graph $K_{1,n-1}$, and complete graphs can also be complete bipartite graphs depending on the choice of $n$. But the $T_n$ type of bridge graph is different from the first two types of graphs. We noticed the lower bound of the second eigenvalues of $T_n^m$ and $T_n^{2\times k}$ and $T_{n\times}^{l}$ are all dependent on $\log(n+1)$. Also, the upper bound is still asymptotically dependent on $\frac{1}{n}$. \newline From the above proofs, we noticed that the test vector method is a very good technique for upper bound of the eigenvalues. This is because theorem 3.2 enables us to find a test vector which is orthogonal to the first eigenvector, and from theorem 3.3 we know that the first eigenvector is $\bold{1}$. It's also important to use theorem 6.4, which bounded general bridge graphs $B_{n\times l}^2$. Now we are curious what will happen if we construct a bridge graph like $B_{n\times \infty}^2$. It will not be the same as the case when $l$ is finite because we can't count the vertices one by one anymore. However, we still want to know if the results are somewhat similar.\newline Our future work will be constructing infinite bridge graphs. Now we can start from some basic definition and related theorems. \begin{definition} An infinite graph $G=(V,E)$ is a graph which has a countably infinite number of vertices. Infinite Bridge graphs $G_{n\times \infty}^m$ are constructed by using path graphs, $P_m$ with $n\geq 2$, and gluing together a countably infinite number of identical finite graphs on each end of the paths. Usually we can find an invertible map from vertex set $V$ to $\mathbb{Q}$. We only discuss unweighted graphs. \end{definition} For infinite graphs we cannot use adjacency matrices anymore. Now we are seeking a substitution of matrices to associate our graph with an operator. Definitions of operators related to infinite graphs are mentioned by other authors like Bojann Mohar\cite{b4} and Dragos M. Cvetokvic\cite{b3}, and Ayadi Hela\cite{b5} also defined the Laplacian Operator. But since we are only interested in infinite bridge graphs, we will use a different definition than other authors. The following is an example formed by attaching a countably infinite number of $K_8$ graphs together: \begin{center} \includegraphics[scale = 0.5]{T_inf.jpg} \end{center} \begin{definition} The space $\ell^2(\mathbb{N} \times \mathbb{N})$ is defined as the space of sequences $\{x_{i,j}\}_{i.j \in \mathbb{N}}$ such that $$\sum_{\mathbb{N} \times \mathbb{N}} |x_{i,j}|^2 < \infty.$$ \end{definition} \begin{definition} The adjacency operator $M$ of a weighted graph $G=(V,E)$ is defined as operator with the following entries $$ M(a,b) = \begin{dcases} 1 \quad (a,b)\in E\\ 0 \quad (a,b)\notin E. \end{dcases} $$ Notice that $\{M(a,b)\}_{a,b \in \mathbb{N}}$ forms a sequence. \end{definition} \begin{definition} The degree operator $D$ of a graph $G=(V,E)$ is a diagonal matrix whose entries are given by $$ D(a,b) = \begin{dcases} d(a)\quad &a=b\\ 0 \quad &a\neq b. \end{dcases} $$ Like before, $\{D(a,b)\}_{a,b \in \mathbb{N}}$ forms a sequence. \end{definition} \begin{definition} The graph laplacian operator $L$ of a graph $G$ is defined to be $$ L =M - D, $$ where the subtraction operation is subtracting corresponding elements in each sequence. \end{definition} \begin{theorem} The adjacency operator, degree operator and laplacian operator are all linear operators. \begin{proof} The proof is straightforward from the definition of the operators. \end{proof} \end{theorem} \begin{theorem} For the laplacian operator $L_G: \mathcal{X} \rightarrow \mathcal{Y}$ with $\mathcal{X}, \mathcal{Y} \subset \ell^2(\mathbb{N})$, where $G$ is a bridge graph, $L$ is a well defined mapping. \begin{proof} We notice that for $a\in V_G$, we have $$Lx(a)=\sum_{b\in N(a)}(x(a)-x(b)).$$ Since we assume that $x\in \ell^2(\mathbb{N})$, we know there is an $M$ such that $(\sum_{a=1}^{\infty} \|a\|_2)^{\frac{1}{2}}<M$ for some $M > 0$. Since $G$ is a bridge graph, then we know that for every vertex $a$, then $a$ has a finite number of vertices in its neighborhood. Hence we have \begin{align*} \| y\|_2 &=\| L x \|_2\\ &=\left(\sum_{a=1}^{\infty}\sum_{b\in N(a)}(x(a)-x(b))^2\right)^{\frac{1}{2}}\\ &\le \left(\sum_{a=1}^{\infty}m \cdot \max_{b\in N(a)}\left\{\|a\|_2^2,\|b\|_2^2\right\}\right)^{\frac{1}{2}}\\ &\le \left(\sum_{a=1}^{\infty} m^2 \|a\|_2^2\right)^{\frac{1}{2}}\\ &=m\| L x \|_2 \\ &< mM. \end{align*} Hence from the second last line of the above equations we get that Laplacian operator $L_G$ is a bounded operator and $y\in \ell^2(\mathbb{N})$. \end{proof} \end{theorem} The following definitions are from Elias M. Stein and Rami Shakarchi\cite{b2}. \begin{definition} It is well known that $\ell^2(\mathbb{N})$ is a Hilbert space. Therefore, it admits an inner product. For vector $x,y\in \ell^2(\mathbb{N})$, the inner product of $x$ and $y$ is defined as $$(x,y)= \sum_{n \in \mathbb{N}} x_n y_n.$$ \end{definition} \begin{definition} For a graph operator $T$, if we have $$ T\phi=\lambda \phi, $$ then we call $\lambda$ the eigenvalue and $\phi$ the eigenvector corresponding to eigenvalue $\lambda$ for operator $T$. \end{definition} \begin{definition} We say $\lambda\in \mathbb{R}$ is in the spectrum of $A$ if $A-\lambda I$ has no bounded inverse. The spectrum is denoted by $\sigma({A})$ where $\sigma({A}) \subset \mathbb{R}$, and the resolvent set for $A$ is $\rho({A})=\mathbb{R}\setminus\sigma({S})$. \end{definition} Our future work will be about the spectrum of Laplacian operator. \end{section} \newpage
{ "timestamp": "2023-01-03T02:06:27", "yymm": "2202", "arxiv_id": "2202.11875", "language": "en", "url": "https://arxiv.org/abs/2202.11875", "abstract": "The Bridge graph is a special type of graph which are constructed by connecting identical connected graphs with path graphs. We discuss different types of bridge graphs $B_{n\\times l}^{m\\times k}$ in this paper. In particular, we discuss the following: complete-type bridge graphs, star-type bridge graphs, and full binary tree bridge graphs. We also bound the second eigenvalues of the graph Laplacian of these graphs using methods from Spectral Graph Theory. In general, we prove that for general bridge graphs, $B_{n\\times l}^2$, the second eigenvalue of the graph Laplacian should be between $0$ and $2$, inclusive. In the end, we talk about future work on infinite bridge graphs. We created definitions and found the related theorems to support our future work about infinite bridge graphs.", "subjects": "Combinatorics (math.CO)", "title": "Characterizing Spectral Properties of Bridge", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9740426405416754, "lm_q2_score": 0.8221891239865619, "lm_q1q2_score": 0.8008472653525177 }
https://arxiv.org/abs/1308.0861
Bounds of incidences between points and algebraic curves
We prove new bounds on the number of incidences between points and higher degree algebraic curves. The key ingredient is an improved initial bound, which is valid for all fields. Then we apply the polynomial method to obtain global bounds on $\mathbb{R}$ and $\mathbb{C}$.
\section{introduction} The Szemer\'{e}di--Trotter theorem \cite{szemeredi1983extremal} says that for a finite set, $L$, of lines and a finite set, $P$, of points in $\mathbb{R}^2$, the number of incidences is less than a constant times $|P|^{\frac{2}{3}} |L|^{\frac{2}{3}} + |P| + |L|$. There have been several generalizations of this theorem. For example, Pach and Sharir \cite{pach1998number} allow simple curves that have $k$ degrees of freedom and multiplicity-type $C$: (i) for any $k$ distinct points there are at most $C$ curves in $L$ that pass through them, and (ii) any two distinct curves in $L$ have at most $C$ intersection points. This is summarized in the following theorem: \begin{thm}[Pach--Sharir 98]\label{PSthm} For a finite set $P$ of points in $\mathbb{R}^2$ and a finite set $L$ of simple curves which have $k$ degrees of freedom and multiplicity-type $C$. The number of incidences $|\mathcal{I}(P,L)|:=|\{(p,l)\in P\times L, p\in l\}|$ satisfies \begin{equation} |\mathcal{I}(P,L)|\lesssim_{C, k} |P|^{\tfrac{k}{2k-1}}|L|^{\tfrac{2k-2}{2k-1}}+|P|+|L|. \end{equation} \end{thm} \subsection*{Notation} We use the asymptotic notation $X=\OO(Y)$ or $X\lesssim Y$ to denote the estimate $X\leq CY$ for some constant $C$. If we need the implicit constant $C$ to depend on additional parameters, then we indicate this by subscripts. For example, $X=\OO_{d}(Y)$ or $X\lesssim_{d} Y$ means that $X\leq C_{d}Y$ for some constant $C_{d}$ that depends on $d$. The main result of this paper is an improvement to Theorem \ref{PSthm} when $L$ is a set of higher degree algebraic curves. Let $L$ be a finite set of algebraic curves of degree $\leq d$ in $\mathbb{R}^{2}$, any two of which do not share a common irreducible component. Let $P$ be a finite set of distinct points in $\mathbb{R}^{2}$. By B\'{e}zout's theorem, there is at most one curve in $L$ that goes through a subset of $P$ of size $d^{2}+1$. In the notation introduced by Pach and Sharir, a degree $d$ algebraic curve has $d^2+1$ degrees of freedom and multiplicity-type $d^2$. However, one may wonder whether $d^2 +1$ is a misleading definition of the degrees of freedom since generically $A:= {d+2 \choose 2} -1$ points determine a degree $d$ algebraic curve and $A\leq d^{2}+1$ when $d\geq 3$. This suggests that Theorem ~\ref{PSthm} may still hold for degree $d$ curves with the ``generic degree of freedom" $A$. Indeed, we prove that this is the case. \begin{thm}\label{main theorem} Let $d$ be a positive integer, $A={d+2\choose 2} -1$, $L$ a finite set of degree $\leq d$ algebraic curves in $\mathbb{R}^{2}$ such that any two distinct algebraic curves do not share a common irreducible component, and $P$ a finite set of points in $\mathbb{R}^{2}$. Then, \begin{equation}\label{estimatemainthm} |\mathcal{I}(P,L)|\lesssim_{d} |P|^{\tfrac{A}{2A-1}}|L|^{\tfrac{2A-2}{2A-1}}+|P|+|L|. \end{equation} \end{thm} This gives a better bound than Theorem ~\ref{PSthm} for degree $d$ algebraic curves when $d\geq 3$. It gives the same bound when $d=1$ or $d=2$. We generalize Theorem ~\ref{main theorem} to algebraic curves parametrized by an algebraic variety. \begin{definition} Given an integer $d\geq 1$ and a field $\F$, consider the Veronese embedding $\nu_{d}: \PP^{2}\rightarrow \PP^{A}$ given by:$$\nu_{d}: [x, y, z]\mapsto [x^{d},\dots, x^{i}y^{j}z^{k},\dots, z^{d}]_{i+j+k=d}.$$ We identify a degree $d$ curve in $\PP^{2}$ with the preimage of a hyperplane in $\PP^{A}$ or a point in the dual space $(\PP^{A})^{*}$. We say a subset $\M $ of the space of degree $d$ polynomials $\subset S_{d}$ is \emph{parametrized by an algebraic variety} $M$ if it is the preimage of $M\subset (\PP^{A})^{*}$, and we define the dimension of $\M$ to be $\dim M$. \end{definition} A consequence of the Theorem ~\ref{main theorem} is the following: \begin{cor}\label{variety} Given an integer $d\geq 1$, $A={d+2\choose 2}-1$, and a subset $\mathcal{M}$ of the space of degree $d$ polynomials is parametrized by an algebraic variety $M$ of dimension $\leq k$. Let $P$ be a finite set of points in $\mathbb{RP}^2$, and $L$ a finite subset of $\M$ such that no two curves in $L$ share a common irreducible component. Then, \begin{equation} |\mathcal{I}(P,L)|\lesssim_{\mathcal{M}}|P|^{\tfrac{k}{2k-1}}|L|^{\tfrac{2k-2}{2k-1}}+|P|+|L|. \end{equation} \end{cor} This generalization is helpful when we consider a family of curves with special properties, for example, parabolas, circles and a family of curves passing through a common point. The proof of Theorem ~\ref{main theorem} is a standard application of the polynomial method (see for example \cite{dvir2009size} and \cite{guth2010erdos}) with the following new initial bound. For an exposition of the polynomial method we refer the reader to \cite{kaplan2012simple}. \begin{lem}[Initial bound]\label{trivial bound} Let $\mathbb{F}$ be a field. Given an integer $d\geq 1$, $A={d+2\choose 2} -1$, a finite set of algebraic curves, $L$, in $\mathbb{F}^{2}$ of degree $\leq d$ such that any two distinct curves in $L$ do not share a common component, and a finite set of points, $P$, in $\mathbb{F}^{2}$. Then, \begin{equation} |\mathcal{I}(P,L)|\lesssim_{d} |P|^A+|L|.\label{initial bound} \end{equation} \end{lem} Combining (\ref{initial bound}) with Solymosi and Tao's polynomial method and induction on the number of points \cite{solymosi2012incidence}, we obtain the following incidence theorem on complex space: \begin{thm}\label{complex epsilon} Given an integer $d\geq 1$, $A={d+2\choose 2} -1$, a finite set of points $P$ in $\mathbb{C}^{2}$, a finite set of algebraic curves, $L$, in $\mathbb{C}^{2}$ of degree $\leq d$ such that any two curves: (i) do not share a common irreducible component; (ii) intersect transversally at smooth points. Then, for any $\epsilon>0$, \begin{equation}\label{estimateepsilonleft} |\mathcal{I}(P,L)|\lesssim_{\epsilon,d} |P|^{\tfrac{A}{2A-1}+\epsilon}|L|^{\tfrac{2A-2}{2A-1}}+|P|+|L| \end{equation} and \begin{equation}\label{estimateepsilonright} |\mathcal{I}(P,L)|\lesssim_{\epsilon,d} |P|^{\tfrac{A}{2A-1}}|L|^{\tfrac{2A-2}{2A-1}+\epsilon}+|P|+|L|. \end{equation} \end{thm} \subsection*{Acknowledgements} We would like to thank Larry Guth for encouraging us to work on the problem. He suggested that we try: improving the initial bound and working on algebraic subsets. We also thank Alex Townsend for reading drafts and for the suggestions. \section{Proof of Theorem ~\ref{main theorem}} \subsection{The initial bound} In this subsection we prove the initial bound by double counting. \begin{definition} Given a point $p\in \PP^{2}$, let $H_{p}$ denote the corresponding hyperplane in $\PP^{A*}$ via the Veronese embedding and dual. Given a finite set of points $\Gamma=\{p_{1},\dots,p_{n}\}$, we define $m_{d}(\Gamma)=\dim(\cap H_{p_{i}})$, which characterizes the dimension of curves passing through all the points in $\Gamma$. In particular, if $m_{d}(\Gamma)=0$, there is at most one curve of degree $d$ that passes through $\Gamma$. \end{definition} \begin{proof}[Proof of Lemma \ref{trivial bound}] We may remove curves containing fewer than $d^{2}+1$ points by adding $\OO(|L|)$ to the bound. Now we assume that each curve contains more than $d^{2}+1$ points of $P$. Fix a curve $l\in L$. We call an $A$-tuple $\Gamma'$ \emph{good} if it is a subset of $\Gamma\subset l$ with $|\Gamma|=d^{2}+1$ and $m_{d}(\Gamma')=m_{d}(\Gamma)$. For any $(d^{2}+1)$-tuple $\Gamma\in l\cap P$, there exists a good $A$-tuple $\subset \Gamma$ since $L$ is parametrized by $\PP^{A}$. Since $\cap_{p\in \Gamma}H_{p}$ and $\cap_{p\in \Gamma'}H_{p}$ have the same dimension and $\cap_{p\in \Gamma}H_{p}\subseteq\cap_{p\in \Gamma'}H_{p}$, the two vector subspaces are in fact the same. In other words, any curve in $L$ passing through $\Gamma'$ must pass through all of $\Gamma$. Since curves in $L$ do not have a common irreducible component, by B\'{e}zout's theorem, every set of $d^{2}+1$ points determines a unique curve in $L$. Hence, every good $A$-tuple determines a unique curve in $L$. There are at least ${|l\cap P| \choose d^2+1}/{ |l\cap P| - A\choose d^{2}+1-A}$ distinct good $A$-tuples $\Gamma'$ determining $l$. The number of good $A$-tuples determines the number of points in $l\cap P$ because ${|l\cap P| \choose d^2+1}/{ |l\cap P| - A\choose d^{2}+1-A}= \OO(|l\cap P|^{A})\gtrsim|l\cap P|$. On the other hand, the number of $A$-tuples is ${|P|\choose A}= \OO(|P|^{A})$. Then, $$|\mathcal{I}(P,L)| = \sum_{l \in L} |l\cap P|\lesssim |P|^{A}+|L|,$$ where the $|L|$ term comes from the first step when we deleted curves with fewer than $d^{2}+1$ points from $P$. \end{proof} \subsection{Polynomial method}\label{polynomialpartitioningsection} Now we can apply the polynomial method to the initial bound and conclude the proof of Theorem \ref{main theorem}. We shall use the following polynomial partitioning proposition (see, for example, Theorem 4.1 in \cite{guth2010erdos}): \begin{prop}\label{cell decomposition} Let $P$ be a finite set of points in $\mathbb{R}^m$ and let $D$ be a positive integer. Then, there exists a nonzero polynomial $Q$ of degree at most $D$ and a decomposition \[\mathbb{R}^{m}=\{Q=0\}\cup U_{1}\cup\cdots \cup U_{M}\] into the hypersurface $\{Q=0\}$ and a collection $U_{1},\ldots ,U_{M}$ of open sets (which we call \emph{cells}) bounded by $\{Q=0\}$, such that $M =\OO_{m}( D^m)$ and that each cell $U_{i}$ contains $\OO_{m}(|P|/D^{m})$ points. \end{prop} \begin{proof}[Proof of Theorem \ref{main theorem}] Applying Proposition ~\ref{cell decomposition}, we find a polynomial $Q$ of degree $D$ (to be chosen later) that partitions $\mathbb{R}^{2}$ into $M$ cells: \[\mathbb{R}^{2}=\{Q=0\}\cup U_{1}\cup\cdots \cup U_{M},\] where $M=\OO( D^2)$ and $U_{1},\ldots ,U_{M}$ are open sets bounded by $\{Q=0\}$ and $P_{i}=U_{i}\cap P$. Let $L_{i}$ be the set of curves that have non-empty intersection with $U_{i}$. Then $|P_{i}|=\OO(|P|/D^{2})$. By B\'{e}zout's theorem we see every curve meets $\{Q=0\}$ at no more than $d\cdot D\leq \OO_{d}(D)$ components. Fix a curve $l\in L$, by Harnack's curve theorem~\cite{harnack1876ueber}, the curve itself has $\OO_d (1)$ connected components. Moreover, since a component of $l$ is either contained in $U_{i}$ for some $i$ or it must meet the partition surface $\{Q=0\}$, $l$ can only meet $\OO_{d}(D)$ many $U_{i}$'s. Thus, we have the inequality $\sum |L_{i}|\leq D|L|$. We may assume that every curve is irreducible, otherwise we can replace the curve with its irreducible components and the cardinality of curves increase by only a constant factor. Let $P_{cell}$ denote the points of $P$ in cells $U_{i}$ and let $P_{alg}$ denote those on the partition surface $\{Q=0\}$. Similarly, let $L_{alg}$ denote those curves that belong to $\{Q = 0 \}$ and let $L_{cell}$ be the union of the other curves. We deduce by Lemma \ref{trivial bound} and H\"{o}lder's inequality: \begin{align}\label{polypartitionestimate} |\mathcal{I}(P,L)|&= |\mathcal{I}(P_{cell}, L_{cell}) |+ |\mathcal{I}(P_{alg},L_{cell}) |+|\mathcal{I}(P_{alg}, L_{alg})|\nonumber\\ &\lesssim_{d} \sum_i (|P_{i}|^{A}+|L_{i}|) + D|L_{cell}|+|\mathcal{I}(P_{alg}, L_{alg})|\nonumber\\ &\lesssim_{d} |P|^{A}D^{-2(A-1)}+D|L|+|\mathcal{I}(P_{alg}, L_{alg})|. \end{align} In addition, we may assume that $ |P|^{1/2}\leq |L|\leq |P|^{A}$, otherwise (\ref{estimatemainthm}) already holds either by Lemma \ref{trivial bound} or by another initial bound $|\mathcal{I} (P, L)| \lesssim |P| + |L|^2$ (every two curves intersect at at most $\OO(1)$ points). In this case, we may choose $D =\OO_{d}( |P|^{\tfrac{A}{2A-1}} |L|^{-\tfrac{1}{2A-1}})$ and $D\leq|L|/2$. Then, the first two terms on the right-hand side of (\ref{polypartitionestimate}) are $O_{d}( |P|^{\tfrac{A}{2A-1}}|L|^{\tfrac{2A-2}{2A-1}})$. Since $|L_{alg}| \leq D \leq \frac{|L|}{2}$, we can perform a dyadic induction on $|L|$. By repeating the above process on $L_{alg}$, we obtain the following: \begin{align} |\mathcal{I}(P,L)|&\lesssim_{d} \sum_{ 0\leq i \leq \log(|L|/|P|^{1/2})} |P|^{\tfrac{A}{2A-1}}(2^{-i}|L|)^{\tfrac{2A-2}{2A-1}} + |P|+|L|\nonumber\\&\lesssim_{d}|P|^{\tfrac{A}{2A-1}}|L|^{\tfrac{2A-2}{2A-1}}+|P|+|L|.\nonumber \end{align} The induction stops when $i> \log(|L|/|P|^{1/2})$ because when $2^{-i}L\lesssim |P|^{1/2}$ the number of incidences in the $i$-th step is bounded by $\OO( |P| )$. This proves (\ref{estimatemainthm}). \end{proof} \section{An estimate for parametrized curves} We prove an initial bound with parametrized curves, which implies Corollary \ref{variety}. We first state two propositions that we will need to prove the initial bound. \begin{prop}(see \cite{hartshorne1977algebraic}, Ch I, Exercise 1.8)\label{dimension diminute} If $V$ is an $r$-dimensional variety in $\mathbb{F}^{d}$, and $P :\mathbb{F}^{d}\rightarrow V$ is a polynomial which is not identically zero on $V$, then every component of $V\cap \{P=0\}$ has dimension $r-1$. \end{prop} \begin{prop}(see \cite{fulton1984intersection}, Section 2.3)\label{refine Bezout} Let $V_{1},\ldots V_{s}$ be subvarieties of $\mathbb{FP}^{N}$, and let $Z_{1},\ldots ,Z_{r}$ be the irreducible components of $V_{1}\cap\cdots \cap V_{r}$. Then, \[\sum_{i=1}^{r} \deg(Z_{i})\leq \prod_{j=1}^{s} \deg(V_{j}).\] \end{prop} \begin{lem}\label{algebraicsettrivialbound} With the same setting and notation as in Corollary \ref{variety}, we have \begin{equation} |\mathcal{I}(P,L)|\lesssim_{\mathcal{M}}|P|^{k}+|L|. \end{equation} \end{lem} \begin{proof} Without loss of generality we assume that $\mathbb{F}$ is an algebraically closed field, every curve in $L$ is irreducible, of degree $d$, and contains more than $d^{2}+1$ points from $P$. A curve $l\in L$ corresponds to a point of intersection $\cap_{p\in l\cap P}H_{p}\cap M$. Comparing with the proof of Lemma \ref{trivial bound}, it suffices to prove that for every $(d^{2}+1)$-tuple $\Gamma$ in $l\cap P$, there is a $\Gamma' \subseteq \Gamma$ such that $\cap_{p\in \Gamma'} H_{p}\cap M$ contains at most $O_{\mathcal{M}} (1)$ curves and $|\Gamma'|=k$. This follows from Proposition~\ref{dimension diminute} and Proposition~\ref{refine Bezout} above. Indeed, by iterately applying Proposition ~\ref{dimension diminute}, we can choose $\Gamma'$ such that $| \Gamma'|=k$ and $\cap_{p\in\Gamma'}H_{p}\cap\mathcal{M}$ has dimension $0$. By Proposition \ref{refine Bezout}, the cardinality of $\cap_{p\in\Gamma'}H_{p}\cap\mathcal{M}$ is bounded by a constant depending on $k$ and $\deg \mathcal{M}$. \end{proof} \section{A theorem on the complex field with $\epsilon$} In this section, we follow the approach of \cite{solymosi2012incidence} to sketch a proof of Theorem \ref{complex epsilon}. The idea is to partition the point set $P$ with a polynomial of constant degree and then use induction on the size of $P$. In other words, the degree does not depend on the size of $P$ and $L$. With an $\epsilon$ loss in the exponents, one can perform induction on $|P|$, which controls incidences in the complement of the partition surface (in the cells). Here, a constant degree implies constant complexity, and we can use dimension reduction to estimate incidences on the partition surface. \begin{proof}[Proof of Theorem \ref{complex epsilon}] In our proof, $C$ is a constant depending only on $d$ and $\epsilon$ which may vary from place to place. $C_0, C_1$ and $C_2$ are positive constants to be chosen later, where $C_{0}, C_1 >2$ are sufficiently large depending on $d$ and $\epsilon$, and $C_{2}$ is sufficiently large depending on $C_{1}$, $C_{0}$, $d$ and $\epsilon$. We do an induction on $|P|$. Suppose for $|P'|\leq \tfrac{|P|}{2}$ and $|L'| \leq |L|$, we have by the induction hypothesis, \begin{equation}\label{hypothesis} |\mathcal{I}(P',L')|\leq C_{2}|P'|^{\tfrac{A}{2A-1}+\epsilon}|L'|^{\tfrac{2A-2}{2A-1}}+C_{0}(|P'|+|L'|). \end{equation} Our goal is to prove \begin{equation}\label{inductiveclaim} |\mathcal{I}(P,L)|\leq C_{2}|P|^{\tfrac{A}{2A-1}+\epsilon}|L|^{\tfrac{2A-2}{2A-1}}+C_{0}(|P|+|L|). \end{equation} We apply Proposition \ref{cell decomposition} to $D = C_{1}$ on $\mathbb{C}^2 \simeq \mathbb{R}^{4}$ and obtain the following partition: \begin{equation} \mathbb{R}^{4}=\{Q=0\}\cup U_{1}\cup\cdots \cup U_{M}. \end{equation} Here, $Q: \mathbb{R}^{4}\rightarrow\mathbb{R}$ has degree at most $C_{1}$, $M =\OO( C_{1}^{4})$, and $|P_{i}| = |P \cap U_{i}| = O(\tfrac{|P|}{C_{1}^{4}})\leq \tfrac{|P|}{2}$. We denote $L_{i}$ to be the set of curves in $L$ with nonempty intersection with $U_{i}$. Thus, by the induction hypothesis we have, \begin{align} |\mathcal{I}(P_{i},L_{i})| \leq & C_{2}|P_{i}|^{\tfrac{A}{2A-1}+\epsilon}| L_{i}|^{\tfrac{2A-2}{2A-1}}+C_{0}(|P_{i}| + |L_{i}|)\nonumber\\ \leq & C [C_{2}C_{1}^{-4(\tfrac{A}{2A-1}+\epsilon)}|P|^{\tfrac{A}{2A-1}+\epsilon}|L_{i}|^{\tfrac{2A-2}{2A-1}}+C_{0}(\tfrac{|P|}{C_1^4}+|L_{i}|)]. \end{align} For $l$ belonging to some $L_{i}$, we apply a result in real algebraic geometry that implies the number of connected components of $l\setminus \{Q=0\}$ is at most $\OO_{d} (C_{1}^{2})$ (see Theorem A.2 of \cite{solymosi2012incidence}, \cite{milnor1964betti},\cite{petrovskii1949topology},\cite{thom1965homologie}). We deduce that \begin{equation} \sum_{i=1}^{M}|L_{i}| \leq CC_{1}^{2}|L|. \end{equation} Adding up $|\mathcal{I}(P_i,L_{i})|$ and applying H\"{o}lder's inequality, we obtain \begin{align} |\mathcal{I}(P_{cell},L_{cell})| = & \sum_{i=1}^{M}|\mathcal{I}(P_{i},L_{i})|\nonumber\\ \leq & C(C_{2}C_{1}^{-4(\tfrac{A}{2A-1}+\epsilon)}|P|^{\tfrac{A}{2A-1}+\epsilon}(\sum_{i=1}^{M}|L_{i}|^{\tfrac{2A-2}{2A-1}})+C_0 (|P|+C_{1}^{2}|L|))\nonumber\\ \leq & C(C_{1}^{-4\epsilon}C_{2}|P|^{\tfrac{A}{2A-1}+\epsilon}|L|^{\tfrac{2A-2}{2A-1}}+C_{0}(|P|+C_{1}^{2}|L|)). \end{align} Now we recall the two trivial bounds (\ref{L2+P}) explained in the case of real space and by Lemma~\ref{trivial bound}, we have: \begin{equation}\label{L2+P} |\mathcal{I}(P,L)|\lesssim_{d} |L|^{2}+|P|,~~~~~ |\mathcal{I}(P,L)|\lesssim_{d} |P|^{A}+|L|. \end{equation} Thus, one may assume that $|P|^{\tfrac{1}{2}}\lesssim_{d}|L|\lesssim_{d}|P|^{A}$, otherwise we immediately have $|\mathcal{I}(P,L)|\lesssim_{d}|P|+|L|$ and it suffices to choose $C_0$ larger than the implicit constant. With this assumption, we have \begin{equation}\label{PcellLcell} |\mathcal{I}(P_{cell},L_{cell})| \leq C(C_{1}^{-4\epsilon}C_{2}+C_{0}(|P|^{-\epsilon}+C_{1}^{2}|P|^{-\epsilon}))|P|^{\tfrac{A}{2A-2}+\epsilon}|L|^{\tfrac{2A-2}{2A-1}}. \end{equation} If the following inequality is given \begin{equation}\label{PalgL} \mathcal{I}(P_{alg}, L)\lesssim_{C_1}|P|^{\tfrac{A}{2A-1}}|L|^{\tfrac{2A-2}{2A-1}} + |P|+|L|, \end{equation} then $ \lesssim_{C_1} |P|^{\tfrac{A}{2A-1}}|L|^{\tfrac{2A-2}{2A-1}}$ by assumption. Hence, we can combine this with (\ref{PcellLcell}) and a careful choice of $C_0, C_1$ and $C_2$ gives us (\ref{inductiveclaim}). When $|P| = 1$, (\ref{inductiveclaim}) is trivial, and we obtain (\ref{estimateepsilonleft}). For (\ref{estimateepsilonright}) the argument is similar and is omitted. Finally, (\ref{PalgL}) follows from Proposition ~\ref{inductionondim} below when $r=3$, $D=C_{1}$ and $\Sigma=\{Q=0\}$. \end{proof} \begin{prop}\label{inductionondim} Let $P$ and $L$ be as in Theorem ~\ref{complex epsilon}, $0\leq r\leq 3$, and let $\Sigma$ be a subvariety in $\mathbb{C}^2 \simeq \mathbb{R}^{4}$ of (real) dimension $\leq r$ and of degree $\leq D$. Then, \begin{equation} \mathcal{I}(P\cap \Sigma, L) \lesssim_{D} |P|^{\tfrac{A}{2A-1}}|L|^{\tfrac{2A-2}{2A-1}}+|P|+|L|. \end{equation} \end{prop} \begin{proof} When $r=0$, $\Sigma$ is a single point and the inequality trivially holds. When $r=1$, we decompose $\Sigma=\Sigma_{1}\cup\Sigma_{2}$, where every component of $\Sigma_{1}$ belongs to some curve in $L$ and $\Sigma_{2}$ do not have a common component with curves in $L$. Then, $|\mathcal{I}(P\cap \Sigma_{1})|\lesssim_{D}|P|$ and $|\mathcal{I}(P\cap \Sigma_{2})|\lesssim_{D, d}|L|$. Now we deal with the case $r=2$. By an algebraic geometry result (see, for example, Corollary 4.5 in \cite{solymosi2012incidence}), one can decompose $\Sigma$ into smooth points on subvarieties \[\Sigma = \Sigma^{smooth}\cup\Sigma_{i}^{smooth},\] where $\Sigma_{i}$'s are subvarieties of $\Sigma$ of dimension $\leq 1$ and of degree $\OO_{D}(1)$. The number of $\Sigma_{i}$'s is at most $\OO_{D}(1)$. It suffices to bound $\mathcal{I}(P\cap\Sigma^{smooth}, L)$. If $l_{1}, l_{2} \in \Sigma$ intersect at $p\in \Sigma^{smooth}$, by considering the tangent space and the transverse assumption, we know that $p$ is a singular point of $l_{1}$ or $l_{2}$. Since each curve has $\OO_d (1)$ singular points, we obtain \begin{equation}\label{SigmaLalg} \mathcal{I}(P\cap\Sigma^{smooth}, L_{alg}) \leq |P| + \OO_d (1)|L|. \end{equation} It remains to estimate $\mathcal{I}(P\cap\Sigma^{smooth}, L_{cell})$. If $l$ does not belong to $\Sigma$, then by Corollary 4.5 of \cite{solymosi2012incidence} the intersection of $l$ and $\Sigma$ can be decomposed as $l\cap\Sigma=\cup_{j=0}^{J (l)}l_{j}$ for some $J(l) \leq \OO_{D}(1)$, where $l_{j}$ is an algebraic variety of dimension $\leq 1$ and of degree $\OO_D (1)$ for each $1\leq j\leq J(l)$. Let $\mathcal{I}_{l, j}$ denote the set $\{p \in P: p \in l_{j}\}$, we obtain \begin{equation} |\mathcal{I}(P\cap \Sigma^{smooth}, L_{cell})|\leq \sum_{l, j: j \leq J(l)} |\mathcal{I}_{l, j}|. \end{equation} If $l_{j}$ is not the union of $\OO_D (1)$ points, then $l_{j}$ belongs to a unique $l$ because distinct curves in $L$ do not share a common component. By taking a generic projection from $\mathbb{R}^4$ to $\mathbb{R}^2$, we can apply arguments in the proof of Theorem ~\ref{main theorem}: use the initial bound given by $L$ and $P$, then apply the polynomial method. When $r=3$, we can repeat the proof of $r=2$ assuming that the bound holds for $r\leq 2$. \end{proof} We also have the following corollary for complex curves parametrized by an algebraic variety: \begin{cor}\label{complex variety} Given a finite point set $P\in \mathbb{C}^{2}$, an integer $d\geq 1$, $A={d+2\choose 2} -1$ and a subset $\mathcal{M}\in S_d$ parametrized by an algebraic variety of dimension $\leq k$. Let $L$ be a finite subset of $\mathcal{M}$ such that any two distinct curves of $L$ do not share a common component and intersect transversally at smooth points. Then, for any sufficiently small $\epsilon>0$, \begin{equation}\label{estimatecomplexepsilonleft} |\mathcal{I}(P,L)|\lesssim_{\epsilon, \mathcal{M}} |P|^{\tfrac{k}{2k-1}+\epsilon}|L|^{\tfrac{2k-2}{2k-1}}+|P|+|L| \end{equation} and \begin{equation}\label{estimatecomplexepsilonright} |\mathcal{I}(P,L)|\lesssim_{\epsilon, \mathcal{M}} |P|^{\tfrac{k}{2k-1}}|L|^{\tfrac{2k-2}{2k-1}+\epsilon}+|P|+|L|. \end{equation} \end{cor} \bibliographystyle{amsalpha}
{ "timestamp": "2015-03-31T02:19:36", "yymm": "1308", "arxiv_id": "1308.0861", "language": "en", "url": "https://arxiv.org/abs/1308.0861", "abstract": "We prove new bounds on the number of incidences between points and higher degree algebraic curves. The key ingredient is an improved initial bound, which is valid for all fields. Then we apply the polynomial method to obtain global bounds on $\\mathbb{R}$ and $\\mathbb{C}$.", "subjects": "Combinatorics (math.CO)", "title": "Bounds of incidences between points and algebraic curves", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9822877023336243, "lm_q2_score": 0.8152324938410784, "lm_q1q2_score": 0.8007928532428634 }
https://arxiv.org/abs/1504.03029
The covering radius of randomly distributed points on a manifold
We derive fundamental asymptotic results for the expected covering radius $\rho(X_N)$ for $N$ points that are randomly and independently distributed with respect to surface measure on a sphere as well as on a class of smooth manifolds. For the unit sphere $\mathbb{S}^d \subset \mathbb{R}^{d+1}$, we obtain the precise asymptotic that $\mathbb{E}\rho(X_N)[N/\log N]^{1/d}$ has limit $[(d+1)\upsilon_{d+1}/\upsilon_d]^{1/d}$ as $N \to \infty $, where $\upsilon_d$ is the volume of the $d$-dimensional unit ball. This proves a recent conjecture of Brauchart et al. as well as extends a result previously known only for the circle. Likewise we obtain precise asymptotics for the expected covering radius of $N$ points randomly distributed on a $d$-dimensional ball, a $d$-dimensional cube, as well as on a 3-dimensional polyhedron (where the points are independently distributed with respect to volume measure). More generally, we deduce upper and lower bounds for the expected covering radius of $N$ points that are randomly and independently distributed on a metric measure space, provided the measure satisfies certain regularity assumptions.
\section{Introduction and Notation} The purpose of this paper is to obtain asymptotic results for the expected value of the covering radius of $N$ points $X_N=\{x_1, x_2,\ldots, x_N\}$ that are randomly and independently distributed with respect to a given measure $\mu$ over a metric space $(\mathcal{X}, m)$. By the \emph{covering radius} $\rho(X_N, \mathcal{X})$ (also known as the \emph{mesh norm }or \emph{fill radius}) of the set $X_N$ with respect to $\mathcal{X}$, we mean the radius of the largest neighborhood centered at a point of $\mathcal{X}$ that contains no points of $X_N$; more precisely, $$ \rho(X_N, \mathcal{X}):=\sup_{y\in \mathcal{X}}\inf_j m(y, x_j). $$ Our focus is on the limiting behavior as $N \to \infty$ of the expected value $\mathbb{E}\rho(X_N, \mathcal{X}).$ \ The covering radius of a discrete point set is an important characteristic that arises in a variety of contexts. For example, it plays an essential role in determining the accuracy of various numerical approximation schemes such as those involving radial basis techniques (see, e.g. \cite{FW}, \cite{MNPW}). Another area where the covering radius arises is in ``1-bit sensing'', i.e., the problem of approximating an unknown vector (signal) $x\in K$ from knowledge of $m$ numbers $\textup{sign}\langle x, \theta_j \rangle$, $j=1,\ldots, m$, where the vectors $\theta_j$ are selected independently and randomly on a sphere; see discussion after Corollary \ref{nets} for details. With regard to asymptotics for the expected value of the covering radius, of particular interest is the case where $\mathcal{X}$ is the unit sphere $\mathbb{S}^d$ in $\mathbb{R}^{d+1}$ and the metric is Euclidean distance in $\mathbb{R}^{d+1}$. In \cite{BSR}, Bourgain, Sarnak and Rudnick study local statistics of certain spherical point configurations derived from normalized sums of squares of integers. Their investigation focuses on whether such configurations exhibit features of randomness, and for this purpose they study various local statistics, including the covering radius of random points on $\mathbb{S}^d$. They prove that this radius is bounded from above by $N^{-1/d +o(1)}$ as $N \to \infty.$\ For $d=1$, i.e. the unit circle, it is shown in \cite {DN} by using order statistics, that for $N$ points independently and randomly distributed with respect to arclength on the circle, $$ \lim_{N \to \infty}\mathbb{E}\rho(X_N,\mathbb{S}^1)\left[ \frac{N}{\log N}\right] =\pi. $$ Up to now, there has been no extension of this result to higher-dimensional spheres where the order statistics approach is more elusive. Based on a heuristic argument and numerical experiments, Brauchart et al. \cite{BHS} have conjectured that the appropriate extension of the circle case is the following: \begin{equation}\label{sphereequiv1} \lim_{N\to\infty} \mathbb{E}\rho(X_N, \mathbb{S}^d)\cdot \left[\frac{N}{\log N }\right]^{1/d}=\left(\frac{(d+1)\upsilon_{d+1}}{\upsilon_d}\right)^{1/d}=\left(2\sqrt{\pi}\frac{\Gamma(\frac{d+2}2)}{\Gamma(\frac{d+1}2)}\right)^{1/d} ,% \end{equation} where $\upsilon_d:=\frac{\pi^{d/2}}{\Gamma(1+d/2)}$ is the volume of a $d$-dimensional unit ball in $\mathbb{R}^d$, and the points of $X_N$ are randomly and independently distributed with respect to surface measure on $\mathbb{S}^d$ (more precisely, $d$-dimensional Hausdorff measure $\mathcal{H}_d).$ Their conjecture is also consistent with a result of H. Maehara \cite{M} who obtained probabilistic estimates for the size of random caps that cover the sphere $\mathbb{S}^2$. He showed that with asymptotic probability one, random caps with radii that are a constant factor larger than the expected radii will cover the sphere, whereas this asymptotic probability becomes zero when the random caps all have radii that are a factor smaller. However, his results fall short of providing a sharp asymptotic for the expected covering radius (in addition, his methods do not readily generalize to other smooth manifolds). As discussed in Section 3, our results for the sphere cannot be directly derived from Maehara's; however, his results are a direct consequence of our Corollary \ref{corsphere}. The main goal of this article is to provide a proof of \eqref{sphereequiv1} and its various generalizations. We remark that for any compact metric space $(\mathcal{X}, m)$ with $\mathcal{X}$ having finite $d$-dimensional Hausdorff measure, there exists a positive constant $C$ such that for any $Y_N=\{y_1,\ldots, y_N\} \subset \mathcal{X}$, there holds \begin{equation}\label{lowerbound} \rho_N:=\rho(Y_N,\mathcal{X}) \geq \frac{C}{N^{1/d}},\quad N \geq 1. \end{equation} Indeed, a lemma of Frostman (see, e.g. Theorem 8.17 in \cite{Mat}) implies the existence of a finite positive measure $\mu$ on $\mathcal{X}$ for which $\mu(B(x,r)) \leq (2r)^d$ for all $x \in \mathcal{X}$ and all $0<r\leq \operatorname{diam}(\mathcal{X}),$ where $B(x,r)$ denotes the ball centered at $x$ having radius $r$. Consequently, $$0<\mu(\mathcal{X}) \leq \sum_{i=1}^N \mu(B(y_i,\rho_N) )\leq N(2\rho_N)^d, $$ which verifies \eqref{lowerbound}. Thus, as also remarked in \cite{BSR} and made more explicit by \eqref{sphereequiv1}, randomly distributed points have relatively good covering properties, differing from the optimal by a factor of $(\log N)^{1/d}$.\ The outline of this paper is as follows. In Section 2, we state our probabilistic and expected covering radius estimates for general compact metric spaces, where the points are randomly distributed with respect to a measure satisfying certain regularity conditions. Results for compact subsets of Euclidean space are given in Section 3, including sharp asymptotic results for randomly distributed points with respect to Hausdorff measure on rectifiable curves, smooth surfaces, bodies with smooth boundaries, $d$-dimensional cubes, and $3$-dimensional polyhedra. The proofs of our stated results are provided in Section 5 utilizing properties established in Section 4 for a commonly arising probability function.\ We conclude this section with a listing of some notational conventions and terminology that will be utilized throughout the paper. \begin{itemize} \item We denote by $B(x,r)$ a closed ball in the metric space $(\mathcal{X}, m);$ more precisely, $B(x,r):=\{y\in \mathcal{X}\colon m(y,x)\leqslant r\}$. For $d$-dimensional balls in Euclidean space we write $B_d(x,r).$ \item For a positive finite Borel measure $\mu$ supported on a set $\mathcal{X}$, we say that a point $x$ is {\it randomly distributed over $\mathcal{X}$ with respect to $\mu$}, if it is distributed with respect to the probability measure $\mu/\mu(\mathcal{X})$; i.e., for any Borel set $K$ it holds that $\P(x\in K) = \mu(K)/\mu(\mathcal{X})$. \item For a positive integer $s\leqslant d$, we denote by $\mathcal{H}_s$ the $s$-dimensional Hausdorff measure on the Euclidean space $\mathbb{R}^d$ with the Euclidean metric, normalized by $\mathcal{H}_s([0,1]^s)=1$. Thus, $\mathcal{H}_s(E)=\frac{\pi^{s/2}}{2^s\Gamma(1+s/2)}\mathcal{H}^s(E)$, where $\mathcal{H}^s$ is the Hausdorff measure defined in \cite{Fal}. \item If $K$ is a subset of the Euclidean space $\mathbb{R}^d$, we always equip it with the Euclidean metric $m(x,y)=|x-y|.$ \item The symbols $c_1, c_2, \ldots$, and $C_1, C_2, \ldots$ shall denote positive constants that may differ from one inequality to another. These constants never depend on $N$. \end{itemize} \section{Main Theorems for Metric Spaces}\label{sectionmetric} Throughout this section, we assume that $(\mathcal{X}, m)$ is a metric space, $\mu$ is a finite positive Borel measure supported on $\mathcal{X}$, and $X_N=\{x_1, \ldots, x_N\}$ is a set of $N$ points, independently and randomly distributed over $\mathcal{X}$ with respect to $\mu$. Our theorems provide estimates for the probability and expected values of the covering radius $\rho(X_N, \mathcal{X})$ when the measure $\mu$ satisfies certain regularity conditions described by a function $\Phi$. \begin{theorem}\label{metricabove} Suppose $\Phi$ is a continuous non-negative strictly increasing function on $(0, \infty)$ satisfying $\Phi(r)\to 0$ as $r\to 0^+$. If there exists a positive number $r_0$ such that $\mu(B(x,r))\geqslant \Phi(r)$ holds for all $x\in \mathcal{X}$ and every $r<r_0$, then there exist positive constants $c_1$, $c_2$, $c_3$, and $\alpha_0$ such that for any $\alpha>\alpha_0$ we have \begin{equation}\label{metricaboveprob} \P\left[\rho(X_N, \mathcal{X})\geqslant c_1\Phi^{-1}\left(\frac{\alpha\log N}{N}\right)\right]\leqslant c_2N^{1-c_3\alpha}. \end{equation} If, in addition, $\Phi$ satisfies $\Phi(r)\leqslant r^\sigma$ for all small $r$ and some positive number $\sigma$, then there exist positive constants $c_1, c_2$ such that \begin{equation}\label{metricaboveexpect} \mathbb{E}\rho(X_N, \mathcal{X}) \leqslant c_1 \Phi^{-1}\left(c_2 \frac{\log N }{N}\right). \end{equation} \end{theorem} A lower bound for the expected covering radius is given in our next result. \begin{theorem}\label{metricbelow} Let $\Phi$ be a continuous non-negative strictly increasing function on $(0, \infty)$ satisfying $\Phi(r)\to 0$ as $r\to 0^+$ and the strict doubling property; i.e., for some constants $C_1, C_2>1$ and any small $r$ it holds that $C_1\Phi(r)\leqslant\Phi(2r)\leqslant C_2\Phi(r)$. If there exists a subset $\mathcal{X}_1\subset \mathcal{X}$ with the following two properties: \begin{enumerate}[label={\upshape(\roman*)}] \item $\mu(\mathcal{X}_1)>0$; \item there exist positive numbers $r_0$ and $c$ such that for any $x\in \mathcal{X}_1$ and every $r<r_0$ the regularity condition $c\Phi(r)\leqslant \mu(B(x,r))\leqslant \Phi(r)$ holds, \end{enumerate} then there exist positive constants $c_1$ , $c_2$, and $c_3$ such that \begin{equation}\label{metricbelowprob} \P\left[\rho(X_N, \mathcal{X})\geqslant c_1\Phi^{-1}\left(\frac{c_2\log N-c_3\log\log N}N\right)\right]=1-o(1), \; \; N\to \infty. \end{equation} Consequently, there exist positive constants $c_1$ and $c_2$ such that \begin{equation}\label{metricbelowexpect} \mathbb{E}\rho(X_N, \mathcal{X}) \geqslant c_1 \Phi^{-1}\left(c_2 \frac{\log N}{N}\right). \end{equation} \end{theorem} Combining Theorems \ref{metricabove} and \ref{metricbelow} we deduce the following. \begin{corollary}\label{twosided} Assume the function $\Phi$ is continuous non-negative, strictly increasing, strictly doubling, and that there exist positive numbers $r_0$ and $\sigma$ such that $\Phi(r)\leqslant r^\sigma$ for every $r<r_0$. If for some positive constants $c, C$, any $x\in \mathcal{X}$ and every $r<r_0$ we have \begin{equation}\label{regularity} c\Phi(r)\leqslant \mu(B(x, r))\leqslant C\Phi(r), \end{equation} then there exist positive constants $c_1, c_2, c_3, c_4$ such that for any $\varepsilon>0$ there is a number $N(\varepsilon)$ such that for any $N>N(\varepsilon)$ we have \begin{equation}\label{distribineq} \P\left[c_1\Phi^{-1}\left(c_2\frac{\log N}{N}\right ) \leqslant \rho(X_N, \mathcal{X})\leqslant c_3\Phi^{-1}\left(c_4\frac{\log N}{N}\right)\right] > 1-\varepsilon. \end{equation} Moreover, there exist positive constants $C_1, C_2, C_3, C_4$ such that \begin{equation}\label{expectineq} C_1\Phi^{-1}\left(C_2\frac{\log N}{N}\right) \leqslant \mathbb{E}\rho(X_N, \mathcal{X})\leqslant C_3\Phi^{-1}\left(C_4\frac{\log N}{N}\right). \end{equation} \end{corollary} For recent estimates similar to \eqref{distribineq} and \eqref{expectineq} for the spherical cap discrepancy of random points on the unit sphere $\mathbb{S}^2\subset \mathbb{R}^3$, see Theorems $9$ and $10$ in \cite{ABD}. An important class of sets in $\mathbb{R}^d$ to which Corollary \ref{twosided} applies are described in the following definition. \begin{defin} We call a set $\mathcal{X}\subset \mathbb{R}^d$ {\it $s$-regular} if the condition \eqref{regularity} holds for $\mu=\mathcal{H}_s$ and $\Phi(r)=r^s$; i.e., for some positive constants $r_0, c$, and $C$ there holds \begin{equation}\label{sregset} cr^s\leqslant \mathcal{H}_s(B_d(x,r)\cap \mathcal{X})\leqslant Cr^s \; \; \mbox{for any $x\in \mathcal{X}$ and every $r<r_0$}. \end{equation} \end{defin} \begin{zamech} Examples of sets in Euclidean space for which Corollary \ref{twosided} holds include a cube $[0,1]^d$, a rectifiable curve $\Gamma\subset \mathbb{R}^d$, the unit sphere $\mathbb{S}^{d-1}\subset \mathbb{R}^d$, or any $s$-regular set $\mathcal{X}\subset \mathbb{R}^d$. Furthermore, the results of the Corollary \ref{twosided} hold not only for $\Phi(r)=r^s$, but for more general regularity functions, such as $\Phi(r)=r^\alpha \log^\beta\left(1/r\right)$, with $\alpha>0$ and $\beta\geqslant 0$. In particular, Corollary \ref{twosided} applies for the ``middle $1/3$'' Cantor set $\mathcal{C}$ in $[0,1]$ with $\textup{d}\mu=\mathbbm{1}_{\mathcal{C}}\textup{d}\mathcal{H}_{\log2/\log3}$. We remark that for $\mu$-a.e. point $x\in \mathcal{C}$ we have $$ \liminf_{r\to 0^+}\frac{\mu(B_1(x,r)\cap \mathcal{C})}{r^{\log2/\log3}}\not=\limsup_{r\to 0^+}\frac{\mu(B_1(x,r)\cap \mathcal{C})}{r^{\log2/\log3}}; $$ i.e., at $\mu$-a.e. point $x$ of $\mathcal{C}$ the density of $\mu$ at $x$ does not exist, which essentially precludes obtaining a sharp asymptotic for $\mathbb{E}\rho(X_N, \mathcal{C})$ (compare with \eqref{uniform} below). However, Corollary \ref{twosided} provides the two-sided estimate $$ c_1 \left(\frac{\log N}{N}\right)^{\log 3/\log 2} \leqslant \mathbb{E}\rho(X_N, \mathcal{C}) \leqslant c_2 \left(\frac{\log N}{N}\right)^{\log 3/\log 2}. $$ \end{zamech} \begin{zamech} The condition in Theorem \ref{metricabove} that $\mu(B(x,r))\geqslant \Phi(r)$ for every $x\in \mathcal{X}$ is essential. Indeed, if we consider the set $\mathcal{X}=[0,1]\cup \{2\}$ with $\mu$ Lebesgue measure, then $\mu(B_1(x,r))\geqslant r$ for $x\in \mathcal{X}\setminus \{2\}$. However, we have $\P\left[\rho(X_N, \mathcal{X})\geqslant 1\right]=1$, and so $\mathbb{E}\rho(X_N, \mathcal{X})\geqslant 1$. The reason that inequality \eqref{metricaboveexpect} fails in this case is that for the point $x=2$ we have $\mu(B_1(x,r))=0$ for small values of $r$. However, Theorem \ref{metricabove} does apply if $\mu=m_{[0,1]}+\alpha\delta_2$, where $m_{[0,1]}$ is Lebesgue measure on $[0,1]$, $\delta_2$ is the unit point mass at $x=2$, and $\alpha>0$. In this case we get $$ \mathbb{E}\rho(X_N, \mathcal{X})\leqslant {C(\alpha)}\cdot \frac{\log N}N. $$ In fact, repeating the proofs from Sections \ref{abovvvve} and \ref{sectionfrombelow} (with $K_1=[0,1]$), we obtain $$ \lim_{N\to \infty} \mathbb{E}\rho(X_N, \mathcal{X}) \cdot \frac{N}{\log N} = \frac{1+\alpha}{2} \; \; \mbox{for any $\alpha>0$}. $$ \end{zamech} The above results have immediate consequences for $\varepsilon$-nets. Since different definitions of an ``$\varepsilon$-net'' occur in the literature, the terminologies that we use are made precise in what follows. \begin{defin} A subset $A$ of a metric space $(\mathcal{X}, m)$ is called an {\it $\varepsilon$-net} (or {\it $\varepsilon$-covering}) if, for any point $y\in \mathcal{X}$, there exists a point $x\in A$ such that $m(x,y)\leqslant \varepsilon$. Equivalently, $A$ is an $\varepsilon$-net if $\rho(A, \mathcal{X})\leqslant \varepsilon$. \end{defin} \begin{defin} A subset $A$ of a metric space $(\mathcal{X}, m)$ with a positive Borel measure $\mu$ is called a {\it measure $\varepsilon$-net} if any ball $B(y,r)$ with $\mu(B(y,r))\geqslant \varepsilon$ intersects $A$. \end{defin} We remind the reader that on $\mathbb{S}^d$ with $\mu$ surface area measure $\mathcal{H}_d$, the minimal $\varepsilon$-net has cardinality $c\varepsilon^{-d}$ (for the proof see, for example, Lemma $5.2$ in \cite{V}), while the minimal measure $\varepsilon$-net has cardinality $c\varepsilon^{-1}$. \begin{corollary}\label{nets} If $\Phi$ and $\mu$ are as in the first part of Theorem \ref{metricabove}, then there exists a positive constant $c_1$ such that for any number $\alpha$ there is a positive constant $C_\alpha$ for whic $$ \P\left[X_N \; \mbox{is an $\varepsilon$-net}\right] \geqslant 1-N^{-\alpha}, \; \; \; \mbox{for} \; \varepsilon=c_1\Phi^{-1}\left(C_\alpha\frac{\log N}N\right). $$ Furthermore, if the function $\Phi$ is doubling, and the measure $\mu$ satisfies the condition \eqref{regularity}, then for any positive number $\alpha$ there exists a positive constant $C_\alpha$ such that $$ \P\left[X_N \; \mbox{is a measure $\varepsilon$-net}\right] \geqslant 1-N^{-\alpha}, \; \; \; \mbox{for} \; \varepsilon=C_\alpha\frac{\log N}N. $$ \end{corollary} By way of illustration, suppose for simplicity that $\Phi(r)=Cr^d$ for some positive constant $C$ and $\varepsilon=\left[(\log N)/N\right]^{1/d}$, which implies that $N$ is of the order $\varepsilon^{-d}\log(1/\varepsilon)$. Then, from the first part of Corollary \ref{nets}, if we take $C_1\varepsilon^{-d}\log\left(1/\varepsilon\right)$ random points, we get an $\varepsilon$-net ($\varepsilon$-covering) with high probability. The cardinality of an $\varepsilon$-covering of a set $K\subset \mathbb{S}^d$ plays an important role in ``1-bit compressed sensing''. The estimates for the number $m$ of random vectors $\{\theta_j\}_{j=1}^m$, essential to approximate an unknown signal $x\in K$ from knowledge of $m$ ``bits'' $\textup{sign}\langle x,\theta_j\rangle$ involve finding an $\varepsilon$-covering of the set $K$ with $\log(N(K, \varepsilon))\leqslant C\varepsilon^{-2}w(K)$, where $N(K, \varepsilon)$ is the cardinality of the covering, and $w$ is the so-called ``mean width'' of $K$. As can be seen from our results, for many sets $K$ a random set of $C\varepsilon^{-d}\log(1/\varepsilon)$ points satisfies this condition with high probability. For further discussion, see \cite{PV1}, \cite{PV2}. \section{Expected Covering Radii for Subsets of Euclidean Space}\label{sectioneuclid} In some cases we are able to ``glue'' upper and lower estimates together to obtain sharp asymptotic results. For this purpose we state the following definitions. \begin{defin}\label{flat} Let $s$ be a positive integer, $s\leqslant d$. Suppose $K$ is a compact $s$-dimensional set in $\mathbb{R}^d$ with the Euclidean metric. We call $K$ an {\it asymptotically flat $s$-regular} set if for any $x\in K$ it holds that \begin{equation}\label{uniform} r^{-s}\mathcal{H}_s(B_d(x,r)\cap K)\rightrightarrows \upsilon_s \;\; \mbox{as $r\to 0^+$}, \end{equation} where the convergence is uniform in $x$, and $\upsilon_s$ is the volume of the $s$-dimensional unit ball $B_s(0,1)$. We call $K$ a {\it quasi-nice $s$-regular} set if \begin{enumerate} [(i)] \item $K$ is countably $s$-rectifiable; i.e., $K$ is of the form $\bigcup_{j=1}^\infty f_j(E_j) \cup G$, where $\mathcal{H}_s(G)=0$ and where each $f_j$ is a Lipschitz function from a bounded subset $E_j$ of $\mathbb{R}^s$ to $\mathbb{R}^d$; \item There exist positive numbers $c, C, r_0$ such that for any $x\in K$ and any $r<r_0$ the $s$-regularity condition holds: $c r^s\leqslant \mathcal{H}_s(B_d(x,r)\cap K)\leqslant C r^s$; \item There is a finite set $T\subset K$ such that for any $r<r_0$ and $y\in K\setminus \bigcup_{x_t\in T}B_d(x_t, r)$ it holds that $\mathcal{H}_s(B_d(y,r)\cap K)\geqslant \upsilon_s r^s$. \end{enumerate} \end{defin} We remark that the appearance of the constant $\upsilon_s$ in the above definitions is quite natural. Indeed, if $K$ is a countably $s$-rectifiable compact set and $0<\mathcal{H}_s(K)<\infty$, then for $\mathcal{H}_s$-almost every point $x\in K$ the following holds: $r^{-s}\mathcal{H}_s(B_d(x,r)\cap K)\to \upsilon_s$ as $r\to 0^+$. For the details see the Theorem 17.6 in \cite{Mat} or Theorem 3.33 in \cite{Fal}. Thus, if any uniform limit in \eqref{uniform} exists, then it must equal $\upsilon_s$. For asymptotically flat $s$-regular and quasi-nice $s$-regular sets we deduce the following precise asymptotics for the expected covering radius as well as its moments. \begin{theorem}\label{manifolds} Suppose $K\subset \mathbb{R}^d$ is an asymptotically flat $s$-regular or a quasi-nice $s$-regular set for integer $s\leqslant d$. Then for $X_N=\{x_1, \ldots, x_N\}$ a set of $N$ independently and randomly distributed points over $K$ with respect to the measure $\textup{d}\mu:=\mathbbm{1}_K\cdot \textup{d}\mathcal{H}_s / \mathcal{H}_s(K)$, and any $p \geq 1,$ \begin{equation}\label{asymp} \lim_{N\to \infty} \mathbb{E}[\rho(X_N, K)^p]\cdot \left[\frac{N}{\log N}\right]^{p/s}=\left(\frac{\mathcal{H}_s(K)}{\upsilon_s}\right)^{p/s}. \end{equation} \end{theorem} Important examples of asymptotically flat $s$-regular sets are given in the following result, which includes the verification of the conjecture of Brauchart et al. in \cite{BDSSWW} for the expected covering radius of randomly distributed points on the unit sphere. \begin{corollary}\label{corsphere} Suppose $K$ is a closed $C^{(1,1)}$ $s$-dimensional embedded submanifold of $\mathbb{R}^d$; i.e., $0<\mathcal{H}_s(K)<\infty$ and, for any embedding $\varphi$, all its first partial derivatives exist and are uniformly Lipschitz. Then $K$ is an asymptotically flat $s$-regular manifold, and thus for $N$ points independently and randomly distributed over $K$ with respect to $\textup{d}\mu=\mathbbm{1}_K \cdot \textup{d}\mathcal{H}_s/\mathcal{H}_s(K)$, equation \eqref{asymp} holds. In particular, if $K=\mathbb{S}^d$ is a unit sphere in $\mathbb{R}^{d+1}$ and $p \geq 1$, then \begin{equation}\label{sphereequiv} \lim_{N\to\infty} \mathbb{E}[\rho(X_N, \mathbb{S}^d)^p]\cdot \left[\frac{N}{\log N }\right]^{p/d}=\left(\frac{(d+1)\upsilon_{d+1}}{\upsilon_d}\right)^{p/d}=\left(2\sqrt{\pi}\frac{\Gamma(\frac{d+2}2)}{\Gamma(\frac{d+1}2)}\right)^{p/d} . % \end{equation} Thus \eqref{sphereequiv1} holds. \end{corollary} As a consequence of the corollary, we shall deduce in Section 5 the result of Maehara mentioned in the Introduction. \begin{corollary}[Maehara \cite{M}]\label{maehara} Suppose $X_N=\{x_1, \ldots, x_N\}$ is a set of $N$ points, independently and randomly distributed over the unit sphere $\mathbb{S}^d$ with respect to $\textup{d}\mu=\mathbbm{1}_{\mathbb{S}^d} \cdot \textup{d}\mathcal{H}_d/\mathcal{H}_s(\mathbb{S}^d))$ and set $$ Z_N:=\rho(X_N, \mathbb{S}^d)\cdot \left(\frac{\upsilon_d}{(d+1)\upsilon_{d+1}}\cdot \frac{N}{\log N}\right)^{1/d}. $$ Then $Z_N$ converges in probability to 1 as $N \to \infty$; i.e., for each $\epsilon > 0,$ \begin{equation}\label{Mae} \lim_{N \to \infty} \mathbb{P}(|Z_N-1|\geq\epsilon)=0. \end{equation} \end{corollary} \begin{zamech} We remark that our results for $\mathbb{S}^d$ do not directly follow from \eqref{Mae}. Maehara's result implies that the bounded sequence $$ p_N(t):=\P(Z_N\geqslant t)\to \mathbbm{1}_{[0,1]}(t) \; \; \mbox{for a.e. $t>0$}; $$ however, since the range of $t$ is $[0, \infty)$, the constant function $1$ is not integrable, and we cannot apply the Lebesgue dominated convergence theorem to get $\mathbb{E} Z_N = \int_0^{\infty} p_N(t)dt \to 1$. \end{zamech} The next corollary gives an example of a quasi-nice $1$-regular set. \begin{corollary}\label{corcurve} Suppose $\gamma$ is a rectifiable curve in $\mathbb{R}^d$(i.e., $0<\mathcal{H}_1(\gamma)<\infty$ and $\gamma$ is a continuous injection of a closed interval of $\mathbb{R}$). If $X_N$ denotes a set of $N$ points independently and randomly distributed over $\gamma$ with respect to $\textup{d}\mu:=\mathbbm{1}_\gamma\cdot \textup{d}\mathcal{H}_1/\mathcal{H}_1(\gamma)$, then $\gamma$ is a quasi-nice $1$-regular set, and for any $p\geqslant 1$ \begin{equation}\label{curveequiv} \lim_{N\to \infty} \mathbb{E}[\rho(X_N, \gamma)^p]\cdot \left[\frac{N}{\log N}\right]^p=\left(\frac{\mathcal{H}_1(\gamma)}{2}\right)^p. \end{equation} \end{corollary} Next we deal with the following problem: suppose $A\subset \mathbb{R}^d$ is a $d$-dimensional set, but the condition $$ \mathcal{H}_d(A\cap B_d(x,r))\geqslant \upsilon_d r^{d} $$ fails for a certain number of points $x\in A$ and the limit \eqref{uniform} in the Definition \ref{flat} is not uniform. Such situations arise for sets with boundary, which include the unit ball $B_d(0,1)$ and the unit cube $[0,1]^d$. The case of the ball is included in the next theorem, while the case of the cube is studied in the Theorem \ref{cube}. \begin{theorem}\label{theoremball} Let $d\geqslant 2$ and $K\subset \mathbb{R}^d$ a set that satisfies the following conditions. \begin{enumerate}[label={\upshape(\roman*)}] \item $K$ is compact and $0<\mathcal{H}_d(K)<\infty$; \item $K=\textup{clos}(K_0)$, where $K_0$ is an open set in $\mathbb{R}^d$ with $\partial K_0 = \partial K$; \item The boundary $\partial K$ of $K$ is a $C^2$ smooth $(d-1)$-dimensional embedded submanifold of $\mathbb{R}^d$. \end{enumerate} Let $X_N=\{x_1, \ldots, x_N\}$ be a set of $N$ points, independently and randomly distributed over $K$ with respect to $\textup{d}\mu=\mathbbm{1}_K \cdot \textup{d}\mathcal{H}_d / \mathcal{H}_d(K)$. Then for any $p\geqslant 1$ \begin{equation}\label{smoothequiv} \lim_{N\to \infty}\mathbb{E}[\rho(X_N, K)^p]\cdot \left[\frac{N}{\log N}\right]^{p/d}=\left(\frac{2(d-1)}d\cdot \frac{\mathcal{H}_d(K)}{\upsilon_d}\right)^{p/d}. \end{equation} In particular, for the unit ball, \begin{equation}\label{ballequiv} \lim_{N\to \infty}\mathbb{E}[\rho(X_N, B_d(0,1))^p]\cdot \left[\frac{N}{\log N}\right]^{p/d}=\left(\frac{2(d-1)}d\right)^{p/d}. \end{equation} \end{theorem} \begin{zamech} We see that in the case $d=2$ we have $2(d-1)/d=1$, and so the constant on the right-hand side of \eqref{smoothequiv} coincides with the constant for smooth closed manifolds, see \eqref{asymp}. However, when $d>2$ we have $2(d-1)/d>1$; thus this constant becomes bigger than for smooth closed manifolds. \end{zamech} The next two propositions deal with cases when the boundary of the set is not smooth. For simplicity, we formulate them for a cube $[0,1]^d$ and a polyhedron in $\mathbb{R}^3$. However, the proof can be applied to other examples, such as cylinders. \begin{prop}\label{cube} Suppose $d\geqslant 2$ and $[0,1]^d$ is the $d$-dimensional unit cube. Let $\textup{d}\mu=\mathbbm{1}_{[0,1]^d}\cdot \textup{d}\mathcal{H}_d$. If $X_N=\{x_1, \ldots, x_N\}$ is a set of $N$ points, independently and randomly distributed over $[0,1]^d$ with respect to $\mu$, then for any $p\geqslant 1$ \begin{equation}\label{cubeequiv} \lim_{N \to \infty} \mathbb{E}[\rho(X_N, [0,1]^d)^p]\cdot \left[\frac{N}{\log N }\right]^{p/d}=\left(\frac{2^{d-1}}{d\upsilon_d}\right)^{p/d}. \end{equation} \end{prop} \begin{prop Suppose $P$ is a polyhedron in $\mathbb{R}^3$ of volume $V(P)$. Let $X_N=\{x_1, \ldots, x_N\}$ be a set of $N$ points, independently and randomly distributed over $P$ with respect to $\textup{d}\mu=\mathbbm{1}_P\cdot \textup{d}\mathcal{H}_3/V(P)$. If $\theta$ is the smallest angle at which two faces of $P$ intersect, then for any $p\geqslant 1$ \begin{equation}\label{polyhequiv1} \lim_{N \to \infty} \mathbb{E}[\rho(X_N, P)^p]\cdot \left[\frac{N}{\log N}\right]^{p/3}=\left(\frac{2\pi V(P)}{3\theta \upsilon_3}\right)^{p/3} = \left(\frac{V(P)}{2\theta}\right)^{p/3}, \;\; \mbox{if $\theta\leqslant \frac{\pi}{2}$}; \end{equation} \begin{equation}\label{polyhequiv2} \lim_{N \to \infty} \mathbb{E}[\rho(X_N, P)^p]\cdot \left[\frac{N}{\log N}\right]^{p/3} = \left(\frac{V(P)}{\pi}\right)^{p/3}, \;\; \mbox{if $\theta\geqslant \frac{\pi}{2}$}. \end{equation} \end{prop} In the theorems up to now we dealt with measures $\mu$ on sets $\mathcal{X}$ satisfying for all $x\in \mathcal{X}$ the condition $cr^s\leqslant \mu(B(x,r)\cap \mathcal{X})\leqslant Cr^s$ (i.e., the regularity function $\Phi$ was the same for all points of $\mathcal{X}$); only the values of best constants $c, C$ differed for points $x$ deep inside $\mathcal{X}$ from those near the boundary. We now give an example of a measure for which the regularity function parameter $s$ depends upon the distance to the boundary. \begin{prop Consider the interval $[-1,1]$ and the measure $\textup{d}\mu=\frac{dx}{\pi\sqrt{1-x^2}}$. Let $X_N=\{x_1, \ldots, x_N\}$ be a set of $N$ points, independently and randomly distributed over $[-1,1]$ with respect to $\mu$. Define $$\hat{\rho}(X_N, [0,1]):=\sup_{y\in [1-\frac{1}{N^a}, 1]} \inf_j |y-x_j|, \;\;\;\;\; \tilde{\rho}(X_N, [0,1]):=\sup_{y\in [-1+\frac{1}{N^a}, 1-\frac{1}{N^a}]} \inf_j |y-x_j|. $$ \begin{enumerate}[label={\upshape(\roman*)}] \item If $a=2$, then there exist positive constants $c_1$ and $c_2$ such that \begin{equation}\label{arcsinineq1} \frac{c_1}{N^2}\leqslant \mathbb{E}\hat{\rho}(X_N, [0,1]) \leqslant \frac{c_2}{N^2}. \end{equation} \item If $0<a<2$, then there exist positive constants $c_1$ and $c_2$ such that \begin{equation}\label{arcsinineq2} \frac{c_1\log N}{N^{1+\frac{a}2}}\leqslant \mathbb{E}\hat{\rho}(X_N, [0,1]) \leqslant \frac{c_2\log N}{N^{1+\frac{a}2}}. \end{equation} \item For any $a>0$ there exist positive constants $c_1$ and $c_2$ such that \begin{equation}\label{arcsinineq3} \frac{c_1\log N}{N}\leqslant \mathbb{E}\tilde{\rho}(X_N, [0,1]) \leqslant \frac{c_2\log N}{N}. \end{equation} \end{enumerate} \end{prop} Observe that if we stay away from the endpoints $\pm1$, the measure $\mu$ acts as the Lebesgue measure, and thus the order of the expectation of the covering radius is $(\log N)/N$. However, when we are close to the points $\pm 1$ (where ``close'' depends on $N$), the measure $\mu$ acts somewhat like the Hausdorff measure $\mathcal{H}_{1/2}$, and we get a different order for the covering radius. \section{An auxiliary function}\label{auxproofs} The proofs of the results stated in Sections \ref{sectionmetric} and \ref{sectioneuclid} rely heavily on the properties of the following function. For three positive numbers $N, n, m$, with $m$ and $N$ being integers and $m\leqslant n\leqslant N$, set \begin{equation}\label{functionf} f(N, n, m):=\sum\limits_{k=1}^m (-1)^{k+1}\binom{m}{k} \left(1-\frac{k}{n}\right)^N. \end{equation} The useful fact about the function $f(N, n,m)$ is the following. \begin{lemma} Suppose $X_N=\{x_1, \ldots, x_N\}$ is a set of $N$ points independently and randomly distributed on a set $\mathcal{X}$ with respect to a Borel probability measure $\mu$. Let $B_1, \ldots, B_m$ be disjoint subsets of $\mathcal{X}$ each of $\mu$-measure $1/n$. Then \begin{equation}\label{oneisempty} \P\big(\exists k \colon B_k\cap X_N=\emptyset\big)=f(N, n, m). \end{equation} \end{lemma} \begin{proof} We use well-known formula that, for any $m$ events $A_1, \ldots, A_m$, \begin{equation}\label{probunion} \P\left(\bigcup_{k=1}^m A_j\right) = \sum\limits_{k=1}^m (-1)^{k+1} \sum\limits_{(j_1, \ldots, j_k)} \P(A_{j_1}\cap A_{j_2}\cap \cdots \cap A_{j_k}), \end{equation} where the integers $j_1, \ldots, j_k$ are distinct. Let the event $A_i$ occur if the set $B_i$ does not intersect $X_N$. Then for any $k$-tuple $(j_1, \ldots, j_k)$ the event $A_{j_1}\cap \cdots \cap A_{j_k}$ occurs if the points $x_1, \ldots, x_N$ are in the complement of the union $B_{j_1}\cup \cdots \cup B_{j_k}$; i.e., $x_1, \ldots, x_N$ are in a set of measure $1-k/n$. We see that for any $k$-tuple the probability of this event is equal to $\left(1-k/n \right)^N$. Moreover, there are exactly $\binom{m}{k}$ such $k$-tuples. Therefore, $$ \sum\limits_{(j_1, \ldots, j_k)} \P(A_{j_1}\cap \cdots \cap A_{j_k}) =\binom{m}{k} \left(1-\frac{k}n\right)^N, $$ and \eqref{oneisempty} follows from \eqref{probunion}. \end{proof} For the lower bounds in Theorems \ref{metricbelow} and \ref{manifolds} we will need the following estimate on the function $f(N,n,m)$. \begin{lemma} For any three numbers $0<m\leqslant n\leqslant N$, such that $m$ and $N$ are integers, \begin{equation}\label{estimateoff} f(N, n,m)\geqslant 1-\left[1-\left(1-\frac{1}n\right)^N\right]^m-\frac{N}{n^2}\cdot \frac{m(m-1)}2 \cdot \left(1-\frac1n\right)^{2(N-1)}\cdot \left[1+\left(1-\frac1n\right)^{N-1}\right]^{m-2}. \end{equation} \end{lemma} \begin{proof} Notice first that for $k\geqslant 1$ and $0\leqslant x \leqslant 1$ we have $$ 1-kx\leqslant (1-x)^k \leqslant 1-kx + \frac{k(k-1)}2 x^2. $$ Thus, for $x=1/n$, we get $$ \left(1-\frac1n\right)^k-\frac{k(k-1)}2\frac{1}{n^2} \leqslant 1-\frac{k}{n}\leqslant \left(1-\frac1n\right)^{k} $$ Suppose $(1-\frac1n)^k\geqslant \frac{k(k-1)}2\frac{1}{n^2}$. Using the inequality $$ a^N - (a-b)^N =b \cdot (a^{N-1}+(a-b)a^{N-2}+\cdots + (a-b)^{N-1})\leqslant N\cdot b \cdot a^{N-1}, \; \mbox{if $a>b>0$}, $$ we get \begin{align}\label{brekekek} \left(1-\frac kn\right)^N& \geqslant \left(\left(1-\frac1n\right)^k-\frac{k(k-1)}2\frac{1}{n^2}\right)^N \\ & \geqslant \left(1-\frac1n\right)^{kN }- N\cdot \frac{k(k-1)}2\frac{1}{n^2} \cdot \left(1-\frac{1}n\right)^{k(N-1)}. \notag \end{align} Suppose now that $(1-\frac1n)^k< \frac{k(k-1)}2\frac{1}{n^2}$. Then $$ \left(1-\frac1n\right)^{kN }- N\cdot \frac{k(k-1)}2\frac{1}{n^2} \cdot \left(1-\frac{1}n\right)^{k(N-1)} = \left(1-\frac1n\right)^{k(N-1)}\left(\left(1-\frac1n\right)^k - N\frac{k(k-1)}2\frac{1}{n^2}\right)<0, $$ so as in inequality \eqref{brekekek} for $k\leqslant n$, $$ \left(1-\frac kn\right)^N\geqslant \left(1-\frac1n\right)^{kN }- N\cdot \frac{k(k-1)}2\frac{1}{n^2} \cdot \left(1-\frac{1}n\right)^{k(N-1)} $$ also holds. Therefore, \begin{multline* f(N,n,m)=\sum\limits_{\text{$k$ odd, $k\leqslant m$}} \binom{m}{k} \left(1-\frac{k}n\right)^N - \sum\limits_{\text{$k$ even, $k\leqslant m$}} \binom{m}{k}\left(1-\frac{k}n\right)^N \geqslant \\ \sum\limits_{\text{$k$ odd}} \binom{m}{k}\left[\left(1-\frac1n\right)^{kN }- N\cdot \frac{k\left(k-1\right)}2\frac{1}{n^2} \cdot \left(1-\frac{1}n\right)^{k\left(N-1\right)}\right] - \sum\limits_{\text{$k$ even}} \binom{m}{k}\left(1-\frac1n\right)^{kN } \geqslant \end{multline*} \begin{equation}\label{koaks1} \sum\limits_{k=1}^m \left(-1\right)^{k+1}\binom{m}{k}\left(1-\frac1n\right)^{kN } - \frac{N}{n^2}\sum\limits_{k=0}^m \binom{m}{k}\frac{k\left(k-1\right)}2 \cdot \left(1-\frac{1}n\right)^{k\left(N-1\right)}. \end{equation} The first sum in \eqref{koaks1} is equal to $1-(1-(1-\frac1n)^N)^m$. To calculate the second sum we notice that $$ \frac{m(m-1)}{2}x^2 (1+x)^{m-2}=\frac{1}{2}x^2 ((1+x)^m)'' = \sum\limits_{k=0}^m \binom{m}k \cdot \frac{k(k-1)}2 x^{k}. $$ Thus, for $x=(1-\frac1n)^{N-1}$ we get $$ \sum\limits_{k=0}^m \binom{m}k \frac{k(k-1)}2 \cdot \left(1-\frac{1}n\right)^{k(N-1)} = \frac{m(m-1)}2 \left(1-\frac1n\right)^{2(N-1)}\cdot \left(1+\left(1-\frac1n\right)^{N-1}\right)^{m-2}. $$ Combining the above estimates we obtain \eqref{estimateoff}. \end{proof} With the help of \eqref{estimateoff} we can deduce some asymptotic properties of $f(N,n,m)$ as $N\to \infty$. \begin{lemma}\label{lemmaforf} Let $N$ be a positive integer and $n,m$ be numbers satisfying $1\leqslant m\leqslant n\leqslant N$. Further, let $\kappa_n$ denote constants depending on $n$ such that $0<c_1\leqslant\kappa_n\leqslant c_2$ for all $n$. \begin{enumerate}[label={\upshape(\roman*)}] \item If $m=\left\lfloor \kappa_n n \right\rfloor$ and $c_2\leqslant 1$, then there exists a number $\alpha$ such that for $n= \frac{N}{\log N-\alpha \log\log N}$ we have $f(N,n,m)\to 1$ as $N\to \infty$. \item If $d>1$ and $m=\left\lfloor \kappa_n n^{\frac{d-1}{d}}\right\rfloor$, then there exists a number $\alpha$ such that for $n= \frac{N}{\frac{d-1}d\log N-\alpha\log\log N}$ we have $f(N, n, m)\to 1$ as $N\to \infty$. \item If $d>1$ and $m=\left\lfloor \kappa_n n^{\frac{1}{d}}\right\rfloor$, then there exists a number $\alpha$ such that for $n= \frac{N}{\frac{1}d\log N-\alpha\log\log N}$ we have $f(N, n, m)\to 1$ as $N\to \infty$. \end{enumerate} \end{lemma} \begin{proof} We prove only part (i) since the proofs of the second and third parts are similar. In what follows, to simplify the displays, we omit the symbol for the integer part. If $a_N$ and $b_N$ are two sequences of positive numbers, we write $a_N\sim b_N$ to mean $a_N/b_N\to 1$ as $N\to \infty$. For our choice of $n$ in part (i) we have $$ \left(1-\frac{1}{n}\right)^N \sim \exp\left(-\frac{N}n\right) \sim \frac{(\log N)^\alpha}{N}. $$ Thus, $$ \left(1-\left(1-\frac{1}{n}\right)^N\right)^{\kappa_n n} \sim \left(1-\frac{(\log N)^\alpha}{N}\right)^{\frac{\kappa_n N}{\log N-\alpha\log\log N}}\sim \exp\left(-\frac{\kappa_n (\log N)^\alpha}{\log N-\alpha\log\log N}\right). $$ If $\alpha>1$, then the last expression tends to zero. Moreover, \begin{align*} \frac{N}{n^2}\cdot \frac{m(m-1)}2 &\left(1-\frac1n\right)^{2(N-1)}\cdot \left(1+\left(1-\frac1n\right)^{N-1}\right)^{m-2} \\ &\sim \frac{\kappa_n^2}2\cdot\frac{(\log N)^{2\alpha}}{N}\cdot \left(1+\frac{(\log N)^\alpha}{N}\right)^{\frac{\kappa_n N}{\log N-\alpha\log\log N}} \\ &\sim \frac{\kappa_n^2}2\cdot\frac{(\log N)^{2\alpha}}{N}\cdot \exp\left(\kappa_n(\log N)^{\alpha-1}\right). \end{align*} For $\alpha=3/2$ (actually, any $0<\alpha<2$ will work) the last expression is comparable to $$ \frac{(\log N)^3}{N}\exp(\kappa_n(\log N)^{\frac12}), $$ which tends to zero as $N$ tends to infinity. Thus from \eqref{estimateoff} we deduce that $\liminf_{N\to \infty} f(N,n,m) \geqslant~ 1$. However, since $f(N,n,m)$ is equal to a certain probability, we have that $f(N, n, m)\leqslant ~1$, and so $\lim_{N\to \infty}f(N,n,m)=1$. \end{proof} \section{Proofs}\label{proofs} \subsection{Preliminary objects}\label{sectionprelim} Fix a compact set $\mathcal{X}_0$ with a metric $m$. For any large positive number $n$ let $\mathcal{E}_n(\mathcal{X}_0)$ be a maximal set of points such that for any $y,z\in \mathcal{E}_n$ we have $m(y,z)\geqslant 1/n$. Then for any $x\in \mathcal{X}_0$ there exists a point $y\in \mathcal{E}_n$ such that $m(x,y)\leqslant 1/n$ (otherwise we can add $x$ to $\mathcal{E}_n$, which contradicts its maximality). In what follows we will clearly indicate the set $\mathcal{X}_0$, and then just write $\mathcal{E}_n$. \subsection{Proof of the Theorem \ref{metricabove}} Recall that $(\mathcal{X}, m)$ is a metric space, and $B(x,r)$ denotes a closed ball (in the metric $m$) with center $x\in \mathcal{X}$ and radius $r$. Put $\mathcal{E}_N:=\mathcal{E}_N(\mathcal{X})$ and note that \begin{equation}\label{cardbound} \mu(\mathcal{X})\geqslant \sum\limits_{x\in \mathcal{E}_n}\mu\left(B\left(x, \frac{1}{3n}\right)\right) \geqslant \textup{card}(\mathcal{E}_n) \Phi(1/(3n)). \end{equation} Suppose now that $X_N=\{x_1, \ldots, x_N\}$ is a set of $N$ random points, independently distributed over $\mathcal{X}$ with respect to the measure $\mu$. We denote its covering radius by $$ \rho(X_N):=\rho(X_N, \mathcal{X}). $$ Suppose $\rho(X_N)> \frac{2}n$. Then there exists a point $y\in \mathcal{X}$ such that $X_N\cap B\left(y, \frac{2}n\right)=\emptyset$. Choose a point $x\in \mathcal{E}_n$ such that $m(x,y)<\frac{1}n$. Then $B(x, \frac{1}{n})\subset B(y, \frac{2}n)$, and so the ball $B(x, \frac{1}{n})$ (and thus $B(x, \frac{1}{3n})$) does not intersect $X_N$. Therefore, \begin{align}\label{aboveprobbb} \P\left(\rho(X_N)\geqslant \frac{2}{n}\right) &\leqslant \P\left(\exists x\in \mathcal{E}_n\colon B(x, 1/(3n))\cap X_N=\emptyset\right) \\ & \leqslant \textup{card}(\mathcal{E}_n)\cdot \left(1- \frac{\Phi(\frac1{3n})}{\mu(\mathcal{X})}\right)^N. \notag \end{align} We now choose $n$ to be such that $\frac{1}{3n}=\Phi^{-1}(\frac{\alpha \log N}{N})$. There exists such an $n$ since $\Phi$ is continuous and $\Phi(r)\to 0$ as $r\to 0^+$. Then utilizing the upper bound for $\textup{card}(\mathcal{E}_n)$ from \eqref{cardbound}, we deduce that for some $C>0$ we have $$ \P\left[\rho(X_N)\geqslant \frac{2}n\right] \leqslant C \frac{N}{\log N}\cdot N^{-C\alpha}, $$ which concludes the proof of the estimate \eqref{metricaboveprob}. To establish the estimate \eqref{metricaboveexpect}, notice that since for small values of $r$ we have $\Phi(r)\leqslant r^\sigma$, it follows that for small $r$ and $D=\frac1\sigma$ we have $\Phi^{-1}(r)\geqslant r^D$. Choose $\alpha$ so large that $N^{1-C\alpha}=o(N^{-D})$ as $N\to \infty$. Then $$ \mathbb{E}\rho(X_N) \leqslant \frac{2}{n} + C\operatorname{diam}(\mathcal{X})\cdot o(N^{-D}) =6\Phi^{-1}(\frac{\alpha \log N}{N})+o(N^{-D}). $$ Finally, since $\Phi^{-1}(\frac{\alpha\log N}{N})\geqslant \Phi^{-1}(N^{-1}) \geqslant N^{-D}$, inequality \eqref{metricaboveexpect} follows. \hfill $\Box$ \subsection{Proof of the Theorem \ref{metricbelow}}\label{proofofmetricbelow} Let $\mathcal{E}_n:=\mathcal{E}_n(\mathcal{X}_1)$, where $\mathcal{X}_1$ is as in the hypothesis. Notice that \begin{equation}\label{cardbound2} 0<\mu(\mathcal{X}_1) \leqslant \sum\limits_{x\in \mathcal{E}_n} \mu\left(B(x, \frac{1}{n})\right) \leqslant \textup{card}(\mathcal{E}_n)\Phi\left(\frac{1}n\right). \end{equation} An estimate as in \eqref{cardbound} together with the doubling property of $\Phi$ imply that $$ \mu(\mathcal{X})\geqslant c \cdot \textup{card}(\mathcal{E}_n) \Phi\left(\frac1{3n}\right) \geqslant \tilde{c} \cdot \textup{card}(\mathcal{E}_n)\Phi\left(\frac1n\right). $$ Thus, $\tau_n:=\textup{card}(\mathcal{E}_n)\cdot \Phi(1/n)$ satisfies $0<c_1<\tau_n<c_2$ for some constants $c_1$ and $c_2$ independent of $n$. Clearly if a ball $B(x, \frac{1}{3n})$ does not intersect $X_N$, then $\rho(X_N)=\rho(X_N, \mathcal{X})\geqslant \frac{1}{3n}$. Thus $$ \P\left(\rho(X_N)\geqslant \frac{1}{3n}\right)\geqslant \P\left(\exists x\in \mathcal{E}_n\colon B(x, 1/(3n))\cap X_N = \emptyset \right). $$ Notice that the balls $B(x, \frac{1}{3n})$ are disjoint for $x\in \mathcal{E}_n$, and their $\mu$-measure is comparable to $t:=\Phi(\frac{1}{n})$. Next we claim that for every $x\in \mathcal{E}_n$ there exists a constant $c_x\leqslant 1$ such that the balls $B(x, c_x \frac{1}{3n})$ have the same measure $c_0 \Phi(\frac{1}{n})=c_0 t$, and moreover that the uniform estimate $c_x > c>0$ holds for some constant $c$. To see this, take two points $x_1, x_2\in \mathcal{X}_1$ and assume that the balls $B_i := B(x_i, r)$, $i=1,2$ are disjoint. Suppose $\mu(B_1)<\mu(B_2)$. Define the function $\varphi(s):=\mu(B(x_2, s\cdot r))$. The strict doubling property of $\Phi$ implies $$ \mu(B_1)\geqslant c\Phi(r) \geqslant c\cdot C_1^k \Phi(r/2^k). $$ Choose $k$ such that $c\cdot C_1^k \geqslant 1$. Then $$ \mu(B_1)\geqslant \Phi(r/2^k) \geqslant \varphi(2^{-k}). $$ Thus, $\varphi(2^{-k})\leqslant\mu(B_1)<\mu(B_2)=\varphi(1)$. By continuity of $\varphi$ we see that there exists a constant $c_{x_2}$ such that $\mu(B(x_2, c_{x_2}r))=\mu(B(x_1, r))$. Notice that $c_{x_2} \geqslant 2^{-k}=:c_0$, where $k$ depends only on the constants $c, C_1$ from Theorem \ref{metricbelow} and not on $x_1, x_2$, or $r$. Applying this procedure to all balls $B(x, 1/(3n))$, $x\in \mathcal{E}_n$, and using the fact that $\textup{card}(\mathcal{E}_n)=\tau_n /t$, we obtain \begin{align}\label{blahblahblah} \P\left(\rho(X_N)\geqslant \frac{c_0}3 \Phi^{-1}(t)\right)&\geqslant \P\left(\mbox{one of $\frac{\tau_n}t$ disjoint balls of measure $c_0t$ is disjoint from $X_N$}\right) \notag \\&= f(N, \frac{1}{c_0t}, \frac{\tau_n}t) = f(N, \frac1{c_0t}, \frac{\kappa_n}{c_0t}), \end{align} where $\kappa_n:=c_0\tau_n$ and $f$ is given in \eqref{functionf}. If necessary, we can decrease the size of $c_0$ so that $\kappa_n \leqslant 1$ for $n$ large. As we have seen in Lemma \ref{lemmaforf}(i), there exists a number $\alpha$ such that if $$\frac1{c_0t}= \frac{N}{\log N-\alpha\log\log N},$$ then $f(N, \frac{1}{c_0t}, \frac{\kappa_n}{c_0t})\to 1$ as $N\to \infty$. Thus, for any sufficiently large number $N$ we have $$ \P\left(\rho(X_N)\geqslant \frac{c_0}3 \Phi^{-1}(\frac{\log N-\alpha\log\log N}{c_0 N})\right)\geqslant 1-o(1), \; N\to \infty, $$ which is the desired inequality \eqref{metricbelowprob}. Moreover, for large values of $N$ we have $\log N-\alpha\log\log N\geqslant \frac{1}2\log N$; thus $$ \mathbb{E} \rho(X_N)\geqslant c_1 \Phi^{-1}(c_2 \frac{\log N}N), $$ which proves inequality \eqref{metricbelowexpect}. \hfill $\Box$ \subsection{Estimates from above for asymptotically flat sets}\label{abovvve} Let $K$ be an asymptotically flat $s$-regular subset of $\mathbb{R}^d$ and put $$ \rho(X_N)=\rho(X_N, K), \qquad \varepsilon_N:=\frac{1}{\log N}. $$ In order to deduce sharp asymptotic results we first improve our estimates from above by considering a better net of points. For each $N>4$ let $\mathcal{E}_{n/\varepsilon_N}:=\mathcal{E}_{n/\varepsilon_N}(K)$. From estimates similar to \eqref{cardbound} and \eqref{cardbound2} we see that $\textup{card}(\mathcal{E}_n)$ is comparable to $\left(n/\varepsilon_N\right)^s$ independently of $N$. Suppose $\rho(X_N)> \frac{1}{n}$. Then, since $K$ is compact, for some $y\in K$ we have $B_d(y,\frac{1}{n})\cap X_N=\emptyset$, and thus there exists a point $x\in \mathcal{E}_{n/\varepsilon_N}$ such that $B_d(x, \frac{1-\varepsilon_N}{n})\cap X_N=\emptyset$. We fix a number $\delta$, $0<\delta<1$, and take $n$ so large that $$ \mathcal{H}_s\left(B_d(x, \frac{1-\varepsilon_N}n)\cap K\right)\geqslant (1-\delta)\upsilon_s \frac{(1-\varepsilon_N)^s}{n^s} \geqslant (1-\delta)\upsilon_s \frac{1-s\varepsilon_N}{n^s}. $$ As in \eqref{aboveprobbb}, \begin{equation}\label{probmanifoldabove} \P\left(\rho(X_N)> \frac{1}{n}\right)\leqslant C \left(\frac{n}{\varepsilon_N}\right)^s \left(1-\frac{1}{\mathcal{H}_s(K)}(1-\delta)\upsilon_s\frac{1-s\varepsilon_N}{n^s}\right)^N. \end{equation} Fix a number $A>0$ and choose $$ n_1 := \left(\frac{(1-\delta)\upsilon_s}{\mathcal{H}_s(K)} \frac{N}{\log N+A\log\log N}\right)^{1/s}. $$ Then with $n=n_1$ in \eqref{probmanifoldabove} we get for all $N$ large, \begin{equation}\label{loglog} \P\left(\rho(X_N)> \frac{1}{n_1}\right) \leqslant C \cdot N (\log N )^{s-1} e^{-(1-s/\log N)(\log N+A\log\log N)}. \end{equation} Recall that $C$ does not depend on $N$. Thus if $A$ and $N$ are sufficiently large, it follows that \begin{equation}\label{problessthanlog} \P\left(\rho(X_N)> \frac{1}{n_1}\right) \leqslant \frac{1}{\log N}. \end{equation} Furthermore, if we plug $n=n_2 := \left(\frac{N}{B\log N}\right)^{1/s}$ in \eqref{probmanifoldabove} we get for sufficiently large $B$ \begin{equation}\label{problessthanpower} \P\left(\rho(X_N)> \frac{1}{n_2}\right) \leqslant N^{-p/s - 1}. \end{equation} With $\textup{d}\mu=\mathbbm{1}_K \textup{d}\mathcal{H}_s/\mathcal{H}_s(K)$, we make use of the formula \begin{multline}\label{superest} \mathbb{E}[\rho(X_N)^p] = \int_{K^N} \rho(X_N)^p \textup{d}\mu(x_1)\ldots \textup{d}\mu(x_N) = \int\limits_{\rho(X_N)\leqslant1/n_1} \rho(X_N)^p\textup{d}\mu(x_1)\ldots \textup{d}\mu(x_N) + \\ \int\limits_{1/n_1 < \rho(X_N)\leqslant 1/n_2} \rho(X_N)^p\textup{d}\mu(x_1)\ldots \textup{d}\mu(x_N) + \int\limits_{\rho(X_N)> 1/n_2}\rho(X_N)^p\textup{d}\mu(x_1)\ldots \textup{d}\mu(x_N) \leqslant \\ \frac1{n_1^p} + \frac{1}{n_2^p}\cdot \P\left(\rho(X_N)> \frac{1}{n_1}\right) + (\textup{diam}(K))^p \cdot \P\left(\rho(X_N)> \frac1{n_2}\right). \end{multline} From \eqref{problessthanlog}, \eqref{problessthanpower}, and the definitions of $n_1$ and $n_2$, we obtain \begin{multline}\label{superest2} \mathbb{E}[\rho(X_N)^p] \leqslant \left(\frac{\log N+A\log\log N}{N}\right)^{p/s} \cdot \left(\frac{\mathcal{H}_s(K)}{\upsilon_s}\right)^{p/s} \cdot (1-\delta)^{-p/s} + C\left(\frac{\log N}{N}\right)^{p/s} \frac{1}{\log N} \\ +CN^{-p/s-1}. \end{multline} Therefore, for any $\delta$ with $0<\delta<1$, $$ \limsup_{N \to \infty} \mathbb{E}[\rho(X_N)^p]\cdot \left(\frac{N}{\log N}\right)^{p/s} \leqslant (1-\delta)^{-p/s}\cdot \left(\frac{\mathcal{H}_s(K)}{\upsilon_s}\right)^{p/s}, $$ and consequently \begin{equation}\label{flatabovelastformula} \limsup_{N \to \infty} \mathbb{E}[\rho(X_N)^p] \cdot \left(\frac{N}{\log N}\right)^{p/s} \leqslant \left(\frac{\mathcal{H}_s(K)}{\upsilon_s}\right)^{p/s}. \end{equation} \hfill $\Box$ \subsection{Estimate from above for quasi-nice sets}\label{abovvvve} Let $K$ be a quasi-nice $s$-regular subset of $\mathbb{R}^d$, and again set $\varepsilon_N:=1/\log N$ and $\mathcal{E}_{n/\varepsilon_N}:=\mathcal{E}_{n/\varepsilon_N}(K)$, where $n/\varepsilon_N\to \infty$ as $N\to \infty$. Since the set $T$ from part (iii) of Definition \ref{flat} is finite, the regularity condition (ii) implies $$ \mathcal{H}_s\left(\bigcup_{x\in T}B_d(x, r)\right) \leqslant C \cdot \textup{card}(T) \cdot r^s = C_1 r^s, \; \; 0<r<r_0. $$ Suppose $y_1, \ldots, y_k \in \mathcal{E}_{n/\varepsilon_N} \cap \bigcup_{x\in T}B_d(x, \frac{1-\varepsilon_N}n)$. Then the balls $B_d(y_j, \frac{\varepsilon_N}{3n})$ are disjoint and $B_d(y_j, \frac{\varepsilon_N}{3n})\subset \bigcup_{x\in T}B_d(x, \frac{1+\varepsilon_N}n)$ for $j=1,\ldots,k$. The chain of inequalities $$ C_1\left(\frac{1+\varepsilon_N}n\right)^s \geqslant \mathcal{H}_s\left(\bigcup_{x\in T}B_d(x, \frac{1+\varepsilon_N}n)\right)\geqslant \sum\limits_{j=1}^k \mathcal{H}_s\left(B_d(y_j, \frac{\varepsilon_N}{3n})\right) \geqslant c\cdot k\cdot (\frac{\varepsilon_N}n)^s $$ implies that $k\leqslant C_2/\varepsilon_N^s$, and $C_2$ does not depend on $N$. Further, if $y\in \mathcal{E}_{n/\varepsilon_N}\setminus\bigcup_{x\in T}B_d(x, \frac{1-\varepsilon_N}n)$, then $\mathcal{H}_s\left(B_d(y, \frac{1-\varepsilon_N}n)\right)\geqslant \upsilon_s \left(\frac{1-\varepsilon_N}n\right)^s$. As we have seen in \eqref{probmanifoldabove}, $\P(\rho(X_N)> 1/n)$ is bounded from above by the probability that for some $y\in \mathcal{E}_{n/\varepsilon_N}$ we have $B_d\left(y, \frac{1-\varepsilon_N}n\right)\cap X_N=\emptyset$. Taking into account that $\textup{card}(\mathcal{E}_{n/\varepsilon_N}) \leqslant C_3(n/\varepsilon_N)^s$, we obtain \begin{multline*} \P\left(\rho(X_N)> \frac1n\right) \leqslant \P\bigg(\mbox{one of $\leqslant\frac{C_2}{\varepsilon_N^s}$ balls of measure $\geqslant\frac{c_1}{n^s}$ is disjoint from $X_N$ or} \\ \mbox{one of $\leqslant C_3\left(\frac{n}{\varepsilon_N}\right)^s$ balls of measure $\geqslant\frac{\upsilon_s(1-\varepsilon_N)^s}{n^s}$ is disjoint from $X_N$}\bigg). \end{multline*} This last probability is bounded from above by $$ \frac{C_2}{\varepsilon_N^s}\left(1-\frac{c_1}{n^s}\right)^N + C_4\left(\frac{n}{\varepsilon_N}\right)^s\left(1-\frac{1}{\mathcal{H}_s(K)}\frac{\upsilon_s (1-\varepsilon_N)^s}{n^s}\right)^N. $$ As in the preceding proof, if $$ n_1=\left(\frac{\upsilon_s}{\mathcal{H}_s(K)} \frac{N}{\log N+A\log\log N}\right)^{1/s}, $$ then, for $N$ large, $$ C_4\left(\frac{n_1}{\varepsilon_N}\right)^s\cdot\left(1-\frac{1}{\mathcal{H}_s(K)}\frac{\upsilon_s (1-\varepsilon_N)^s}{n_1^s}\right)^N \leqslant \frac{C_5}{\log N}. $$ Furthermore notice that if $C_6$ is sufficiently large, then $$ \frac{C_2}{\varepsilon_N^s}\left(1-\frac{c_1}{n_1^s}\right)^N \leqslant C_6(\log N)^s N^{-c_2}, \;\; N\to \infty. $$ Repeating estimates \eqref{superest} and \eqref{superest2}, we obtain \begin{equation}\label{lastinquasinice} \limsup_{N \to \infty} \mathbb{E}[\rho(X_N)^p] \cdot \left(\frac{N}{\log N}\right)^{p/s} \leqslant \left(\frac{\mathcal{H}_s(K)}{\upsilon_s}\right)^{p/s}. \end{equation} \hfill $\Box$ Note that \eqref{lastinquasinice} holds whether or not $K$ is countably $s$-rectifiable; it requires only that properties (ii) and (iii) of Definition \ref{flat} hold. \subsection{Estimate from below for quasi-nice sets}\label{sectionfrombelow} For the proof of Theorem \ref{manifolds}, it remains in view of inequalities \eqref{flatabovelastformula} and \eqref{lastinquasinice}, to establish \begin{equation}\label{liminfbigger} \liminf_{N \to \infty} \mathbb{E}[\rho(X_N)^p] \cdot \left(\frac{N}{\log N}\right)^{p/s} \geqslant \left(\frac{\mathcal{H}_s(K)}{\upsilon_s}\right)^{p/s} \end{equation} for asymptotically flat and quasi-nice $s$-dimentional manifolds $K$. Since by the H\"older inequality we have $$ \liminf_{N\to \infty}\mathbb{E}[\rho(X_N)^p] \cdot \left[\frac{N}{\log N}\right]^{p/s} \geqslant \left(\liminf_{N\to \infty}\mathbb{E} \rho(N_N)\cdot \left[\frac{N}{\log N}\right]^{1/s}\right)^p, $$ it is enough to prove \eqref{liminfbigger} for $p=1$. If $K$ is quasi-nice, then $K$ is countably $s$-rectifiable ($s$ is an integer) and $0<\mathcal{H}_s(K)<\infty$; thus as previously remarked, the following holds for $\mathcal{H}_s$-almost every point $x\in K$: $$ r^{-s}\cdot \mathcal{H}_s(B_d(x,r)\cap K)\to \upsilon_s, \;\; r\to 0^+. $$ Fix a number $\delta$ with $0<\delta<1$ and define $r_n:=1/n$ and $q_n:=\left(\frac{1-\delta}{1+\delta}\right)^{1/s}\cdot 1/n$, where $\{n\}$ is a given countable sequence tending to infinity. By Egoroff's theorem, there exists a set $K_1=K_1(\delta)\subset K$ with $\mathcal{H}_s(K_1)>\frac{1}{2}\mathcal{H}_s(K)$ on which the above limit is uniform for radii $r$ equal to $r_n$ and $q_n$. That is, \begin{equation}\label{uniforr} r^{-s}\mathcal{H}_s(B_d(x,r)\cap K)\rightrightarrows \upsilon_s, \;\; r=r_n \; \mbox{or} \; r=q_n, \;\; n\to \infty. \end{equation} This means that there exists a large number $n(\delta)$, such that for any $n>n(\delta)$ we have, for every $x\in K_1$, \begin{align} &(1-\delta)\upsilon_s r_n^s \leqslant \mathcal{H}_s(B_d(x,r_n)\cap K) \leqslant (1+\delta)\upsilon_sr_n^s, \label{tilitili1} \\ &(1-\delta)\upsilon_sq_n^s \leqslant \mathcal{H}_s(B_d(x,q_n)\cap K) \leqslant (1+\delta)\upsilon_s q_n^s = (1-\delta)\upsilon_s r_n^s. \label{tilitili2} \end{align} Recalling the notation of Section \ref{sectionprelim}, we set $\mathcal{E}_{n/2}:=\mathcal{E}_{n/2}(K_1)$. Then, as in the preceding sections, there exist positive constants $c_1$ and $c_2$ (independent of $n$) such that $c_1 n^s \leqslant \textup{card}(\mathcal{E}_{n/2}) \leqslant c_2 n^s$ where, for the lower bound, we use $$ 0<\mathcal{H}_s(K_1)\leqslant \mathcal{H}_s\left(\bigcup_{x\in \mathcal{E}_{n/2}} (B_d(x, 2/n)\cap K)\right) \leqslant C\cdot \textup{card}(\mathcal{E}_{n/2})(2/n)^s. $$ Thus, $\tau_n:=\textup{card}(\mathcal{E}_{n/2})/n^s$ satisfies $0<c_1\leqslant \tau_n \leqslant c_2$. Clearly, if for some $x\in \mathcal{E}_{n/2}$ the ball $B_d(x, \frac1n)$ is disjoint from $X_N$, then $\rho(X_N)\geqslant \frac 1n$. Thus, for a given $\delta>0$ and sufficiently large $n$ we have a family $\{B_d(x, 1/n)\cap K \colon x\in \mathcal{E}_{n/2}(K_1)\}$ of $\tau_n n^s$ balls (relative to $K$) with disjoint interiors of radius $1/n$ and $\mathcal{H}_s$-measure between $(1-\delta)\upsilon_s/n^s$ and $(1+\delta)\upsilon_s /n^s$. For a fixed $x\in \mathcal{E}_{n/2}(K_1)$, define $\varphi(s):=\mathcal{H}_s(B(x, s/n)\cap K)$. Then $\varphi(1)\geqslant (1-\delta)\upsilon_s/n^s$. On the other hand, inequalities \eqref{tilitili2} imply $$ \varphi\left(\left(\frac{1-\delta}{1+\delta}\right)^{1/s}\right) \leqslant (1-\delta)\upsilon_s/n^s. $$ \noindent Thus, there is a number $c_x=c_{x,n}$, with $c_x\geqslant (\frac{1-\delta}{1+\delta})^{1/s}$, such that $\varphi(c_x)=(1-\delta) \upsilon_s/n^s$. That is, there exists a new family $\{B_d(x, c_x/n)\cap K\colon x\in \mathcal{E}_{n/2}(K_1)\}$, with $c_x \geqslant (\frac{1-\delta}{1+\delta})^{1/s}$, and the sets $B_d(x, c_x/n)\cap K$ all have the same $\mathcal{H}_s$ measure, namely $(1-\delta)\upsilon_s/n^s$. As in \eqref{blahblahblah}, it follows that \begin{align}\label{onemoreestimate} \P\left(\rho(X_N)\geqslant \left(\frac{1-\delta}{1+\delta}\right)^{1/s}\frac1{n}\right) &\geqslant f\left(N, \frac{\mathcal{H}_s(K)n^s}{(1-\delta)\upsilon_s}, \tau_n n^s\right)\\ &= f\left(N, \frac{\mathcal{H}_s(K)n^s}{(1-\delta)\upsilon_s}, \kappa_n \cdot\frac{\mathcal{H}_s(K)n^s}{(1-\delta)\upsilon_s}\right),\notag \end{align} where $$ \kappa_n:=\tau_n \cdot \frac{(1-\delta)\upsilon_s}{\mathcal{H}_s(K)}. $$ It is easily seen that $$ \mathcal{H}_s(K)\geqslant \tau_n n^s \cdot \frac{(1-\delta) \upsilon_s}{n^s} = \mathcal{H}_s(K)\kappa_n; $$ thus $\kappa_n \leqslant 1$. Part (i) of Lemma \ref{lemmaforf} therefore implies that the sequence in \eqref{onemoreestimate} tends to $1$ as $N\to \infty$ if (for suitable $\alpha$) we have $$ \frac{(1-\delta) \upsilon_s}{\mathcal{H}_s(K)n^s} = \frac{\log N-\alpha\log\log N}{N}, $$ which is equivalent to \begin{equation}\label{newdefinofn} n := \left[\frac{(1-\delta) \upsilon_s}{\mathcal{H}_s(K)}\cdot \frac{N}{\log N-\alpha\log\log N}\right]^{1/s}. \end{equation} We take $N$ so large that $n$ exceeds $n(\delta)$, which ensures that the inequalities \eqref{tilitili1}--\eqref{tilitili2} hold. From \eqref{onemoreestimate} we obtain $$ \mathbb{E}\rho(X_N)\geqslant \left(\frac{1-\delta}{1+\delta}\right)^{1/s}\frac{1}{n} \cdot f\left(N, \frac{\mathcal{H}_s(K)n^s}{(1-\delta)\upsilon_s}, \tau_n n^s\right). $$ Using the definition of $n$ in \eqref{newdefinofn}, we get \begin{multline}\label{manifbelowlastformula} \mathbb{E} \rho(X_N) \cdot \left[\frac{N}{\log N}\right]^{1/s} \geqslant \\ \left[\frac{N}{\log N}\right]^{1/s}\cdot \left(\frac{1-\delta}{1+\delta}\right)^{1/s} \cdot \left[\frac{\mathcal{H}_s(K)}{(1-\delta) \upsilon_s}\cdot \frac{\log N-\alpha\log\log N}{N}\right]^{1/s} \cdot f\left(N, \frac{\mathcal{H}_s(K)n^s}{(1-\delta)\upsilon_s}, \tau_n n^s\right), \end{multline} and passing to the $\liminf$ as $N\to \infty$ yields $$ \liminf_{N \to \infty} \mathbb{E}\rho(X_N) \cdot \left(\frac{N}{\log N}\right)^{1/s} \geqslant \left(\frac{1}{1+\delta}\right)^{1/s} \cdot \left[\frac{\mathcal{H}_s(K)}{\upsilon_s}\right]^{1/s}. $$ Recalling that $\delta$ can be taken arbitrarily small, we obtain \eqref{liminfbigger} for quasi-nice sets. For asymptotically flat sets the same (but even simpler) argument applies. \hfill $\Box$ \subsection{Proof of Corollary \ref{maehara}} Recall that $$ Z_N=\rho(X_N, \mathbb{S}^d)\cdot \left(\frac{\upsilon_d}{(d+1)\upsilon_{d+1}}\cdot \frac{N}{\log N}\right)^{1/d}. $$ Corollary \ref{corsphere} implies that $\mathbb{E} Z_N \to 1$ and $\mathbb{E}[Z_N^2]\to 1$; thus $\mathbb{E}[(Z_N-1)^2] = \mathbb{E}[Z_N^2]-2\mathbb{E} Z_N + 1 \to 0$. The Chebyshev inequality then implies $$ \P(|Z_N-1|>\varepsilon) \leqslant \frac{\mathbb{E}[(Z_N-1)^2]}{\varepsilon^2} \to 0, $$ which completes the proof. \hfill $\Box$ \subsection{Proof of the Corollaries \ref{corsphere} and \ref{corcurve}} It is well known that a closed $C^{(1,1)}$ manifold is an asymptotically flat set, and a rectifiable curve is a quasi-nice $1$-dimensional set. For the first fact, we refer the reader to a textbook on Riemannian geometry, for instance, \cite[Chapters 5--10]{BurIv}. The second fact can be deduced from \cite[Section 3.2]{Fal}. \subsection{Proof of the Theorem \ref{theoremball}: estimate from above} The proof of the theorem is similar to the proof for asymptotically flat sets. However, we need to take into account that the limit \eqref{uniform} is not equal to $\upsilon_d$ for points on the boundary. We use properties (ii) and (iii) of $K$ to obtain \begin{align} &r^{-d}\mathcal{H}_d(B_d(x,r)\cap K) \rightrightarrows \frac12 \upsilon_d, \; \; r\to 0, \;\; x\in \partial K; \label{fact1}\\ & x\in K, \; \textup{dist}(x, \partial K)>r \Rightarrow \mathcal{H}_d(B_d(x,r)\cap K)=\mathcal{H}_d(B_d(x,r))=\upsilon_d r^d; \label{fact2} \\ & \forall \delta>0 \; \exists r(\delta)>0 \colon \forall r<r(\delta), \forall x\in K\colon \mathcal{H}_d(B_d(x,r)\cap K)\geqslant (\frac12-\delta)\upsilon_d r^d. \label{fact3} \end{align} For the details, we refer the reader to Lee, \cite[Chapter 5]{L} For large $N$, set $\mathcal{E}_{n/\varepsilon_N}:=\mathcal{E}_{n/\varepsilon_N}(K)$ and $\varepsilon_N:=1/\log N$, where $n(N)$ is a sequence such that $n\asymp (N/\log N)^{1/d}$. We now fix a number $\delta$ with $0<\delta<1/2$. Notice that if $x\in \mathcal{E}_{n/\varepsilon_N}$ and $\textup{dist}(x, \partial K) > (1-\varepsilon_N)/n$, then $$\mathcal{H}_d(B_d(x,(1-\varepsilon_N)/n)\cap K)=\upsilon_d \left((1-\varepsilon_N)/n\right)^d;$$ if $x\in \mathcal{E}_{n/\varepsilon_N}$ and $\textup{dist}(x, \partial K) \leqslant (1-\varepsilon_N)/n$ then, for large enough $n$, $$ \mathcal{H}_d(B_d(x, (1-\varepsilon_N)/n)\cap K)\geqslant (\frac12-\delta)\upsilon_d ((1-\varepsilon_N)/n)^d. $$ On considering disjoint balls (relative to $K$) of radius $\varepsilon_N/(3n)$ and using that $$ \mathcal{H}_d(\{x\colon \textup{dist}(x, \partial K)\leqslant (1-\frac23\varepsilon_N)/n\}) \leqslant C_1/n, $$ we deduce, as in \eqref{cardbound}, that $$ \textup{card}\left\{x\in \mathcal{E}_{n/\varepsilon_N}\colon \operatorname{dist}(x, \partial K)\leqslant \frac{1-\varepsilon_N}n\right\} \leqslant C_2\frac{n^{d-1}}{\varepsilon_N^d}. $$ Therefore, for large enough $n$, we get \begin{multline}\label{ballabovelong} \P\left(\rho(X_N)> \frac1n\right) \leqslant \P\left(\exists x\in \mathcal{E}_{n/\varepsilon_N}\colon B_d\left(x, \frac{1-\varepsilon_N}n\right)\cap K \cap X_N =\emptyset\right) \leqslant \\ C_2\frac{n^{d-1}}{\varepsilon_N^d} \left(1- \frac{(1/2-\delta)\upsilon_d}{\mathcal{H}_d(K)}\left(\frac{1-\varepsilon_N}{n}\right)^d\right)^N + C_3\frac{n^d}{\varepsilon_N^d} \left(1- \frac{\upsilon_d}{\mathcal{H}_d(K)} \left(\frac{1-\varepsilon_N}{n}\right)^d\right)^N. \end{multline} Repeating the estimates \eqref{problessthanlog}--\eqref{flatabovelastformula} with $$ n_1:=\left( \frac{(1/2-\delta)\upsilon_d}{\mathcal{H}_d(K)} \cdot \frac{N}{\frac{d-1}d \log N + A\log\log N}\right)^{1/d}, $$ and $$ n_2:=\left(\frac{N}{B\log N}\right)^{1/d}, $$ where $A$ and $B$ are sufficiently large, we obtain, after letting $\delta\to 0^+$, the estimate $$ \limsup_{N\to \infty}\mathbb{E}[\rho(X_N)^p] \left( \frac{N}{\log N}\right)^{p/d} \leqslant \left(\frac{2(d-1)}d \cdot \frac{\mathcal{H}_d(K)}{\upsilon_d}\right)^{p/d}. $$\hfill $\Box$ \subsection{Proof of the Theorem \ref{theoremball}: estimate from below}\label{sectionballbelow} We repeat the proof from the Section \ref{sectionfrombelow}, but now we will place our net $\mathcal{E}$ only on the boundary $\partial K$. Namely, put $\mathcal{E}_{n/2}:=\mathcal{E}_{n/2}(\partial K)$. Since $\partial K$ is a smooth $d-1$-dimensional submanifold, we see that $\textup{card}(\mathcal{E}_{n/2})=\tau_n n^{d-1}$ with $0<c_1<\tau_n <c_2$. Moreover, from \eqref{fact1} we obtain as in \eqref{uniforr} that $$ r^{-d}\mathcal{H}_d(B_d(x,1/n)\cap K) \rightrightarrows \frac12 \upsilon_d/n^d, \; \; r=r_n \; \mbox{or} \; r=q_n, \; n\to \infty, $$ uniformly for $x\in \mathcal{E}_{n/2}$. The remainder of the proof just involves repeating the estimates \eqref{onemoreestimate}--\eqref{manifbelowlastformula}, using part (ii) of Lemma \ref{lemmaforf}. \hfill $\Box$ \subsection{Estimate from above for the cube $[0,1]^d$} The proof is similar to the case of the bodies with smooth boundary. The only change we need to make is to the formula \eqref{fact1}. Namely, if a point $x$ lies on a $(d-k)$-dimensional edge of the cube, then $\mathcal{H}_d(B_d(x,r)\cap [0,1]^d) \asymp 2^{-k}\upsilon_d r^d$. Moreover, $\mathcal{H}_d(B_d(x,r)\cap [0,1]^d) = 2^{-k}\upsilon_d r^d$ for points $x$ on the $(d-k)$-dimensional edge that are at distance larger than $r$ from all $(d-k-1)$-dimensional edges. Thus, if we consider a set $\mathcal{E}_{n/\varepsilon_N}:=\mathcal{E}_{n/\varepsilon_N}([0,1]^d)$, we have for any $k=0,\ldots, d$ at most $C_k n^{d-k}/\varepsilon_N^d$ points $x\in \mathcal{E}_{n/\varepsilon_N}$ with $\mathcal{H}_d(B_d(x, (1-\varepsilon_N)/n)\cap [0,1]^d)\geqslant 2^{-k}\upsilon_d((1-\varepsilon_N)/n)^d$ . In particular, if $k=d$ we have only finitely many such points $x\in \mathcal{E}_{n/\varepsilon_N}$; and if $k=d-1$, we have no more than $Cn/\varepsilon_N^d$ such points. We now repeat the estimates \eqref{problessthanlog}--\eqref{flatabovelastformula} and \eqref{ballabovelong} with $$ n_1 := \left(2^{-(d-1)}\cdot d\cdot\upsilon_d\cdot \frac{N}{\log N+A\log\log N}\right)^{1/d}. $$\hfill $\Box$ \subsection{Estimate from below for the cube $[0,1]^d$} The proof is almost identical to the proof in the Section \ref{sectionballbelow}; the only difference is that now we take $\mathcal{E}_{n/2}:=\mathcal{E}_{n/2}(L)$, where $L$ is a $1$-dimensional edge of the cube $[0,1]^d$. To complete the analysis we appeal to part (iii) of Lemma \ref{lemmaforf}. \hfill $\Box$ \subsection{Estimates for a polyhedron in $\mathbb{R}^3$} The estimates here are the same as for the unit cube $[0,1]^d$. The only difference is that, for points $x\in L$, where $L$ is the edge where two faces intersect at angle $\theta$, we have, if $x$ is far enough from the vertices of $P$: $$ \mathcal{H}_3(B(x,r)\cap P) = \frac{\theta}{2\pi} \cdot \upsilon_3 \cdot r^3. $$ Consequently, for $k=0,1,2,3$ we have at most $a_k n^{3-k}/\varepsilon_N^3$ points $x\in \mathcal{E}_{n/\varepsilon_N}(P)$ with $\mathcal{H}_3(B_3(x, (1-\varepsilon_N)/n)\cap P)\geqslant c_k\upsilon_3((1-\varepsilon_N)/n)^3$, where $a_0=1$, $a_1=1/2$, and $a_2=\theta/(2\pi)$. In the case $\theta\leqslant \pi/2$, one needs to choose $$ n_1:=\left(\frac{2\theta}{V(P)}\cdot \frac{N}{\log N + A\log\log N}\right)^{1/3}, $$ and in the case $\theta\geqslant \pi/2$, one needs to choose $$ n_1:=\left(\frac{\pi}{V(P)}\cdot \frac{N}{\log N + A\log\log N}\right)^{1/3}. $$ For the estimate from above, consider $\mathcal{E}_{n/2}(L)$ and repeat the estimates for the cube. \hfill $\Box$ \subsection{Estimates for $\textup{\textmd{d}}\mu=\frac{\textup{\textmd{d}}x}{\sqrt{1-x^2}}$} We remind the reader that $\hat{\rho}(X_N)=\hat{\rho}(X_N, [0,1])=\sup_{y\in [1-\frac{1}{N^2}, 1]} \inf_j |y-x_j|$, where $x_j$, $j=1,\ldots,N$, are randomly and independently distributed over $[0,1]$ with respect to $\mu$. \subsubsection{Case $a=2$} Suppose that an interval $I_\alpha:=[1-\frac{\alpha}{N^2}, 1]$ is disjoint from $X_N$ for some $\alpha>1$. Then we get $$ \hat{\rho}(X_N)\geqslant \frac{\alpha-1}{N^2}. $$ We notice that if $\alpha<C_1\log^2(N)$, and $N$ is sufficiently large, then $$ \mu(I_\alpha)\leqslant C_2 \frac{\sqrt{\alpha}}{N}. $$ Therefore, if $\alpha$ is some number greater than $1$, $$ \P\left(\hat{\rho}\geqslant \frac{\alpha-1}{N^2}\right)\geqslant \left(1-C_2\frac{\sqrt{\alpha}}N\right)^N \geqslant C_3. $$ Consequently, $$ \mathbb{E}\hat{\rho}\geqslant \frac{C_4}{N^2}, $$ where $C_4 = C_3(\alpha-1)$. For the estimate from above, notice that $\mu(I_\alpha)\geqslant \sqrt{\alpha}/(\sqrt{2}\pi N)$. Assuming $\hat{\rho}(X_N)\geqslant \frac{\alpha}{N^2}$, we get that the distance from $1$ to any $x_j$ exceeds $\alpha/N^2$, and thus the interval $[1-\frac{\alpha}{N^2}, 1]$ is disjoint from $X_N$. The probability of this event is less than $$ \left(1-C_5\frac{\sqrt{\alpha}}{N}\right)^N \leqslant e^{-C_5\sqrt{\alpha}}. $$ Thus, for any $\alpha$, $1<\alpha<N^2$, it follows that $$\P\left(\hat{\rho}(X_N)\geqslant \frac{\alpha}{N^2}\right)\leqslant e^{-C_5\sqrt{\alpha}}. $$ In particular, for sufficiently large $C_6$ we have $$\P\left(\hat{\rho}(X_N)\geqslant \frac{C_6\log^2(N)}{N^2}\right)\leqslant N^{-3}. $$ Therefore, $$ \mathbb{E}\hat{\rho}(X_N)\leqslant \frac{1}{N^2} + \sum\limits_{\alpha=1}^{C_6\log^2(N)}\frac{\alpha+1}{N^2}e^{-C_5\sqrt{\alpha}} + N^{-3}. $$ It is easy to see that the latter expression is bounded by $C_7/N^2$, which completes the proof for this case. \subsubsection{Case $0<a<2$} We again notice that, if $\alpha$ is a number and $I=[\alpha, \alpha+\varepsilon]\subset [1-\frac{1}{N^a},1]$ is an interval of length $\varepsilon$, then $$ \mu(I)=\int\limits_{\alpha}^{\alpha+\varepsilon} \frac{dt}{\pi\sqrt{1-t^2}}\geqslant \frac{1}{\pi} \frac{\varepsilon}{\sqrt{1-\alpha^2}}\geqslant \frac{1}{\pi} \frac{\varepsilon}{\sqrt{1-(1-\frac{1}{N^a})^2}}\geqslant C_1 \varepsilon N^{\frac{a}2}. $$ Now consider $n$ intervals of length $\frac{1}{nN^a}$ (and thus having $\mu$-measure $\mu$ greater than $\frac{C_1}{nN^{\frac{a}2}}$) inside $[1-\frac{1}{N^a}, 1]$. As we have seen before, if $\hat{\rho}(X_N)>\frac{2}{nN^a}$, then for some $y\in [1-\frac{1}{N^a}, 1]$ the interval of length $\frac{2}{nN^a}$ centered at $y$ is disjoint from $X_N$; thus one of the fixed intervals of length $\frac{1}{nN^a}$ is disjoint from $X_N$. Consequently, $$ \P\left(\hat{\rho}(X_N)\geqslant \frac2{nN^a}\right)\leqslant n\left(1-\frac{C_1}{nN^{\frac{a}{2}}}\right)^N. $$ With $$ n:=\frac{N^{1-\frac{a}{2}}}{A\log N}, $$ where $A$ large enough, we get $$ \P\left(\hat{\rho}(X_N)\geqslant \frac2{nN^a}\right)\leqslant n\left(1-\frac{C_1}{nN^{\frac{a}{2}}}\right)^N \leqslant N^{-3}. $$ Therefore, $$ \mathbb{E}\hat\rho \leqslant C\frac{\log N}{N^{1+\frac{a}2}} + N^{-3}, $$ which finishes the estimate from above. \bigskip For the estimate from below we notice that if $I=[\alpha, \alpha+\varepsilon]\subset [1-\frac{1}{N^a},1-\frac{1}{2N^a}]$, then $$ \mu(I)\leqslant C_2 \frac{\varepsilon}{\sqrt{1-(1-\frac{1}{2N^a})^2}}\leqslant C_2\frac{\varepsilon}{N^\frac{a}{2}}. $$ Take $n$ intervals in $[1-\frac{1}{N^a},1-\frac{1}{2N^a}]$ of length comparable to $\frac{1}{nN^a}$ and having equal $\mu$-measures $C_3\frac{1}{nN^\frac{a}2}$ (notice that if we are allowed to take such intervals near $1$, then the best measure we can get is $\frac{1}{\sqrt{nN^a}}$). If one of them is disjoint from $X_N$, then $\hat{\rho}(X_N)\geqslant \frac{C_4}{nN^a}$. Thus, $$ \P\left(\hat\rho(X_N)\geqslant \frac{C_4}{nN^a}\right)\geqslant f\left(N, nN^{\frac{a}2}/C_3, n\right). $$ It is easy to see that if we take $$ n:=\frac{N^{1-\frac{a}2}}{A\log N-B\log\log N} $$ for suitable $A$ and $B$, then the latter expression tends to one. Recall that $0<a<2$. Therefore, for large values of $N$ we have $$ \P\left(\hat\rho(X_N) \geqslant C_4\frac{\log N}{N^{1+\frac{a}2}}\right) \geqslant \frac 12, $$ which completes the proof for this case. \subsubsection{The estimate for $\tilde\rho$} For the estimate from above simply notice that for any interval $I$ we have $\mu(I)\geqslant |I|$. For the estimate from below take the interval $[-\frac12, \frac12]$. For any interval $I\subset [-\frac12, \frac12]$ we have $\mu(I)\leqslant C|I|$, and thus the estimate from below runs as usual. \hfill $\Box$
{ "timestamp": "2015-04-14T02:09:57", "yymm": "1504", "arxiv_id": "1504.03029", "language": "en", "url": "https://arxiv.org/abs/1504.03029", "abstract": "We derive fundamental asymptotic results for the expected covering radius $\\rho(X_N)$ for $N$ points that are randomly and independently distributed with respect to surface measure on a sphere as well as on a class of smooth manifolds. For the unit sphere $\\mathbb{S}^d \\subset \\mathbb{R}^{d+1}$, we obtain the precise asymptotic that $\\mathbb{E}\\rho(X_N)[N/\\log N]^{1/d}$ has limit $[(d+1)\\upsilon_{d+1}/\\upsilon_d]^{1/d}$ as $N \\to \\infty $, where $\\upsilon_d$ is the volume of the $d$-dimensional unit ball. This proves a recent conjecture of Brauchart et al. as well as extends a result previously known only for the circle. Likewise we obtain precise asymptotics for the expected covering radius of $N$ points randomly distributed on a $d$-dimensional ball, a $d$-dimensional cube, as well as on a 3-dimensional polyhedron (where the points are independently distributed with respect to volume measure). More generally, we deduce upper and lower bounds for the expected covering radius of $N$ points that are randomly and independently distributed on a metric measure space, provided the measure satisfies certain regularity assumptions.", "subjects": "Probability (math.PR)", "title": "The covering radius of randomly distributed points on a manifold", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9766692298333415, "lm_q2_score": 0.8198933293122506, "lm_q1q2_score": 0.80076458648489 }
https://arxiv.org/abs/2302.06290
Automorphisms of valued Hahn groups
Hahn groups endowed with the canonical valuation play a fundamental role in the classification of valued abelian groups. In this paper we study the group of valuation (respectively order) preserving automorphisms of a Hahn group $G$. Under the assumption that $G$ satisfies some lifting property, we prove a structure theorem decomposing the automorphism group into a semidirect product of two notable subgroups. We characterise a class of Hahn groups satisfying the aforementioned lifting property. For some special cases we provide a matrix description of the automorphism group.
\section{Introduction}\label{sec:introduction} In \cite{conrad}, Conrad describes the group of order preserving automorphisms of a Hahn sum as a certain group of matrices. Droste and G\"obel \cite{droste-goebel} extend this approach to balanced Hahn sum (see Definition \ref{def-hahn-sum-prod}). Hofberger \cite{hofbergerthesis} describes the group of automorphisms of Hahn fields as a semidirect product of so-called internal and external automorphisms. Building up on Hofberger's work we completed the description of the group of valuation preserving automorphisms of Hahn fields \cite{ours-fields}. In this paper we extend the approach of \cite{ours-fields} to the study of $ v\text{-}\Aut G $ and $ o\text{-}\Aut G $ for a Hahn group $ G $. We introduce the notion of (canonical) lifting property for a general Hahn group and, based on \cite{KKS-rayner-structures}, we prove that this is satisfied by a relatively large class of Hahn groups. We then decompose $ v\text{-}\Aut G $ (resp. $ o\text{-}\Aut G $) for Hahn groups satisfying the canonical lifting property. This extends the main result obtained in \cite{droste-goebel} (for the special case of balanced Hahn sums), to this large class of Hahn groups. The paper is organised as follows. In Section~\ref{sec:definitions-and-notation} we provide the definitions and notations that will be needed. In Section~\ref{sec:the-lifting-property} we introduce the important notion of \email{lifting property}. Subsection~\ref{sec:lifting-property-with-respect-to-the-rank} focusses on order preserving automorphisms and on the lifting property with respect to the rank (Definition~\ref{def:lifting-property-rank}). Theorem~\ref{lifting-principal-convex-sbgps} characterises automorphisms of the rank that lift by means of principal convex subgroups. We use this result to show that the lifting property with respect to the rank is not preserved by taking lexicographic sums (Example~\ref{es:lifting-not-preserved-lex-sums}). In Subsection~\ref{section-lifting-skeleton} we study valuation preserving automorphisms and we introduce the lifting property with respect to the skeleton (Definition~\ref{def:lifting-property-skeleton}). The main result of the section is Theorem~\ref{hofberger-groups}, which provides a semidirect product decomposition of the $ v\text{-}\Aut G $ of a Hahn group $ G $ satisfying the lifting property w.r.t. the skeleton. In Subsection~\ref{sec:rayner-groups}, based on \cite{KKS-rayner-structures}, we introduce Rayner groups (Definition~\ref{def:rayner-group}) and characterise those that satisfy the lifting property w.r.t. the skeleton (Theorem~\ref{prop:rayner-groups-canonical-stable}). Section~\ref{sec:matrices} focusses on Hahn sums, providing a description of the group of order preserving automorphism as a group of matrices, generalising results of \cite{conrad} and \cite{droste-goebel}. Proposition~\ref{prop:oAutGamma-G-direct product} gives a decomposition of the subgroup of order preserving automorphisms that induce the identity on the rank while Theorem~\ref{thm:DG-semidirect-enhanced} provides the matrix group representation. \section{Definitions and notation}\label{sec:definitions-and-notation} Let $ \Gamma $ be a (totally) ordered set and for all $ \gamma\in\Gamma $ let $ A_\gamma $ be an (divisible ordered) abelian group. For an element $ a = (a_\gamma)_{\gamma\in\Gamma} \in \prod_{\gamma\in\Gamma}A_\gamma $ we define the \emph{support} of $ a $ to be the set $ \supp a = \{\gamma\in\Gamma : a_\gamma\neq 0\} $. \begin{defn}[Hahn group] \label{def-hahn-sum-prod} \begin{enumerate}[(i)] \item The \emph{Hahn sum} of the $ A_\gamma $'s, denoted by $ \coprod_{\gamma\in\Gamma}A_\gamma $ is the set of elements of the full cartesian product with finite support: \[ \coprod_{\gamma\in\Gamma}A_\gamma = \left\lbrace a\in\prod_{\gamma\in\Gamma}A_\gamma : |\supp a|<\infty \right\rbrace. \] \item The \emph{Hahn product} of the $ A_\gamma $'s, denoted by $ \HH_{\gamma\in\Gamma}A_\gamma $ is the set of elements of the full direct product with well ordered support: \[ \HH_{\gamma\in\Gamma}A_\gamma = \left\lbrace a\in\prod_{\gamma\in\Gamma}A_\gamma : \supp a \text{ is well ordered} \right\rbrace. \] The Hahn sum is a subgroup of the Hahn product. \item A group comprised between them will be called a \emph{Hahn group}. \item If for all $ \gamma\in\Gamma $ we have $ A_\gamma\simeq A $ for a given abelian group $A$ we say that $ G $ is a \emph{balanced Hahn group}. \end{enumerate} \end{defn} \begin{notation} We will use the following notation for the elements of $ G $: \[ \sum_{\gamma\in\Gamma}a_\gamma\1_\gamma := (a_\gamma)_{\gamma\in\Gamma}. \] \end{notation} \begin{defn}\label{der:skeleton}\label{def:rank} Let $ \coprod_{\gamma\in\Gamma}A_\gamma \leq G \leq \HH_{\gamma\in\Gamma}A_\gamma $ be a Hahn group. \begin{enumerate}[(i)] \item The pair $ [\Gamma;\{A_\gamma:\gamma\in\Gamma\}] $ will be called the \emph{skeleton} of $ G $ and denoted by $ S(G) $. \item If $ G $ is balanced, with $ A_\gamma\simeq A $ for all $ \gamma\in\Gamma $, we will write $ S(G) = [\Gamma; A] $. \item In the sequel we will also refer to $ \Gamma $ as the \emph{rank of} $ G $ (sometimes denoted by $ \rk G $) and to the groups $ A_\gamma $ as the \emph{components of} $ G $. \end{enumerate} \end{defn} \begin{defn} Let $ G_1,G_2 $ be Hahn groups and let $ S(G_i) = [\Gamma_i:\{ A^i_\gamma : \gamma\in\Gamma_i \}],\ i=1,2 $ be the respective skeletons. An \emph{isomorphism} $ \tau:S(G_1)\rightarrow S(G_2) $ consists of an isomorphism $ \tau_\Gamma:\Gamma_1\rightarrow\Gamma_2 $ of ordered sets, and for all $ \gamma\in\Gamma_1 $ an isomorphism $ \tau_\gamma\colon A^1_\gamma\rightarrow A^2_{\tau_\Gamma(\gamma)} $ of ordered groups. We will also write $ \tau = [\tau_\Gamma;\{\tau_\gamma:\gamma\in\Gamma_1\}]. $ An automorphism of $ S(G) $, for a Hahn group $G$, is hence defined in the obvious way and we will denote by $ \Aut S(G) $ the group of automorphisms of $ S(G) $ under composition. \end{defn} Notice that, an automorphism of the skeleton is, in particular, an automorphism of the rank. However, there can be automorphism of the rank that are not part of any automorphism of the skeleton: \begin{es} Let $ G = \coprod_\mathbb Z A_n $ where $ A_{2n} = \mathbb Q $ and $ A_{2n+1} = \mathbb R $ for all $ n\in\mathbb Z $. Then $\sigma_\mathbb Z\colon n\mapsto n+1 $ is an automorphism of the chain $ \mathbb Z $ that cannot appear in any automorphism of the skeleton, as there is no isomorphism between $ A_n $ and $ A_{n+1} $, for any $ n $. \end{es} For the remaining of this section, let us fix a Hahn group $ G $ with skeleton $ S(G) = [\Gamma;\{A_\gamma:\gamma\in\Gamma\}] $. \begin{defn}\label{def:valuation-and-ordering} Since all elements of $G$ have well ordered support, we can define a valuation with value set $ \Gamma $ by \[ v_{\min}\colon G \longrightarrow \Gamma\cup\{\infty\},\quad v_{\min}(0)=\infty,\ v_{\min}(g) = \min\supp(g) \text{ for } g\neq 0. \] We will call $ v_{\min} $ the \emph{canonical valuation on} $ G $ and if the context is clear we will simply denote it by $ v $. \end{defn} Unless otherwise stated we will always consider $ G $ endowed with its canonical valuation $ v $. \begin{defn}\label{def:val-pres} An automorphism $ \sigma $ of $ G $ is said to be \emph{valuation preserving} if there exists a chain automorphism $ \sigma_{\Gamma} $ of $ \Gamma $ such that, for all $ g\in G $, we have $ \sigma_{\Gamma}(v(g)) = v(\sigma(g)) $. The valuation preserving automorphisms for ma group under composition that we will denote by $ v\text{-}\Aut G $. \end{defn} \begin{rmk}[{\cite[Lemma~0.1]{salma-monograph}}] \label{rmk:induced-autom} An automorphism $ \sigma\in\Aut G $ is valuation preserving if and only if the map $ \sigma_\Gamma\colon\Gamma\rightarrow\Gamma,\ v(x)\mapsto v(\sigma(x)) $ is a well defined automorphism of chains. We will say that $ \sigma_{\Gamma} $ is the chain automorphism \emph{induced on $ \Gamma $ by $ \sigma $}.\qed \end{rmk} \begin{defn}\label{def:val-pres-iso} More generally, let $ G_1,G_2 $ be Hahn groups with skeletons $ S(G_i) = [\Gamma_i:\{ A^i_\gamma : \gamma\in\Gamma_i \}],\ i=1,2 $ and let $ v_1,v_2 $ be the canonical valuations on $ G_1 $ and $ G_2 $, respectively. An isomorphism $ \rho\colon G_1\to G_2 $ is said to be \emph{valuation preserving} if there exists an isomorphism of chains $ \rho_\Gamma\colon\Gamma_1\to\Gamma_2 $ such that, for all $ g\in G_1 $ we have $ \rho_\Gamma(v_1(g)) = v_2(\rho(g)) $. \end{defn} \begin{lemma}[{\cite[Lemma~0.2]{salma-monograph}}] \label{lemma:induced-autom-skeleton} Let us keep the notation of Definition~\ref{def:val-pres-iso}. A valuation preserving isomorphism $ \sigma\colon G_1\rightarrow G_2 $ of Hahn groups naturally induces an isomorphism of the skeletons $ [\sigma_\Gamma;\{\sigma_\gamma:\gamma\in\Gamma_1\}] \colon S(G_1)\rightarrow S(G_2) $ by: $ \sigma_\Gamma(v_1(g)) = v_2(\sigma(g)) $ for all $ v_1(g)\in\Gamma_1 $ and for each $ \gamma\in\Gamma_1 $ and each $ g_\gamma\in A^1_\gamma $ set $ \sigma_\gamma(g_\gamma) = \sigma(g)_{\sigma_\Gamma(\gamma)} $, for some $ g\in G_1 $ with $ v_1(g)=\gamma $. In particular, a valuation preserving automorphism $ \sigma\inv\text{-}\Aut G $ induces an automorphism $ \sigma_S = [\sigma_\Gamma;\{\sigma_\gamma:\gamma\in\Gamma\}]\in\Aut S(G) $ of the skeleton where $ \sigma_\Gamma\ino\text{-}\Aut \Gamma $ is defined as in Remark~\ref{rmk:induced-autom} and $ \sigma_\gamma\colon A_\gamma\to A_{\sigma_\Gamma(\gamma)} $ is given by $ \sigma_\gamma(a) = (\sigma(a\1_\gamma))_{\sigma_\Gamma(\gamma)} $. \qed \end{lemma} \begin{defn}\label{def:lex-order}\label{def:oAut} Assume all the components $ A_{\gamma} $ to be ordered groups and denote by $ <_\gamma $ the order relation on $ A_\gamma $. then we can order $ G $ lexicographically by setting, for all $ g \in G $ \[g>_{\lex}0\iff g_{v(g)}>_{v(g)}0. \] An automorphism $ \sigma $ is said to be \emph{order preserving} if for all $ g\in G $ we have $ g>0\Rightarrow \sigma(g)>0 $. The order preserving automorphisms also form a group under composition which will be denoted by $ o\text{-}\Aut G $. \end{defn} \begin{rmk}\label{rmk:oAut<vAut} Assume all the components $ A_\gamma $ to be ordered \textit{archimedean} groups. Then an order preserving automorphism of $ G $ is automatically valuation preserving. So we have $ o\text{-}\Aut G\leq v\text{-}\Aut G\leq \Aut G. $\qed \end{rmk} \section{The lifting property}\label{sec:the-lifting-property} Let $ G $ be a Hahn group with skeleton $ S(G) = [\Gamma;\{A_\gamma:\gamma\in\Gamma\}] $. In Section~\ref{sec:definitions-and-notation} we saw that a valuation preserving automorphism $ \sigma\inv\text{-}\Aut G $ induces an automorphism $\sigma_S\in\Aut S(G) $ of the skeleton (Lemma~\ref{lemma:induced-autom-skeleton}) and, in particular, an automorphism $ \sigma_\Gamma\ino\text{-}\Aut \Gamma $ of the rank (Remark~\ref{rmk:induced-autom}). In this section we will study under what conditions, given an automorphism of the skeleton (resp. rank) we can find an automorphism of the group that induces it. In Section~\ref{sec:lifting-property-with-respect-to-the-rank} we will study automorphisms of the chain $ \Gamma $ that lift to \emph{order preserving} automorphisms of $ G $ and in Section~\ref{section-lifting-skeleton} automorphisms of $ S(G) $ that lift to \emph{valuation preserving} automorphisms of $ G $. \subsection{Lifting from the rank to $ o\text{-}\Aut G $}\label{sec:lifting-property-with-respect-to-the-rank} Let $G$ be a Hahn group with skeleton $ S(G) = [\Gamma;\{A_\gamma:\gamma\in\Gamma\}] $ and let $ \bbG = \HH_\Gamma A_\gamma $ be the corresponding Hahn product. In this section we will assume that all the components $ A_\gamma $ of the skeleton of $ G $ be archimedean ordered abelian groups and we will consider the induced lexicographic order on $ G $ (Definition~\ref{def:lex-order}). Let $\sigma \in o\text{-}\Aut G$. Then, by Remark~\ref{rmk:oAut<vAut}, $ \sigma\inv\text{-}\Aut G $ and therefore, it induces an automorphism $\sigma_{\Gamma}\ino\text{-}\Aut \Gamma $ given by $\sigma_{\Gamma}(v(g)) = v(\sigma(g))$ (Remark~\ref{rmk:induced-autom}). This gives rise to a group homomorphism \begin{equation}\label{eq:phi-rank} \Phi_\Gamma:o\text{-}\Aut G \longrightarrow o\text{-}\Aut \Gamma,\ \sigma\longmapsto \sigma_\Gamma. \end{equation} It is straightforward to verify that this is a group homomorphism. \begin{defn} \label{def:rank-int} The kernel $ \ker\Phi_\Gamma $ of the map \eqref{eq:phi-rank} is a normal subgroup of $ o\text{-}\Aut G $ that we call the \emph{subgroup of internal o-automorphisms} of $G$. We will denote it by $\Into\text{-}\Aut G$. \end{defn} \begin{defn}\label{def:lifting-property-rank} We say that an automorphism $\tau \in o\text{-}\Aut\Gamma$ \emph{lifts to $ G $} if there exists an automorphism $\sigma\ino\text{-}\Aut G$ such that $\Phi_\Gamma(\sigma) = \tau$. If the map $ \Phi_\Gamma $ defined in \eqref{eq:phi-rank} admits a section, i.e., an injective group homomorphism $ \Psi_\Gamma\colon o\text{-}\Aut \Gamma\hookrightarrow o\text{-}\Aut G $ such that $ \Phi_\Gamma\circ\Psi_\Gamma = \id_{o\text{-}\Aut \Gamma} $, we say that $ G $ has the \emph{lifting property with respect to $ \Psi_\Gamma $}. \end{defn} \begin{defn} \label{def:rank-ext} Assume $ G $ has the lifting property with respect to a fixed section $ \Psi_\Gamma $ of $ \Phi_\Gamma $. The subgroup $ \Psi_\Gamma(o\text{-}\Aut G) $ of $ o\text{-}\Aut G $ will be called the subgroup of $ \Psi_\Gamma $\emph{-external automorphisms of $G$} and denoted by $\pExt{\Psi_\Gamma}o\text{-}\Aut G$. \end{defn} \begin{defn}\label{def:can-lift-prop-rank} Assume $ G $ is balanced, say $ A_\gamma = A $ for all $ \gamma $. Then for the Hahn product $ \bbG = \HH_\Gamma A $ there exists always the \emph{canonical section} \[ \Psi^c_\Gamma\colon o\text{-}\Aut \Gamma\to o\text{-}\Aut \bbG,\ \tau\longmapsto \tilde\tau \] where $ \tilde\tau $ is defined by \[ \tilde\tau\left( \sum_\gamma g_\gamma\1_\gamma\right) = \sum_\gamma g_\gamma\1_{\tau(\gamma)}. \] We call $ \tilde\tau $ the \emph{canonical lift of $ \tau $ to $ \bbG $.} \noindent If, moreover, $ \tilde\tau(G) = G $ we say that $ \tau $ \emph{lifts canonically} to an automorphism of $ G $ and call $ \tilde\tau|_G $ the \emph{canonical lift} of $ \tau $ to $ G $. If all automorphisms of $ \Gamma $ lift canonically we say that $ G $ has the \emph{canonical lifting property with respect to} $ \Gamma $. \end{defn} Now we give a characterisation of the automorphisms of the rank $ \Gamma $ that lift to an order-preserving automorphism or $ G $ by means of its principal convex subgroups (Definition~\ref{def:princ-convex-sbg}). We do this in Theorem~\ref{lifting-principal-convex-sbgps}, after proving a preliminary lemma. \begin{defn}\label{def:princ-convex-sbg} For an element $ \gamma\in\Gamma $, the \emph{principal convex subgroup of $ G $ associated to $ \gamma $} is the subgroup $ C_\gamma := \{x\in G : v(x)\geq \gamma \} $. \end{defn} \begin{lemma} \label{h-commutes-C} Let $\sigma\inv\text{-}\Aut G$ and let $g\in G$ and $C_g := \{x\in G:\ v(x)\geq v(g)\}$. Then we have: $$\sigma(C_g) = C_{\sigma(g)} = \{x:\ v(x) \geq v(\sigma(g))\}.$$ \begin{proof} Let $ \sigma_\Gamma$ be the automorphism of the chain $\Gamma$ induced by $\sigma$. Let $x\in C_g$. Then $$v(x)\geq v(g) \Rightarrow v(\sigma(x)) = \sigma_\Gamma (v(x)) \geq \sigma_\Gamma (v(g)) = v(\sigma(g))$$ so $\sigma(x) \in C_{\sigma(g)}$, that is $\sigma(C_g) \subseteq C_{\sigma(g)}$. Vice versa, let $x\in C_{\sigma(g)}$. Then $$ v(x) \geq v(\sigma(g)) \Rightarrow v(\sigma^{-1}(x)) = \tilde \sigma^{-1}(v(x)) \geq \sigma_\Gamma^{-1}(v(\sigma(g)))= v(\sigma^{-1}(\sigma(g)))=v(g) $$ because $\sigma_\Gamma^{-1}$ is also an automorphism of $\Gamma$. So $\sigma^{-1}(x) \in C_g$, that is $x\in \sigma(C_g)$. Hence $\sigma(C_g) \supseteq C_{\sigma(g)}$, which completes the proof. \end{proof} \end{lemma} \begin{theorem} \label{lifting-principal-convex-sbgps} Let $\tau\ino\text{-}\Aut\Gamma$. Then $\tau$ lifts to an automorphism $\sigma \in o\text{-}\Aut G$ if and only if, for all $\gamma\in\Gamma$ there is an isomorphism $$ \sigma_\gamma\colon C_\gamma \overset{\sim}{\longrightarrow} C_{\tau(\gamma)} $$ such that \begin{equation} \label{cond-gamma-delta} \gamma>\delta \Rightarrow \sigma_\delta|_{C_\gamma}=\sigma_\gamma. \end{equation} \end{theorem} \begin{proof} Assume that $\tau$ lifts to $ \sigma\ino\text{-}\Aut G $. Let $\gamma\in\Gamma$ and choose $g\in G$ such that $v(g) = \gamma$. Recall that, by definition, $v(\sigma(g)) = \tau (v(g))$. Then Lemma~\ref{h-commutes-C} gives $$ C_\gamma =C_g = \{x:\ v(x)\geq \gamma = v(g)\}\simeq C_{\sigma(g)} = \{x:\ v(x) \geq \tilde \sigma(v(g)) = \tau(\gamma)\} = C_{\tau(\gamma)}. $$ where the isomorphism is simply given by $\sigma$, which is order preserving, so condition \eqref{cond-gamma-delta} is satisfied. Vice versa, suppose that for all $\gamma\in\Gamma$ there is an isomorphism $$ \sigma_\gamma\colon C_\gamma \overset{\sim}{\longrightarrow} C_{\tau(\gamma)} $$ such that $\gamma>\delta \Rightarrow \sigma_\gamma(C_\gamma)\subset \sigma_\delta(C_\delta)$. We want to construct a lift $\sigma \in o\text{-}\Aut G$ of $\tau$. We define it as follows: for $x\in G$ with $v(x) = \gamma$, set $\sigma(x):=\sigma_\gamma(x)$. First let us show that $\sigma$ is a well defined group morphism. Let $x,y \in G$. We have two cases:\\ \textbf{Case 1.} $v(x)\neq v(y)$, and we can assume $v(x)<v(y)$. Then $v(x+y)=v(x) =:\gamma$. So we have $$ \sigma(x-y) = \sigma_\gamma (x-y) = \sigma_\gamma(x) - \sigma_\gamma(y) = \sigma(x) - \sigma_{v(y)}(y) = \sigma(x) - \sigma(y)$$ where in the second-last equality we used condition~\eqref{cond-gamma-delta}.\\ \textbf{Case 2.} $v(x) = v(y) = \delta$. Then $v(x+y) =: \gamma \geq \delta$. Then $$ \sigma(x-y) = \sigma_\gamma(x-y) = \sigma_\delta(x-y) = \sigma_\delta(x) - \sigma_\delta(y) = \sigma(x) - \sigma(y). $$ Again, for the second equality we used condition~\eqref{cond-gamma-delta}. To show that $\sigma$ is an automorphism, we define its inverse. Similarly to what we did to define $\sigma$ we define $\sigma^{-1}$ as follows: for $x\in G$ with $v(x) = \tau(\gamma)$ we set $\sigma^{-1}(x) = \sigma_\gamma^{-1}(x) $. Checking that this is a well defined group morphism is done the same way as it was done for $\sigma$. To see that this is indeed the inverse of $ \sigma $, let $x\in G$ with $v(x)=\gamma$. Recall that $\sigma_\gamma(x) \in C_{\tau(\gamma)}$ and $\sigma_\gamma^{-1}: C_{\tau(\gamma)}\rightarrow C_\gamma$. So $$ \sigma^{-1}(h(x)) = \sigma^{-1}(\sigma_\gamma(x)) = \sigma_\gamma^{-1}(\sigma_\gamma(x)) = x $$ and similarly $\sigma(\sigma^{-1}(x)) = x$. So $\sigma$ is an automorphism of $G$, which completes the proof. \end{proof} The following example shows that, in general, the lifting property is not preserved after taking lexicographic sums. \begin{es}\label{es:lifting-not-preserved-lex-sums} Let $\Gamma = \mathbb Q$ and $G = (\HH_{q<0}\mathbb Q)\sqcup \mathbb Q\sqcup (\coprod_{q>0}\mathbb Q)$. So $\Gamma = \rk G = \mathbb Q$. Consider the automorphism of the chain $\mathbb Q$ defined by $\tilde h (q)= q+2$ and assume that $\tilde h$ lifts to an automorphism $h$ of $G$. Let $g\in G$ be such that $v(g) = -1/2$. So $v(h(g)) = \tilde h(-\frac{1}{2})=\frac{3}{2}>1$. By Lemma~\ref{h-commutes-C} we have $C_g \simeq C_{h(g)}$. Now $C_{h(g)} = \{x:\ v(x) \geq \frac{3}{2}\}\subseteq\coprod_{q>0}\mathbb Q$ is countable. But $C_g = \{x:\ v(x)\geq -\frac{1}{2}\}$ is uncountable. To prove this, we show that the uncountable set $\{0,1\}^\mathbb N$ of functions $f:\mathbb N\rightarrow\{0,1\}$ can be embedded into $C_g$. We construct the embedding as follows: to each function $f:\mathbb N\rightarrow\{0,1\}$ we associate the element $f^* = (f^*_q)_{q<0}\in\HH_{q<0}\mathbb Q$ defined by $$ f^*_q = \begin{cases} f(n) \text{ if } q = -\frac{1}{2+n}\\ 0 \text{ otherwise.} \end{cases} $$ First, notice that, for all $f\in \{0,1\}^\mathbb N$, the element $f^*$ belongs to $C_g$: indeed we have $\supp f^* \subseteq \{-\frac{1}{2+n}|n\in\mathbb N\}$ which is well ordered, so $f^*\in\HH_{q<0}\mathbb Q$ and $v(f^*)\geq -\frac{1}{2}$. Finally, the map $f\mapsto f^*$ is injective, for if $f\neq g$ then there exists $n\in\mathbb N$ such that $$f^*_{-\frac{1}{2+n}} = f(n)\neq g(n) = g^*_{-\frac{1}{2+n}}.$$\qed \end{es} Last example shows that, even if two Hahn groups have the lifting property, after we take the lexicographic sum not all automorphism of the skeleton will lift. The next proposition characterises those that do. \begin{prop}\label{prop:condition-lifting-preserved-lex} Let $G_1,G_2$ be two groups with the lifting property (with respect to the rank) and consider $G =G_1\coprod G_2$. Then $\rk G=\rk G_1 + \rk G_2$. Let $\tau$ be an automorphism of $\rk G$ such that $\tau|_{\rk G_i}\ino\text{-}\Aut(\rk G_i) $ for $i=1,2$. Then $\tau$ lifts to an automorphism of $G$. \end{prop} \begin{proof} Let $ \Gamma:= \rk G $ and $ \Gamma_i :=\rk G_i $ for $ i=1,2 $. Let $ \tau \ino\text{-}\Aut \Gamma $ be such that $ \tau_i :=\tau|_{\Gamma_i}\in o\text{-}\Aut\Gamma_i $ for $ i=1,2 $. Since both $ G_i $'s have the lifting property, let $ \sigma_i\inv\text{-}\Aut G_i $ be a lift of $ \tau_i $ and define $ \sigma\colon G\rightarrow G $ through $ \sigma(g_1\1_1 + g_2\1_2):=\sigma_1(g_1)\1_1+\sigma_2(g_2)\1_2 $. It is straightforward to verify that this is an automorphism of $ G $. \end{proof} \subsection{Lifting from the skeleton to $ v\text{-}\Aut G $} \label{section-lifting-skeleton} As a generalisation of the problem discussed in Section~\ref{sec:lifting-property-with-respect-to-the-rank}, we can consider the \emph{lifting property with respect to the skeleton}, and ask under what conditions an automorphism of the skeleton of a Hahn group lifts to an automorphism of the group. Recall that (Lemma~\ref{lemma:induced-autom-skeleton}) a valuation preserving automorphism $ \sigma\inv\text{-}\Aut G $ induces an automorphism $ \sigma_S \in\Aut S(G) $ of the skeleton. This gives rise to a map, that we will denote by $ \Phi_S $, in analogy to the one in \eqref{eq:phi-rank}: \begin{equation}\label{eq:phi-skeleton} \Phi_S\colon v\text{-}\Aut G \longrightarrow \Aut S(G),\ \sigma\longmapsto[\sigma_\Gamma;\{\sigma_{\gamma}:\gamma\in\Gamma\}]. \end{equation} It is straightforward to verify that $ \Phi_S $ is a group homomorphism. \begin{defn}\label{def:internal} The kernel of the map $ \Phi_S $ defined in \eqref{eq:phi-skeleton} is a normal subgroup of $ v\text{-}\Aut G $ called \emph{the group of internal v-automorphisms} and denoted by $ \Intv\text{-}\Aut G $. If the context is clear we will drop the ``$ v $'' from the notation and terminology. \end{defn} \begin{rmk}\label{rmk:internal-by-valuation} An automorphism $\sigma\inv\text{-}\Aut G$ is internal if and only if, for all $g\in G$, we have \begin{equation} \label{eq:internal-by-valuation} v(g)=v(\sigma(g))\quad\text{ and }\quad g_{v(g)}=\sigma(g)_{v(g)}. \end{equation} Indeed, conditions \eqref{eq:internal-by-valuation} are equivalent to $ \Phi(\sigma)=\id_{S(G)} $. \qed \end{rmk} \begin{defn}\label{def:lifting-property-skeleton} We say that an automorphism $\tau \in \Aut S(G)$ \emph{lifts to $ G $} if there exists an automorphism $\sigma\inv\text{-}\Aut G$ such that $\Phi_S(\sigma) = \tau$. If the map $ \Phi_S $ defined in \eqref{eq:phi-skeleton} admits a section, i.e., an injective group homomorphism $ \Psi_S\colon \Aut S(G)\hookrightarrow v\text{-}\Aut G $ such that $ \Phi_S\circ\Psi_S = \id_{\Aut S(G)} $, we say that $ G $ has the \emph{lifting property with respect to $ \Psi_S $}. \end{defn} \begin{defn} \label{def:skeleton-ext} Assume $ G $ has the lifting property with respect to a fixed section $ \Psi_S $ of $ \Phi_S $. The subgroup $ \Psi_S(\Aut S(G)) $ of $ v\text{-}\Aut G $ will be called the subgroup of $ \Psi_S $\emph{-external automorphisms of $G$} and denoted by $\pExt{\Psi_S}\Aut G$. If no confusion can arise we will just write $ \Ext\Aut G $. \end{defn} \begin{rmk} If $ G $ has the lifting property with respect to a given section $ \Psi_S $ then the homomorphism $ \Phi_S $ defined in \eqref{eq:phi-skeleton} is surjective and every $ \tau\in\Aut S(G) $ lifts to an automorphism $ \sigma\inv\text{-}\Aut G $. The first isomorphism theorem thus yields \[ \Aut S(G)\simeq\frac{v\text{-}\Aut G}{\Int\Aut G}. \] \qed\end{rmk} In \cite[Section~3]{ours-fields} we generalise a result of Hofberger \cite[Satz~2.2]{hofbergerthesis}, providing a decomposition of the group of valuation preserving automorphisms of a Hahn field with the lifting property into the semidirect product of its subgroups of internal and external automorphisms. Here we generalise this result further to Hahn groups with the lifting property. \begin{theorem} \label{hofberger-groups} Let $G$ be a Hahn group with the lifting property with respect to a given section $ \Psi_S $ of $ \Phi_S $. Then we can decompose $ v\text{-}\Aut G $ into the (inner) semidirect product of its subgroups of internal and external automorphisms: $$v\text{-}\Aut G = \Int\Aut G \rtimes \Ext\Aut G.$$ \end{theorem} \begin{proof} Let us consider the sequence $$ \xymatrix{ \Int\Aut G \ar[r]^-\iota & v\text{-}\Aut G \ar[r]^-{\Phi_S} & \Aut S(G) \ar@{-->}@/^1pc/[l]^{\Psi_S} } $$ where $\iota$ is the canonical embedding. By definition of $ \Int\Aut G $, we have $\im \iota = \ker\Phi$, so the sequence is exact. Therefore (see \cite[Theorem~3.3]{conrad-note-splitting}) we have \[ v\text{-}\Aut G = \im \iota \rtimes \im \Psi = \ker\Phi\rtimes \im\Psi = \Int\Aut G \rtimes \Ext\Aut G. \] \end{proof} Now that we have established a decomposition result, we want to further investigate the two subgroups $\Int\Aut G$ and $\Ext\Aut G$. The latter is isomorphic to $ \Aut S(G) $. The former is a bit more mysterious. \subsection{Rayner groups and the canonical lifting property}\label{sec:rayner-groups} In Section~\ref{section-lifting-skeleton} we introduced the lifting property with respect to the skeleton. Now we are going to define a stronger lifting property and study some distinguished classes of Hahn groups that fulfil it. \begin{defn}\label{def:can-lift-prop-skeleton} For the Hahn product $ \bbG = \HH_\Gamma A_\gamma $ there exists always the \emph{canonical section} \[ \Psi^c_S\colon \Aut S(G)\to v\text{-}\Aut \bbG,\ \tau = [\tau_\Gamma;\{\tau_\gamma : \gamma\in\Gamma \}]\longmapsto \tilde\tau \] where $ \tilde\tau $ is defined by \[ \tilde\tau\left( \sum_\gamma g_\gamma\1_\gamma\right) = \sum_\gamma \tau_\gamma(g_\gamma)\1_{\tau_\Gamma(\gamma)}. \] We call $ \tilde\tau $ the \emph{canonical lift of $ \tau $ to $ \bbG $.} \noindent If, moreover, $ \tilde\tau(G) = G $ we say that $ \tau $ \emph{lifts canonically} to an automorphism of $ G $ and call $ \tilde\tau|_G $ the \emph{canonical lift of $ \tau $ to $ G $}. If all automorphisms of $ S(G) $ lift canonically we say that $ G $ has the \emph{canonical lifting property with respect to} $ S(G) $. \end{defn} As announced above, we are going to introduce a special class of Hahn groups that have the canonical lifting property. A family $ {\mathcal F} $ of subsets of $ \Gamma $ is said to be a \emph{group family with respect to} $ \Gamma $ (cf. \cite[Paragraph 2]{rayner1968}) if the following properties are satisfied: \begin{enumerate}[(i)] \item the elements of $ {\mathcal F} $ are well ordered subsets of $ \Gamma $; \item $ A,B\in{\mathcal F} \Rightarrow A\cup B\in{\mathcal F} $; \item $ A\in{\mathcal F}\wedge B\subseteq A \Rightarrow B\in{\mathcal F} $. \end{enumerate} \begin{prop}[{\cite[Theorem~1]{rayner1968}}]\label{prop:rayner-family-yield-groups} If $ {\mathcal F} $ is a group family then the set $ \bbG({\mathcal F}) $ of elements of $ \bbG $ whose supports belong to $ {\mathcal F} $ is a subgroup of $ \bbG $. \end{prop} \begin{proof} By property (i) all elements of $ {\mathcal F} $ are well ordered so they can occur as supports of elements of $ \bbG $. Since for all $ g\in \bbG $ we have $ \supp g = \supp(-g) $ then $ g\in\bbG({\mathcal F}) $ implies $ -g\in\bbG({\mathcal F}) $. If $ g,h\in\bbG({\mathcal F}) $, then $ \supp(g+h)\subseteq \supp g \cup \supp h $: the latter is in $ {\mathcal F} $ by (ii) and hence by (iii) the former is too. So $ g+h\in\bbG({\mathcal F}) $. It follows then that also $ 0\in\bbG({\mathcal F}) $ so $ \bbG({\mathcal F}) $ is indeed a subgroup of $ \bbG $. \end{proof} \begin{defn}\label{def:rayner-group} Groups obtained as in Proposition~\ref{prop:rayner-family-yield-groups} will be called \emph{Rayner groups}. These generalise an analogous notion (Rayner fields) introduced by Rayner in \cite{rayner1968} in the context of Hahn fields. Further \emph{Rayner structures} are introduced and studied in \cite{KKS-rayner-structures}. \end{defn} The next result gives a necessary and sufficient condition for a Rayner group to satisfy the canonical lifting property. \begin{theorem}\label{prop:rayner-groups-canonical-stable} Let $ \bbG $ be the Hahn product over the skeleton $ [\Gamma;\{ A_\gamma:\gamma\in\Gamma \}] $. Let $ {\mathcal F} $ be a group family with respect to $ \Gamma $. Then the Rayner group $ G = \bbG({\mathcal F}) $ has the canonical lifting property with respect to the skeleton if and only if $ {\mathcal F} $ is stable under the action of $ \Aut S(G) $, that is, for all $ [\bar\tau;\{ \tau_\gamma:\gamma\in\Gamma \}]\in\Aut S(G) $ and for all $ T\in{\mathcal F} $ we have $ \bar\tau(T)=T $. \end{theorem} \begin{proof} Assume that $ {\mathcal F} $ be stable under $ \Aut S(G) $ and let $ g = \sum g_\gamma\1_\gamma\in G $. Let also $ \tau=[\bar\tau;\tau_\gamma] $ be an automorphism of $ S(G) $. Then $ \supp\left( \sum\tau_\gamma(g_\gamma)\1_{\bar\tau(\gamma)}\right) = \{ \bar\tau(\gamma):\gamma\in\supp g \} = \bar\tau(\supp g) $ which, by our assumption, belongs to $ {\mathcal F} $. Thus $ \sum\tau_\gamma(g_\gamma)\1_{\bar\tau(\gamma)} \in G $ and so $ G $ has the canonical lifting property with respect to the skeleton. Vice versa, assume that $ G $ has the canonical lifting property and let $ T\in{\mathcal F} $. Let $ g = \sum g_\gamma\1_\gamma \in G $ be such that $ \supp(g)=T $. Then, for all $ [\bar\tau;\tau_\gamma] $ we have $ h:= \sum\tau_\gamma(g_\gamma)\1_{\bar\tau(\gamma)}\in G $, so in particular $ \supp h \in {\mathcal F} $. So $ {\mathcal F} $ is stable under action of $ \Aut S(G) $. \end{proof} \begin{defn} Let $ \kappa $ be an infinite regular cardinal and let $ {\mathcal F}_\kappa $ be the family of all well ordered subsets of $ \Gamma $ of cardinality smaller than $ \kappa $. The $ \kappa $-bounded subgroup of $ \bbG $ consists of all elements of $ \bbG $ with support of cardinality less than $ \kappa $, and will be denoted by $ \bbG_\kappa $. \noindent Notice that the Hahn sum and the Hahn product are special cases of $ \kappa $-bounded subgroups, with $ \kappa = \aleph_0 $ and $ \kappa = |\bbG|^+ $ respectively. \end{defn} \begin{cor}\label{cor:k-bdd-have-canonical-lifting} Let $ \kappa $ be an infinite regular cardinal. Then $ \bbG_\kappa $ has the canonical lifting property. In particular, $ \bbG $ and $ \coprod_\Gamma A_\gamma $ have the canonical lifting property. \end{cor} \begin{proof} By Proposition~\ref{prop:rayner-groups-canonical-stable} it suffices to show that the family $ {\mathcal F}_\kappa $ of well ordered subsets of $ \Gamma $ is stable under $ \Aut S(G) $. Let $ [\bar\tau;\{ \tau_\gamma:\gamma\in\Gamma \}]\in\Aut S(G) $ and let $ T $ be a well ordered subset of $ \Gamma $ of cardinality less than $ \kappa $. Then $ \bar\tau(T) $ will have the same cardinality as $ T $, because $ \bar\tau $ is an automorphism of $ \Gamma $. Since it is also order preserving then $ \bar\tau(T) $ will also be well ordered, hence $ \bar\tau(T)\in{\mathcal F}_\kappa $. \end{proof} Analogous results to the ones contained in this section hold for Rayner fields: see \cite[Section~3.2]{ours-fields}. \section{Automorphism groups as groups of matrices}\label{sec:matrices} In this section we are going to generalise and sharpen results of Conrad \cite{conrad} and Droste-G\"obel \cite{droste-goebel}. The first describes the group of order preserving automorphisms of a Hahn sum as a certain group of triangular matrices. The latter obtain a semidirect product decomposition, which is different from the one we gave in Theorem~\ref{hofberger-groups}. Let $ \Gamma $ be a chain, $ \{A_\gamma:\gamma\in\Gamma\} $ a family of ordered abelian groups and $ G = \coprod_{\gamma\in\Gamma}A_\gamma $ the Hahn sum. Let $ \End A_\gamma $ be the ring of endomorphisms of $ A_\gamma $ (with pointwise addition and composition). For all $ \alpha,\beta\in\Gamma $ let $ H_{\alpha\beta} = \Hom(A_\beta,A_\alpha) $ be the group of homomorphisms from $ A_\beta $ into $ A_\alpha$ (with pointwise addition). Let $ \Delta $ be the set of all $ \Gamma\times\Gamma $-matrices $ (\sigma_{\alpha\beta}) $ where \begin{enumerate}[(i)] \item $ \sigma_{\alpha\alpha}\in\End A_\alpha $; \item $ \sigma_{\alpha\beta}\in H_{\alpha\beta} $; \item for every $ \beta $ and for all $ a\in A_\beta $ we have $ \sigma_{\alpha\beta}(a)=0 $ for all but finitely many $ \alpha $. \end{enumerate} Then $ \Delta $ forms a ring with respect to the usual matrix addition and multiplication (condition (iii) ensures that the product be well defined). \begin{rmk}\label{rmk:isom-End-Delta} There is an isomorphism between the rings $ \End G $ and $\Delta $. Indeed, for $ a=\sum a_\gamma\1_\gamma\in G $ and $ (\sigma_{\alpha\beta})\in\Delta $ we can consider the row vector $ (a_\gamma) $ and multiply it on the left to get \begin{equation}\label{eq:formula-matrix-multiplication} (a_\gamma)(\sigma_{\alpha\beta}) = \left(\sum_{\alpha\in\supp(a)} \sigma_{\alpha\beta}(a_\alpha)\right)_{\beta\in\Gamma} =: (b_\beta)_{\beta\in\Gamma} \end{equation} where the sum on the right hand side makes sense because the $ \supp (a) $ is finite. Then it is clear that the map $ \sigma: \sum a_\gamma\1_\gamma\mapsto \sum b_\gamma\1_\gamma $ is an endomorphism of $ G $ induced by $ (\sigma_{\alpha\beta}) $. Vice versa, if $ \sigma\in\End G $ and $ a=\sum a_\gamma\1_\gamma\in G $ let $ b=\sum b_\gamma\1_\gamma := \sigma(a) $. Then for all $ \alpha\in\Gamma $ define $ \sigma_{\alpha\beta}(a_\alpha):= b_\beta $. Since $ b\in G $ then $ \supp b $ is finite and therefore $ \sigma_{\alpha\beta} $ satisfies condition (iii) above. Hence $ (\sigma_{\alpha\beta})\in \Delta $ and it is easy to verify that the correspondence $ \sigma \mapsto (\sigma_{\alpha\beta}) $ is an isomorphism of $ \End G $ onto $ \Delta $. Thus the group $ \Aut G $ corresponds isomorphically to the group of units in $ \Delta $. Here by $ \Aut G $ we mean \emph{all} automorphisms of $ G $ as an abelian group, order (reps. valuation) preserving or otherwise. \qed \end{rmk} Now we see how we can characterise the order preserving automorphisms. Let us consider the multiplicative monoid $ T $ of all matrices $ (\sigma_{\alpha\beta}) \in\Delta $ such that \begin{enumerate}[(i')] \item $ (\sigma_{\alpha\beta}) $ is lower triangular; \item for all $ \alpha\in\Gamma $, $ \sigma_{\alpha\alpha}\ino\text{-}\Aut A_\alpha $. \end{enumerate} The order we consider on $ G $ is the lexicographic one, as always (cf. Definition~\ref{def:valuation-and-ordering}). \begin{lemma}\label{lemma:conrad-lemma0} A matrix $ (\sigma_{\alpha\beta})\in T $ induces, via the correspondence described in Remark~\ref{rmk:isom-End-Delta}, an order preserving endomorphism $ \sigma $ on $ G $. In particular, an invertible matrix in $ T $ induces an order preserving automorphism of $ G $. \end{lemma} \begin{proof} Let $ a = \sum_\gamma a_\gamma\1_\gamma >0 $ with $ v(a)=\gamma $. Then \begin{align*} \sigma(a) = (a)_\alpha(\sigma_{\alpha\beta}) &= \left(\sum_{\alpha\geq\gamma}\sigma_{\alpha\beta}(a_\alpha) \right)_{\beta\in\Gamma} \\&= \left(\sum_{\alpha\geq\gamma}\sigma_{\alpha\beta}(a_\alpha) \right)_{\beta\geq\gamma} = \sigma_{\gamma\gamma}(a_\gamma)+\left(\sum_{\alpha\geq\gamma}\sigma_{\alpha\beta}(a_\alpha) \right)_{\beta>\gamma}>0 \end{align*} where the third equality holds because $ \sigma_{\alpha\beta} $ is lower triangular and the final inequality is due to the fact that, by (ii'), $ \sigma_{\gamma\gamma} $ is order preserving. \end{proof} \begin{cor}\label{cor:conrad-lemma0} Let $ U $ denote the group of units (the invertible matrices) in $ T $. Then $ U $ embeds into $ o\text{-}\Aut G $. \end{cor} \begin{rmk} From the proof of Lemma~\ref{lemma:conrad-lemma0} it is clear that all $ o $-automorphisms of $ G $ induced by the elements $ (\sigma_{\alpha\beta})\in U $ induce the identity on $ \Gamma $. If, moreover, such an automorphism also induces the identity on each component then it will be an internal automorphism. Again, from the formula in the proof of Lemma~\ref{lemma:conrad-lemma0}, we see that this happens if and only if $ \sigma_{\alpha\alpha}=1 $ for all $ \alpha\in\Gamma $. \qed \end{rmk} \begin{prop}\label{prop:oAutGamma-G-direct product} Let us denote by $ o\text{-}\Aut_\Gamma G $ the automorphisms of $ G $ that induce the identity on the chain $ \Gamma $. Similarly, let $ \Aut_\Gamma S(G) $ denote the group of automorphisms of the skeleton whose component on $ \Gamma $ is the identity. Then we have \[ o\text{-}\Aut_\Gamma G \simeq \Int\Aut G \rtimes \Aut_\Gamma S(G). \] It follows, in particular, that $ o\text{-}\Aut_\Gamma G $ forms a group. \end{prop} \begin{proof} Let $ U^1 $ be the subgroup of $ U $ given by all the \emph{unitriaugular} matrices of $ T $ (all the diagonal entries are 1). Then $ \Int\Aut G \simeq U^1 $. When we multiply two diagonal matrices, the elements on the diagonal of the product are the products of the corresponding diagonal entries. Hence $ U^1 $ is a normal subgroup of $ U $. Moreover, every element $ (\sigma_{\alpha\beta})\in U $ is the product of an element in $ U^1 $ and a diagonal matrix: namely $ (\sigma_{\alpha\beta})=(\sigma_{\alpha\beta})^1(\sigma_{\alpha\beta})^d $ where $ (\sigma_{\alpha\beta})^1 $ is obtained by $ (\sigma_{\alpha\beta}) $ replacing all diagonal elements by 1 and $ (\sigma_{\alpha\beta})^d $ is the diagonal matrix with the diagonal entries of $ (\sigma_{\alpha\beta}) $ on its diagonal. And finally, if we denote by $ U^d $ the subgroup of diagonal matrices in $ U $ we have $ U^d\cap U^1=\{ 1 \} $. Hence $ U = U^1\rtimes U^d $. \end{proof} The last proposition gives us a matrix description of the group of automorphisms that induce the identity on $ \Gamma $. But we know that every automorphism of the skeleton (there might be some where the component on $ \Gamma $ is not trivial) lifts to an automorphism of $ G $, because Hahn sums have the (canonical) lifting property. So, in order to completely describe $ o\text{-}\Aut G $, we need to take into account the lifts of automorphisms of the skeleton with non trivial component on $ \Gamma $. Let $ \Aut_\Gamma'S(G) $ be a complement of $ \Aut_\Gamma S(G) $ in $\Aut S(G) $. It consists exactly of all those automorphisms of the skeleton that are not the identity on $ \Gamma $, plus the identity $ \id_{S(G)} $. Then this is a subgroup. We have: \begin{theorem}\label{thm:DG-semidirect-enhanced} We have \[ o\text{-}\Aut G = \left( U^1 \rtimes U^d \right) \rtimes \Aut_\Gamma'S(G). \] \end{theorem} \begin{proof} By proposition~\ref{prop:oAutGamma-G-direct product} we have $ o\text{-}\Aut_\Gamma G \simeq \Int\Aut G \times \Aut_\Gamma S(G) \simeq U^1 \rtimes U^d $ hence it suffices to show that $ o\text{-}\Aut G \simeq o\text{-}\Aut_\Gamma G \rtimes \Aut_\Gamma'S(G) $. We notice immediately that the two groups have trivial intersection and that $ o\text{-}\Aut_\Gamma G $ is normal inside $ o\text{-}\Aut G $. To show that they generate $ o\text{-}\Aut G $ we take an automorphism $ \sigma\ino\text{-}\Aut G $. This will induce an automorphism $ \tau\in\Aut S(G) $ which might or might not fix $ \Gamma $. In any case, if we lift $ \tau $ to a $ \tilde{\tau}\ino\text{-}\Aut G $ we certainly have that $ \sigma\tilde\tau^{-1} $ induces the identity on $ \Gamma $. So $ \sigma = (\sigma\tilde\tau^{-1})\tilde\tau $. \end{proof} Theorem~\ref{thm:DG-semidirect-enhanced} improves \cite[Corollary~3.2]{droste-goebel} in two ways: it provides a generalisation from the balanced case to that of a Hahn sum over an arbitrary skeleton and gives a more precise description of $ o\text{-}\Aut G $ by identifying the internal automorphisms within the group of matrices $ U $. \addcontentsline{toc}{section}{\refname} \printbibliography \end{document}
{ "timestamp": "2023-02-28T02:07:23", "yymm": "2302", "arxiv_id": "2302.06290", "language": "en", "url": "https://arxiv.org/abs/2302.06290", "abstract": "Hahn groups endowed with the canonical valuation play a fundamental role in the classification of valued abelian groups. In this paper we study the group of valuation (respectively order) preserving automorphisms of a Hahn group $G$. Under the assumption that $G$ satisfies some lifting property, we prove a structure theorem decomposing the automorphism group into a semidirect product of two notable subgroups. We characterise a class of Hahn groups satisfying the aforementioned lifting property. For some special cases we provide a matrix description of the automorphism group.", "subjects": "Group Theory (math.GR); Commutative Algebra (math.AC); Logic (math.LO)", "title": "Automorphisms of valued Hahn groups", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9879462211935647, "lm_q2_score": 0.8104789155369048, "lm_q1q2_score": 0.8007095819617434 }
https://arxiv.org/abs/math/0610708
Families of linear semigroups with intermediate growth
Methods from additive number theory are applied to construct families of finitely generated linear semigroups with intermediate growth.
\section{Growth functions of finitely generated semigroups} Let $\mathcal{S}$ be a finitely generated semigroup and let $A$ be a set of generators for $\mathcal{S}$. Every element $x \in \mathcal{S}$ can be written as a word with letters from the set $A$, that is, as a finite product of elements of $A$. The \emph{length} of $x$ with respect to $A$, denoted $\ell_A(x)$, is the number of letters in the shortest word that represents $x$. Note that $\ell_A(x) = 1$ if and only if $x \in A$. We shall assume that \ensuremath{\mathcal S}\ contains an identity element 1 and that $1 \notin A$. We define $\ell_A(1) = 0.$ Let $\ensuremath{ \mathbf N } = \{1,2,3,\ldots\}$ denote the set of positive integers and $\ensuremath{ \mathbf N }_0 = \ensuremath{ \mathbf N } \cup \{0\}$ the set of nonnegative integers. For every $n \in \ensuremath{ \mathbf N }_0$, let $\lambda_A(n)$ denote the number of elements of \ensuremath{\mathcal S}\ of length exactly $n$. We define the \emph{growth function} $\gamma_{A}(n)$ of $\mathcal{S}$ with respect to $A$ by \[ \gamma_{A}(n) = \ensuremath{\text{card}}(\{x\in \mathcal{S} : \ell_A(x) \leq n\}) = \sum_{m=0}^n \lambda_A(m). \] The function $\gamma_A(n)$ is an increasing function that counts the number of elements of \ensuremath{\mathcal S}\ of length at most $n$. If $\ensuremath{\text{card}}(A) = k $, then, for all nonnegative integers $n$, \[ \lambda_A(n) \leq k^n \] and \[ \gamma_A(n) \leq \begin{cases} n+1 & \text{if $k=1$}\\ \frac{k^{n+1}-1}{k-1} & \text{if $k>1$.} \end{cases} \] The semigroup $\mathcal{S}$ has \emph{polynomial growth} with respect to the generating set $A$ if there exist positive numbers $c$ and $d$ such that $\gamma_A(n) \leq cn^d$ for all sufficently large $n$. In this case, $\log \gamma_A(n) \leq \log c + d\log n$ and so \[ \limsup_{n\rightarrow\infty} \frac{\log \gamma_A(n)}{\log n} < \infty. \] If $\mathcal{S}$ does not have polynomial growth, then \[ \limsup_{n\rightarrow\infty} \frac{\log \gamma_A(n)}{\log n} = \infty. \] We say that $\mathcal{S}$ has \emph{superpolynomial growth} if \[ \lim_{n\rightarrow\infty} \frac{\log \gamma_A(n)}{\log n} = \infty. \] The semigroup $\mathcal{S}$ has \emph{exponential growth} with respect to the generating set $A$ if there exists $\theta > 1$ such that $\gamma_A(n) \geq \theta^n$ for all sufficently large $n$. In this case, $\log \gamma_A(n) \geq n \log \theta$ and so \[ \liminf_{n\rightarrow\infty} \frac{\log \gamma_A(n)}{n} > 0. \] If \ensuremath{\mathcal S}\ does not have exponential growth, then \[ \liminf_{n\rightarrow\infty} \frac{\log \gamma_A(n)}{n} = 0. \] We say that the $\mathcal{S}$ has \emph{subexponential growth} if \[ \lim_{n\rightarrow\infty} \frac{\log \gamma_A(n)}{n} = 0. \] The semigroup $\mathcal{S}$ has \emph{intermediate growth} with respect to $A$ if the growth of $\mathcal{S}$ is both superpolynomial and subexponential. We review some standard facts about growth functions. Let $A$ and $B$ be finite generating sets for $\mathcal{S}$, and let \[ c_A = \max(\ell_A(b) : b \in B ) \] and \[ c_B = \max(\ell_B(a) : a \in A ). \] Then \[ \gamma_{B}(n) \leq \gamma_{A}(c_An) \] and \[ \gamma_{A}(n) \leq \gamma_{B}(c_B n). \] These inequalities imply that the growth function $\gamma_A(n)$ is polynomial, superpolynomial, subexponential, or exponential if and only if the growth function $\gamma_B(n)$ is, respectively, polynomial, superpolynomial, subexponential, or exponential. \begin{lemma} \label{intermed:lemma:fone-to-one} Let $\mathcal{S}$ and $\mathcal{T}$ be finitely generated semigroups and let $f: S \rightarrow T$ be an injective semigroup homomorphism. If $A$ is any finite generating set for \ensuremath{\mathcal S}, then there is a finite generating set $B$ for \ensuremath{\mathcal T}\ such that \[ \gamma^{(S)}_A(n) \leq \gamma^{(T)}_B(n) \] for all $n \in \ensuremath{ \mathbf N }_0.$ If the growth of $\mathcal{T}$ is polynomial or subexponential, then the growth of $\mathcal{S}$ is, respectively, polynomial or subexponential. If the growth of $\mathcal{S}$ is superpolynomial or exponential, then the growth of $\mathcal{T}$ is, respectively, superpolynomial or exponential. \end{lemma} \begin{proof} Let $A$ be a finite generating set for $\mathcal{S}$, let $B'$ be any finite generating set for $\mathcal{T}$, and let $B = B' \cup f(A).$ If $\ell_A(s) = m$, then $s=a_1\cdots a_m$ for some $a_1,\ldots,a_m \in A$. Since $f(s) = f(a_1)\cdots f(a_m)$, it follows that $\ell_B(f(s)) \leq m.$ Since $f$ is one-to-one, it follows that $\gamma^{(S)}_A(n) \leq \gamma^{(T)}_B(n)$. This inequality implies the statements about growth rates. \end{proof} \begin{lemma} \label{intermed:lemma:subsemi} Let $\mathcal{S}$ be a finitely generated subsemigroup of a finitely generated semigroup $\mathcal{T}$. If the growth of $\mathcal{T}$ is polynomial or subexponential, then the growth of $\mathcal{S}$ is, respectively, polynomial or subexponential. If the growth of $\mathcal{S}$ is superpolynomial or exponential, then the growth of $\mathcal{T}$ is, respectively, superpolynomial or exponential. \end{lemma} \begin{proof} This follows immediately from Lemma~\ref{intermed:lemma:fone-to-one}. \end{proof} \begin{lemma} \label{intermed:lemma:fonto} Let $\mathcal{S}$ and $\mathcal{T}$ be finitely generated semigroups and let $f: S \rightarrow T$ be a surjective semigroup homomorphism. If $A$ is any finite generating set for \ensuremath{\mathcal S}, then $B = \{f(a): a\in A\}$ is a finite generating set for \ensuremath{\mathcal T}, and \[ \gamma^{(S)}_A(n) \geq \gamma^{(T)}_B(n) \] for all $n \in \ensuremath{ \mathbf N }_0$. If the growth of $\mathcal{S}$ is polynomial or subexponential, then the growth of $\mathcal{T}$ is, respectively, polynomial or subexponential. If the growth of $\mathcal{T}$ is superpolynomial or exponential, then the growth of $\mathcal{S}$ is, respectively, superpolynomial or exponential. \end{lemma} \begin{proof} Let $y \in \ensuremath{\mathcal T}.$ Since $f$ is onto, there exists $x \in \ensuremath{\mathcal S}$ such that $f(x) = y.$ If $\ell_A^{(S)}(x) = m,$ then there is a sequence $a_1,\ldots, a_m \in A$ such that $y = f(x) = f(a_1)\cdots f(a_m)$, and so $B = \{f(a) : a \in A\}$ is a generating set for \ensuremath{\mathcal T}. Conversely, if $y \in \ensuremath{\mathcal T}$ and $\ell_B(y) = m$, then there exist $a_1,\ldots,a_m\in A$ such that $y=f(a_1)\cdots f(a_m)$. Let $x=a_1\cdots a_m \in \ensuremath{\mathcal S}$. Then $f(x) = y$ and $\ell_A(x) \leq m.$ This implies that $\gamma^{(T)}_B(n) \leq \gamma^{(S)}_A(n)$ for all $n$, and the growth conditions follow directly from this inequality. \end{proof} The growth of a finitely generated abelian semigroup $\mathcal{S}$ is always polynomial. Indeed, if $A$ is a set of generators for $\mathcal{S}$ with $\ensuremath{\text{card}}(A) = k$, then \[ \lambda_A(n) \leq {n+k-1 \choose k-1} \ll n^{k-1} \] and \[ \gamma_A(n) \leq {n+k \choose k} \ll n^k. \] More precisely, Khovanskii~\cite{khov92,khov95}, Nathanson~\cite{nath00d}, Nathanson and Ruzsa~\cite{nath02a} proved that there is a polynomial $f_A(x)$ with integer coefficients such that $\lambda_A(n) = f_A(n)$ for all sufficiently large integers $n$. It follows that there is a polynomial $F_A(x)$ with integer coefficients such that $\gamma_A(n) = F_A(n)$ for all sufficiently large $n$. The growth of finitely generated free semigroups of rank at least two is always exponential. If \ensuremath{\mathcal S}\ is the free semigroup generated by a set of $k \geq 2$ elements, then $\lambda_A(n) = k^n$ and $\gamma_A(n) = (k^{n+1}-1)/(k-1)> k^n.$ By Lemma~\ref{intermed:lemma:subsemi}, a semigroup that contains a free subsemigroup on two generators has exponential growth. Semigroups of intermediate growth are more difficult to construct. Beljaev, Sesekin, and Trofimov~\cite{belj-sese-trof77} proved that the free semigroup generated by two elements $e$ and $g$ with the relations $e^2 = e$ and $eg^ieg^je = eg^jeg^ie$ for all nonnegative integers $i$ and $j$ is a semigroup of intermediate growth. Okni{\` n}ski~\cite{okni98} constructed two examples of linear semigroups, that is, subsemigroups of the multiplicative semigroup $M_n(\ensuremath{\mathbf Z})$ of $n \times n$ matrices. Let $\ensuremath{\mathcal S}_1$ be the subsemigroup of $M_2(\ensuremath{\mathbf Z})$ generated by the set \[ A = \left\{ \left(\begin{matrix} 1 & 1 \\ 0 & 1 \end{matrix}\right), \left(\begin{matrix} 1 & 0 \\ 1 & 0 \end{matrix}\right) \right\} \] and let $\ensuremath{\mathcal T}_1$ be the subsemigroup of $M_3(\ensuremath{\mathbf Z})$ generated by the set \[ A = \left\{ \left(\begin{matrix} 1 & 1 & 1\\ 0 & 2 & 1\\ 0 & 0 & 1 \end{matrix}\right), \left(\begin{matrix} 1 & 0 & 0 \\ 0 & 0 & 0 \\ 0 & 0 & 1 \end{matrix}\right) \right\}. \] Both $\ensuremath{\mathcal S}_1$ and $\ensuremath{\mathcal T}_1$ are homomorphic images of the Beljaev, Sesekin, and Trofimov semigroup and have intermediate growth. Nathanson~\cite{nath99b} gave a simple number-theoretical proof that $\mathcal{S}^{(1)}$ has intermediate growth, and Lavrik-M\" annlin~\cite{lavr01} computed the growth function of $\mathcal{S}^{(1)}$. Grigorchuk~\cite{grig88} and Shneerson~\cite{shne01a,shne04a,shne05a} have also investigated the growth of semigroups. In this paper we apply ideas from additive number theory to construct families of linear semigroups of intermediate growth that generalize Okni{\` n}ski's examples. \section{Partition functions of sets of positive integers} Results about the intermediate growth of groups and semigroups often use estimates for the asymptotics of partition functions in additive number theory. We review some of these results here. Let $A$ be a set of positive integers and let $p_A(n)$ denote the number of partitions of $n$ into parts belonging to $A$. If $A = \ensuremath{ \mathbf N }$ is the set of all positive integers, then $p(n) = p_{\ensuremath{ \mathbf N }}(n)$ is the classical partition function. If $A$ is a nonempty finite set of relatively prime positive integers and $\ensuremath{\text{card}}(A) = r,$ then \[ p_A(n) = \left( \frac{1}{\prod_{a\in A} a} \right)\frac{n^{r-1}}{(r-1)!} + O\left(n^{r-2}\right) \] (Nathanson~\cite{nath00a} and~\cite[Theorem 15.2]{nath00aa}). Let $p_r(n)$ denote the number of partitions of $n$ into at most $r$ parts, let $\hat{p}_r(n)$ denote the number of partitions of $n$ into exactly $r$ parts, and let $\hat{q}_r(n)$ denote the number of partitions of $n$ into exactly $r$ distinct parts. Since the number of partitions of $n$ into at most $r$ parts is equal to the number of partitions of $n$ into parts belonging to the set $A = \{1,2,\ldots,r\}$, it follows that \[ p_r(n) = \frac{n^{r-1}}{r!(r-1)!} + O\left(n^{r-2}\right) \] and \begin{equation} \label{intermed:pkn} \hat{p}_r(n) = p_r(n)-p_{r-1}(n) = \frac{n^{r-1}}{r!(r-1)!} + O\left(n^{r-2}\right). \end{equation} We note that $p_r(n)$ and $\hat{p}_r(n)$ are increasing functions of $n$, and \begin{equation} \label{intermed:prp} p(n) = \sum_{r=1}^n \hat{p}_r(n). \end{equation} If $n=a_1+\cdots + a_r$ is a partition of $n$ into $r$ distinct positive parts $a_1 > \cdots > a_r$, then $n \geq r(r+1)/2$ and \[ n - \frac{r(r-1)}{2} = (a_1- (r-1)) +( a_2 - (r-2)) + \cdots + (a_2 -1) + a_1 \] is a partition of $n - r(r-1)/2$ into $r$ parts. This identity establishes a bijection between partitions into exactly $r$ distinct parts and partitions into exactly $r$ parts, and so \begin{equation} \label{intermed:qkn} \hat{q}_r(n) = \hat{p}_r(n - r(r-1)/2) = \frac{n^{r-1}}{r!(r-1)!} + O\left(n^{r-2}\right). \end{equation} The set $A$ has asymptotic density $d(A) = \alpha$ if \[ \lim_{n\rightarrow\infty} \frac{1}{n} \sum_{\substack{a \in A\\ a\leq n}} 1 = \alpha. \] If $d(A) = \alpha>0$ and $\gcd(A) = 1$, then \begin{equation} \label{intermed:part} \log p_A(n) \sim c_0\sqrt{\alpha n} \end{equation} where \[ c_0 = 2\sqrt{\frac{\pi^2}{6}}. \] (An elementary proof of~\eqref{intermed:part} is in Nathanson~\cite[Theorem 16.1]{nath00aa}.) In particular, if $A = \ensuremath{ \mathbf N }$, then $\alpha = 1$ and the partition function $p(n) = p_{\ensuremath{ \mathbf N }}(n)$ satisfies the Hardy-Ramanujan asymptotic estimate \[ \log p(n) \sim c_0\sqrt{ n}. \] If $A$ is the set of odd positive integers, then $\alpha = 1/2$ and $\log p_A(n) \sim c_0\sqrt{n/2}.$ Let $q(n)$ denote the number of partitions of $n$ into distinct positive integers. Since the number of partitions of an integer into distinct parts is equal to the number of partitions into odd parts, it follows that \[ \log q(n) \sim c_0\sqrt{n/2}. \] \section{A condition for subexponential growth} \begin{lemma} \label{intermed:lemma:gijcondition} Let $\mathcal{S}$ be a semigroup generated by a set $A = \{e,g\}$, where $e$ is an idempotent. The set \[ \mathcal{S}_0 = e\mathcal{S}e = \{exe:x\in \mathcal{S}\} \] is a subsemigroup of $\mathcal{S}$ with identity $e$, and is generated by the set $\{ eg^ke: k = 0,1,2,3,\ldots \}$. Define $g^0=1.$ Then \[ \mathcal{S} = \{g^i\}_{i=0}^{\infty} \cup \bigcup_{i,j=0}^{\infty} g^{i}\mathcal{S}_0g^{j}. \] Suppose that, for all nonnegative integers $i_1,j_1,i_2,j_2$, \begin{equation} \label{intermed:gijcondition} g^{i_1}\mathcal{S}_0g^{j_1} \cap g^{i_2}\mathcal{S}_0g^{j_2} \neq \emptyset \text{ if and only if } (i_1,j_1) = (i_2,j_2). \end{equation} For all nonnegative integers $i$ and $j$, let \[ \gamma_{i,j}(n) = \ensuremath{\text{card}}\{y\in g^i\ensuremath{\mathcal S}_0 g^j : \ell_A(y) \leq n\} \] and let $\gamma_0(n) = \gamma_{0,0}(n).$ Then \[ \gamma_{i,j}(n) \leq \gamma_0(n-i-j). \] \end{lemma} \begin{proof} Let $x \in g^i\ensuremath{\mathcal S}_0g^j.$ By condition~\eqref{intermed:gijcondition}, every representation of $x$ as a word in $e$ and $g$ must be of the form $x = g^iy'g^j$ for some $y' \in \ensuremath{\mathcal S}_0$. Therefore, \[ \ell_A(x) = \min\{i+j+\ell_A(y') : y' \in \ensuremath{\mathcal S}_0 \text{ and } x = g^iy'g^j \} = i + j + \ell_A(y) \] for some $y \in \ensuremath{\mathcal S}_0$. If $\ell_A(x) = m$, then $\ell_A(y) = m-i-j$ and so $\gamma_{i,j}(n) \leq \gamma_0(n-i-j).$ \end{proof} \begin{theorem} \label{intermed:theorem:subexponential} Let $\mathcal{S}$ be a semigroup generated by a set $A = \{e,g\}$, where $e$ is an idempotent, and let $\mathcal{S}_0 = e\mathcal{S}e.$ Suppose that \begin{enumerate} \item[(i)] For all nonnegative integers $k_1$ and $k_2$, \[ eg^{k_1}eg^{k_2}e = eg^{k_2}eg^{k_1}e \] \item[(ii)] For all nonnegative integers $i_1,j_1,i_2,j_2$, \[ g^{i_1}\mathcal{S}_0g^{j_1} \cap g^{i_2}\mathcal{S}_0g^{j_2} \neq \emptyset \text{ if and only if } (i_1,j_1) = (i_2,j_2). \] \end{enumerate} Then the semigroup $\mathcal{S}$ has subexponential growth. \end{theorem} \begin{proof} If $e=1$, then $\ensuremath{\mathcal S} = \ensuremath{\mathcal S}_0 = \{g^i\}_{i=0}^{\infty}$ and condition~(ii) is not satisfied. Thus, we can assume that $e\neq 1$. If $1 \in \ensuremath{\mathcal S}_0$, then there exist positive integers $k_1,\ldots,k_r$ such that $eg^{k_1}eg^{k_2}e \cdots eg^{k_r}e=1.$ Multiplying this identity by $e$, we obtain $e=1$, which is absurd. Therefore, $1 \notin \ensuremath{\mathcal S}_0$. Let $\gamma_0(n)$ denote the number of elements $y \in \mathcal{S}_0$ such that $\ell_A(y) \leq n.$ If $y \in \mathcal{S}_0$ and $\ell_A(y) = m \leq n$, then conditions~(i) and~(ii) imply that either $y=e$ or there exist positive integers $r$ and $k_1 \geq \cdots \geq k_r$ such that \[ y = eg^{k_1}eg^{k_2}e \cdots eg^{k_r}e \] and \[ \ell_A(y) = k_1+k_2+\cdots + k_r+r+1 = m. \] Thus, to every $y \in \ensuremath{\mathcal S}_0$ with $\ell_A(y) =m$ there are associated a positive integer $r$ and a partition of $m-r-1$ into exactly $r$ parts, and so the number of elements $y \in \mathcal{S}_0$ of length exactly $m$ is \[ \lambda_0(m) = \gamma_0(m)-\gamma_0(m-1) \leq \sum_{r=1}^m \hat{p}_r(m-r-1) \leq \sum_{r=1}^m \hat{p}_r(m) = p(m) \] by~\eqref{intermed:prp}. Since $1 \notin \ensuremath{\mathcal S}_0$, we have \[ \gamma_0(n)= \sum_{m=1}^n \lambda_0(m) \leq \sum_{m=1}^n p(m) \leq np(n). \] Let $\gamma_{i,j}(n)$ denote the growth function of the set $g^{i}\mathcal{S}_0g^{j}$ with respect to the generating set $A$. By Lemma~\ref{intermed:lemma:gijcondition}, \[ \gamma_{i,j}(n) \leq \gamma_0(n-i-j). \] By condition~(ii), if $g^i = g^j$, then $g^i\ensuremath{\mathcal S}_0 g^i = g^j \ensuremath{\mathcal S}_0 g^j$ and so $i = j.$ Therefore, for every $n \geq 0$ there is exactly one element in the set $\{g^i\}_{i=0}^{\infty}$ of length $n$, and \begin{align*} \gamma_A(n) & \leq \sum_{i=0}^{n} \sum_{j=0}^{n}\gamma_{i,j}(n) + n+1 \\ & \leq \sum_{i=0}^{n} \sum_{j=0}^{n}\gamma_{0}(n-i-j) + n +1\\ & \leq (n+1)^2\gamma_{0}(n) + n+1 \\ & \leq 2(n+1)^3 p(n). \end{align*} From the asymptotic formula~\eqref{intermed:part} for the partition function, we obtain \[ \log \gamma_A(n) \leq \log 2(n+1)^3 + \log p(n) \ll \sqrt{n} \] and so \[ \lim_{n\rightarrow\infty} \frac{\log \gamma_A(n)}{n} = 0. \] Thus, the growth function $\gamma_A(n)$ is subexponential. \end{proof} \section{Sequences with many partition products} Let $W=\{w_k\}_{k=1}^{\infty}$ be a sequence of elements in a semigroup $(\mathcal{X},\ast)$ such that \[ w_{k_1}\ast w_{k_2} = w_{k_2} \ast w_{k_1} \] for all positive integers $k_1,k_2.$ To every finite sequence of positive integers $k_1, k_2, \ldots, k_r$ we associate the element $w_{k_1}\ast w_{k_2} \ast \cdots \ast w_{k_r} \in \ensuremath{\mathcal X}$. For every integer $r \geq 1$, let $W_r(n)$ denote the subset of $\mathcal{X}$ associated to partitions of positive integers not exceeding $n$ into exactly $r$ parts, that is, $w \in W_r(n)$ if and only if there is a sequence of positive integers $k_1 \geq k_2 \geq \cdots \geq k_r$ such that $k_1 + k_2 + \cdots + k_r \leq n$ and $w = w_{k_1}\ast w_{k_2} \ast\cdots \ast w_{k_r}$. We define \[ \Phi(r) = \liminf_{n\rightarrow\infty} \frac{\log \ensuremath{\text{card}}(W_r(n))}{\log n} . \] The sequence $\{w_k\}_{k=1}^{\infty}$ has \emph{many partition products in $W$} if \[ \lim_{r\rightarrow\infty} \Phi(r) = \infty. \] Here are some examples of sequences with many partition products. \begin{theorem} \label{intermed:theorem:Ud} For every positive integer $d$, let $\ensuremath{\mathcal U}_d$ be the multiplicative semigroup of all positive integers $u$ such that $u \equiv 1\pmod{d}.$ The sequence \[ W = \{dk+1\}_{k=1}^{\infty} \] has many partition products. \end{theorem} \begin{proof} Let $r \geq 1$ and let $k_1 \geq k_2 \geq \cdots \geq k_r$ be a sequence of positive integers with $k_1+k_2+\cdots + k_r \leq n.$ Associated to this sequence is the integer $(dk_1+1)(dk_2+1)\cdots (dk_r+1) \in W_r(n).$ Let \ensuremath{\mathcal K}\ denote the set of positive integers $k$ such that $dk+1$ is prime, and let $K(t)$ count the number of elements $k\in \ensuremath{\mathcal K}$ with $k\leq t.$ If $k_1 \geq k_2 \geq \cdots \geq k_r$ and $j_1 \geq j_2 \geq \cdots \geq j_r$ are distinct sequences of elements of \ensuremath{\mathcal K}\ such that $k_1 \leq n/r$ and $j_1 \leq n/r$, then $k_1+k_2+ \cdots + k_r \leq n$ and $j_1+j_2+\cdots + j_r \leq n$, hence $(dk_1+1)(dk_2+1)\cdots (dk_r+1) \in W_r(n)$ and $(dj_1+1)(dj_2+1)\cdots (dj_r+1) \in W_r(n)$. By the fundamental theorem of arithmetic, $(dk_1+1)(dk_2+1)\cdots (dk_r+1) \neq (dj_1+1)(dj_2+1)\cdots (dj_r+1)$, and so \[ \ensuremath{\text{card}}(W_r(n)) \geq {K(n/r) + r-1 \choose r} \geq \frac{K(n/r)^r}{r!}. \] Let $\pi(x,d,1)$ denote the number of prime numbers $p \leq x$ such that $p\equiv 1\pmod{d}.$ Let \[ x = \frac{dn}{r} + 1. \] Then $dk+1 \leq x$ is prime if and only if $k \leq n/r$. By the prime number theorem for arithmetic progressions, for sufficiently large $n$ we have \[ K(n/r) = \pi(x,d,1) \gg \frac{x}{\log x} \gg \frac{n}{r\log n} \] where the implied constants depend only on $d$. Therefore, \[ \ensuremath{\text{card}}(W_r(n)) \geq \frac{K(n/r)^r}{r!} \gg \frac{n^r}{r!r^r(\log n)^r} \] and so \[ \log\left( \ensuremath{\text{card}}(W_r(n))\right) \gg r\log n -\log (r!r^r) - r\log\log n \] and \[ \Phi(r) = \liminf_{n\rightarrow\infty} \frac{ \log\left( \ensuremath{\text{card}}(W_r(n))\right) }{\log n} \gg r. \] It follows that \[ \lim_{r\rightarrow\infty} \Phi(r) = \infty \] and the sequence $\{dk+1\}_{k=1}^{\infty}$ has many partition products. This completes the proof. \end{proof} In additive number theory, a sequence $\{b_k\}_{k=k_0}^{\infty}$ contained in an additive abelian semigroup is called a $\hat{B}_r$-sequence if sums of $r$ distinct terms of the sequence are distinct, that is, if $k_1 < k_2< \cdots < k_r$ and $j_1 < j_2 < \cdots < j_r$, and if \[ b_{k_1} + b_{k_2} + \cdots + b_{k_r} = b_{j_1} + b_{j_2} + \cdots + b_{j_r} \] then $k_i = j_i$ for $i=1,\ldots,r.$ The sequence is called a $\hat{B}_{\infty}$-sequence if all finite sums of distinct elements of the set are distinct, that is, if $k_1<k_2< \cdots < k_r$ and $j_1 < j_2 < \cdots < j_s$, and if \[ b_{k_1} + b_{k_2} + \cdots + b_{k_r} = b_{j_1} + b_{j_2} + \cdots + b_{j_s} \] then $r=s$ and $k_i = j_i$ for $i=1,\ldots,r.$ If $\{b_ k\}_{k=k_0}^{\infty}$ is a sequence of positive real numbers such that $\sum_{k=k_0}^{\ell-1} b_k < b_{\ell}$ for all $\ell > k_0$, then the sequence is a $\hat{B}_{\infty}$-sequence. In particular, a $\hat{B}_{\infty}$-sequence is a $\hat{B}_r$-sequence for all $r \geq 1.$ \begin{lemma} \label{intermed:lemma:ctr} Let $c,c_1,d,$ and $t$ be positive real numbers with $t > 2$, and let $\{e_k\}_{k=1}^{\infty}$ be a sequence of real numbers such that $|e_k| \leq c_1k^d$ for all $k \geq 1$. There is an integer $k_0$ such that the sequence \[ \{ ct^k + e_k\}_{k=k_0}^{\infty} \] is a strictly increasing $\hat{B}_{\infty}$-sequence. \end{lemma} \begin{proof} Let $b_k = ct^k + e_k$ for $k \geq 1.$ Since $t>2$ and $|e_k| \leq c_1k^d$, there is an integer $k_0$ such that \[ 0 < b_k < b_{k+1} \] for all $k \geq k_0$, and also \begin{equation} \label{intermed:partprodineq} \frac{c_1(t-1)(k+1)^{d+1}}{c(d+1)(t-2)} < t^k \end{equation} for all $k > k_0.$ Let $\ell > k_0.$ Using~\eqref{intermed:partprodineq} and the inequality \[ \sum_{k=1}^{\ell} k^d < \frac{(\ell +1)^{d+1}}{d+1} \] we obtain \begin{align*} \sum_{k=k_0}^{{\ell}-1} b_k & = c\sum_{k=k_0}^{{\ell}-1}t^k + \sum_{k=k_0}^{{\ell}-1}e_k \\ & < \frac{ct^{\ell}}{t-1} + c_1 \sum_{k=k_0}^{{\ell}-1} k^d \\ & < ct^{\ell} - c_1{\ell}^d \\ & \leq b_{\ell}. \end{align*} This completes the proof. \end{proof} \begin{theorem} \label{intermed:theorem:Br} Let $c,c_1d,$ and $t$ be positive real numbers with $t > 2$, and let $\{e_k\}_{k=1}^{\infty}$ be a sequence of real numbers such that $|e_k| \leq c_1k^d$ for all $k \geq 1$. Let $k_0$ be a positive integer such that sequence \[ W = \{ ct^k + e_k\}_{k=k_0}^{\infty} \] is a $\hat{B}_{\infty}$-sequence. Then $W$ has many partial products. \end{theorem} \begin{proof} Let $b_k = ct^k + e_k$ for $k \geq 1$. By Lemma~\ref{intermed:lemma:ctr}, there is an integer $k_0$ such that $W = \{ ct^k + e_k\}_{k=k_0}^{\infty}$ is a strictly increasing $\hat{B}_r$-sequence for every positive integer $r$. To every partition $n = k_1 + \cdots + k_r$ into exactly $r$ distinct parts $ k_1 > k_2 > \cdots > k_r$, we associate the real number $b_{k_1+k_0} + b_{k_2+k_0} + \cdots + b_{k_r+k_0} \in W_r(n+rk_0).$ Since $W$ is a $\hat{B}_r$-sequence, it follows that different partitions are associated to different real numbers, and so, by the partition asymptotic~\eqref{intermed:qkn}, \[ \ensuremath{\text{card}}(W_r(n+rk_0)) \geq \hat{q}_r(n) \gg n^{r-1} \] and \[ \Phi(r) = \liminf_{n\rightarrow\infty}\frac{\log \ensuremath{\text{card}}(W_r(n))}{\log n} \geq r-1. \] This completes the proof. \end{proof} \section{A condition for superpolynomial growth} \begin{theorem} \label{intermed:theorem:superpolynomial} Let $\mathcal{S}$ be a semigroup generated by a set $A = \{e,g\}$, where $e$ is an idempotent. Let $\mathcal{S}_0 = e\mathcal{S}e = \{exe:x\in \mathcal{S}\}.$ Suppose that, for all nonnegative integers $k_1$ and $k_2$, \[ eg^{k_1}eg^{k_2}e = eg^{k_2}eg^{k_1}e. \] Let $\varphi$ be a semigroup homomorphism from $\mathcal{S}_{0}$ into a semigroup $(\ensuremath{\mathcal X},\ast)$ such that the sequence \[ W = \{\varphi(eg^k e)\}_{k=1}^{\infty} \] has many partition products in $\ensuremath{\mathcal X}$. Then the semigroup $\mathcal{S}$ has superpolynomial growth. \end{theorem} \begin{proof} Let $w_k = \varphi(eg^ke)$ for $k = 1,2,\ldots.$ Then \begin{align*} w_{k_1}\ast w_{k_2} & = \varphi(eg^{k_1}e) \ast \varphi(eg^{k_2}e) = \varphi(eg^{k_1}eg^{k_2}e) \\ & = \varphi(eg^{k_2}eg^{k_1}e) = \varphi(eg^{k_2}e) \ast \varphi(eg^{k_1}e) \\ & = w_{k_2}\ast w_{k_1} \end{align*} for all $k_1, k_2 \in \ensuremath{ \mathbf N }_0.$ If $y = eg^{k_1}eg^{k_2}e \cdots eg^{k_r}e \in \mathcal{S}_0$, then \[ \varphi(y) = \varphi\left(eg^{k_1}e\right)\ast \varphi\left(eg^{k_2}e \right)\ast \cdots \ast\varphi\left(eg^{k_r}e\right) = w_{k_1}\ast w_{k_2} \ast \cdots \ast w_{k_r}. \] Let $\gamma_A(n)$ denote the growth function of $\mathcal{S}$ with respect to $A$, and let $\gamma_0(n)$ denote the growth function of $\mathcal{S}_0$ with respect to $A$. Fix a positive integer $r$. If $w \in W_r(n-r-1)$, then there are positive integers $k_1, k_2, \ldots, k_r$ such that \[ w_{k_1} \ast w_{k_2} \ast \cdots \ast w_{k_r} = w \] and \[ k_1 + \cdots + k_r \leq n-r-1. \] Let \[ y = eg^{k_1}eg^{k_2}e\cdots eg^{k_h}e \in \mathcal{S}_0. \] Then \[ \varphi(y) = w_{k_1} \ast w_{k_2} \ast \cdots \ast w_{k_r} = w \] and \[ \ell_A(y) \leq k_1+k_2+\cdots + k_r +r+1 \leq n. \] Thus, to every element $w \in W_r(n-r-1)$ there is at least one element $y\in \mathcal{S}_0$ with $\varphi(y) = w$ and $\ell_A(y) \leq n$, and so \[ \gamma_0(n) \geq \ensuremath{\text{card}}(W_r(n-r-1)). \] It follows that \[ \liminf_{n\rightarrow\infty} \frac{\log\gamma_A(n)}{\log n} \geq \liminf_{n\rightarrow\infty} \frac{\log\gamma_0(n)}{\log n} \geq \liminf_{n\rightarrow\infty} \frac{\log\ensuremath{\text{card}}(W_r(n-r-1))}{\log n} = \Phi(r). \] Since this inequality is true for all positive integers $r$ and $\Phi(r)$ tends to infinity, it follows that \[ \lim_{n\rightarrow\infty} \frac{\gamma_A(n)}{\log n} = \infty \] and so the growth function $\gamma_A(n)$ is superpolynomial. \end{proof} \section{The semigroups $\mathcal{S}^{(d)}$} \begin{theorem} \label{intemed:theorem:Sd} Let $d$ be a positive integer and let $\mathcal{S}^{(d)}$ be the subsemigroup of $M_2(\ensuremath{\mathbf Z})$ generated by the matrices \[ g = \left(\begin{matrix} 1 & d \\ 0 & 1 \end{matrix}\right) \] and \[ e = \left(\begin{matrix} 1 & 0 \\ 1 & 0 \end{matrix}\right). \] The semigroup $\ensuremath{\mathcal S}^{(d)}$ has intermediate growth. \end{theorem} \begin{proof} Multiplying matrices, we obtain \[ eg^ke = (dk+1)e \] and \[ eg^{k_1}eg^{k_2}e = (dk_1+1)(dk_2+1) = eg^{k_2}eg^{k_1}e \] for all nonnegative integers $k_1,k_2.$ It follows that $\ensuremath{\mathcal S}^{(d)}$ satisfies condition~(i) of Theorem~\ref{intermed:theorem:subexponential}, and \[ \mathcal{S}^{(d)}_0 = e\mathcal{S}^{(d)}e = \{ (dk+1)e : k = 0,1,2,\ldots\} = \{ue : u \in \mathcal{U}_d\} \] where $\mathcal{U}_d$ is the multiplicative semigroup of positive integers that are congruent to 1 modulo $d$. For all $u \in \ensuremath{\mathcal U}_d$ and all nonnegative integers $i$ and $j$, \[ g^i ue g^j = u \left(\begin{matrix} 1 & di \\ 0 & 1 \end{matrix}\right) \left(\begin{matrix} 1 & 0 \\ 1 & 0 \end{matrix}\right) \left(\begin{matrix} 1 & dj \\ 0 & 1 \end{matrix}\right) = u\left(\begin{matrix} di +1& dj+d^2 ij \\ 1 & dj \end{matrix}\right). \] Let $u_1,u_2 \in \ensuremath{\mathcal U}_d$ and $i_1,i_2,j_1,j_2 \in \ensuremath{ \mathbf N }_0$. Then \[ g^{i_1} u_1e g^{j_1} = g^{i_2} u_2e g^{j_2} \] if and only if $u_1=u_2$ and $(i_1,j_1) = (i_2,j_2).$ Thus, the semigroup $\ensuremath{\mathcal S}^{(d)}$ also satisfies condition~(ii) of Theorem~\ref{intermed:theorem:subexponential}, and so has subexponential growth. The function \[ \varphi: \mathcal{S}_0^{(d)} \rightarrow \mathcal{U}_d \] defined by \[ \varphi(ue) = u \] is a semigroup homomorphism. Let $w_k = \varphi(eg^ke) = \varphi((dk+1)e) = dk+1$ and $W = \{dk+1\}_{k=1}^{\infty}$. By Theorem~\ref{intermed:theorem:Ud}, the sequence $W$ has many partition products. By Theorem~\ref{intermed:theorem:superpolynomial}, the semigroup $\mathcal{S}^{(d)}$ has superpolynomial growth. This completes the proof. \end{proof} \begin{problem} If $d'$ divides $d$, then $\mathcal{S}^{(d)}$ is a subsemigroup of $\mathcal{S}^{(d')}$. In particular, for every positive integer $d$, the semigroup $\mathcal{S}^{(d)}$ is a finitely generated subsemigroup of the Okni{\` n}ski semigroup $\mathcal{S}^{(1)}$. Let $\mathcal{S}$ be finitely generated subsemigroup of a finitely generated semigroup $\mathcal{S}$. How are the growth rates of $\mathcal{S'}$ and $\mathcal{S}$ related? \end{problem} \section{Subsemigroups of $M_3(\ensuremath{ \mathbf{R} })$} For $k \geq 0$ and $m \geq 1$, let $F_k(x_1,\ldots,x_m)$ be the symmetric function of degree $k$ in $m$ variables defined by \[ F_k(x_1,\ldots,x_m) = \sum_{\substack{(i_1,\ldots,i_m)\in \ensuremath{ \mathbf N }_0^m \\ i_1+i_2+\cdots + i_m=k}} x_1^{i_1}\cdots x_m^{i_m}. \] Define $F_k(x_1,\ldots,x_m)=0$ for $k<0$. If $k \geq 1$ and $m \geq 2$, then \[ F_k(x_1,\ldots,x_m) = x_1F_{k-1}(x_1,x_2,\ldots,x_m) + F_k(x_2,\ldots,x_m). \] In particular, \begin{equation} \label{intermed:symmetric2} F_k(x_1,x_2) = x_1F_{k-1}(x_1,x_2) + x_2^k \end{equation} and \begin{equation} \label{intermed:symmetric3} F_k(x_1,x_2,x_3) = x_1F_{k-1}(x_1,x_2,x_3) + F_k(x_2,x_3). \end{equation} \begin{lemma} \label{intermed:lemma:semigroupT} Consider the upper triangular matrices \[ g = \left(\begin{matrix} s & v & w \\ 0 & t & u \\ 0 & 0 & r \end{matrix}\right) \] and \[ e = \left(\begin{matrix} 1 & 0 & 0 \\ 0 & 0 & 0 \\ 0 & 0 & 1 \end{matrix}\right) \] with coefficients in a ring. For all $k \geq 0,$ define \begin{align*} u_k & = uF_{k-1}(r,t) \\ v_k & = vF_{k-1}(s,t) \\ w_k & = wF_{k-1}(r,s) + uv F_{k-2}(r,s,t). \end{align*} Then \begin{equation} \label{intermed:e2e} e^2=e \end{equation} \begin{equation} \label{intermed:matrixgk} g^k = \left(\begin{matrix} s^k & v_k & w_k \\ 0 & t^k & u_k \\ 0 & 0 & r^k \end{matrix}\right) \end{equation} \begin{equation} \label{intermed:matrix-egke} eg^{k}e = \left( \begin{matrix} s^k & 0 & w_{k} \\ 0 & 0& 0 \\ 0 & 0 & r^k \end{matrix} \right) \end{equation} and \begin{equation} \label{intermed:matrix-egk1k2e} eg^{k_1}eg^{k_2}e = \left(\begin{matrix} s^{k_1+k_2} & 0 & s^{k_1}w_{k_2}+r^{k_2}w_{k_1} \\ 0 & 0& 0 \\ 0 & 0 & r^{k_1+k_2} \end{matrix}\right). \end{equation} For $k_1,k_2\in \ensuremath{ \mathbf N }_0$, \begin{equation} \label{intermed:k1k2commute} eg^{k_1}eg^{k_2}e = eg^{k_2}eg^{k_1}e \end{equation} if and only if \begin{equation} \label{intermed:k1k2commutebecause} s^{k_1}w_{k_2}+r^{k_2}w_{k_1} = s^{k_2}w_{k_1}+r^{k_1}w_{k_2}. \end{equation} If $r=s=1$, then~\eqref{intermed:k1k2commute} holds for all $k_1,k_2\in \ensuremath{ \mathbf N }_0.$ \end{lemma} \begin{proof} Identity~\eqref{intermed:matrixgk} holds for $k=0$ and $k=1$. If the formula is true for some $k \geq 1$, then~\eqref{intermed:symmetric2} and~\eqref{intermed:symmetric3} imply that \begin{align*} g^{k+1} & = \left(\begin{matrix} s^k & v_k & w_k \\ 0 & t^k & u_k \\ 0 & 0 & r^k \end{matrix}\right) \left(\begin{matrix} s & v & w \\ 0 & t & u \\ 0 & 0 & r \end{matrix}\right) \\ & = \left(\begin{matrix} s^{k+1} & s^kv+v_kt & s^kw+v_ku+w_kr\\ 0 & t^{k+1} & t^ku+u_kr \\ 0 & 0 & r^{k+1} \end{matrix}\right) \\ & = \left(\begin{matrix} s^{k+1} & v_{k+1} & w_{k+1}\\ 0 & t^{k+1} & u_{k+1} \\ 0 & 0 & r^{k+1} \end{matrix}\right) \end{align*} and identity~\eqref{intermed:matrixgk} follows by induction on $k$. Identities~\eqref{intermed:matrix-egke} and~\eqref{intermed:matrix-egk1k2e} follow by matrix multiplication, and~\eqref{intermed:k1k2commute} and~\eqref{intermed:k1k2commutebecause} follow from inspection of~\eqref{intermed:matrix-egk1k2e}. This completes the proof. \end{proof} \begin{theorem} \label{intermed:theorem:semigroupT} Let $t,u,v,w$ be real numbers with $t > 2$. Let $\mathcal{T}$ be the subsemigroup of $M_3(\ensuremath{ \mathbf{R} })$ generated by the matrices \[ e = \left(\begin{matrix} 1 & 0 & 0 \\ 0 & 0 & 0 \\ 0 & 0 & 1 \end{matrix}\right) \] and \[ g = \left(\begin{matrix} 1 & v & w \\ 0 & t & u \\ 0 & 0 & 1 \end{matrix}\right) \] The semigroup \ensuremath{\mathcal T}\ has intermediate growth. \end{theorem} \begin{proof} For every $k\in \ensuremath{ \mathbf N }$, let \[ t_k = 1 + t + t^2 + \cdots + t^{k-1} = \frac{t^k-1}{t-1} \] and \[ w_k = \frac{uv t^k}{(t-1)^2} + \left(w-\frac{uv}{t-1}\right)k - \frac{uv}{(t-1)^2}. \] By Lemma~\ref{intermed:lemma:semigroupT}, we have $e^2=e$, \[ g^k = \left( \begin{matrix} 1 & vt_k & w_k \\ 0 & t^k & ut_k \\ 0 & 0 & 1 \end{matrix} \right) \] \[ eg^ke = \left( \begin{matrix} 1 & 0 & w_k \\ 0 & 0 & 0 \\ 0 & 0 & 1 \end{matrix} \right) \] and \[ eg^{k_1}eg^{k_2}e = eg^{k_2}eg^{k_1}e \] for $k_1,k_2\in \ensuremath{ \mathbf N }_0$. Let $\ensuremath{\mathcal T}_0 = e\ensuremath{\mathcal T} e.$ Then $\ensuremath{\mathcal T}_0$ consists of all matrices of the form \[ eg^{k_1}eg^{k_2}e\cdots eg^{k_r}e = \left( \begin{matrix} 1 & 0 & w_{k_1}+w_{k_2}+\cdots + w_{k_r} \\ 0 & 0 & 0 \\ 0 & 0 & 1 \end{matrix} \right) \] where $k_1,\ldots, k_r$ is a finite sequence of nonnegative integers. For $i,j \in \ensuremath{ \mathbf N }_0$, the set $g^i\ensuremath{\mathcal T}_0 g^j$ consists of all matrices of the form \begin{equation} \label{intermed:eMatrix} g^ieg^{k_1}eg^{k_2}e\cdots eg^{k_r}e g^j= \left( \begin{matrix} 1 & vt_j & w_i + w_j + w_{k_1}+w_{k_2}+\cdots + w_{k_r} \\ 0 & 0 & ut_i \\ 0 & 0 & 1 \end{matrix} \right) \end{equation} Since $\{t_i\}_{i=0}^{\infty}$ is a strictly increasing sequence of real numbers, formula~\eqref{intermed:eMatrix} implies that for all $i_1,i_2,j_1,j_2 \in \ensuremath{ \mathbf N }_0$, \[ g^{i_1}\ensuremath{\mathcal T}_0 g^{j_1} \cap g^{i_2}\ensuremath{\mathcal T}_0 g^{j_2} \neq \emptyset \] if and only if \[ (i_1,j_1) = (i_2,j_2). \] Thus, the semigroup \ensuremath{\mathcal T}\ satisfies conditions~(i) and~(ii) of Theorem~\ref{intermed:theorem:subexponential} and so has subexponential growth. Define the function $\varphi:\ensuremath{\mathcal T}_0 \rightarrow \ensuremath{ \mathbf{R} }$ by \[ \varphi(eg^ke) = w_k = ct^k + e_k \] where $c = uv/(t-1)^2$ and $e_k = (w-uv/(t-1))k -c = O(k)$. Then $\varphi$ is a homomorphism from $\ensuremath{\mathcal T}_0$ into the additive group of real numbers. By Theorem~\ref{intermed:theorem:Br}, there is an integer $k_0$ such that the sequence $W = \{w_k\}_{k=k_0}^{\infty}$ has many partial products. By Theorem~\ref{intermed:theorem:superpolynomial}, the semigroup \ensuremath{\mathcal T}\ has superpolynomial growth. This completes the proof. \end{proof} \emph{Acknowledgements.} I wish to thank David Newman, Kevin O'Bryant, and Lev Shneerson for helpful discussions about this work. \def$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2]{% \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2}
{ "timestamp": "2006-10-24T21:34:04", "yymm": "0610", "arxiv_id": "math/0610708", "language": "en", "url": "https://arxiv.org/abs/math/0610708", "abstract": "Methods from additive number theory are applied to construct families of finitely generated linear semigroups with intermediate growth.", "subjects": "Group Theory (math.GR); Number Theory (math.NT)", "title": "Families of linear semigroups with intermediate growth", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9879462183543603, "lm_q2_score": 0.8104789063814617, "lm_q1q2_score": 0.8007095706155427 }
https://arxiv.org/abs/1712.08666
Modular periodicity of the Euler numbers and a sequence by Arnold
For any positive integer $q$, the sequence of the Euler up/down numbers reduced modulo $q$ was proved to be ultimately periodic by Knuth and Buckholtz. Based on computer simulations, we state for each value of $q$ precise conjectures for the minimal period and for the position at which the sequence starts being periodic. When $q$ is a power of $2$, a sequence defined by Arnold appears, and we formulate a conjecture for a simple computation of this sequence.
\section{Introduction} \label{sec:introduction} The sequence of Euler up/down numbers $(E_n)_{n\geq0}$ is the sequence with exponential generating series \begin{equation} \sum_{n=0}^{\infty} \frac{E_n}{n!} x^n = \sec x + \tan x. \end{equation} It is referenced as sequence A000111 in~\cite{OEIS} and its first terms are \[ 1, 1, 1, 2, 5, 16, 61, 272, 1385, 7936, 50521, 353792, 2702765,\ldots \] The numbers $E_n$ were shown by Andr\'e~\cite{A79} to count up/down permutations on $n$ elements (see Section~\ref{sec:powerof2}). Knuth and Buckholtz~\cite{KB67} proved that for any integer $q\geq1$, the sequence $(E_n \mod q)_{n\geq 0}$ is ultimately periodic. For any $q \geq 1$ we define : \begin{itemize} \item $s(q)$ to be the minimum number of terms one needs to delete from the sequence $(E_n \mod q)_{n \geq 0}$ to make it periodic ; \item $d(q)$ to be the smallest period of the sequence $(E_n \mod q)_{n \geq s(q)}$. \end{itemize} For example, the sequence $(E_n \mod 3)$ starts with \[ 1,1,1,2,2,1,1,2,2,1,1,2,2,\ldots \] so one might expect to have $s(3)=1$ and $d(3)=4$. Clearly $s(1)=0$ and $d(1)=1$. In the remainder of this paper, we formulate precise conjectures for the values of $s(q)$ and $d(q)$ for any $q\geq2$. \subsection*{Organisation of the paper} In Section~\ref{sec:oddprime} we reduce the problem to the case when $q$ is a prime power and we conjecture the values of $s(q)$ and $d(q)$ when $q$ is an odd prime power. In Section~\ref{sec:powerof2} we conjecture the values of $s(q)$ and $d(q)$ when $q$ is a power of $2$, after having introduced the Entringer numbers and a sequence defined by Arnold describing the $2$-adic valuation of the Entringer numbers. In Section~\ref{sec:Arnold}, we provide a simple construction which conjecturally yields the Arnold sequence. \section{Case when $q$ is not a power of $2$} \label{sec:oddprime} The following lemma implies that it suffices to know the values of $s(q)$ and $d(q)$ when $q$ is a prime power in order to know the values of $s(q)$ and $d(q)$ for any $q \geq 2$. \begin{lem} Fix $q \geq 2$ and write its prime number decomposition as \begin{equation} q=\prod_{i=1}^k p_i^{\alpha_i}, \end{equation} where $k\geq1$, $p_1,\ldots,p_k$ are distinct prime numbers and $\alpha_1,\ldots,\alpha_k$ are positive integers. Then \begin{align} s(q)&=\max_{1 \leq i \leq k} s(p_i^{\alpha_i}) \\ d(q)&=\lcm (d(p_1^{\alpha_1}),\ldots,d(p_k^{\alpha_k})). \end{align} \end{lem} The proof is elementary and uses the Chinese remainder theorem. When $q$ is an odd prime power, Knuth and Buckholtz~\cite{KB67} found the following : \begin{thm}[\cite{KB67}] \label{thm:KB} Let $p$ be an odd prime number. \begin{enumerate} \item If $p \equiv 1 \mod 4$, then \begin{equation*} d(p)=p-1. \end{equation*} \item If $p \equiv 3 \mod 4$, then \begin{equation*} d(p)=2p-2. \end{equation*} \item For any $k\geq1$, \begin{equation*} s(p^k)\leq k. \end{equation*} \item For any $k\geq 2$, \begin{equation*} d(p^k) | p^{k-1} d(p). \end{equation*} \end{enumerate} \end{thm} We conjecture the following for the exact values of $s(q)$ and $d(q)$ when $q$ is an odd prime power : \begin{conj} \label{conj:oddprime} Let $p$ be an odd prime number. \begin{enumerate} \item For any $k\geq1$, \begin{equation*} s(p^k)=k. \end{equation*} \item For any $k\geq 2$, \begin{equation*} d(p^k)=p^{k-1} d(p). \end{equation*} \end{enumerate} \end{conj} Conjecture~\ref{conj:oddprime} is supported by Mathematica simulations done for all odd prime powers $q<1000$. \section{Entringer numbers and case when $q$ is a power of $2$} \label{sec:powerof2} Formulating a conjecture analogous to Conjecture~\ref{conj:oddprime} for powers of $2$ requires to define, following Arnold~\cite{A91}, a sequence describing the behavior of the $2$-adic valuation of the Entringer numbers. \subsection{The Seidel-Entringer-Arnold triangle} The Entringer numbers are a refined version of the Euler numbers, enumerating some subsets of up/down permutations. For any $n\geq0$, a permutation $\sigma\in\mathcal S_n$ is called \emph{up/down} if for any $2\leq i \leq n$, we have $\sigma(i-1)<\sigma(i)$ (resp. $\sigma(i-1)>\sigma(i)$) if $i$ is even (resp. $i$ is odd). Andr\'e~\cite{A79} showed that the number of up/down permutations on $n$ elements is $E_n$. For any $1 \leq i \leq n$, the \emph{Entringer number} $e_{n,i}$ is defined to be the number of up/down permutations $\sigma\in\mathcal S_n$ such that $\sigma(n)=i$. The Entringer numbers are usually displayed in a triangular array called the Seidel-Entringer-Arnold triangle, where the numbers $(e_{n,i})_{1 \leq i\leq n}$ appear from left to right on the $n$-th line (see Figure~\ref{fig:SEAtriangle}). \begin{figure}[htpb] \[ \begin{matrix} &&&&1&&&& \\ &&&&&&&& \\ &&&0&&1&&& \\ &&&&&&&& \\ &&1&&1&&0&& \\ &&&&&&&& \\ &0&&1&&2&&2& \\ &&&&&&&& \\ 5&&5&&4&&2&&0 \\ \end{matrix} \] \caption{First five lines of the Seidel-Entringer-Arnold triangle.} \label{fig:SEAtriangle} \end{figure} The Entringer numbers can be computed using the following recurrence formula (see for example~\cite{S97}). For any $n\geq2$ and for any $1\leq i \leq n$, we have \begin{equation} e_{n,i}= \begin{cases} \sum_{j<i} e_{n-1,j} &\text{ if } $n$ \text{ is even} \\ \sum_{j\geq i} e_{n-1,j} &\text{ if } $n$ \text{ is odd} \end{cases}. \end{equation} \subsection{Arnold's sequence} Replacing each entry of the Seidel-Entringer-Arnold triangle by its $2$-adic valuation, we obtain an infinite triangle denoted by $T$ (see Figure~\ref{fig:2valuationtriangle}). \begin{figure}[htpb] \[ \begin{matrix} &&&&0&&&& \\ &&&&&&&& \\ &&&\infty&&0&&& \\ &&&&&&&& \\ &&0&&0&&\infty&& \\ &&&&&&&& \\ &\infty&&0&&1&&1& \\ &&&&&&&& \\ 0&&0&&2&&1&&\infty \\ \end{matrix} \] \caption{First five lines of the triangle $T$ of $2$-adic valuations of the Entringer numbers.} \label{fig:2valuationtriangle} \end{figure} We read this triangle $T$ diagonal by diagonal, with diagonals parallel to the left boundary. For any $i\geq1$, denote by $D_i$ the $i$-th diagonal of the triangle $T$ parallel to the left boundary. For example $D_1$ starts with $0,\infty,0,\infty,0,\ldots$. For any $i\geq1$, denote by $m_i$ the minimum entry of diagonal $D_i$. Arnold~\cite{A91} observed that the further away one moves from the left boundary, the higher the $2$-adic valuation of the Entringer numbers becomes. In particular, he observed (without proof) that the sequence $(m_i)_{i\geq1}$ was weakly increasing to infinity. He defined the following sequence : for any $k \geq 1$, \[ u_k:=\max \left\{i\geq1 | m_i < k \right\}. \] In other words, $u_k$ is the number of diagonals containing at least one entry that is not zero modulo $2^k$. The sequence $(u_k)_{k\geq1}$ is referenced as the sequence A108039 in OEIS~\cite{OEIS} and its first few terms are given in Table~\ref{tab:firstterms}. \begin{table}[htbp] \centering \begin{tabular}{|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|} \hline $k$ & 1& 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 & 11 & 12 & 13 & 14 & 15 & 16 & 17 & 18 \\ \hline $u_k$ & 2 & 4 & 4 & 4 & 8 & 8 & 8 & 8 & 10 & 12 & 12 & 16 & 16 & 16 & 16 & 16 & 18 & 20 \\ \hline \end{tabular} \caption{The first few values of $u_k$.} \label{tab:firstterms} \end{table} Note that the first few terms given by Arnold were incorrect, because the entry $4$ appeared four times, whereas it should be appearing only three times. We also remark that we cannot define any sequence analogous to $(u_k)$ when studying the $p$-adic valuations of the Entringer numbers for odd primes $p$. Indeed, the $p$-adic valuation $0$ seems to appear in diagonals of arbitrarily high index. \subsection{Case when $q$ is a power of $2$} Using the sequence $(u_k)_{k\geq1}$, we formulate the following conjecture for $s(q)$ and $d(q)$ when $q$ is a power of $2$ : \begin{conj} \label{conj:powersoftwo} For any $k\geq 1$, we have \begin{equation} s(2^k)=u_k. \end{equation} Furthermore, if $k\geq1$ and $k\neq2$, we have \begin{equation} d(2^k)= 2^k. \end{equation} Finally, we have $d(4)=2$. \end{conj} Numerical simulations performed on Mathematica for $k\leq 12$ support Conjecture~\ref{conj:powersoftwo}. \section{Construction of Arnold's sequence} \label{sec:Arnold} In this section we provide a construction which conjecturally yields Arnold's sequence $(u_k)_{k\geq1}$. We denote by $\mathbb Z_+$ the set of nonnegative integers and we denote by \[ S:=\bigsqcup_{d\geq1}\mathbb Z_+^d \] the set of all finite sequences of nonnegative integers. We define a map $f:S\rightarrow S$, which maps each $\mathbb Z_+^d$ to $\mathbb Z_+^{2d}$, as follows. Fix $\underline{x}=(x_1,\ldots,x_d)\in S$. If all the $x_i$'s are equal to $x_d$, we set \[ f(\underline{x})=(x_d,\ldots,x_d,2x_d,\ldots ,2x_d), \] where $x_d$ and $2x_d$ both appear $d$ times on the right-hand side. Otherwise, define \[ s:=\max \left\{ 1 \leq i \leq d-1 | x_i \neq x_d \right\} \] and set \[ f(\underline{x})=(x_1,\ldots,x_d,x_1+x_d,\ldots,x_{s-1}+x_d,2x_d,\ldots,2x_d), \] where $2x_d$ appears $d-s+1$ times on the right-hand side. For example, we have \begin{equation} f((2,4,4,4))=(2,4,4,4,8,8,8,8) \end{equation} and \begin{equation} f(2,4,4,4,8,8,8,8)=(2,4,4,4,8,8,8,8,10,12,12,16,16,16,16,16). \end{equation} By iterating this function $f$ indefinitely, one produces an infinite sequence : \begin{lem} Fix $d\geq1$ and $\underline{x}\in\mathbb Z_+^d$. There exists a unique (infinite) sequence $(X_k)_{k\geq 1}$ such that for any $k\geq1$ and for any $n\geq \log_2 (k/d)$, $X_k$ is the $k$-th term of the finite sequence $f^n(\underline{x})$. \end{lem} This infinite sequence is called the \emph{$f$-transform} of $\underline{x}$. The lemma follows from the observation that for any $\ell\geq1$ and for any $\underline{y}\in \mathbb Z_+^\ell$, $\underline{y}$ and $f(\underline{y})$ have the same first $\ell$ terms. We can now formulate a conjecture about the construction of the sequence $(u_k)_{k\geq1}$ : \begin{conj} \label{conj:Arnold} Arnold's sequence $(u_k)_{k\geq1}$ is the $f$-transform of the quadruple $(2,4,4,4)$. \end{conj} Conjecture~\ref{conj:Arnold} is supported by the estimation on Mathematica of $u_k$ for every $k\leq 512$. \section*{Acknowledgments} The author acknowledges the support of the Fondation Simone et Cino Del Duca. \bibliographystyle{alpha}
{ "timestamp": "2017-12-27T02:01:24", "yymm": "1712", "arxiv_id": "1712.08666", "language": "en", "url": "https://arxiv.org/abs/1712.08666", "abstract": "For any positive integer $q$, the sequence of the Euler up/down numbers reduced modulo $q$ was proved to be ultimately periodic by Knuth and Buckholtz. Based on computer simulations, we state for each value of $q$ precise conjectures for the minimal period and for the position at which the sequence starts being periodic. When $q$ is a power of $2$, a sequence defined by Arnold appears, and we formulate a conjecture for a simple computation of this sequence.", "subjects": "Combinatorics (math.CO); Number Theory (math.NT)", "title": "Modular periodicity of the Euler numbers and a sequence by Arnold", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9850429112114063, "lm_q2_score": 0.8128673269042767, "lm_q1q2_score": 0.8007091981224226 }
https://arxiv.org/abs/1602.05298
On critical points of random polynomials and spectrum of certain products of random matrices
In the first part we study critical points of random polynomials. We choose two deterministic sequences of complex numbers,whose empirical measures converge to the same probability measure in complex plane. We make a sequence of polynomials whose zeros are chosen from either of sequences at random. We show that the limiting empirical measure of zeros and critical points agree for these polynomials. As a consequence we show that when we randomly perturb the zeros of a deterministic sequence of polynomials, the limiting empirical measures of zeros and critical points agree. This result can be interpreted as an extension of earlier results where randomness is reduced. Pemantle and Rivin initiated the study of critical points of random polynomials. Kabluchko proved the result considering the zeros to be i.i.d. random variables.In the second part we deal with the spectrum of products of Ginibre matrices. Exact eigenvalue density is known for a very few matrix ensembles. For the known ones they often lead to determinantal point process. Let $X_1,X_2,...,X_k$ be i.i.d matrices of size $n \times n$ whose entries are independent complex Gaussian random variables. We derive the eigenvalue density for matrices of the form $Y_1.Y_2....Y_n$, where each $Y_i = X_i$ or $X_i^{-1}$. We show that the eigenvalues form a determinantal point process. The case where $k=2$, $Y_1=X_1,Y_2=X_2^{-1}$ was derived earlier by Krishnapur. The case where $Y_i =X_i$ for all $i=1,2,...,n$, was derived by Akemann and Burda. These two known cases can be obtained as special cases of our result.
\chapter{Critical points of random polynomials} \label{ch:criticalpoints4} \section{Introduction}In this chapter we will investigate the distribution of the critical points in relation to the zeros of a polynomial. For a holomorphic function $f:\mathbb{C} \rightarrow \mathbb{C}$ a point $z \in \mathbb{C}$ is called a critical point of $f$ if $f'(z)=0$. The oldest known result relating the zeros and critical points of a polynomial is Gauss-Lucas theorem, which states that the critical points of any polynomial with complex coefficients lie inside the convex hull formed by the zeros of the polynomial. \begin{theorem}[Gauss-Lucas; See Chapter 2, Theorem 6.1 in~\cite{marden}]\label{guass-lucas} Let $P$ be a non-constant complex polynomial then the zeros of $P'$ are contained in the convex hull formed by the zeros of $P$. \end{theorem} In general nothing more can be said. Our interest is in dealing with sequences of polynomials, usually randomness included, with increasing degrees. We consider the case in which the point cloud made from the zeros of these polynomials will converge to a probability measure in the complex plane. We want to understand the behaviour of critical points of these polynomials. We recall the definition of weak convergence. \begin{definition} For a sequence of probability measures, $\{\mu_n\} \text{ and } \mu$ on $\text{$\mathbb{C}$}$, we say that $\mu_n \xrightarrow{w} \mu$ \textit{weakly}, if for any $ f\in C_c^{\infty}(\text{$\mathbb{C}$})$, we have $\lim\limits_{n \rightarrow \infty}\int_{X}^{}fd\mu_n = \int_{X}^{}fd\mu$. \end{definition} The following definition formalizes the notion of point cloud converging to a probability measure. Here the point cloud being the collection of terms from a sequence of complex numbers. \begin{definition}We say a sequence of complex numbers $\{a_n\}_{n\geq1}$ to be \textit{$\mu$-distributed} if its empirical measures $\frac{1}{n}\sum_{k=1}^{n}\delta_{a_k}$ converge weakly to the probability measure $\mu$. \end{definition} In the next section we study the critical points of deterministic sequence of polynomials. We show that if all the zeros are confined in regions that are well separated, then the critical points also confine to these regions. Then, in subsequent sections we discuss a few examples in which the limiting measures of zeros and of critical points do not agree. Later we give a brief overview of existing results in the literature where a sequence of random polynomials are considered. In all these cases it was shown that the limiting measures of zeros and critical points agree. In the Section 2.3 we show the results we have obtained and discuss their consequences. In the first result we construct the zeros sequence, for each term choose the term from one of the two deterministic sequences at random. We construct a sequence of polynomials whose zeros are the terms of this sequence. We show that the limiting empirical measures of zeros and critical points of these polynomials agree. We then state the corollaries of this result, where we perturb this sequence randomly and show that the limiting measure of zeros and critical points agree. In our second result we consider a random rational function which can be used to get a generalized derivative and show that the limiting distribution of zeros and poles agree. As a corollary of this we show that if we choose a random subsequence of a deterministic sequence then the limiting measures of zeros and critical points agree. In the last section we prove the corollaries mentioned earlier. We defer the proofs of the theorems to the next chapter. We will recall a well known proof of Gauss-Lucas theorem. If $z_1,z_2,\dots,z_n$ are the roots of the polynomial $P$, then for some $c$, $P(z)=c(z-z_1)(z-z_2)\dots(z-z_n)$. Define $L(z):=\frac{P'(z)}{P(z)}=\sum_{k=1}^{n}\frac{1}{z-z_k}$. If $z$ is a zero of $P'$ and not equal to any of the $z_i$s, then $L(z)=0$. Hence, \[ \sum_{k=1}^{n}\frac{1}{z-z_k}=0 \text{\hspace{12 pt} or, \hspace{12 pt} } \sum_{k=1}^{n}\frac{\overline{z}-\overline{z}_k}{|z-z_k|^2}=0. \] Therefore if $L(z)=0$, then $z$ satisfies, \[ z=\frac{\sum_{k=1}^{n}\frac{1}{|z-z_k|^2}z_k}{\sum_{k=1}^{n}\frac{1}{|z-z_k|^2}}. \] In the above equation $z$ is expressed as a convex combination of $z_k$s. In the other case where $z$ is one of the $z_k$s, it trivially true that $z$ is in the convex hull formed by $z_1,z_2,\dots,z_n$. For a different proof of this theorem the reader can refer to \cite[Chapter-2, Theorem 6.1]{marden}. After the proof of Gauss-Lucas theorem, there have been several results concerning critical points and zeros of polynomials. Interested reader may see the references in \cite{pemantle}. Several conjectures on the same can be found in \cite{mardenconjectures} and \cite{borcea}. A brief survey on results connecting zeros and critical points of polynomials can be found in \cite{sury}. In the proof of Gauss-Lucas theorem we have defined a function $L(z)=\sum_{k=1}^{n}\frac{1}{z-z_k}$. It is interpreted as potential at the point $z$ due to unit charges present at points $z_1,z_2,\dots,z_n$. In studying critical points this potential function plays a key role. All the results in this chapter are obtained by analyzing this function. \section{Critical points of a sequence of deterministic polynomials.} Consider a sequence of deterministic polynomials. In the case where all the zeros of these polynomials are in a bounded convex set, Gauss-Lucas theorem asserts that the critical points of these polynomials lie inside the same set. In this context we state a well known related result by Walsh. \begin{theorem}[J.L.Walsh ~\cite{walsh}] Let $C_1$, $C_2$ be disks with centres $c_1$, $c_2$ and radii $r_1$, $r_2$. Let $P$ be a polynomial of degree $n$ with all its zeros in $C_1\cup C_2$, say $n_1$ zeros in $C_1$ and $n_2$ zeros in $C_2$. Then $P$ has all its critical points in $C_1\cup C_2 \cup C_3$, where $C_3$ is the disk with centre $c_3$ and radius $r_3$ given by \[ c_3=\frac{n_1c_2+n_2c_1}{n}, \hspace{5 pt} r_3=\frac{n_1r_2+n_2r_1}{n}. \] Furthermore, if $C_1$, $C_2$ and $C_3$ are pairwise disjoint, then $C_1$ contains $n_1-1$ critical points, $C_2$ contains $n_2-1$ critical points and $C_3$ contains $1$ critical point. \end{theorem} Inspired by the above theorem of Walsh, we derive the following result. Consider a sequence of polynomials whose zeros are in well separated clusters (say for example separated unit disks). Further assume that the number of zeros in these sets grow proportionately. Under these assumptions we show that in a neighbourhood of each of these sets the number of critical points differ from the number of zeros by at most a constant number. \begin{theorem}\label{walsh_general} Let $S_1,S_2,\dots, S_k$ be pairwise disjoint bounded convex sets in the complex plane. Assume that $\mbox{diam}(S_i)\leq 1$, for $i=1,2,\dots,k$ and the distance of separation between any two $S_i,S_j$, for $i\neq j$, is at least $5k$. Define the sequence of polynomials $\{P_n\}_{n\geq 1}$ as $P_n(z):=\prod\limits_{j=1}^{k}\prod\limits_{i=1}^{n}(z-z^{(j)}_i)$, where $z_i^{(j)} \in S_j$ for $j=1,\dots, k$ and $i=1,\dots, n$. Then, for $\epsilon > \frac{3-\sqrt{5}}{2}$ we have a constant $c(k,\epsilon)$, such that the number of zeros of $P_n'(z)$ in the $S_i^{\epsilon}$ is at least $n-c(k,\epsilon)$ for any $i=1,\dots,k$. \end{theorem} \begin{proof} We will estimate the number of critical points of $P_n(z)$ in $S_1^\epsilon$ the $\epsilon$ neighbourhood of $S_1$. Let the diameter of the sets $S_i$ be at most $d$ for any $i \in \{1,2,\dots,k\}$. Assume that the separation between any two $S_i,S_j$ is at least $d_s$, where $d_s>0$. In the course of this proof we will substitute the values $d=1$ and $d_s=5k$ as given in the statement of the theorem. For $\epsilon$ small enough, we know that there are $n$ zeros of $P_n(z)$ in $S_1^{\epsilon}$. Argument principle computes the difference between zeros and poles of a meromorphic function in a domain by evaluating a certain integral on the boundary of the domain. We will use argument principle to estimate the critical points of $P_n(z)$ in $S_1^{\epsilon}$. A version of argument principle is stated below. Let $f: U \rightarrow\mathbb{C}$ be a meromorphic function on a simply connected domain $U$ and let $C$ be a rectifiable simple closed curve in $U$. Assume that $f$ does not vanish on $C$. Then \begin{equation} \frac{1}{2\pi i}\oint_{C}\frac{f'(z)}{f(z)}dz = N_Z\left(f,C\right)-N_P\left(f,C\right)\label{argumentprinciple} \end{equation} where $N_Z\left(f,C\right)$ and $N_P\left(f,C\right)$ are the number of zeros and poles of $f$ enclosed by the curve $C$. Define $L_n(z):=\frac{P_n'(z)}{P_n(z)}=\sum\limits_{j=1}^{k}\sum\limits_{s=1}^{n}\frac{1}{z-z_s^{(j)}}$ and notice that the zeros and poles of $L_n(z)$ are zeros of $P_n'(z)$ and $P_n(z)$ respectively. We shall apply the argument principle for the function $L_n(z)$ for the boundary curve $\gamma_1^\epsilon$ obtained from $\partial S_1^{\epsilon}$. Assume that there are no zeros of $P_n'(z)$ on the curve $\gamma_1^\epsilon$. Hence applying the formula \eqref{argumentprinciple} to $L_n(z)$ we get, \begin{align} |N_Z\left(L_n,\gamma_1^\epsilon\right)-N_P\left(L_n,\gamma_1^\epsilon\right)| & = \bigg|\frac{1}{2\pi}\oint\limits_{\gamma_1^\epsilon}\frac{L_n'(z)}{L_n(z)}dz\bigg| & \text{(or)} \\ |N_Z\left(P_n,\gamma_1^\epsilon\right)-n| & \leq \frac{1}{2\pi}\oint\limits_{\gamma_1^\epsilon}\bigg|\frac{L_n'(z)}{L_n(z)}\bigg||dz| \label{lineintegral} \end{align} For the above integrand we will give an upper bound for the numerator and a lower bound for the denominator on the curve $\gamma_1^\epsilon$. Because $|z-z_s^{(j)}|\geq\epsilon$ and from the triangle inequality, for $z$ on the curve $\gamma_1^\epsilon$ we get, \begin{equation} |L_n'(z)|=\Biggl|-\sum\limits_{j=1}^{k}\sum\limits_{s=1}^{n}\frac{1}{\left(z-z^{(j)}_s\right)^2}\Biggr| \leq \sum\limits_{j=1}^{k}\sum\limits_{s=1}^{n}\frac{1}{\bigl|z-z^{(j)}_s\bigr|^2} \leq \frac{kn}{\epsilon^2}.\label{numerator} \end{equation} Similarly for $L_n(z)$ using triangle inequality we get, \begin{align} |L_n(z)|=\left|\sum\limits_{j=1}^{k}\sum\limits_{s=1}^{n}\frac{1}{z-z^{(j)}_s}\right| & \geq \left|\sum\limits_{s=1}^{n}\frac{1}{z-z^{(1)}_s}\right|- \left|\sum\limits_{j=2}^{k}\sum\limits_{s=1}^{n}\frac{1}{z-z^{(j)}_s}\right|\label{eqn:lemma0:1} \\ \end{align} The first term in the right most expression in \eqref{eqn:lemma0:1} is invariant under multiplication by $e^{i\theta}$. Therefore for any $\theta \in [0,2\pi)$ we have, \begin{align} |L_n(z)| &\geq\left|\sum\limits_{s=1}^{n}\frac{e^{i\theta}}{z-z^{(1)}_s}\right|-\left|\sum\limits_{j=2}^{k}\sum\limits_{s=1}^{n}\frac{1}{z-z^{(j)}_s}\right|,\\ &\geq \left|\sum\limits_{s=1}^{n}\Im\left(\frac{e^{i\theta}}{z-z^{(1)}_s}\right)\right|- \left|\sum\limits_{j=2}^{k}\sum\limits_{s=1}^{n}\frac{1}{z-z^{(j)}_s}\right|, \\ & \geq \frac{n\epsilon}{(d+\epsilon)^2}-\frac{(k-1)n}{d_s-\epsilon}. \end{align} because $S_1^\epsilon$ is a convex set, there is a line passing through $z\in \partial S_1^\epsilon$ (from separating hyperplane theorem) such that all the points $z_s^{(1)}$ lie on the same side of this line. Let $\theta$ be the angle made by this line with real axis, then all the terms in the $\sum\limits_{s=1}^{n}\Im{\frac{e^{i\theta}\left(\overline z- \overline z_s^{(1)}\right)}{\left|z-z_s^{(1)}\right|^2}}$ have the same sign and the absolute value of the numerators is atleast $\epsilon$ and denominators with at most $(d+\epsilon)^2$. Similarly $\bigg|\sum\limits_{j =2}^{k}\sum\limits_{s=1}^{n}\frac{1}{z-z_s^{(j)}}\bigg| \leq \sum\limits_{j=2}^{k}\sum\limits_{s=1}^{n}\frac{1}{|z-z_s^{(j)}|}$, and for $j\neq 1$ the denominator in the previous expression $|z-z_s^{(j)}|$ is at least $d_s-\epsilon$, because $|z-z_s^{(j)}| \geq d(S_1^\epsilon,S_j)\geq d(S_1,S_j)-\epsilon \geq d_s-\epsilon$. For $z \in \partial S_1^\epsilon$, we obtain \begin{equation} \left|L_n(z)\right| \geq \frac{n\epsilon}{(d+\epsilon)^2}-\frac{(k-1)n}{d_s-\epsilon}. \label{denominator} \end{equation} By substituting $d=1$, for the right hand side of \eqref{denominator} to be positive, we need $d_s$ to satisfy \begin{equation} d_s>\epsilon+\frac{(k-1)(1+\epsilon)^2}{\epsilon}.\label{eqn:d_sandepsilon} \end{equation} For any choice of $d_s\geq 5k-2$ and $\epsilon \in [\frac{3-\sqrt{5}}{2},\frac{3+\sqrt{5}}{2}]$, the inequality \eqref{eqn:d_sandepsilon} is satisfied. For these choices of variables we have, \begin{align} |N_Z\left(P_n,\gamma_1^\epsilon\right)-n| & \leq \frac{1}{2\pi}\oint\limits_{\gamma_1^\epsilon}\bigg|\frac{L_n'(z)}{L_n(z)}\bigg||dz|,\\ & \leq \frac{1}{2\pi}\oint\limits_{\gamma_1^\epsilon}\dfrac{\frac{kn}{\epsilon^2}}{\frac{n\epsilon}{(1+\epsilon)^2}-\frac{(k-1)n}{d_s-\epsilon}}|dz|,\\ &\leq \frac{1+2\epsilon}{2\epsilon^2}\dfrac{k}{\frac{\epsilon}{(1+\epsilon)^2}-\frac{(k-1)}{d_s-\epsilon}} =: c(k,\epsilon,d_s). \label{eqn:integral} \end{align} The inequality \eqref{eqn:integral} is obtained from the fact $p\leq \pi d$, where $p$ is the perimeter of the convex set and $d$ is the diameter of the convex set. Substituting $d_s=5k$ we obtain that $|N_Z(L_n,\gamma_1^\epsilon)-N_P(L_n,\gamma_1^\epsilon)|\leq c(k,\epsilon)$. By the choice of $\epsilon$, we have proved the Theorem for $\epsilon \in [\frac{3-\sqrt{5}}{2},\frac{3+\sqrt{5}}{2}]$. For $\epsilon>\frac{3+\sqrt{5}}{2}$, $S_1^\epsilon$ contains $\frac{3+\sqrt{5}}{2}$-neighbourhood of $S_1$, hence number of critical points in $S_1^\epsilon$ is at least $n-c(k,\frac{3+\sqrt{5}}{2})$. Choosing $c(k,\epsilon)=c(k,\frac{3+\sqrt{5}}{2})$, the Theorem is proved for $\epsilon>\frac{3+\sqrt{5}}{2}$. \end{proof} \begin{remark} In Theorem \ref{walsh_general} we have assumed all $S_i$ have equal number of zeros of $P_n$. Instead we may assume that the number of zeros $P_n$ in $S_i$ be $c_in$ for constants $c_1,c_2,\dots,c_k$ and obtain a similar result. The same ideas in the above proof can be used to prove this result. \end{remark} It was raised by Pemantle and Rivin whether it is true that if the limiting measure of zeros of the sequence of polynomials converging to a probability measure $\mu$, then the limiting measure of critical points also converge to $\mu$. We will see in the forthcoming example that this is indeed false. For convenience we will introduce the following notation. For any polynomial $P$, denote $Z(P)$ to be the multi-set of zeros of $P$ and $\text{$\mathscr{M}$}(P)$ to be the uniform probability measure on $Z(P)$. Here we will construct sequence of polynomials for which the limiting measure of zeros and critical points do not agree. The most commonly quoted~\cite{pemantle} sequence of polynomials in this regard is $P_n(z)=z^n-1$. In this case the limiting zero measure is the uniform probability measure on $S^1$ and the limiting critical point measure is the Dirac measure at origin. We generalize the above stated example and construct new set of examples for which the limiting measures of zeros and critical points are different. \begin{eg} Observe that if a polynomial has all zeros real, then all its critical points have to be real and are interlaced between the zeros of the polynomial. Consider the polynomial $P_n(z)=(z-a_1^n)(z-a_2^n)\dots(z-a_k^n)$, where $a_1,a_2,\dots,a_k$ are real numbers such that $0 < a_1 < a_2 < \dots < a_k$. Define the sequence of polynomials to be $Q_n(z)=P_n(z^n)$, then $Q_n'(z)=nz^{n-1}P_n'(z^n)$. The zero set of $Q_n$ is \[Z(Q_n)=\bigcup\limits_{j=1}^{k}\bigcup\limits_{\ell=1}^{n}\{a_je^{ 2\pi i\frac{\ell}{n}}\}.\] Where as the zero set of $Q_n'$ is \[Z(Q_n')=\left(\bigcup\limits_{j=1}^{k-1}\bigcup\limits_{\ell=1}^{n} \{b_{j,n}^{\frac{1}{n}}e^{ 2\pi i\frac{\ell}{n}}\}\right)\bigcup\{0,0,\dots,0\},\] where $b_{1,n},b_{2,n},\dots,b_{k-1,n}$ are the zeros of the polynomial $P_n'(z)$. The probability measure $\text{$\mathscr{M}$}(Q_n')$ has mass $\frac{n-1}{kn-1}$ at $0$, hence its limiting measure will have mass $\frac{1}{k}$ at $0$. On the other hand the probability measure $\text{$\mathscr{M}$}(Q_n)$ is supported on $\bigcup\limits_{j=1}^{k}a_jS^1$. Hence the limiting measures do not agree. \end{eg} \begin{eg} For the second class of examples choose a sequence of complex numbers all containing in a disk of radius $r$ around 0. Let the sequence be $\{a_n\}_{n\geq1}$ and $|a_n|\leq r$. Make a sequence of polynomials using the terms of this sequence as its zeros. Define $P_n(z)=(z-a_1)(z-a_2)\dots(z-a_n)$. Using these define the polynomials $Q_n(z)=\int_{0}^{z}P_n(w)dw+(2r+d)^{n+1}$, where $d>0$. Notice that $Q_n'(z)=P_n(z)$. We will show that the polynomial $Q_n(z)$ does not vanish in the disk $\textbf{$\mathbb{D}$}_r$. For this observe \begin{align} \min\limits_{z \in \text{$\mathbb{D}$}_r}|Q_n(z)| & \geq |(2r+d)^n-\max\limits_{z \in \text{$\mathbb{D}$}_r}|\int_{0}^{z}P_n(w)dw||,\\ & \geq |(2r+d)^{n+1}-r\max\limits_{z\in \text{$\mathbb{D}$}_r}|P_n(z)||,\\ & =|(2r+d)^{n+1}-r(2r)^n| >0. \end{align} Therefore $Q_n(z)$ does not vanish in the disk $\text{$\mathbb{D}$}_r$. Hence for all large $n$ the zeros of $Q_n(z)$ are outside the disk $\text{$\mathbb{D}$}_{2r+d}$. Assuming that $Q_n(z)$ has a limiting zero measure, the support of the limiting zero measure of $Q_n(z)$ is disjoint from the support of the limiting zero measure of $P_n(z)$. \end{eg} \begin{eg} We will illustrate a more concrete example based on the above technique. Choose a polynomial $P$, whose zeros are in $\text{$\mathbb{D}$}_r$, where $r<1$. Define $Q_n(z)=P^n(z)-1$, then $Q_n'=nP^{n-1}(z)P'(z)$. If $z$ is a zero of $Q_n(z)$, then it satisfies $P^n(z)=1$, or $|P(z)|=1$. Therefore the limiting zero measure of $Q_n(z)$ is supported on the boundary of the lemniscate $\{z:|P(z)|\leq1\}$ of the polynomial $P$. The limiting zero measure for the sequence $\{Q_n\}_{n \geq 1}$ exists because $Q_n$ is the $nk$-th Chebyshev polynomial of the lemniscate of $P$. Hence the limiting zero measure is the equilibrium measure for the domain $\{z:|P(z)|\leq1\}$. For a detailed discussion on the relation between Chebyshev polynomials and equilibrium measures, the reader can refer Chapter 5 in \cite{ransford}. On the other side, if $z_1,z_2,\dots,z_k$ are the roots of the polynomial $P$, then the limiting zero distribution of $Q_n'$ will be $\frac{1}{k}\sum\limits_{i=1}^{k}\delta_{z_i}$. Hence the limiting measures of zeros and critical points of the given sequence of polynomials do not agree. \end{eg} In this context we quote the question posed by Pemantle and Rivin in \cite{pemantle}. \begin{question}\label{Pemantle_question} When are the zeros of $P_n'$ stochastically similar to the zeros of $P_n$? \end{question} \section{Critical points of random polynomials.} To tackle the Question \ref{Pemantle_question}, $P_n$s can be considered to be random. The study of critical points of random polynomials through random zeros was initiated by Pemantle and Rivin in \cite{pemantle}. They considered a sequence of random polynomials whose zeros are i.i.d. with law $\mu$ having finite 1-energy and proved that the empirical law of critical points converge weakly to the same probability measure $\mu$. A similar result for probability measures supporting on $S^1$ was proved by Subramanian \cite{sneha}. Kabluchko in \cite{kabluchko} proved the result without any assumption on $\mu$. Before stating the above mentioned results we recall the modes of convergence for random measures. \begin{definition}\label{modes of convergence} Let $\mbox{\textbf{M}}(\text{$\mathbb{C}$})$ be the set of probability measures on the complex plane, equipped with \textit{weak topology}. Let $\{\mu_n\}_{n\geq1}$ be a sequence in $\mbox{\textbf{M}}(\text{$\mathbb{C}$})$ and $\mu \in \mbox{\textbf{M}}(\text{$\mathbb{C}$})$ we say, \begin{itemize} \item $\mu_n \xrightarrow{w} \mu$ in probability if $\lim\limits_{n\rightarrow\infty}\Pr(\mu_n \in N_\mu)=1$ for any neighbourhood $N_\mu$ of $\mu$, \item $\mu_n \xrightarrow{w} \mu$ almost surely if $\Pr(\lim\limits_{n\rightarrow\infty}\mu_n \in N_\mu)=1$ for any neighbourhood $N_\mu$ of $\mu$. \end{itemize} \end{definition} We now give precise statements of the results in \cite{kabluchko} and \cite{pemantle}. \begin{definition} Define the \textit{p-energy of $\mu$} to be \[ \mathcal{E}_\text{$p$}(\mu):=\left(\int\limits_{\mathbb{C}}\int\limits_{\mathbb{C}}\text{$\frac{1}{|z-w|^p}d\mu(z)d\mu(w)$}\right)^\textbf{$\frac{1}{p}$} \] \end{definition} \begin{theorem}[Pemantle-Rivin ~\cite{pemantle}]\label{pemantle-rivin}Let $X_1,X_2,\dots $ be a sequence of i.i.d random variables from the probability measure $\mu$. Assume that $\mu$ has finite 1-energy. Let $P_n(z)=(z-X_1)(z-X_2)\dots(z-X_n)$, then the critical points measure $\text{$\mathscr{M}$}(P_n')\xrightarrow{w}\mu$ almost surely. \end{theorem} One limitation of the above result is that it is not applicable to probability measures that are supported on 1-dimensional subsets of the complex plane. But the result can be easily verified for probability measures supported on real line. By Rolle's theorem the critical points are interlaced between the roots of the polynomial. Hence the L\'{e}vy distance between the zeros measure $\text{$\mathscr{M}$}(P_n)$ and critical points measure $\text{$\mathscr{M}$}(P_n')$ is at most $\frac{1}{n}$. On the other side the zeros measure $\text{$\mathscr{M}$}(P_n)$ has a limiting measure $\mu$ which is the probability measure from which the random variables are drawn. Combining the previous two observations the result follows. Pemantle and Rivin in \cite{pemantle} conjectured that the statement of Theorem \ref{pemantle-rivin} is true without any assumptions on $\mu$. \begin{conjecture}[Pemantle-Rivin \cite{pemantle}]\label{pemantle_conjecture} Let $X_1,X_2,\dots$ be i.i.d. random variables distributed according to a probability measure $\mu$ and $P_n(z):=(z-X_1)(z-X_2)\dots(z-X_n)$. Then $\text{$\mathscr{M}$}(P_n')\xrightarrow{w}\mu$ almost surely. \end{conjecture} Kabluchko proved the conjecture of Pemantle and Rivin in a weak form. \begin{theorem}[Kabluchko~\cite{kabluchko}]\label{kabluchko} Let $X_1, X_2 ,\dots$ be i.i.d. random variables distributed according to a probability measure $\mu$ and $P_n(z):=(z-X_1)(z-X_2)\dots(z-X_n)$. Then $\text{$\mathscr{M}$}(P_n')\xrightarrow{w}\mu$ in probability. \end{theorem} Further results concerning critical points and zeros of random polynomials are discussed below. In \cite{cheung} the authors (Pak-Leong Cheung, Tuen Wai Ng, Jonathan Tsai and SCP Yam) prove that the empirical law of zeros of the higher derivatives for the polynomial whose zeros are i.i.d. with law $\mu$ supported in $S^1$ converge to the same probability measure $\mu$. In \cite{cheung} the authors also obtain similar results for the zeros of generalized derivatives of polynomials. Similar results for critical points of characteristic polynomials of random matrix ensembles (Haar distributed on $O(n)$, $SO(n)$, $U(n)$, $Sp(n)$) are proved in \cite{orourke} by O'Rourke. In this section we will present two results concerning the zeros and critical points of the sequence of random polynomials. In the previous section we have seen examples of polynomials for which the limiting empirical distribution of zeros and critical points do not agree. Where as if the zeros of the polynomial are chosen to be i.i.d. random variables, then the statement holds \cite{kabluchko}. These results bridge the gap between the two scenarios, i.e., we reduce the randomness in choosing the zeros and show that the statement holds. In Theorem \ref{thm1} we will start with two sequences of complex numbers which are asymptotically distributed according to a same probability measure. We also assume that the two sequences are sufficiently different (precise conditions are stated in the theorem). Then we construct a sequence of random numbers, whose terms are chosen independently at random from the corresponding terms of either of the sequences. If we make a sequence of polynomials whose zeros are the terms of the obtained random sequence, then the limiting measure of the critical points of this sequence of polynomials will agree with that of the limiting measure of the sequences we started with. We prove the result for a specific class of sequences which we call as \textit{log-Ces\'{a}ro-bounded} which is defined as follows. \begin{definition}\label{def:log-cesaro} We say a sequence of complex numbers $\{a_n\}_{n\geq1}$ to be \textit{log-Ces\'{a}ro-bounded} if the Ces\'{a}ro means of the positive part of their logarithms are bounded i.e., the sequence $\{\frac{1}{n}\sum_{k=1}^{n}\log_+|a_k|\}$ is bounded. \end{definition} \begin{eg} Any bounded sequence is a log-Ces\'{a}ro bounded. \end{eg} \begin{theorem}\label{thm1} Let $\{a_k\}_{k\geq1}$ and $\{b_k\}_{k\geq1}$ be two $\mu$-distributed and log-Ces\'{a}ro bounded sequences of complex numbers. Additionally assume that, $a_k \neq b_k$ for infinitely many $k$. Define the sequence of independent random variables $\xi_k$ such that $\xi_k = a_k $ or $b_k$ with equal probability, for $k\geq1$. Define the polynomials $P_n(z):=(z-\xi_1)(z-\xi_2)\dots(z-\xi_n)$. Then, $\text{$\mathscr{M}$}(P_n)\xrightarrow{w}\mu$ almost surely and $\text{$\mathscr{M}$}(P_n')\xrightarrow{w}\mu$ in probability. \end{theorem} For the assertion of the above theorem to hold, it is necessary to assume that the two sequences differ in infinitely many terms. Suppose not, we may choose one of the sequence to be a sequence for which the assertion of the theorem doesn't hold. Since both the sequences differ only in finitely many terms, the resulting sequence will be same as that of the sequence for which the assertion doesn't hold, with non zero probability. Hence with positive probability the statement of the Theorem \ref{thm1} doesn't hold. The log-Ces\'{a}ro boundedness on the sequences is assumed to enable the proof. We don't have any strong reason for either of the cases whether it is necessary or not. The Theorem \ref{thm1} can be used to obtain corollaries of the following form. Choose a deterministic sequence which is $\mu$-distributed and perturb each of its term by a random variable with diminishing variances. It can be obtained that the empirical measure of the critical points of the polynomial, made from the perturbed sequence also converge to the same limiting probability measure $\mu$. \begin{corollary}\label{Symmetric perturbations} Let $\{u_n\}_{n\geq1}$ be a $\mu$-distributed sequence and log-Ces\'{a}ro bounded sequence. Let $\{v_n\}_{n\geq1}$ be the sequence such that $v_n=u_n+\sigma_nX_n$, where $X_n$s are i.i.d random variables satisfying $X_n\stackrel{d}{=}-X_n$, $\ee{\text{$|X_n|$}}<\infty$ and $\sigma_n \downarrow 0$, $\sigma_n\neq0$. Define the polynomial $P_n(z):=(z-v_1)(z-v_2)\dots(z-v_n)$. Then, $\text{$\mathscr{M}$}(P_n)\xrightarrow{w}\mu$ almost surely and $\text{$\mathscr{M}$}(P_n')\xrightarrow{w}\mu$ in probability. \end{corollary} \begin{remark} In Corollary \ref{Symmetric perturbations}, we may choose the random variables $X_n$s to have complex Gaussian distribution or uniform distribution on unit disk centred at $0$. In the case of complex Gaussian distributed random variables we get the result for unbounded perturbations and in the case of uniformly distributed random variables the perturbations are bounded. \end{remark} It is an easy fact (Page 15 in \cite{manjubook}) that if $\{X_n\}_{n \geq 1}$ is a sequence of i.i.d random variables that are not identically $0$ such that $\ee{\textbf{$\log_+|X_1|$}}<\infty$, then $\limsup\limits_{n \rightarrow \infty}|X_n|^\frac{1}{n}=1$. A special case of Theorem \ref{kabluchko} can be obtained as a corollary of the Theorem \ref{thm1}. The special case being the one in which the probability measure $\mu$ in consideration has bounded $\log_+$-moment. \begin{corollary}\label{corollary4:kabluchko} Let $\mu$ be any probability measure on $\mathbb{C}$ satisfying $\int\limits_{\mathbb{C}}\log_+|z|d\mu(z)<\infty$. Let $X_1,X_2,\dots, X_n$ be i.i.d random variables distributed according to $\mu$. Define the polynomials $P_n(z):=(z-X_1)(z-X_2)\dots(z-X_n)$. Then, $\text{$\mathscr{M}$}(P_n)\xrightarrow{w}\mu$ almost surely and $\text{$\mathscr{M}$}(P_n')\xrightarrow{w}\mu$ in probability. \end{corollary} Let $\{u_n\}_{n\geq1}$ be $\mu$-distributed sequence and $\{v_n\}_{n\geq1}$ be $\nu$-distributed sequence and both are log-Ces\'{a}ro bounded. We replace the terms in the first sequence with those of the second sequence each with probability $p>0$. Let the random sequence be $\{\xi_n\}_{n\geq1}$, define $P_n(z)=(z-\xi_1)(z-\xi_2)\dots(z-\xi_n)$. Then, the limiting empirical measures of zeros and critical points of the polynomial $P_n$ will agree. We state this as the following proposition. \begin{proposition}\label{corollary5:thm1} Let $\{u_k\}_{k\geq1}$ be a $\mu$-distributed sequence of complex numbers and $\{v_k\}_{k\geq1}$ be a $\nu$-distributed sequence of complex numbers. Assume that both the sequences $\{u_k\}_{k\geq1}$ and $\{v_k\}_{k\geq1}$ are log-Ces\'{a}ro bounded and $u_k\neq v_k$ for infinitely many $k$. For $i\geq1$ define the sequence of independent random variables to be $\xi_i = u_i $ with probability $p$ and $v_i$ with probability $1-p$, where $0 < p<1$. Define the polynomials $P_n(z):=(z-\xi_1)(z-\xi_2)\dots(z-\xi_n)$. Then, $\text{$\mathscr{M}$}(P_n)\xrightarrow{w} p\mu+(1-p)\nu$ almost surely and $\text{$\mathscr{M}$}(P_n')\xrightarrow{w} p\mu+(1-p)\nu$ in probability. \end{proposition} We have seen examples of sequence of polynomials for which the limiting measure of zeros and critical points do not agree. Consider the case of the sequence of polynomials whose $n$-th term is $P_n(z)=z^n-1$. We have seen that the limiting measure of zeros is the uniform probability measure on $S^1$ and that of critical points is dirac measure at $0$. The zeros of $P_n$ are the $n$-th roots of unity. The zeros are symmetrical and balanced in many respects. Removing any of these zeros can disturb this symmetry and be considered as a perturbation asymptotically. Define the sequence to be $\{Q_n\}_{n\geq1}$, where \[Q_n(z)=\frac{P_{n+1}(z)}{z-1}=z^n+z^{n-1}+\dots+1.\] It will be shown that the limiting zero measure of the sequence $\{Q_n\}_{n\geq1}$ is the uniform probability measure on $S^1$. The derivative of these polynomials is \[Q_n'(z)=nz^{n-1}+(n-2)z^{n-1}+\dots+1=\frac{nz^{n+1}-(n+1)z^n+1}{(z-1)^2}.\] We will show that the limiting zero measure of $(z-1)^2Q_n'(z)$ $(=nz^{n+1}-(n+1)z^n+1)$ is the uniform probability measure on $S^1$ which in turn gives the limiting zero measure for $Q_n'$. Fix any $r>1$, there is $N_r$ such that whenever $n>N_r$, for $|z|\geq r$ we have, \[|nz^{n+1}-(n+1)z^n+1|\geq|n|z|^{n+1}+1-(n+1)|z|^n|>0.\] Similarly fix any $r<1$, there is $N_r$ such that whenever $n>N_r$, for $|z|\leq r$ we have, \[|nz^{n+1}-(n+1)z^n+1|=|z|^{n+1}\bigg|n-\frac{n+1}{z}+\frac{1}{z^{n+1}}\bigg|\geq|z|^{n+1}\bigg|\big|\frac{n+1}{z}-n\big|-\frac{1}{|z|^{n+1}}\bigg|>0.\] Hence the limiting zero measure of the sequence $\{Q_n'\}_{n\geq1}$ is supported on $S^1$. If we show that asymptotically the angular distribution of the zeros of $Q_n'$ is uniform on $[0,2\pi)$, then it follows that the limiting zero measure of the sequence $\{Q_n'\}_{n\geq1}$ is the uniform probability measure on $S^1$. To show this we use a bound of Erd\"{o}s-Turan for the discrepancy between a probability measure and uniform measure on $S^1$. We will sate the inequality in the case where the two measures are counting probability measure zeros of polynomial and uniform probability measure on $S^1$. \begin{theorem}[Erd\"{o}s-Turan~\cite{erdos-turan}] Let $\{a_k\}_{0\leq k\leq N}$ be a sequence of complex numbers such that $a_0a_N\neq 0$ and let, \[P(z)=\sum\limits_{k=0}^{N}a_kz^k.\] Then, \[\bigg|\frac{1}{N}\nu_N(\theta,\phi)-\frac{\phi-\theta}{2\pi}\bigg|^2\leq\frac{C}{N}\log\bigg|\frac{\sum_{k=0}^{N}|a_k|}{\sqrt{|a_0a_N|}}\bigg|,\] for some constant $C$ and $\nu_N(\theta,\phi):=\#\{z_k:\theta\leq\arg(z_k)<\phi\}$, where $z_1,z_2,\dots,z_N$ are zeros of $P(z)$. \end{theorem} Applying the above inequality for the polynomial $(z-1)^2Q_n'(z)$, we get \[\bigg|\frac{1}{n}\nu_n(\theta,\phi)-\frac{\phi-\theta}{2\pi}\bigg|^2\leq\frac{C}{n}\log\bigg|\frac{2n+2}{\sqrt{n}}\bigg|\stackrel{n\rightarrow \infty}{\longrightarrow}0.\] Therefore the limiting zero measure of $Q_n'$ is uniform probability measure on $S^1$ which agrees with the limiting zero measure of $Q_n$. As an application of the forthcoming theorem we will see that if we choose random subsequence from a $\mu$-distributed sequence, then the limiting distribution of zeros and critical points agree for the polynomials made from this random sequence. The next result (Theorem \ref{thm2}) deals with counting the zeros and pole of a random rational function. The random rational function is defined as $L_n(z)=\sum\limits_{k=1}^{n}\frac{a_k}{z-z_k}$. In a special case where $\sum_{k=1}^{n}a_k=n$ and $a_k>0$ for every $k=1,2,\dots,n$, it is called generalized Sz.-Nagy derivative. For a classical derivative all $a_k$s are equal to $1$. It is mentioned in \cite{rahmanbook} that the motivation in studying generalized derivative is that many of the results for classical derivatives extend to the generalized derivatives. \begin{theorem}\label{thm2} Let $a_1, a_2, \dots$ be i.i.d. random variables satisfying $\ee{\text{$|a_1|$}}\text{$< \infty$}$. Let $\{z_n\}_{n\geq 1}$ be a sequence satisfying that for Lebesgue a.e. $z \in \mathbb{C}$ there exists a compact set $K_z$ with $d(z,K_z)>0$ such that there are infinitely many $z_k$'s in $K_z$, and there is a point $\omega$ that is not a limit point of $z_n$'s. Define $L_n(z):= \frac{a_1}{z-z_1}+\frac{a_2 }{z-z_2}\dots+\frac{a_n}{z-z_n} $. Then $\frac{1}{n}\Delta \log(|L_n(z)|)\rightarrow 0$ in probability, in the sense of distributions. \end{theorem} In the statement of the above theorem, there is a mention of the sequence $\{z_n\}_{n\geq1}$ satisfying that for Lebesgue a.e. $z \in \mathbb{C}$ there exists a compact set $K_z$ with $d(z,K_z)>0$ such that there are infinitely many $z_k$'s in $K_z$, and there is a point $\omega$ that is not a limit point of $z_n$'s. Several classes of sequences satisfy this condition. For example any bounded sequence or any sequence that is not dense and $\mu$-distributed for appropriate $\mu$ satisfies this condition. \begin{remark} In Theorem \ref{thm2}, let $L_n(z)=\frac{Q_n(z)}{P_n(z)}$. Where $Q_n(z)$ id defined to be the generalized derivative of the polynomial $P_n$. Then Theorem \ref{thm2} asserts that $\frac{1}{n}\Delta\log|L_n(z)|\rightarrow0$, which in turn imply that $\text{$\mathscr{M}$}(Q_n)-\text{$\mathscr{M}$}(P_n)\rightarrow 0$ in the sense of distributions. If we assume that the sequence $\{z_k\}_{k\geq1}$ is $\mu$-distributed then it follows that the limiting measure of critical points converge to $\mu$. \end{remark} As an application of previous Theorem \ref{thm1} we choose a $\mu$-distributed deterministic sequence and perturb it randomly and show that the empirical distribution of critical points is also $\mu$. Instead here we choose a random subsequence of a $\mu$-distributed sequence and show that the corresponding result holds. We state this result as the following corollary. \begin{corollary}\label{corollary:thm2} Let $\{z_n\}_{n\geq1}$ be a $\mu$-distributed sequence that is not dense in $\mathbb{C}$, for a $\mu$ which is not supported on the whole complex plane. Choose a subsequence $\{z_{n_k}\}_{k \geq 1}$ at random that is, each of $z_n$ is part of subsequence with probability $p<1$ independent of others. Define the polynomials $P_k(z):=(z-z_{n_1})(z-z_{n_2})\dots(z-z_{n_k})$. Then, $\text{$\mathscr{M}$}(P_k)\xrightarrow{w}\mu$ almost surely and $\text{$\mathscr{M}$}(P_k')\xrightarrow{w}\mu$ in probability. \end{corollary} \section{Proofs of corollaries and Proposition \ref{corollary5:thm1}.} In Corollary \ref{Symmetric perturbations} we deal with perturbations of a $\mu$-distributed sequence. We expect that the perturbed sequence will also have the same limiting probability measure as of the original sequence. It is formally stated and proved in the following lemma. \begin{lemma}\label{lemma:perturb_limit_measure} Let $\{a_n\}_{n\geq1}$ be a $\mu$-distributed sequence, $\sigma_n\downarrow0$ and $X_1,X_2,\dots$ are i.i.d. random variables. Then, $\{a_n+\sigma_nX_n\}_{n\geq1}$ is a $\mu$-distributed sequence almost surely. \end{lemma} \begin{proof} It is enough to show that for any $f\in C_c^\infty(\mathbb{C})$, \[\frac{1}{n}\sum\limits_{k=1}^{n}\left(f(a_k)-f(a_k+\sigma_kX_k)\right)\rightarrow 0,\] almost surely. Fix $\epsilon>0$, choose $M$ such that $\Pr(|X_n|>M)<\epsilon.$ Then, \begin{align} \frac{1}{n}|\sum\limits_{k=0}^{n}(f(a_k)-f(a_k+\sigma_kX_k)| & \leq \frac{1}{n}\sum\limits_{k=1}^{n}|(f(a_k)-f(a_k+\sigma_kX_k))\text{$\mathbbm{1}$}\{|X_k|>M\}|\\&+\frac{1}{n}\sum\limits_{k=1}^{n}|(f(a_k)-f(a_k+\sigma_kX_k))\text{$\mathbbm{1}$}\{|X_k|\leq M\}|,\\ &\leq\frac{ 2||f||_\infty}{n} \sum\limits_{k=1}^{n}\text{$\mathbbm{1}$}\{|X_k|>M\}+\frac{1}{n}\sum\limits_{k=1}^{n}|\sigma_kX_k|||f'||_\infty,\\ &\leq \frac{ 2||f||_\infty}{n} \sum\limits_{k=1}^{n}\text{$\mathbbm{1}$}\{|X_k|>M\} + \frac{M||f'||_\infty}{n}\sum\limits_{k=1}^{n}\sigma_k. \label{eqn:perturb_limit_measure} \end{align} Using law of large numbers and $\sigma_n\downarrow0$in the above equation \ref{eqn:perturb_limit_measure} we have \[ \lim\limits_{n\rightarrow\infty}\frac{1}{n}\bigg|\sum\limits_{k=0}^{n}(f(a_k)-f(a_k+\sigma_kX_k)\bigg| \leq 2||f||_\infty\epsilon. \] Because $\epsilon>0$ is arbitrary, we get $\lim\limits_{n\rightarrow \infty}\frac{1}{n}\sum\limits_{k=1}^{n}\left(f(a_k)-f(a_k+\sigma_kX_k)\right)=0$. \end{proof} The main idea in proving the corollaries is that we condition the random sequences suitably, so that the resulting sequences satisfy the hypothesis of the Theorem \ref{thm1} and then apply to obtain the result. More formally, say we condition the sequence on the event $E$. Assume the conditioned sequence can be realized as a random sequence which satisfies the hypothesis of Theorem \ref{thm1}. Let $\nu_n^E$ be the empirical measure of the critical points of the degree-$n$ polynomial formed by conditioned sequence. Fix $\epsilon>0$, then \begin{align} \Pr\left(d(\nu_n,\mu)>\epsilon\right) &=\ee{\text{$\text{$\mathbbm{1}$}\{d(\nu_n,\mu)>\epsilon\}$}},\\ &=\ee{\cee{\text{$\text{$\mathbbm{1}$}\{d(\nu_n,\mu)>\epsilon\}$}}{\text{$E$}}},\\ &=\ee{\text{$\text{$\mathbbm{1}$}\{d(\nu_n^E,\mu)>\epsilon\}$}}. \label{eqn:convergence} \end{align} But from the assumption made above, for every $\epsilon>0$ we have, \[\ee{\text{$\text{$\mathbbm{1}$}\{d(\nu_n^E,\mu)>\epsilon\}$}}=\text{$\Pr\left(d(\nu_n^E,\nu)>\epsilon\right)\xrightarrow{n \rightarrow \infty}0$}.\] Applying the dominated convergence theorem to \eqref{eqn:convergence} it follows that, for everey $\epsilon>0$ \[\Pr\left(d(\nu_n,\mu)>\epsilon\right)\xrightarrow{n \rightarrow \infty} 0.\] Therefore it is justified that to show the convergence of probability measures, it is enough to show the convergence of conditioned probability measures almost surely. To invoke the hypothesis of the Theorem \ref{thm1} we need to show that the perturbed sequence is also log-Ces\'{a}ro bounded. It will be proved in the following lemma. We will use the following inequalities, whenever required. \begin{align} \log_+|ab| & \leq \log_+|a| + \log_+|b| \label{logplusprod}\\ \log_-|ab| & \leq \log_-|a| +\log_-|b| \label{logminusprod}\\ \log_+|a_1+a_2+ \dots + a_n| & \leq \log_+|a_1| + \log_+|a_2|+ \dots +\log_+|a_n| +\log(n)\label{logplussum} \end{align} \begin{remark} The inequality \eqref{logplussum} is obtained by using the inequalities $|a_1+\dots+a_n|\leq |a_1|+\dots+|a_n| \leq n\max\limits_{i\leq n}{|a_i|}$ and $\log_+(\max\limits_{i\leq n}|a_i|)\leq \log_+|a_1|+\dots+\log_+|a_n|.$ \end{remark} \begin{lemma}\label{lemma:log-cesaro-bounded} Let $\{a_n\}_{n\geq1}$ be a sequence that is log-Ces\'{a}ro bounded and $\{b_n\}_{n\geq1}$ be a sequence such that $b_n=a_n+\sigma_nX_n$, $\sigma_n\downarrow0$ and $X_1,X_2,\dots$ are i.i.d. random variables with $\ee{\text{$\log_+|X_1|$}}<\infty$. Then the sequence $\{b_n\}_{n\geq1}$ is also log-Ces\'{a}ro bounded. \end{lemma} \begin{proof} \begin{align} \frac{1}{n}\sum\limits_{k=1}^{n}\log_+|b_k| & \leq \frac{1}{n}\sum\limits_{k=1}^{n}(\log_+(|a_k|+|a_k-b_k|)),\\ &\leq \frac{1}{n}\sum\limits_{k=1}^{n}\log_+|a_k|+\frac{1}{n}\sum\limits_{k=1}^{n}\log_+|\sigma_kX_k|+\log(2)\\ &\leq \frac{1}{n}\sum\limits_{k=1}^{n}\log_+|a_k|+\frac{1}{n}\sum\limits_{k=1}^{n}\log_+|\sigma_k|+\frac{1}{n}\sum\limits_{k=1}^{n}\log_+|X_k|+\log(2) \label{eqn:lemma:log-cesaro:1} \end{align} The sequence $\{\frac{1}{n}\sum_{k=1}^{n}\log_+|\sigma_k|\}_{n\geq1}$ goes to $0$, because $\lim\limits_{n\rightarrow0}\sigma_n=0$. Using law of large numbers and the fact that $\ee{\text{$\log_+|X_1|$}}<\infty$, the sequence $\{\frac{1}{n}\sum_{k=1}^{n}\log_+|X_k|\}_{n\geq1}$ is bounded almost surely. Combining \eqref{eqn:lemma:log-cesaro:1} and the above facts we get that the sequence $\frac{1}{n}\sum\limits_{k=1}^{n}\log_+|b_k|$ is bounded. This completes the proof. \end{proof} \begin{proof}[Proof of Corollary \ref{Symmetric perturbations}] Fix $r_n$ and $\theta_n$ for $n \geq 1$. Choose $E=\{w:X_n(w)=\pm r_ne^{i\theta_n} \linebreak\text{ for }n\geq1\}$. Because $X_n$s are symmetric random variables, the $n^{th}$ term of the resulting sequence will be $u_n + \sigma_nr_ne^{i\theta_n}$ or $u_n - \sigma_nr_ne^{i\theta_n}$ with equal probability independent of other terms. Choose $a_n=u_n+\sigma_nr_ne^{i\theta_n}$ and $b_n=u_n-\sigma_nr_ne^{i\theta_n}$. We need to show that almost surely the sequences $\{a_n\}_{n\geq1}$ and $\{b_n\}_{n\geq1}$ satisfy the hypotheses of the Theorem \ref{thm1}. It follows from Lemmas \ref{lemma:perturb_limit_measure} and \ref{lemma:log-cesaro-bounded} the sequences $\{a_n\}_{n \geq 1}$ and $\{b_n\}_{n \geq 1}$ are $\mu$-distributed and log-Ces\'{a}ro bounded almost surely. \end{proof} \begin{proof}[Proof of Corollary \ref{corollary4:kabluchko}] If $\mu$ is a degenerate probability measure then the result is trivial to verify. If $\mu$ is not deterministic then choose two independent sequences of random numbers $\{a_n\}_{n\geq1}$ and $\{b_n\}_{n\geq1}$, where $a_n$s and $b_n$s are i.i.d random numbers obtained from measure $\mu$. Choose $X_n= a_n$ or $b_n$ with equal probability independent of other terms, then $\{X_n\}_{n\geq1}$ is a sequence of i.i.d random variables distributed according to probability measure $\mu$. Using the hypothesis $\int\limits_{\mathbb{C}}\log_+|z|d\mu(z)<\infty$ and applying law of large numbers for the random variables $\{\log_+|X_n|\}_{n \geq 1}$, we get that the sequences $\{a_n\}_{n\geq1}$ and $\{b_n\}_{n\geq1}$ are log-Ces\'{a}ro bounded almost surely. Therefore the constructed sequences satisfy the hypothesis of the Theorem \ref{thm1}. \end{proof} Before proving Proposition \ref{corollary5:thm1} we will prove the following lemma which give the limiting empirical measure of random sequence whose terms are drawn from either of two deterministic sequences. \begin{lemma}\label{random_seq_limit} Let $\{a_k\}_{k\geq1}$ and $\{b_k\}_{k\geq1}$ be two sequences which are $\mu$ and $\nu$ distributed respectively. Define a random sequence $\{\xi_k\}_{k\geq1}$, where $\xi_k=a_k$ with probability $p$ and $\xi_k=b_k$ with probability $1-p$. Then $\mu_n=\frac{1}{n}\sum\limits_{k=1}^{n}\delta_{\xi_k}$ weakly converge to $\lambda=p\mu+(1-p)\nu$ almost surely. \end{lemma} \begin{proof} It is enough to show that for any open set $U\subset\mathbb{C}$, $ \frac{1}{n}\sum\limits_{k=1}^{n}\text{$\mathbbm{1}$}\left\{\xi_k \in U\right\}$ converge to $\lambda(U) $ almost surely. But from a version of law of large numbers we know that if $X_1,X_2,\dots$ are independent random variables (not necessarily identical), then \[ \frac{1}{n}\sum\limits_{k=1}^{n}\left(X_k-\ee{\text{$X_k$}}\right)\xrightarrow{a.s}0 \] provided that $\sum\limits_{k=1}^{\infty}\frac{1}{k^2}\var{X\textbf{$_k$}}<\infty$. Applying this to the random variables $\textbf{$\mathbbm{1}$}\left\{\xi_k \in U\right\}$ we get that $\frac{1}{n}\sum\limits_{k=1}^{n}\text{$\mathbbm{1}$}\left\{\xi_k \in U\right\}$ converge to $\lambda(U)$ almost surely. \end{proof} \begin{proof}[Proof of Proposition \ref{corollary5:thm1}] For $i\geq1$ choose a sequence of independent random variables to be $a_k$ which assumes values $u_k $ with probability $p$ and $v_k$ with probability $1-p$. Independent of this sequence choose another sequence of independent random variables whose terms are $b_k = u_k $ with probability $p$ and $v_k$ with probability $1-p$. The terms of the sequences $\{a_n\}_{n\geq1}$ and $\{b_n\}_{n\geq1}$ satisfy, \begin{align*} \log_+|a_n|\leq \log_+|u_n|+\log_+|v_n|,\\ \log_+|b_n|\leq \log_+|u_n|+\log_+|v_n|. \end{align*} Because the sequences $\{u_n\}_{n\geq1}$ and $\{v_n\}_{n\geq1}$ are log-Ces\'{a}ro bounded, it follows from above inequalities, the sequences $\{a_n\}_{n\geq1}$ and $\{b_n\}_{n\geq1}$ are also log-Ces\'{a}ro bounded. Therefore, from the above arguments and Lemma \ref{random_seq_limit} the sequences $\{a_n\}_{n\geq1}$ and $\{b_n\}_{n\geq1}$ satisfy the hypothesis of the Theorem \ref{thm1} almost surely. Hence the corollary is proved. \end{proof} \begin{proof}[Proof of Corollary \ref{corollary:thm2}] Choose $a_1,a_2,\dots$ be i.i.d $\mbox{Bernoulli}(p)$ random variables. Let $\{k_n\}_{n\geq1}$ be a random sequence such that $a_{k_n}=1$ and $a_\ell=0$ whenever $\ell \notin \{k_1,k_2,\dots\}$. Define $L_n^{(1)}(z)=L_{k_n}(z)=\frac{P_n'(z)}{P_n(z)}$. It is enough to show that $\frac{1}{n}\Delta\log|L_{k_n}(z)|\rightarrow 0$ in probability. The sequences $\{a_n\}_{n\geq1}$ and $\{z_n\}_{n\geq1}$ satisfy the hypothesis of the Theorem \ref{thm2}. Therefore $\frac{1}{n}\Delta \log|L_n(z)|\rightarrow 0$ in probability. Because $\{L_{k_n}^{(1)}(z)\}_{n\geq1}$ is a subsequence of $\{L_n(z)\}_{n\geq1}$ it follows that $\frac{1}{k_n}\Delta \log|L_{k_n}^{(1)}(z)|\rightarrow 0$ in probability. Because $k_n$ is a negative binomial random variable with parameters $(n,p)$, we have $\frac{k_n}{n}\rightarrow p$ almost surely. Therefore, $\frac{1}{n}\Delta \log|L_{k_n}^{(1)}(z)|\rightarrow 0$ in probability. \end{proof} In the next chapter we provide proofs for both the Theorems \ref{thm1} and \ref{thm2}. \chapter{Determinantal point processes from product of random matrices.} \label{ch:ginibreproduct2} \section{Introduction} A Ginibre matrix is a random matrix whose entries are i.i.d. complex Gaussian random variables. In this chapter we derive exact eigenvalue density for certain products of random matrices. In this section we give a over view of the results where the exact eigenvalue density was obtained. Then we state our result and show that the earlier results were special cases of our result. In the next section we give a brief discussion about generalized Schur decomposition, which will used as a transformation to derive the eigenvalue density. In the later section we compute the Jacobian for this transformation. We complete the proof in the last section. We will now recall a well known fact (Theorem 4.5.5 in \cite{manjubook}, Lemma 4 in \cite{soshnikovsurvey}) about the determinantal point process on the complex plane. Let the vector $(z_1,z_2,\dots,z_n)$ be a random vector in $\text{$\mathbb{C}$}^n$ having density proportional to $\prod\limits_{i<j}|z_i-z_j|^2$ w.r.t a measure $\mu^{\otimes n}$. Notice that by doing column operations on the matrix $$\left[ \begin{smallmatrix} \phi_0(z_1) & \phi_1(z_1) & \dots & \phi_{n-1}(z_1) \\ \phi_0(z_2) & \phi_1(z_2) & \dots & \phi_{n-1}(z_2) \\ \vdots & \vdots & \ddots & \vdots \\ \phi_0(z_n) & \phi_1(z_n) & \dots & \phi_{n-1}(z_n) \end{smallmatrix}\right], $$ where $\phi_i$s are orthonormal polynomial w.r.t measure $\mu$, we get \[ \det\left[ \begin{smallmatrix} \phi_0(z_1) & \phi_1(z_1) & \dots & \phi_{n-1}(z_1) \\ \phi_0(z_2) & \phi_1(z_2) & \dots & \phi_{n-1}(z_2) \\ \vdots & \vdots & \ddots & \vdots \\ \phi_0(z_n) & \phi_1(z_n) & \dots & \phi_{n-1}(z_n)\end{smallmatrix} \right] \left[ \begin{smallmatrix} \phi_0(z_1) & \phi_1(z_1) & \dots & \phi_{n-1}(z_1) \\ \phi_0(z_2) & \phi_1(z_2) & \dots & \phi_{n-1}(z_2) \\ \vdots & \vdots & \ddots & \vdots \\ \phi_0(z_n) & \phi_1(z_n) & \dots & \phi_{n-1}(z_n)\end{smallmatrix} \right]^* = c_n\prod\limits_{i<j}|z_i-z_j|^2, \] for some constant $c_n$. On simplification the left hand side of the above equation can be written as $\det[((\textbf{$\mathbb{K}$}(z_i,z_j)))_{1\leq i,j \leq n}]$, where $\textbf{$\mathbb{K}$}_n(z,w)=\sum\limits_{i=0}^{n-1}\phi_i(z)\overline{\phi}_i(w)$. Therefore the entries of the vector $(z_1,z_2,\dots,z_n)$ form a determinantal point process with kernel given by \[\textbf{$\mathbb{K}$}_n(z,w)=\sum\limits_{i=0}^{n-1}\phi_i(z)\overline{\phi}_i(w),\] where $\phi_0, \phi_1, \dots, \phi_n$ are the orthonormal polynomials w.r.t measure $\mu$ on complex plane. Ginibre \cite{ginibre} introduced three ensembles of matrices with i.i.d. real, complex and quaternion Gaussian entries respectively without imposing a Hermitian condition. These matrices are called Ginibre matrices in the literature. Here we restrict our attention to matrices with i.i.d. complex Gaussian entries. In \cite{ginibre}, Ginibre derived the eigenvalue density for $n\times n$ matrix with i.i.d. standard complex Gaussian entries. \begin{theorem}[Ginibre ~\cite{ginibre}] Let $A$ be an $n\times n$ matrix with i.i.d standard complex Gaussian entries. Then the eigenvalues of $A$ form a determinantal point process on the complex plane with kernel \begin{equation*} \text{$\mathbb{K}$}_n(z,w) = \sum_{k=0}^{n-1}\frac{(z\bar w)^k}{k!} \end{equation*} w.r.t to background measure $\frac{1}{\pi}e^{-|z|^2}dm(z)$. Equivalently, the vector of eigenvalues has density \[ \frac{1}{\pi^n\prod_{k=1}^{n}k!}e^{-\sum_{k=1}^{n}|z_k|^2}\prod_{i<j}|z_i - z_j|^2 \] w.r.t Lebesgue measure on $\text{$\mathbb{C}$}^n$. \end{theorem} These are the first non-hermitian matrix ensembles for which the exact eigenvalue density is computed. Later Krishnapur \cite{manjunath} showed that the eigenvalues of $A^{-1}B$ form a determinantal point process on the complex plane when $A$ and $B$ are independent random matrices with i.i.d. standard complex Gaussian entries. In random matrix literature this matrix ensemble $A^{-1}B$ is known as \textit{spherical ensemble}. \begin{theorem}[M.Krishnapur ~\cite{manjunath}] Let $A$ and $B$ be i.i.d. $n\times n$ matrix with i.i.d. standard complex Gaussian entries. Then the eigenvalues of $A^{-1}B$ form a determinantal point process on the complex plane with kernel \begin{equation*} \text{$\mathbb{K}$}_n(z,w) = (1+z\overline{w})^{n-1} \end{equation*} w.r.t to background measure $\frac{n}{\pi}\frac{dm(z)}{(1+|z|^2)^{n+1}}$. Equivalently, the vector of eigenvalues has density $$ \frac{1}{n}\left(\dfrac{n}{\pi}\right)^n\prod_{k=1}^{n}{{n-1}\choose k}\prod_{k=1}^{n}\dfrac{1}{(1+|z_k|^2)^{n+1}}\prod_{i<j}|z_i - z_j|^2 $$ w.r.t Lebesgue measure on $\mathbb{C}\text{$^n$}$ \end{theorem} Akemann and Burda \cite{akemann} have derived the eigenvalue density for the product of $k$ independent $n\times n$ matrices with i.i.d. complex Gaussian entries. In this case the joint probability distribution of the eigenvalues of the product matrix is found to be given by a determinantal point process as in the case of Ginibre, but with a weight given by a Meijer $G$-function depending on $k$. Their derivation hinges on the generalized Schur decomposition for matrices and the method of orthogonal polynomials. We shall state that result as the following theorem. \begin{theorem}[Akemann-Burda ~\cite{akemann}] Let $A_1,A_2,\dots,A_n$ be i.i.d. $n\times n$ matrices with i.i.d. standard complex Gaussian entries. Then the eigenvalues of $A_1A_2\dots A_n$ has density (with respect to Lebesgue measure on $\text{$\mathbb{C}$}^{n}$) proportional to \[ \prod_{\ell=1}^{n}\omega(z_{\ell})\prod_{i<j}^{n}|z_i-z_j|^2 \] with a weight function $ \omega(z)$, where \begin{equation}\label{weight0} \omega(z)=\int_{x_1\cdots x_{k} =z}e^{-\sum_{j=1}^{k}|x_{j}|^2} \prod_{j=1}^{k}|x_{j}|^{(n-1)}d\sigma. \end{equation} Here $\sigma$ is the Lebesgue measure restricted to the hyper surface $\{(x_1,x_2,\dots,x_k):x_1\cdots x_{k} =z\}$. \end{theorem} In all the above results after calculating the density of the eigenvalues, it turns out that they form a determinantal point process. Now following the work of Krishnapur \cite{manjunath} on spherical ensembles and the work of Akemann and Burda \cite{akemann} on the product of $k$ independent $n\times n$ Ginibre matrices, it is a natural question to ask, what can be said about the eigenvalues of product of $k$ independent Ginibre matrices when a few of them are inverted? We investigated the case $A=A_1^{\epsilon_1}A_2^{\epsilon_2} \cdots A_k^{\epsilon_k}$, where each $\epsilon_i$ is $+1$ or $-1$ and $A_1, A_2,\ldots, A_k$ are independent matrices with i.i.d. standard complex Gaussian entries, and obtained the density of eigenvalues of $A$. We state the result as the following theorem. \begin{theorem}[K Adhikari, K Saha, NK Reddy, TR Reddy ~\cite{atr}]\label{chap5:thm1} Let $A_1,A_2,\ldots,A_k$ be independent $n\times n$ random matrices with i.i.d. standard complex Gaussian entries. Then the eigenvalues of $A=A_1^{\epsilon_1}A_2^{\epsilon_2} \ldots A_k^{\epsilon_k}$, where each $\epsilon_i$ is $+1$ or $-1$, has density (with respect to Lebesgue measure on $\mathbb{C}\textbf{$^n$}$) proportional to \[ \prod_{\ell=1}^{n}\omega(z_{\ell})\prod_{i<j}^{n}|z_i-z_j|^2 \] with a weight function $ \omega(z)$, where \begin{equation}\label{weight1} \omega(z)=\int_{x_1^{\epsilon_1}\cdots x_{k}^{\epsilon_k} =z}e^{-\sum_{j=1}^{k}|x_{j}|^2} \prod_{j=1}^{k}|x_{j}|^{(1-\epsilon_j)(n-1)}d\sigma. \end{equation} \noindent Here $\sigma$ is the Lebesgue measure restricted to the hyper surface given by $\{(x_1,x_2,\dots,x_k):x_1^{\epsilon_1}\cdots x_{k}^{\epsilon_k} =z\}$. \end{theorem} \begin{remark} From the symmetry of the expressions in the Theorem \ref{chap5:thm1} notice that the density of the eigenvalues of the matrix $A=A_1^{\epsilon_1}A_2^{\epsilon_2} \ldots A_k^{\epsilon_k}$ depends only on $\sum_{i=1}^{k}\epsilon_i$ but not on individual $\epsilon_i$s. \end{remark} \begin{remark} If $k=2$, $\epsilon_1=-1$ and $\epsilon_2=1$, then from \eqref{weight1} we get that \[\omega(z)=\int_{\frac{x_{2}} {x_{1}}=z}e^{- (|x_{1}|^{2}+|x_{2}|^2)} |x_{1 }|^{2(n-1)}d\sigma=C_n \frac{1}{(1+|z|^2)^{(n+1)}}, \] where $\sigma$ is the Lebesgue measure restricted to the hyper surface $\{(x_1,x_2):\frac{x_{2}} {x_{1}}=z\}$ $C_n$ is a constant . Hence the density of the eigenvalues of $A_1^{-1}A_2$ is proportional to \[ \prod_{i=1}^{n}\frac{1}{(1+|z_i|^2)^{n+1}}\prod_{i<j}|z_{i}-z_j|^2. \] From the above expression it is clear that the eigenvalues of $A_1^{-1}A_2$ form a determinantal point process in a complex plane. This result was proved by Krishnapur in \cite{manjunath} through a different approach. \end{remark} \begin{remark} If $\epsilon_i=1$ for $i=1,2,\ldots,k$, then by Theorem \ref{chap5:thm1} it follows that the eigenvalues of $A_1A_2\ldots A_k$ form a determinantal point process. This result is due to Akemann and Burda \cite{akemann}. \end{remark} For proving the Theorem \ref{chap5:thm1}, we will do an appropriate transformation and then integrate the auxiliary variables to obtain the eigenvalue density. We will use generalized Schur decomposition for the matrices $X_1,X_2,\dots,X_k$, where $X_i=A_i^{\epsilon_i}$ for $i=1,2,\dots,k$. In the forthcoming section we will present generalized Schur decomposition appeared in \cite{akemann}. In the subsequent section we compute the Jacobian for this transformation. \section{Generalized Schur decomposition} We will first recall the Schur decomposition and then state generalized Schur decomposition. \paragraph{Schur decomposition:}Let $A$ be a $n\times n$ matrix. Then there exists an unitary matrix $U$, a diagonal matrix $D$ and a strictly upper triangular matrix $T$, such that $A=U(D+T)U^*$. The diagonal elements of $D$ are the eigenvalues of $A$. Further, if the eigenvalues of $A$ are all distinct and we fix the order of their appearance in $D$, then the decomposition is unique up to a conjugation by a diagonal matrix, whose diagonal entries are in $S^1$. The key ingredient in deriving the result \ref{chap5:thm1} is the generalization of the above mentioned Schur product. \paragraph{Generalized Schur decomposition:}Let $A_1,A_2,\dots,A_k$ be $n \times n$ matrices, then there exists unitary matrices $U_1,U_2,\dots,U_k$, diagonal matrices $D_1,D_2,\dots,D_k$ and strictly upper triangular matrices $T_1,T_2,\dots,T_k$ such that they can be decomposed in the following form. \begin{align} A_1 & = U_1(Z_1+T_1)U_2^*,\\ A_2 & = U_2(Z_2+T_2)U_3^*,\\ &\vdots\\ A_k &=U_k(Z_k+T_k)U_1^*. \end{align} To prove this decomposition we first consider the Schur decomposition for the matrix $A_1A_2\dots A_k$. Let it be $A_1A_2\dots A_k=U_1(Z+T)U_1^*$. Now performing the Gram-Schimdt orthogonalization to the columns of the matrix $U_1^*A_1$, we get $U_1^*A_1=(Z_1+T_1)U_2^*$ for some unitary matrix $U_2$, and $Z_1, T_1$ being diagonal and strictly upper triangular matrices respectively. We repeat this process i.e., in the $i^{th}$ step we perform the Gram-Schimdt orthogonalization to the matrix $U_i^*A_i$ to get $U_i^*A_i=(Z_i+T_i)U_(i+1)$. After performing $n-1$ steps it forces that $U_n^*A_n=(Z_n+T_n)U_1^*.$ This decomposition is not unique in general. But, if we incorporate certain conditions on the matrices then it will be unique. Assume that the diagonal entries of $Z_1Z_2\dots Z_k$ are distinct, and appear in a particular order (in particular may choose lexicographical ordering). Observe that replacing $U_i$ with $\Theta_iU_i$, where $\Theta_i$ is a diagonal unitary matrix, we have that $A_i$s assume a similar decomposition. Hence if all the diagonal entries of $U_i$s are non-zero we may assume them to be positive. On assuming these two conditions the decomposition will be unique. Another criterion for the uniqueness of this decomposition is to assume that the first non-zero entry in each row of $U_i$ is non-negative. In the next section while computing Jacobian we assume that all the diagonal entries are positive as the unitary matrices with zero diagonal entries form a null set. Notice that the eigenvalues of $A_1A_2\dots A_k$ are same as that of $Z_1Z_2\dots Z_k$. We will exploit this and use generalized Schur decomposition as the transformation in recovering the eigenvalue density. We will compute the Jacobian for this transformation in the next section. For a more general discussion on this the reader can refer to the appendix in \cite{atr}. \section{Jacobian computation} To obtain eigenvalue density of $A$, we need to do an appropriate change of variables. We do generalized Schur decomposition as mentioned in the previous section. We will compute the Jacobian for this transformation in this section. The computation of Jacobian is on the lines of the computation, given in \cite{manjubook} (Section 6.3, Chapter-6), while deriving the eigenvalue density for Ginibre matrices. Before doing the Jacobian determinant calculation, we state a basic property about wedge product, which will be used repeatedly. If $dy_j=\sum_{k=1}^na_{j,k}dx_k$, for $1\le j\le n$, then using the alternating property $dx\wedge dy=-dy\wedge dx$ it is easy to see that \begin{equation}\label{eqn:wedge:relation} dy_1\wedge dy_2\wedge \ldots\wedge dy_n=\det[((a_{j,k}))_{j,k\le n}]dx_1\wedge x_2\wedge\ldots\wedge dx_n. \end{equation} As a consequence we can see that, if $\underline{x}=(x_1,\dots,x_n)$ is a unitary transformation of $\underline{y}=(y_1,\dots,y_n)$, then \begin{equation}\label{eqn:wedge:relation2} dy_1\wedge d\overline{y}_1\wedge dy_2\wedge d\overline{y}_2 \ldots\wedge dy_n\wedge d\overline{y}_n=dx_1\wedge d\overline{x}_1\wedge dx_2\wedge d\overline{x}_2 \ldots\wedge dx_n\wedge d\overline{x}_n. \end{equation} From generalized Schur decomposition we have that for any matrices $X_1,X_2,\dots,X_k$ in $g\ell (n,\mathbb{C})$ can be written as \begin{align}\label{eqn:gschur} X_1 & = U_1(Z_1+T_1)U_2^*,\\ X_2 & = U_2(Z_2+T_2)U_3^*,\\ &\vdots\\ X_k &=U_k(Z_k+T_k)U_{k+1}^*. \end{align} Where $U_1,U_2,\dots,U_k,U_{k+1}$ are unitary matrices satisfying $U_{k+1}=U_1$, $Z_1,Z_2,\dots,Z_k$ are diagonal matrices and $T_1,T_2,\dots,T_k$ are strictly upper triangular matrices. Because $U_iU_i^*=\text{$\mathbb{I}$}_n$ we have $(dU_i)U_i^*=-U_i(dU_i^*)$, for $i=1,2,\dots,k$. Using this fact and the generalized Schur decomposition \ref{eqn:gschur}, for any $\ell=1,2,\dots,k$ we get, \begin{align} dX_\ell &= (dU_\ell)(Z_\ell+T_\ell)U_{\ell+1}^*+U_\ell (dZ_\ell+dT_{\ell})U_{\ell+1}^*+U_\ell (Z_\ell+T_{\ell})dU_{\ell+1}^*\\ &= (dU_\ell)(Z_\ell+T_\ell)U_{\ell+1}^*+U_\ell (dZ_{\ell}+dT_\ell)U_{\ell+1}^*-U_\ell (Z_\ell+T_\ell)U_{\ell+1}^*(dU_{\ell+1})U_{\ell+1}^*\\ &= U_\ell\left[(U_\ell^*dU_\ell)(Z_\ell+T_\ell)-(Z_\ell+T_\ell)(U_{\ell+1}^*dU_{\ell+1})+dZ_\ell+dT_\ell\right]U_{\ell+1}^*. \end{align} For convenience let us denote $\Lambda_\ell:=U_\ell^*(dX_\ell)U_{\ell+1}$, $\Omega_\ell:=U_\ell^*dU_\ell$ and $S_\ell:=Z_\ell+T_\ell$. Note that $\Lambda_\ell=(\lambda_\ell (i,j))$ and $\Omega_\ell=(\omega_\ell (i,j))$ are $n\times n$ matrices of one forms, and $dS_\ell$ ($=dZ_\ell+dT_\ell$) is an upper triangular matrix of one form. Let $Z_\ell=\mbox{diag}(Z_\ell(1),\linebreak Z_\ell(2),\dots,Z_\ell(n))$, $T_\ell=(t_\ell(i,j))$ and $\Lambda_\ell=(\lambda_\ell(i,j))$. Define, \begin{equation}\label{eqn:lambda_l} \Lambda_\ell=\Omega_\ell S_\ell-S_\ell\Omega_{\ell+1}+dS_\ell. \end{equation} For any unitary matrix $U$ the transformation $X \rightarrow UX$ is a unitary transformation. Therefore from \eqref{eqn:wedge:relation2}, we have \[\bigwedge_{i,j}\left(dX_\ell(i,j)\wedge d\overline{X}_\ell(i,j)\right)= \bigwedge_{i,j}\left(d\lambda_\ell(i,j)\wedge d\overline{\lambda}_\ell(i,j)\right),\] for $\ell=1,2,\dots,k.$ Throughout this computation we will ignore the constants, hence each equality is indeed an equality up to a constant. Because we will be dealing with probability densities the constants can be retrieved by equating the integral to $1$. Expanding the equation \eqref{eqn:lambda_l} we get, \begin{eqnarray}\label{eqn:wedge_nullify} \lambda_\ell (i,j) &=&\sum\limits_{m=1}^{j}S_\ell (m,j)\omega_\ell (i,m)-\sum\limits_{m=i}^{n}S_\ell (i,m)\omega_{\ell+1}(m,j)+dS_\ell (i,j) \\&=&\left\{ \begin{array}{lcr} S_\ell (j,j)\omega_\ell (i,j)-S_\ell (i,i)\omega_{\ell+1}(i,j) \\ +\left[\sum\limits_{m=1}^{j-1}S_\ell (m,j)\omega_\ell (i,m)-\sum\limits_{m=i+1}^{n}S_\ell (i,m)\omega_{\ell+1}(m,j)\right] \mbox{ if } i>j;\\ dS_\ell (i,j)+S_\ell (i,j)\left(\omega_\ell (i,i)-\omega_{\ell+1}(j,j)\right)\\ +\left[\sum\limits_{\substack{m=1\\m\neq i}}^{j}S_\ell (m,j)\omega_\ell (i,m)-\sum\limits_{\substack{m=i+1\\m\neq j}}^{n}S_\ell (i,m)\omega_{\ell+1}(m,j)\right] \mbox{ if }i\leq j.\\ \end{array} \right. \end{eqnarray} To execute the wedge product, we will now arrange $\{\lambda_\ell (i,j),\overline{\lambda}_\ell (i,j)\}$ in a particular order. We will use lexicographic order on the indices associated with $\lambda_\ell (i,j)$. The indices associated with $\lambda_\ell (i,j)$ are $(i,j,\ell)$. The lexicographical order will be taken on $(i,n-j,k-\ell)$. The corresponding conjugate terms will be followed by the term for which it is conjugate. For convenience, we present the ordering as the following table (each row is read from left to right and top row precedes bottom rows). \[ \begin{smallmatrix} \lambda_1(n,1), \overline{\lambda}_1(n,1), \dots, \lambda_k(n,1), \overline{\lambda}_k(n,1),&\dots&, \lambda_1(n,n), \overline{\lambda}_1(n,n), \dots, \lambda_k(n,n), \overline{\lambda}_k(n,n)\\ \lambda_1(n-1,1), \overline{\lambda}_1(n-1,1), \dots, \lambda_k(n-1,1), \overline{\lambda}_k(n-1,1),&\dots&, \lambda_1(n-1,n), \overline{\lambda}_1(n-1,n), \dots, \lambda_k(n-1,n),\overline{\lambda}_k(n-1,n) \\ \vdots &\ddots&\vdots\\ \lambda_1(1,1), \overline{\lambda}_1(1,1), \dots, \lambda_k(1,1), \overline{\lambda}_k(1,1),&\dots&, \lambda_1(1,n), \overline{\lambda}_1(1,n), \dots, \lambda_k(1,n),\overline{\lambda}_k(1,n) \end{smallmatrix} \] Using the fact that $\Omega_\ell$ is skew-hermitian (i.e., $\omega_\ell(i,j)=-\overline{\omega}_\ell(j,i)$), while executing the wedge product, notice that the terms in the square brackets are one forms and have already appeared before in the given ordering. Hence their contribution to the entire product is nullified. In the next couple of paragraphs we will explain in detail about this cancellation, the reader who already got convinced may skip them. If $i>j$, then each of the terms in the square brackets contain either $\omega_\ell(i,m_1)$ or $\omega_{\ell+1}(m_2,j)$, where $m_1<j$ and $m_2>i$. These terms have already appeared in $\lambda_\ell(i,m_1)$ and $\lambda_{\ell+1}(m_2,j)$ respectively, outside the square brackets, which are leading the order we have executed the product. If $i\leq j$, then each of the terms in the square brackets contain either $\omega_\ell(i,m_1)$ or $\omega_{\ell+1}(m_2,j)$, where $m_1\leq j$, $m_1 \neq i$, $m_2>i$ and $m_2\neq j$. For the case $j\geq m_1>i$, by skew hermitian property of $\Omega_\ell$, we have $\omega_\ell(i,m_1)=-\overline{\omega}_\ell(m_1,i)$ which has already appeared in $\overline{\lambda}_\ell(m_1,i)$, outside the square brackets. For the case $m_1<i$, $\omega_\ell(i,m_1)$ has already appeared in $\lambda_\ell(i,m_1)$ outside square brackets. Similarly for the case $m_2<j$,we have $\omega_{\ell+1}(m_2,j)=-\overline{\omega}_{\ell+1}(j,m_2)$ which has appeared outside square brackets in $\overline{\lambda}_{\ell+1}(j,m_2)$. Lastly in the case of $m_2>j$, $\omega_{\ell+1}(m_2,j)$ has already appeared in $\lambda_{\ell+1}(m_2,j).$ Therefore, if we assume \begin{eqnarray} \mu_{\ell}(i,j) &=&\left\{ \begin{array}{lcr} S_\ell (j,j)\omega_\ell (i,j)-S_\ell (i,i)\omega_{\ell+1} (i,j) & \mbox{ if } i>j;\\ dS_\ell (i,j)+S_\ell (i,j)\left(\omega_\ell (i,i)-\omega_{\ell+1} (j,j)\right) & \mbox{ if }i\leq j; \end{array} \right. \end{eqnarray} then, \[\bigwedge_{\ell}^{}\bigwedge_{i,j}^{}\left(\lambda_\ell (i,j)\wedge \overline{\lambda}_\ell (i,j)\right)=\bigwedge_{\ell}^{}\bigwedge_{i,j}^{}\left(\mu_\ell (i,j)\wedge \overline{\mu}_\ell (i,j)\right).\] Recall that $S_\ell=Z_\ell+T_\ell$. For $i>j$, the term $\bigwedge_\ell(\mu_\ell(i,j)\wedge\overline{\mu}_\ell(i,j))$ yields $\big|\prod\limits_{\ell=1}^{k}Z_\ell(i)-\prod\limits_{\ell=1}^{k}Z_\ell(j)\big|^2$. Hence we get, \begin{align}\label{eqn:manifold_nullification_1} \bigwedge_{\ell}\bigwedge_{i,j}|\lambda_\ell(i,j)|^2 =\left(\prod_{i>j}\bigg|\prod\limits_{\ell=1}^{k}Z_\ell(i)-\prod\limits_{\ell=1}^{k}Z_\ell(j)\bigg|^2\right)&\bigwedge_{\ell}\left(\bigwedge_{i>j}|\omega_\ell(i,j)|^2\bigwedge_{i}|dZ_\ell(i)|^2\right)\\\times\bigwedge_{\ell} \bigwedge_{i,j}\big|dT_\ell(i,j)&+T_\ell(i,j)\left(\omega_\ell(i,i)-\omega_{\ell+1}(j,j)\right)\big|^2. \end{align} Note that we have simplified the notation by denoting $|dz|^2:=dz\wedge d\overline{z}$. Consider the set of unitary matrices $\text{$\mathcal{M}$}_\ell=\{U_\ell:U_\ell^*U_\ell=\text{$\mathbb{I}$}_n,U_\ell(i,i)>0\}$. $\text{$\mathcal{M}$}_\ell$ is a sub manifold of dimension $n^2-n$ in the manifold $\mathcal{U}\textbf{$(n)$}$. The dimension of $\text{$\mathcal{M}$}_\ell$ can be obtainded by imposing the constrains $\{U_\ell(i,i)>0;i=1,2,\dots,n\}$, on $\mathcal{U}\textbf{$(n)$}$. Now the product $\omega_\ell(m,m)\wedge\left(\bigwedge_{i>j}|\omega_\ell(i,j)|^2\right)=0$, because $\omega_\ell(i,j)$s are one forms on the manifold $\text{$\mathcal{M}$}_\ell$ whose dimension is $n^2-n$ and the product contains $n^2-n+1$ terms. Hence \eqref{eqn:manifold_nullification_1}, will be reduced to, \begin{align} \bigwedge_{\ell}\bigwedge_{i,j}&|\lambda_\ell(i,j)|^2\\&=\left(\prod_{i>j}\bigg|\prod\limits_{\ell=1}^{k}Z_\ell(i)-\prod\limits_{\ell=1}^{k}Z_\ell(j)\bigg|^2\right)\bigwedge_{\ell}\bigwedge_{i>j}|\omega_\ell(i,j)|^2\bigwedge_{i}\bigwedge_{\ell}|dZ_\ell(i)|^2\bigwedge_{\ell}\bigwedge_{i<j}|dT_\ell(i,j)|^2\label{eqn:manifold_nullification_2}. \end{align} Notice that $\bigwedge_{i>j}|\omega_\ell(i,j)|^2$ is $n^2-n$ form on space of unitary matrices whose diagonal entries are non-negative (or in other words $\text{$\mathcal{U}$}(n)/\text{$\mathcal{U}$}(1)$) and it is invariant under any unitary transformation. Hence the measure induced by this form is Haar measure on $\text{$\mathcal{U}$}(n)/\text{$\mathcal{U}$}(1)$. We will denote it by $|dH(U_\ell)|$. Therefore we have \begin{align} \bigwedge_{\ell}\bigwedge_{i,j}|\lambda_\ell(i,j)|^2=&\left(\prod_{i>j}\bigg|\prod\limits_{\ell=1}^{k}Z_\ell(i)-\prod\limits_{\ell=1}^{k}Z_\ell(j)\bigg|^2\right)\\ &\times\bigwedge_{\ell}|dH(U_\ell)|\bigwedge_{i}\bigwedge_{\ell}|dZ_\ell(i)|^2\bigwedge_{\ell}\bigwedge_{i<j}|dT_\ell(i,j)|^2\label{eqn:manifold_nullification_3}. \end{align} Now that we have the basic ingredients ready, we will proceed for the proof of the theorem in the next section. \section{Proof of Theorem \ref{chap5:thm1}} The density of $(A_1,A_2,\ldots,A_k)$ is proportional to \[ \prod_{\ell=1}^{k}e^{-\Tr( A_{\ell}A_{\ell}^*)}\bigwedge_{\ell=1}^{k}\bigwedge_{i,j=1}^n|dA_{\ell}(i,j)|^2 \] where $|dA_{\ell}(i,j)|^2=dA_{\ell}(i,j)\wedge d\bar{A}_{\ell}({i,j}).$ Through out the proof, we will ignore the proportionality constants where ever present. Since we are dealing with probability densities, the proportionality constants can be recovered by equating the integral of the density to $1$. Let $X_{\ell}=A_{\ell}^{\epsilon_\ell}$ for $\ell=1,2,\ldots,k$. The Jacobian for the transformation $A_\ell \rightarrow A_\ell^{\epsilon_\ell}$ is $|\det(A_\ell)|^{2(\epsilon_\ell-1)n}$. Hence the joint density of $(X_1,X_2,\ldots,X_k)$ is proportional to \begin{align} \prod_{\ell=1}^{k}e^{-\Tr( X_{\ell}^{\epsilon_{\ell}}X_{\ell}^{\epsilon_{\ell}*})}\prod_{\ell=1}^k |\det(X_{\ell})|^{2(\epsilon_{\ell}-1)n} \bigwedge_{\ell=1}^{k}\bigwedge_{i,j=1}^n|dX_{\ell}(i,j)|^2 \label{eqn:dencityofx} \end{align} Using generalized Schur decomposition we have \begin{align} X_{\ell}=U_{\ell}S_{\ell}U_{\ell+1}^*,\;\;\mbox{for $\ell=1,2,\ldots,k,$} \label{eqn:schurdecomposition} \end{align} where $U_{k+1}=U_1$ and $U_{\ell}$ are unitary matrices, $S_{\ell}$ are upper triangular matrices. We write $S_{\ell}$ for $1\leq l \leq k$ as \begin{equation} S_{\ell}=Z_{\ell}+T_{\ell}, \label{eqn:diag+uppertriangular} \end {equation} where $Z_{\ell}=\mbox{diag}(Z_{\ell }(1),Z_{\ell }(2),\ldots,Z_{\ell}(n))$ and $T_{\ell}$ are strictly upper triangular matrices. Now from \eqref{eqn:manifold_nullification_3}, we have \begin{eqnarray} \bigwedge_{\ell=1}^{k}\bigwedge_{i,j=1}^n|dX_{\ell}(i,j)|^2&=&\prod_{i<j}|z_i-z_j|^2\bigwedge_{\ell=1}^k\big(|dH(U_\ell)|\bigwedge_{i\le j}|dS_{\ell}(i,j)|^2\big)\\ \nonumber &=&\prod_{i<j}|z_i-z_j|^2\bigwedge_{\ell=1}^k\bigg(|dH(U_\ell)|\bigwedge_{i<j}|dT_{\ell}(i,j)|^2\bigwedge_{i=1}^n |dZ_\ell(i)|^2\bigg),\label{eqn:{wedge}} \end{eqnarray} where $|dH(U_{\ell})|$ are independent Haar measures on $\mathcal{U}\text{$(n)$}/\mathcal{U}\text{$(1)$}$ and $z_i=\prod _{\ell=1}^k Z_{\ell}(i)$ for $1\leq i \leq n$. Notice that $z_1,z_2,\ldots,z_n$ are the eigenvalues of $X_1X_2\cdots X_k$. Now using \eqref{eqn:schurdecomposition} and \eqref{eqn:{wedge}}, \eqref{eqn:dencityofx} can be written as \begin{align} &\prod_{\ell=1}^{k}\left[e^{-\Tr( S_{\ell}^{\epsilon_{\ell}}S_{\ell}^{\epsilon_{\ell}*})} |\det(S_{\ell})|^{2(\epsilon_{\ell}-1)n}\right]\prod_{i<j}|z_i-z_j|^2\\ &\times \bigwedge_{\ell=1}^k\bigg(|dH(U_\ell)| \bigwedge_{i<j}|dT_{\ell}(i,j)|^2\bigwedge_{i=1}^n |dZ_{\ell}(i)|^2\bigg).\label{eqn:sipmle} \end{align} Now our aim is to integrate out all auxiliary variables from \eqref{eqn:sipmle} to get the density of eigenvalues of $X_1X_2\cdots X_k$. Observe that if $S=Z+T$ where $Z=diag(x_1,x_2,\ldots,x_n)$ and $T$ is a strictly upper triangular matrix, then \[ S^{-1}S^{-1*}=(I+Z^{-1}T)^{-1}Z^{-1}Z^{-1*}(I+Z^{-1}T)^{-1*}.\] Observe that $Z^{-1}T$ is also a strictly upper triangular matrix. If we let $P=Z^{-1}T$, we get $P$ is an upper triangular matrix and \[ |DP|=\prod_{i=1}^{n-1}\frac{1}{|x_{i}|^{2(n-i)}}|DT|, \] where $|DP|=\bigwedge_{i,j}|dP(i,j)|^2$ and $|DT|=\bigwedge_{i,j}|dT(i,j)|^2$. Now replacing $(I+P)^{-1}$ by $Q$, we have $|DQ|=|DP|$ and therefore \[ |DT|=\prod_{i=1}^{n-1}|x_{i}|^{2(n-i)}|DQ|. \] Hence we have \begin{align} e^{-\Tr(S^{-1}S^{-1*})}|DT|&= e^{-\Tr(QZ^{-1}Z^{-1*}Q^*)}\prod_{i=1}^{n-1}|x_{i}|^{2(n-i)}|DQ|\\ &= e^{-\sum_{i=1}^{n}\frac{1}{|x_{i}|^2}(\sum_{j=1}^{i-1}|Q(j,i)|^2+1)}\prod_{i=1}^{n-1}|x_{i}|^{2(n-i)}|DQ|.\label{eqn:simplification} \end{align} Now using \eqref{eqn:simplification} for each $S_{\ell}$, $1\leq \ell \leq k$, we get that the expression \eqref{eqn:sipmle} is proportional to \begin{eqnarray} \prod_{i<j}|z_i-z_j|^2\prod_{\ell=1}^{k}\bigg[e^{-\sum_{i=1}^n(|Z_{\ell}(i)|^{2\epsilon_{\ell}} +\frac{1-\epsilon_{\ell}}{2}|Z_{\ell }(i)|^{2\epsilon_\ell}\sum_{j=1}^{i-1}|Q_{\ell}(j,i)|^2 +\frac{1+\epsilon_{\ell}}{2}\sum_{j=i+1}^n|T_\ell(i,j)|^2)} \\\times\prod_{i=1}^{n-1}|Z_{\ell}(i)|^{(n-i)(1-\epsilon_\ell)} \prod_{i=1}^{n}|Z_{\ell }(i)|^{2(\epsilon_{\ell}-1)n}\bigg]\bigwedge_{\ell=1}^k|dH({U_{\ell}})| |DQ_{\ell}|^{\frac{1-\epsilon_{\ell}}{2}}|DT_{\ell}|^{\frac{1+\epsilon_{\ell}}{2}}|DZ_{\ell}|.\nonumber \end{eqnarray} Now integrating this expression with respect to all variables except the variables of $Z_{\ell}$ for $\ell=1,2,\ldots,k$, we have (omitting the constant) \begin{eqnarray}\label{eqn:result1} \prod_{i<j}|z_i-z_j|^2\prod_{\ell=1}^{k}\left(e^{-\sum_{i=1}^n(|Z_{\ell }(i)|^{2\epsilon_{\ell}})} \prod_{i=1}^{n}|Z_{ \ell }(i)|^{(\epsilon_{\ell}-1)(n+1)}\right)\bigwedge_{\ell=1}^k |DZ_{\ell}|. \end{eqnarray} Now integrating the above expression on the respective hypersurfaces given by \linebreak $\left\{(Z_{1}(i),Z_{2}(i),\dots, Z_{k}(i)):Z_{1}(i)Z_{2}(i)\dots Z_{k}(i)=z_i\right\}$, for $i=1,2,\dots,n$ we get the density of the eigenvalues $(z_1,z_2,\ldots,z_n)$ of $A_1^{\epsilon_1}A_2^{\epsilon_2}\cdots A_k^{\epsilon_k}$. This completes the proof of the theorem. \qed \chapter{Introduction} \label{ch:introduction1} The fundamental theorem of algebra states that every polynomial of degree $n$ of a single variable has, counted with multiplicity, exactly $n$ zeros (roots) in the complex plane. Abel and Galois proved that the roots of any polynomial of degree $5$ or higher cannot be expressed in terms of radicals. Like in many other problems, where exact formulae are not known (or the formulae not amenable to analysis), one may study the statistics of a `typical' polynomial. This is done by equipping a probability measure on the space of polynomials (or by introducing randomness within the polynomial) and choosing a random polynomial according to this measure. There are two natural ways of inducing randomness into polynomials - one by considering the characteristic polynomials of a random matrix and other by choosing the coefficients of polynomials to be random variables. Both cases are of interest. The former leads to the study of random matrices while the latter to the theory of random polynomials. In the theory of random polynomials, the central problem is to understand the behaviour of zeros which is studied by choosing coefficients to be random variables (often independent). The study of random polynomials many times provide us with interesting phenomenon. For example, the zeros of Kac's polynomials, which are defined as random polynomials with i.i.d. complex Gaussian random variables as coefficients, accumulate near the unit circle. Pemantle and Rivin, in \cite{pemantle}, considered a sequence of polynomials whose zeros are i.i.d. complex random variables. They conjectured that the empirical measures of zeros and critical points of these polynomials agree in limit. The first part of this thesis is inspired by Kabluchko's proof \cite{kabluchko} to the problem of Pemantle and Rivin. Here we construct a random sequence of zeros by choosing its terms from two predefined deterministic sequences. It is shown that for the sequence of random polynomials, whose zeros are the terms of the random sequence constructed above, the limiting empirical measures of zeros and critical points agree. This phenomenon fails in general for deterministic sequence of polynomials. Examples of deterministic sequence of polynomials where the limiting measures of zeros and critical points don't agree can be constructed. However, as a consequence of the previous result, it can be shown that if we slightly perturb the zeros of these polynomials, then the limiting measures of zeros and critical points will agree. Hannay in \cite{hannay} and Hanin in \cite{hanin1} observed that the critical points of random polynomials are closely paired with the zeros of the polynomials. In this setting, we study the matching distance between zeros and critical points. When the zeros of the polynomials are all i.i.d. real valued random variables, it is shown that the matching distance between zeros and critical points of these polynomials will remain bounded if the random variables have finite first moment. It is also shown that in limit the bound is exactly the first moment of these random variables. The theory of random matrices deals with the study of eigenvalues of a random matrix. Finding exact eigenvalue density is an important problem in random matrix theory. There are only a handful matrix ensembles for which the exact eigenvalue density is known. For these ensembles, it is often true that the eigenvalues constitute an important class of point processes called determinantal point processes for which a theory and framework is already available for analysis. Of the known ensembles, very few are non-hermitian matrix ensembles. Ginibre, in \cite{ginibre}, derived the eigenvalue density for a matrix whose entries are i.i.d. complex Gaussian random variables. These matrices are called Ginibre matrices since. Krishnapur, in \cite{manjunath} derived the eigenvalue density for $A^{-1}B$, where $A$ and $B$ are independent Ginibre matrices. Akemann and Burda in \cite{akemann} derived the eigenvalue density for random matrices obtained as the product of independent Ginibre matrices. A generalization for these matrices is to consider product of independent matrices where each matrix or its inverse is a Ginibre matrix. In this thesis, the eigenvalue density for these matrices is derived and it is shown that they form a determinantal point process. This result generalizes all the previously mentioned results. Studying the behaviour of real eigenvalues for real random matrices and real zeros for real random polynomials have posed different challenges (due to the lack of conventional symmetries) and simultaneously offered various insights. A different problem on the products of i.i.d. real matrices of fixed size with i.i.d. entries was considered by Lakshminarayan in \cite{arul}. He considered the case when the entries are i.i.d. real Gaussian random variables. He conjectured that the probability of the product of these matrices have all real eigenvalues, converge to $1$ as the size of the product increase to infinity. He established this conjecture for the matrices of size $2 \times 2$. Forrester, in \cite{forrester}, proved this conjecture for any $k\geq1$. It is natural to believe that this phenomenon is universal and hence may hold for any matrix with i.i.d. entries. We show this in a case where the entries of these matrices are distributed according to the probability measure $\mu$ which has an atom. \section{Outline} We now outline the contents of this thesis briefly. This thesis broadly deals with two themes. In the first part we study the zeros and critical points of random polynomials, which is covered in Chapters 2, 3 and 4. \begin{itemize} \item In Chapter 2, a brief history of the results relating critical points and zeros of random polynomials are given. We deal with sequences of deterministic polynomials to provide explicit examples in which the limiting measures of zeros and critical points do not agree. Thereafter, a little randomness is introduced into these polynomials which ensures that the limiting measures agree. We also discuss some of their consequences. \item In Chapter 3, we prove that the limiting measure of zeros and critical points of a sequence of random polynomials agree when a little randomness is introduced. The results stated in Chapter 2 are proved. \item In Chapter 4, we consider the matching problem between zeros and critical points of random polynomials. We show that when the zeros are real and i.i.d. from a given distribution with finite first moment, then the $\ell^1$ matching distance will be finite. We also consider the spacing between the zeros and critical points of the random polynomials. In the case where zeros are i.i.d. exp($\lambda$) random variables, we show that the extremal critical point is much closer to the extremal zero of the random polynomial. \end{itemize} In the second part we study the eigenvalues of certain products of random matrices. This is covered in Chapters 5 and 6. \begin{itemize} \item In Chapter 5, we derive the exact eigenvalue density for $X=X_1^{\epsilon_1}X_2^{\epsilon_2}\dots X_n^{\epsilon_n}$, where $\epsilon_i=\pm1$ and $X_i$ are i.i.d. complex Ginibre matrices. In other words, we derive the eigenvalue density for products of complex Ginibre matrices of fixed size in which some of them are inverted. It is also observed that they form a determinantal point process. \item In Chapter 7, we present a stronger version of the conjecture, by Lakshminarayan in \cite{arul}. We prove the conjecture in a special case when the entries of the matrices are distributed according to $\mu$ which has an atom. \end{itemize} \chapter{Matching between zeros and critical points of random polynomials} \label{ch:matching6} \section{Introduction} In the previous chapters we have seen the behaviour of the point cloud of zeros and critical points in bulk. In this chapter we study the pairing of zeros and critical points. Dennis and Hannay in \cite{hannay}, gave an electrostatic argument to show that the zeros and critical points are closely paired for a generic (random) polynomial of higher degree. In \cite{hanin1} Hanin argued that the critical points and zeros for a random polynomial are mutually paired by computing the covariance between these measures. We will restrict our attention to critical points of polynomials having all real zeros. We choose the real zeros to be i.i.d. random variables. In the next section we show that the sum of distances between zeros and critical points, when paired appropriately, remains bounded under the assumption that the random variables have finite first moment. In the next section we study the spacings between the extremal zeros and critical points when the zeros are i.i.d. exponential or uniform random variables. We prove that the extremal critical point is much closer to the extremal zero than any other zeros. \section{Matching distance between zeros and critical points of random polynomials.} Matching between two sets is defined as follows. Let $U,V$ be two sets of finite and equal cardinality in the complex plane. A matching is a bijection from $U$ to $V$. The concept of matching is used in qualitatively defining distance between two sets of same cardinality. Matching distance is a natural distance to quantify the closeness of the sets of zeros and critical points of a polynomial. There are several notions of matching distance. In this chapter we will deal with $\ell^1$ matching distance, which is defined as, \[ d_1(U,V)=\inf\limits_{\pi\in \text{$\mathfrak{S}$}_n}\sum\limits_{i=1}^{n}|u_i-v_{\pi(i)}|, \] where $U=\{u_1,u_2,\dots,u_n\}$, $V=\{v_1,v_2,\dots,v_n\}$ and $\pi=(\pi(1),\pi(2),\dots,\pi(n))$ is an element in set of permutations of size $n$ denoted by $\text{$\mathfrak{S}$}_n$. To define mapping distance between set of zeros and critical points, we include the element $0$ in the set of critical points. We map the set of critical points to set of zeros so that the sum of the distances between the critical points and zero set is minimized and call that as matching distance. The order statistics for a set of real numbers $\{\alpha_1,\alpha_2,\dots,\alpha_n\}$ are denoted as $\alpha_{(1)}\leq \alpha_{(2)} \leq \dots \leq \alpha_{(n)}$. In the following lemma we will compute the $\ell_1$ matching distance between two sets in the real line. \begin{lemma}\label{matching_distance} Let $X=\{x_1,x_2,\dots,x_n\}$ and $Y=\{y_1,y_2,\dots,y_n\}$ be two sets in the set of real numbers. Then the $\ell_1$ matching distance between $X$ and $Y$ is given by \[ d_1(X,Y)=\sum\limits_{i=1}^{n}|x_{(i)}-y_{(i)}|. \] \end{lemma} \begin{proof} Without loss generality assume that $x_i=x_{(i)}$ and $y_i=y_{(i)}$ for all $i=1,2,\dots,n$. We will show that the matching distance is attained for identity matching. Suppose not, let $\pi$ be the permutation for which the matching distance is attained. Then there is $i<j$ such that $\pi(i)>\pi(j)$. If we tweak the permutation to $\pi'$ by choosing $\pi'(i)=\pi(j)$, $\pi'(j)=\pi(i)$ and $\pi'(\ell)=\pi(\ell)$ for $\ell\neq i,j$, then $\sum\limits_{i=1}^{n}|x_{i}-y_{\pi'(i)}|\leq\sum\limits_{i=1}^{n}|x_{i}-y_{\pi(i)}|$. Repeating this argument, it follows that $\sum\limits_{i=1}^{n}|x_{i}-y_{i}|\leq\sum\limits_{i=1}^{n}|x_{i}-y_{\pi(i)}|$. Hence the matching distance is attained for identity permutation. \end{proof} As an application the above Lemma \ref{matching_distance}, we compute the $\ell_1$ matching distance between the sets of zeros and critical points of a given polynomial. \begin{proposition}\label{matching_zeros_critical_positive} Let $x_1,x_2,\dots,x_n$ be all non-negative numbers. Then the $\ell_1$ matching distance between the sets of zeros and critical points of a polynomial $P_n(z)=(z-x_1)(z-x_2)\dots(z-x_n)$ is given by \[ d_1(Z(P_n),Z(P_n')\cup\{0\})=\frac{1}{n}\sum\limits_{i=1}^{n}x_i. \] \end{proposition} \begin{proof} Let $\eta_1,\eta_2,\dots,\eta_{n-1}$ be the critical points of $P_n$. Because the critical points interlace the zeros of the $P_n$, we have $0\leq x_{(1)}\leq \eta_{(1)} \leq x_{(2)} \leq \dots, \leq \eta_{(n-1)} \leq x_{(n)}$. Applying Lemma \ref{matching_distance}, we get that \begin{equation} d_1(Z(P_n),Z(P_n')\cup\{0\}) = \sum\limits_{i=2}^{n}(x_{(i)}-\eta_{(i-1)})+x_{(1)}=\sum\limits_{i=1}^{n}x_i-\sum\limits_{i=1}^{n-1}\eta_{i}.\label{l_1_dist_eqn} \end{equation} Recall Vieta's formula that if $\alpha_1,\alpha_2,\dots,\alpha_n$ are the roots of the polynomial defined as $P(z)=a_0+a_1z+\dots+a_nz^n$, then \[ \sum\limits_{1\leq i_1<i_2<\dots<i_k\leq n}\alpha_{i_1}\alpha_{i_2}\dots \alpha_{i_k}=(-1)^k\frac{a_{n-k}}{a_n}.\] From Vieta's formula, observe the identity $\sum\limits_{i=1}^{n}x_i = \dfrac{n}{n-1}\sum\limits_{i=1}^{n-1}\eta_i$. Substituting this in \eqref{l_1_dist_eqn}, we get \[ d_1(Z(P_n),Z(P_n')\cup\{0\}) =\sum\limits_{i=1}^{n}x_i-\frac{n-1}{n}\sum\limits_{i=1}^{n}x_i=\frac{1}{n}\sum\limits_{i=1}^{n}x_i. \] \end{proof} \begin{proposition}\label{matching_zeros_critical} Let $x_1,x_2,\dots,x_k$ be negative numbers and $x_{k+1},x_{k+2},\dots,x_{n}$ be non-negative numbers. Then the $\ell_1$ matching distance between the sets of zeros and critical points of a polynomial $P_n(z)=(z-x_1)(z-x_2)\dots(z-x_n)$ is bounded by \[ d_1(Z(P_n),Z(P_n')\cup\{0\})\leq\frac{1}{k}\sum\limits_{i=1}^{k}|x_i|+\frac{1}{n-k}\sum\limits_{i=k+1}^{n}|x_i|. \] \end{proposition} Before proving Proposition \ref{matching_zeros_critical}, we prove the following lemma, which indicates that the critical points move towards right when a new zero is introduced into the polynomial towards the left of all the zeros. \begin{lemma}\label{assistant} Let $\eta_1,\eta_2,\dots,\eta_{n-1}$ be the critical points of the polynomial $P(z)=(z-\alpha_1)(z-\alpha_2)\dots(z-\alpha_n)$. Let $\eta_0',\eta_1',\eta_2',\dots,\eta_{n-1}'$ be the critical points of $Q(z)=(z-\alpha)P(z)$ where $\alpha<\alpha_i$ for $i=1,2,\dots,n$. Then, $\alpha_{(i+1)}-\eta_{(i)}' \leq \alpha_{(i+1)}-\eta_{(i)}$, for $i=1,2,\dots,n$ \end{lemma} \begin{proof} It is enough to show that $\eta_{(i)}\leq\eta_{(i)}'$. Define $L_P(z):=\frac{P'(z)}{P(z)}=\sum\limits_{i=1}^{n}\frac{1}{z-\alpha_i}$ and $L_Q(z):=\frac{Q'(z)}{Q(z)}=\frac{1}{z-\alpha}+L_P(z)$. Both $L_P$ and $L_Q$ are decreasing functions in any interval which does not contain any of the zeros of $P$ and $Q$. Fix an $i\in \{1,2,\dots,n\}$. Because $\eta_{(i)}$ is a critical point of $P$, we have $L_P(\eta_{(i)})=0$ and $L_Q(\eta_{(i)})=\frac{1}{\eta_{(i)}-\alpha}>0$. But $L_Q$ vanishes exactly once in the interval $(\alpha_{(i)},\alpha_{(i+1)})$ at $\eta_{(i)}'$. Combing the facts that $L_Q$ is decreasing in $(\alpha_{(i)},\alpha_{(i+1)})$ and $L_Q(\eta_{(i)})>0$, we get that $\eta_{(i)}\leq\eta_{(i)}'$. \end{proof} \begin{proof}[Proof of Proposition \ref{matching_zeros_critical}] With out loss of generality assume that $x_1\leq x_2\leq\dots x_k \leq0 \leq x_{k+1}\leq \dots x_n$. Let $\eta_1\leq\eta_2\dots \eta_{n-1}$ be the critical points of $P_n$. Factorize the polynomial $P_n$ as $P_n(z)=Q_n(z)R_n(z)$, where $Q_n(z)=(z-x_1)(z-x_2)\dots(z-x_k)$ and $R_n(z)=(z-x_{k+1})(z-x_{k+2})\dots(z-x_{n})$. If $\eta_1'\leq\eta_2'\leq \dots \eta_{k-1}'$ and $\eta_{k+1}'\leq\eta_{k+2}'\leq\dots\eta_{n-1}'$ are critical points of $Q_n$ and $R_n$ respectively, then by repeatedly applying the previous Lemma \ref{assistant} to $Q_n$ and $R_n$, we get $\eta_1-x_1 \leq \eta_1'-x_1, \dots, \eta_{k-1}-x_{k-1}\leq \eta_{k-1}'-x_{k-1}$ and $x_{k+2}-\eta_{k+1} \leq x_{k+2}-\eta_{k+1}', \dots,x_{n}-\eta_{n-1}\leq x_{n}-\eta_{n-1}'.$ The $\ell_1$ matching distances between zeros and critical points of $Q_n$ and $R_n$ are bounded by $\frac{1}{k}\sum\limits_{i=1}^{k}|x_i|$ and $\frac{1}{n-k}\sum\limits_{i=k+1}^{n}|x_i|$ respectively. Therefore we get, \begin{align} d_1(Z(P_n),Z(P_n')\cup\{0\}) &\leq d_1(Z(Q_n),Z(Q_n')\cup\{0\})+d_1(Z(R_n),Z(R_n')\cup\{0\}) \\&= \frac{1}{k}\sum\limits_{i=1}^{k}|x_i|+\frac{1}{n-k}\sum\limits_{i=k+1}^{n}|x_i|. \end{align} Hence the proposition is proved. \end{proof} Notice that Propositions \ref{matching_zeros_critical_positive} and \ref{matching_zeros_critical} are stated for deterministic polynomials. As an application we obtain the matching distance for the polynomials whose zeros are i.i.d. random variables. \begin{theorem}\label{random_matching_distance} Let $X_1,X_2,\dots$ be i.i.d. random variables satisfying $\ee{\text{$|X_1|$}}$, define the polynomial $P_n(z)=(z-X_1)(z-X_2)\dots(z-X_n)$. Then \[ \limsup\limits_{n\rightarrow \infty}d_1(Z(P_n),Z(P_n')\cup\{0\})\leq \ee{\text{$|X_1|$}}. \] Moreover if $X_i$s are non-negative random variables, then \[ \limsup\limits_{n\rightarrow \infty}d_1(Z(P_n),Z(P_n')\cup\{0\})= \ee{\text{$X_1$}}. \] \end{theorem} Range of the zeros of $P_n$ gives a trivial bound for the $\ell_1$ matching distance between the zeros and critical points of $P_n$. If the random variables are all bounded then this $\ell_1$ matching distance remains bounded uniformly for any $n$. Where as if the random variables $X_i$s are unbounded the above Theorem \ref{random_matching_distance} shows that the $\ell_1$ matching distance between the zeros and critical points of $P_n$ remains bounded almost surely uniformly for any $n$. \begin{proof}[Proof of Theorem \ref{random_matching_distance}] The second part of the theorem follows immediately by applying law of large numbers for Proposition \ref{matching_zeros_critical_positive}. For the first part, write $X_i=X_i^+-X_i^-$, where $X_i^+\geq0$ and $X_i^-<0$. Let $k_n=\#\{X_i:X_i<0\}$. Applying Proposition \ref{matching_zeros_critical} we get that $$d_1(Z(P_n),Z(P_n')\cup\{0\})\leq\frac{1}{k_n}\sum\limits_{i=1}^{n}X_i^-+\frac{1}{n-k_n}\sum\limits_{i=1}^{n}X_i^+.$$ Applying law of large numbers for the above we get $$\limsup\limits_{n\rightarrow \infty}d_1(Z(P_n),Z(P_n')\cup\{0\}) \leq \ee{\text{$X_1^-$}}+\ee{\text{$X_1^+$}}=\ee{\text{$|X_1|$}}.$$ \end{proof} \section{Spacings of zeros and critical points of random polynomials.} Theorem \ref{random_matching_distance} shows that even if $X_i$s are unbounded random variables then the $\ell_1$ matching distance between the zeros and critical points remain bounded. This indicates that the extremal critical points stay much closer to one of the zeros than the others. We formalize this in the case of exponential random variables as the following result. \begin{theorem}\label{exponential} Let $X_1,X_2,\dots X_n$ be i.i.d exponential random variables. Let $\eta_{(1)} \leq \eta_{(2)} \leq \dots \leq \eta_{(n-1)}$ be the critical points of the polynomial $P_n(z):=(z-X_1)(z-X_2)\dots(z-X_n)$. Then the following hold true, \begin{itemize} \item[1.] $n\log n(\eta_{(1)}-X_{(1)})\rightarrow 1$ in probability. \item[2.] $n\log n(X_{(n)}-\eta_{(n-1)})\rightarrow 1$ in probability. \end{itemize} \end{theorem} We will use R\'{e}nyi's representation~\cite{boucheron} for the order statistics of exponential random variables while proving Theorem \ref{exponential} and is stated below. \paragraph{R\'{e}nyi's representation for order statistics:} Let $Y_{(1)},Y_{(2)},\dots,Y_{(n)}$ be the order statistics of the sample of i.i.d exponential random variables, then $$(Y_{(1)},Y_{(2)},\dots,Y_{(n)}) \,{\buildrel d \over =}\, \left(\frac{E_n}{n},\frac{E_{n-1}}{n-1}+\frac{E_n}{n},\dots,E_1+\frac{E_2}{2}+\dots+\frac{E_n}{n}\right),$$ where $E_1,E_2,\dots,E_n$ are i.i.d exponential random variables. \begin{proof}[Proof of Theorem \ref{exponential}] Let $L_n(z):= \dfrac{P_n'(z)}{P_n(z)}=\sum\limits_{i=1}^{n}\frac{1}{z-X_i}=\sum\limits_{i=1}^{n}\frac{1}{z-X_{(i)}}$. If $z$ is a critical point of $P_n(z)$ then it satisfies $L_n(z)=0$. Hence from the equation $L_n(z)=0$ we have, \begin{align} \eta_{(1)}-X_{(1)} =& \left(\sum\limits_{i=2}^{n}\frac{1}{X_{(i)}-\eta_{(1)}}\right)^{-1} \\ \leq& \left(\sum\limits_{i=2}^{n}\frac{1}{X_{(i)}-X_{(1)}}\right)^{-1} \end{align} But $X_{(i)}-X_{(1)}=\frac{E_{n-1}}{n-1}+ \dots + \frac{E_{n-i+1}}{n-i+1}$ for all $i=2,3,\dots,n$, where $E_1,E_2,\dots,E_n$ are i.i.d exponential random variables. From R\'{e}nyi's representation it can be noticed that $\frac{E_{n-1}}{n-1}+ \dots + \frac{E_{n-i}}{n-i} \,{\buildrel d \over =}\, Y_{(i)}$ for $i=1,2,\dots,n-1$, where $Y_{(1)},Y_{(2)},\dots,Y_{(n-1)}$ are order statistics of $Y_1,Y_2,\dots,Y_{n-1}$ which are i.i.d exponential random variables. Therefore, \begin{align} \eta_{(1)}-X_{(1)} \,{\buildrel d \over =}\,& \left(\sum\limits_{i=2}^{n}\frac{1}{\frac{E_{n-1}}{n-1}+ \dots + \frac{E_{n-i+1}}{n-i+1}}\right)^{-1}\\ =& \left(\sum\limits_{i=1}^{n-1}\frac{1}{Y_{(i)}}\right)^{-1} = \left(\sum\limits_{i=1}^{n-1}\frac{1}{Y_i}\right)^{-1}.\label{eqn:thm1:1} \end{align} Observe that $\frac{1}{Y_1}$ is regularly varying-1 and applying central limit theorem ( Chapter 2, Theorem 7.7 in \cite{durrett}) for \eqref{eqn:thm1:1} we get, \begin{equation}\label{dist} \frac{(\eta_{(1)}-X_{(1)})^{-1}-n\log n}{n} \overset{d}{\to} R \end{equation} where $R$ has stable-1 distribution. From \eqref{dist} it follows that $n\log n(\eta_{(1)}-X_{(1)})\overset{p}{\to}1.$ The proof of the second statement $n\log n(X_{(n)}-\eta_{(n-1)})\stackrel{p}{\rightarrow} 1$ is similar as that of the first statement. \end{proof} \begin{remark}In the above Theorem \eqref{exponential}, instead of i.i.d exponential random variables we can choose i.i.d uniform random variables and obtain the same result. One may need to use the fact that if $U_1,U_2,\dots,U_n$ are i.i.d uniform random variables, then the order statistics \begin{align} (U_{(1)},U_{(2)}&,\dots,U_{(n)}) \stackrel{d}{=}\\ &\left(\frac{E_1}{E_1+E_2+\dots+E_{n+1}},\frac{E_1+E_2}{E_1+E_2+\dots+E_{n+1}}, \dots,\frac{E_1+E_2+\dots+E_n}{E_1+E_2+\dots+E_{n+1}}\right), \end{align} where $E_1,E_2,\dots,E_{n+1}$ are i.i.d exponential random variables. \end{remark} We believe the same result holds for any random variables whose densities satisfy certain regularity properties. The proof may be following the same idea but using R\'{e}nyi's representation theorem in more general form as given in \cite{boucheron} . \chapter{Poisson process from spectrum of Ginibre matrices} \label{ch:PoissonGinibre3} \section{Introduction} Let $X_n$ be a complex Ginibre matrix of size $n \times n$, whose entries $X(i,j)$ are i.i.d $CN(0,\frac{1}{n})$. Let $\lambda_{1,n},\lambda_{2,n},\dots,\lambda_{n,n}$ be the eigenvalues of $X_n$. It is long known that the point process constituted by the eigenvalues of $X_n$ is a determinantal point process with kernel given by $\textbf{$\mathbb{K}$}(z,w)=\sum\limits_{j=0}^{n-1}\frac{(nz\overline{w})^j}{j!}$ w.r.t the reference measure $\frac{1}{\pi}ne^{-n|z|^2}dm(z)$. Let $\mu_{1,n},\mu_{2,n},\dots,\mu_{n,n}$ be the eigenvalues of $X_n^n$ which are related as $\lambda_{k,n}^n=\mu_{k,n}$. In this chapter we will investigate whether the point processes obtained by $\mu_{k,n}$s will converge to some point process as $n \rightarrow \infty$. For this it is enough to check if the $k^{th}$ joint intensity $\rho_{k,n}'(z_1,z_2,\dots,z_k)$ of $\mu_{\ell,n}$s converge to the corresponding joint intensity of the limiting point process. We will show that the limiting point process is Poisson on $\textbf{$\mathbb{C}$}\backslash \{0\}$ with intensity $f(.)$. As the $k^{th}$ joint intensity of Poisson point process at points $z_1,z_2,\dots,z_k$ is $f(z_1)f(z_2)\dots f(z_k)$, it is enough to show that $\rho_{k,n}'(z_1,z_2,\dots,z_n)$ converge to the same. We will show the computational proof of the above fact by demonstrating the proof for $k=1,2$ and later prove for the general $k$. \section{Preliminaries and results} \begin{theorem}\label{poissonginibre} \end{theorem} \section{Proof of Theorem\ref{poissonginibre}} Let $\omega_n=e^{\frac{2\pi i}{n}}$ is an $n^{th}$root of unity. \subsubsection{Case:k=1} For this case there isn't anything to verify but merely obtain the intensity of the Poisson point process. \begin{align} \rho_{1,n}'(z) & = \lim\limits_{\epsilon \rightarrow 0}\dfrac{\Pr\left(\sum\limits_{i=1}^{n}\text{$\mathbbm{1}$}\left(\mu_{i,n}\in B(z,\epsilon)\right)=1 \right)}{\pi\epsilon^2}\\ & = \lim\limits_{\epsilon \rightarrow 0}\dfrac{\Pr\left(\sum\limits_{i=1}^{n}\sum\limits_{j=1}^{n}\text{$\mathbbm{1}$}\left(\lambda_{i,n}\in B\left(z^{\frac{1}{n}}\omega_n^j,\dfrac{\epsilon}{n|z|^{1-\frac{1}{n} }}+O(\epsilon^2)\right)\right)=1\right)}{\pi\epsilon^2}\\ & = \lim\limits_{\epsilon \rightarrow 0}\dfrac{\Pr\left(\sum\limits_{i=1}^{n}\text{$\mathbbm{1}$}\left(\lambda_{i,n}\in \bigcup\limits_{j=1}^{n}B\left(z^{\frac{1}{n}}\omega_n^j,\dfrac{\epsilon}{n|z|^{1-\frac{1}{n}}}+O(\epsilon^2)\right)\right)=1\right)}{\pi\epsilon^2}\\ & = \lim\limits_{\epsilon \rightarrow 0}\dfrac{\int\limits_{A_n}\rho_{1,n}(w)dm(w)+O(\epsilon^4)}{\pi\epsilon^2} \end{align} Where $A_n=\bigcup\limits_{j=1}^{n}B\left(z^{\frac{1}{n}}\omega_n^j,\dfrac{\epsilon}{n|z|^{1-\frac{1}{n}}}+O(\epsilon^2)\right)$. Hence \begin{align} \rho_{1,n}(z) & = \dfrac{\sum\limits_{i=1}^{n}\rho_{1,n}(z^\frac{1}{n}\omega_n^j)}{n^2|z|^{2(1-\frac{1}{n})}}= \dfrac{K_n(z^\frac{1}{n},z^\frac{1}{n})}{n|z|^{2-\frac{2}{n}}}\\ & = \sum\limits_{i=0}^{n-1}\dfrac{n^i|z|^\frac{2i}{n}}{i!}\dfrac{e^{-n|z|^\frac{2}{n}}}{\pi|z|^{2-\frac{2}{n}}}\\ & = \dfrac{1}{\pi|z|^{2-\frac{2}{n}}}\sum\limits_{i=0}^{n-1}\dfrac{n^i|z|^\frac{2i}{n}}{i!}e^{-n|z|^\frac{2}{n}}\\ & = \frac{1}{\pi |z|^{2-\frac{2}{n}}}\Pr\left(\sum_{i=1}^{n}P_i^{(|z|^\frac{2}{n})}<n\right)\\ & \text{(where $P_1^{(t)},P_2^{(t)},\dots,P_n^{(t)}$ are i.i.d Poisson(t) random variables)} \end{align} Therefore the 1-point intensity of the limiting point process will be $$\rho_1'(z)=\lim\limits_{n\rightarrow\infty}\rho_{1,n}'(z)=\lim\limits_{n\rightarrow\infty}\frac{1}{\pi|z|^2}|z|^\frac{2}{n}\Pr\left(\sum\limits_{i=1}^{n}P_i^{(|z|^\frac{2}{n})}<n\right)=\frac{1}{2\pi|z|^2}$$ \subsubsection{Case:k=2} \begin{align} \rho_{2,n}'(z_1,z_2) & = \lim\limits_{\epsilon\rightarrow 0}\dfrac{\Pr\left(\sum\limits_{i=1}^{n}\text{$\mathbbm{1}$}\left(\mu_{i,n}\in B(z_1,\epsilon)\right)=1 \& \sum\limits_{i=1}^{n}\text{$\mathbbm{1}$}\left(\mu_{i,n}\in B(z_2,\epsilon)\right)=1 \right)}{(\pi\epsilon^2)^2}\\ \end{align} Using the relation $\mu_{i,n}=\lambda_{i,n}^n$ and changing variables we obtain that the events $\{\mu_{i,n}\in B(z_1,\epsilon)\}$ and $\{\lambda_{i,n}\in \bigcup\limits_{j=1}^{n}B(z_1^\frac{1}{n}\omega_n^j,\epsilon_n')\}$ as well as $\{\mu_{i,n}\in B(z_2,\epsilon)\}$ and $\{\lambda_{i,n}\in \bigcup\limits_{j=1}^{n}B(z_2^\frac{1}{n}\omega_n^j,\epsilon_n'')\}$ are identical for some $\epsilon_n'=\frac{\epsilon}{n|z_1|^{1-\frac{1}{n}}}+O(\epsilon^2)$ and $\epsilon_n''=\frac{\epsilon}{n|z_2|^{1-\frac{1}{n}}}+O(\epsilon^2)$. Hence we have \begin{align} \rho_{2,n}'(z_1,z_2)& = \lim\limits_{\epsilon\rightarrow 0} \dfrac{\Pr\left(\sum\limits_{i=1}^{n}\text{$\mathbbm{1}$}\left(\lambda_{i,n}\in \bigcup\limits_{j=1}^{n}B(z_1^\frac{1}{n}\omega_n^j,\epsilon_n')\right)=1 \& \sum\limits_{i=1}^{n}\text{$\mathbbm{1}$}\left(\lambda_{i,n}\in\bigcup\limits_{j=1}^{n} B(z_2^\frac{1}{n}\omega_n^j,\epsilon_n'')\right)=1 \right)}{(\pi\epsilon^2)^2}\\ & = \lim\limits_{\epsilon \rightarrow 0} \dfrac{\int\limits_{A_n\times B_n}\rho_{2,n}(\alpha,\beta)dm(\alpha)dm(\beta)}{(\pi\epsilon^2)^2} \end{align} where $A_n=\bigcup\limits_{j=1}^{n}B(z_1^\frac{1}{n}\omega_n^j,\epsilon_n'),B_n=\bigcup\limits_{j=1}^{n}B(z_2^\frac{1}{n}\omega_n^j,\epsilon_n'')$ and $\rho_{2,n}(.,.)$ is the $2^{nd}$ joint intensity of the eigenvalue process of $X_n$. Therefore, \begin{align} \rho_{2,n}'(z_1,z_2) &= \dfrac{\sum\limits_{i,j=1}^{n}\left(\rho_{2,n}(z_1^{\frac{1}{n}}\omega_n^i,z_2^\frac{1}{n}\omega_n^j)\right)}{n^4|z_1z_2|^{2-\frac{2}{n}}}\\ &= \frac{1}{n^4|z_1z_2|^{2-\frac{2}{n}}}\sum\limits_{i,j=1}^{n}\left(K(z_1^{\frac{1}{n}}\omega_n^j,z_1^{\frac{1}{n}}\omega_n^i)K(z_2^{\frac{1}{n}}\omega_n^j,z_2^{\frac{1}{n}}\omega_n^j)-K(z_1^{\frac{1}{n}}\omega_n^i,z_2^{\frac{1}{n}}\omega_n^j)K(z_2^{\frac{1}{n}}\omega_n^j,z_1^{\frac{1}{n}}\omega_n^i)\right)\\ &= \frac{1}{n^4|z_1z_2|^{2-\frac{2}{n}}}\left(n^2K(z_1^\frac{1}{n},z_1^\frac{1}{n})K(z_2^\frac{1}{n},z_2^\frac{1}{n})-\sum\limits_{i,j=1}^{n}|K(z_1^\frac{1}{n}\omega_n^i,z_2^\frac{1}{n}\omega_n^j)|^2\right) \label{last2} \end{align} We will now show that $\lim\limits_{n\rightarrow\infty}\dfrac{\sum\limits_{i,j=1}^{n}|K(z_1^\frac{1}{n}\omega_n^i,z_2^\frac{1}{n}\omega_n^j)|^2}{n^4|z_1z_2|^{2-\frac{2}{n}}}=0$. \begin{align} \sum\limits_{i,j=1}^{n}|K(z_1^\frac{1}{n}\omega_n^i,z_2^\frac{1}{n}\omega_n^j)|^2 & = n^2e^{-n(|z_1|^\frac{2}{n}+|z_2|^\frac{2}{n})}\sum\limits_{i,j=1}^{n}\left(\sum\limits_{\ell=0}^{n-1}(z_1^\frac{1}{n}\omega_n^i\overline{z_2}^\frac{1}{n}\overline{\omega_n}^{j})^\ell\frac{n^\ell}{\ell!}\right)\left(\sum\limits_{\ell=0}^{n-1}(\overline{z_1}^\frac{1}{n}\overline{\omega_n}^{i}z_2^\frac{1}{n}\omega_n^{j})^\ell\frac{n^\ell}{\ell!}\right)\\ & = n^2e^{-n(|z_1|^\frac{2}{n}+|z_2|^\frac{2}{n})}\sum\limits_{i,j=1}^{n}\left(\sum\limits_{\ell,k=1}^{n-1}\frac{\left(z_1^\frac{\ell}{n}\omega_n^{i\ell}\overline{z_2}^\frac{\ell}{n}\overline{\omega_n}^{j\ell}\overline{z_1}^\frac{k}{n}\overline{\omega_n}^{ki}z_2^\frac{k}{n}\omega_n^{kj}\right)n^{k+\ell}}{k!\ell!}\right)\\ & =n^2e^{-n(|z_1|^\frac{2}{n}+|z_2|^\frac{2}{n})}\sum\limits_{i,j=1}^{n}\left(\sum\limits_{\ell,k=1}^{n-1}\frac{n^{k+\ell}}{k!\ell!}\left(\omega_n^{(i-j)(\ell-k)}\right)\left(z_1^\frac{\ell}{n}\overline{z_1}^\frac{k}{n}z_2^\frac{k}{n}\overline{z_2}^\frac{\ell}{n}\right)\right) \end{align} Let $z_1=r_1e^{i\theta_1}$, $z_2=r_2e^{i\theta_2}$ and $\alpha=e^{i(\theta_1-\theta_2)}$. \begin{align} \sum\limits_{i,j=1}^{n}|K(z_1^\frac{1}{n}\omega_n^i,z_2^\frac{1}{n}\omega_n^j)|^2 & = n^2e^{-n(|z_1|^\frac{2}{n}+|z_2|^\frac{2}{n})}\sum\limits_{i,j=1}^{n}\left(\sum\limits_{\ell,k=0}^{n-1}\frac{n^{k+\ell}}{k!\ell!}\left(\omega_n^{(i-j)(\ell-k)}\right)\alpha^{\ell-k}(r_1r_2)^\frac{\ell+k}{n}\right)\\ & = n^2e^{-n(|z_1|^\frac{2}{n}+|z_2|^\frac{2}{n})}\sum\limits_{\ell\neq k=0}^{n-1}\left(\sum\limits_{i,j=1}^{n}\frac{n^{k+\ell}}{k!\ell!}\left(\omega_n^{(i-j)(\ell-k)}\alpha^{\ell-k}(r_1r_2)^\frac{\ell+k}{n}\right)\right)\\\label{eqn:case2:1} & +n^2e^{-n(|z_1|^\frac{2}{n}+|z_2|^\frac{2}{n})}\sum\limits_{i,j=1}^{n}\left(\sum\limits_{\ell=k=0}^{n-1}\frac{n^{k+\ell}}{k!\ell!}(r_1r_2)^\frac{\ell+k}{n}\right) \end{align} The term \eqref{eqn:case2:1} doesnot contribute to the sum, because for fixed $\ell\neq k$ and $j$ the sum $\sum\limits_{i=1}^{n}\omega_n^{(\ell-k)(i-j)}=0$. Hence, \begin{align} \lim\limits_{n \rightarrow \infty}\frac{1}{n^4|z_1z_2|^{2-\frac{2}{n}}}\sum\limits_{i,j=1}^{n}|K(z_1^\frac{1}{n}\omega_i,z_2^\frac{1}{n}\omega_j)|^2 & = \lim\limits_{n \rightarrow \infty}\frac{n^2e^{-n(r_1^\frac{2}{n}+r_2^\frac{2}{n})}}{n^4(r_1r_2)^{2-\frac{2}{n}}}\sum\limits_{i,j=1}^{n}\left(\sum\limits_{\ell=k=0}^{n-1}\frac{n^{k+\ell}}{k!\ell!}(r_1r_2)^\frac{\ell+k}{n}\right)\\ & =\lim\limits_{n \rightarrow \infty}\frac{1}{\pi^2(r_1r_2)^{2-\frac{2}{n}}} \sum\limits_{\ell=0}^{n-1}\left(\frac{n^\ell e^{-n r_1^{\frac{2}{n}}}r_1^{\frac{2\ell}{n}}}{\ell!}\frac{n^\ell e^{-n r_2^{\frac{2}{n}}}r_2^{\frac{2\ell}{n}}}{\ell!}\right)\\ & \leq \lim\limits_{n \rightarrow \infty}\frac{1}{\pi^2(r_1r_2)^{2-\frac{2}{n}}} \sum\limits_{\ell=0}^{\infty}\left(\frac{n^\ell e^{-n r_1^{\frac{2}{n}}}r_1^{\frac{2\ell}{n}}}{\ell!}\frac{n^\ell e^{-n r_2^{\frac{2}{n}}}r_2^{\frac{2\ell}{n}}}{\ell!}\right)\\ & = \lim\limits_{n\rightarrow\infty}\frac{1}{\pi^2(r_1r_2)^{2-\frac{2}{n}}}\Pr\left(Y_1^{(n)}=Y_2^{(n)}\right) \end{align} Where $Y_1^{(n)}$ and $Y_2^{(n)}$ are independent Poisson random variables with means $nr_1^\frac{2}{n}$ and $nr_2^\frac{2}{n}$ respectively. At the same time we can write each $Y_1^{(n)}$ and $Y_2^{(n)}$ as the sum of n-independent identical Poisson random variables whose means are $r_1^\frac{2}{n}$ and $r_2^\frac{2}{n}$ respectively. In other words $Y_1^{(n)}=P_1^{n}+P_2^{(n)}+\dots +P_n^{(n)}$, $Y_2^{(n)}=Q_1^{(n)}+Q_2^{(n)}+\dots+Q_n^{(n)}$ where $P_i^{(n)}$s and $Q_i^{(n)}$s are independent Poisson random variables with means $r_1^{\frac{2}{n}}$ and $r_2^\frac{2}{n}$ respectively. We will show that $$ \lim\limits_{n\rightarrow\infty}\frac{1}{\pi^2(r_1r_2)^{2-\frac{2}{n}}}\Pr\left(Y_1^{(n)}-Y_2^{(n)}\right)=0 \label{eqn:case2:3} $$ Once we prove \eqref{eqn:case2:3} we will have $$\rho_2'(z_1,z_2)=\lim\limits_{n\rightarrow\infty}\rho_{2,n}'(z_1,z_2)=\lim\limits_{n\rightarrow\infty}\dfrac{n^2K(z_1^\frac{2}{n},z_1^\frac{2}{n})K(z_2^\frac{2}{n},z_2^\frac{2}{n})}{n^4|z_1z_2|^{2-\frac{2}{n}}}=\frac{1}{(2\pi)^2|z_1z_2|^2}=\rho_1'(z_1)\rho_1'(z_2)$$ Now it remains to prove \eqref{eqn:case2:3}. We shall state Lynaponav central limit theorem and show that the triangular array $P_i^{(n)}-Q_i^{(n)}$ satisfies the hypothesis of Lyapunov central limit theorem and use this to show \eqref{eqn:case2:3}. \begin{theorem}[Lyaponov CLT \cite{billingsley}] Let $\{X_{i,n}\}$ be a triangular array of random variables with i.i.d random variables in each row and let $s_n^2=\sum\limits_{i=1}^{n}Var(X_{i,n})$, $S_n=\sum\limits_{i=1}^{n}X_{i,n}$. If $ \lim\limits_{n\rightarrow\infty}\dfrac{\sum\limits_{i=1}^{n}\text{$\mathbb{E}$}|X_{i,n}-\text{$\mathbb{E}$}(X_{i,n})|^{2+\delta}}{s_n^{2+\delta}}=0$ for some $\delta>0$ then as $n$ approaches to infinity the random variables $\dfrac{S_n-\text{$\mathbb{E}$}(S_n)}{s_n}$ converges in distribution to $N(0,1)$. $$\dfrac{S_n-\text{$\mathbb{E}$}(S_n)}{s_n}\xrightarrow{d} N(0,1)$$. \end{theorem} For showing the hypothesis of Lyaponov CLT is satisfied choose $X_{i,n}=P_i^{(n)}-Q_i^{(n)}$. Then $$s_n^4=\left(\sum\limits_{i=1}^{n}Var(X_{i,n})\right)^2=\left(\sum\limits_{i=1}^{n}Var(P_i^{(n)}-Q_i^{(n)})\right)^2=\left(\sum\limits_{i=1}^{n}\left(Var(P_i^{(n)})+Var(Q_i^{(n)})\right)\right)^2=n^2(r_1^\frac{2}{n}+r_2^\frac{2}{n})^2$$ and $$\text{$\mathbb{E}$}\left(P_i^{(n)}+Q_i^{(n)}-\text{$\mathbb{E}$}(P_i^{(n)}+Q_i^{(n)})\right)^4=\left(r_1^\frac{2}{n}+r_2^\frac{2}{n}\right)\left(3(r_1^\frac{2}{n}+r_2^\frac{2}{n})+4\right)$$. Therefore $$ \lim\limits_{n\rightarrow\infty}\dfrac{\sum\limits_{i=1}^{n}\text{$\mathbb{E}$}|X_{i,n}-\text{$\mathbb{E}$}(X_{i,n})|^{2+2}}{s_n^{2+2}} = \lim\limits_{n\rightarrow\infty}\dfrac{n\left(r_1^\frac{2}{n}+r_2^\frac{2}{n}\right)\left(3(r_1^\frac{2}{n}+r_2^\frac{2}{n})+4\right)}{n^2\left(r_1^\frac{2}{n}+r_2^\frac{2}{n}\right)^2}=0 $$ So $\dfrac{Y_1^{(n)}-Y_2^{(n)}-n(r_1^\frac{2}{n}-r_2^\frac{2}{n})}{\sqrt{n(r_1^\frac{2}{n}+r_2^\frac{2}{n})}}\xrightarrow{d}N(0,1)$. Observing that $\lim\limits_{n \rightarrow \infty}\sqrt{n}(r_1^\frac{2}{n}-r_2^\frac{2}{n})=0$ and $\lim\limits_{n\rightarrow \infty}(r_1^\frac{2}{n}+r_2^\frac{2}{n})=2$ we get $\dfrac{Y_1^{(n)}-Y_2^{(n)}}{\sqrt{n}}\xrightarrow{d}N(0,2)$. From the above facts we obtain \begin{align} \lim\limits_{n\rightarrow\infty}\Pr(Y_1^{(n)}-Y_2^{(n)}=0) & \leq\lim\limits_{n\rightarrow\infty}\Pr(|Y_1^{(n)}-Y_2^{(n)}|<1)\\ &=\lim\limits_{n\rightarrow\infty}\Pr\left(\frac{|Y_1^{(n)}-Y_2^{(n)}|}{\sqrt{n}}<\frac{1}{\sqrt{n}}\right) =0 \end{align} Hence we have shown that \eqref{eqn:case2:3} is satisfied and verified the case k=2. \subsubsection{Case:Arbitrary-$k$} For proving the result for $k \geq 2$, we need to Like in the previous cases let us define $z_j=r_j\alpha_j$ for $j=1,2,\dots,k$ where $|\alpha_j|=1$. \begin{align} \rho_{k,n}'(z_1,z_2,\dots,z_k) & =\frac{\sum\limits_{i_1,i_2,\dots,i_k=1}^{n}\rho_{k,n}(z_1^{\frac{1}{n}}\omega_n^{i_1},z_2^{\frac{1}{n}}\omega_n^{i_2},\dots,z_k^{\frac{1}{n}}\omega_n^{i_k})}{n^{2k}|z_1z_2\dots z_k|^{2-\frac{2}{n}}}\\ & = \frac{\sum\limits_{i_1,i_2,\dots,i_k=1}^{n}\det\left(\left(K(z_\ell^{\frac{1}{n}}\omega_n^{i_\ell},z_m^{\frac{1}{n}}\omega_n^{i_m})\right)\right)_{1\leq\ell,m\leq k}}{n^{2k}|z_1z_2\dots z_k|^{2-\frac{2}{n}}}\\ & = \frac{\sum\limits_{i_1,i_2,\dots,i_k=1}^{n}\sum\limits_{\sigma \in S_k} sgn(\sigma) \prod\limits_{s=1}^{k}K(z_s^{\frac{1}{n}}\omega_n^{i_s},z_{\sigma(s)}^{\frac{1}{n}}\omega_n^{i_{\sigma(s)}})}{n^{2k}|z_1z_2\dots z_k|^{2-\frac{2}{n}}}\\ & = \frac{\sum\limits_{i_1,i_2,\dots,i_k=1}^{n}\sum\limits_{\sigma \in S_k} sgn(\sigma) \prod\limits_{s=1}^{k}\sum\limits_{j=0}^{k-1}\frac{n^j}{j!}(r_sr_{\sigma(s)}\alpha_s\overline{\alpha_{\sigma(s)}})^\frac{j}{n}\omega_n^{j(i_s-i_{\sigma(s)})}\times ne^{-(r_s^\frac{2}{n}+r_{\sigma(s)}^\frac{2}{n})/2}}{n^{2k}|z_1z_2\dots z_k|^{2-\frac{2}{n}}}\\ & =\sum\limits_{i_1,i_2,\dots,i_k=1}^{n}\dfrac{n^ke^{-(r_1^\frac{2}{n}+\dots+r_k^\frac{2}{n})}}{n^{2k}\pi^k|z_1z_2\dots z_k|^{2-\frac{2}{n}}} \sum\limits_{\sigma \in S_k} sgn(\sigma)\prod\limits_{s=1}^{k}\left(\sum\limits_{j=0}^{k-1}(r_sr_{\sigma(s)})^\frac{j}{n}\omega_n^{j(i_s-i_{\sigma(s)})}(\alpha_s\overline{\alpha_{\sigma(s)}})^\frac{j}{n}\frac{n^j}{j!}\right)\label{eqn:perm}\\ \end{align} Expanding the product in \ref{eqn:perm} we get \begin{align} \sum\limits_{i_1,i_2,\dots,i_k=1}^{n}\dfrac{n^ke^{-(r_1^\frac{2}{n}+\dots+r_k^\frac{2}{n})}}{n^{2k}\pi^k|z_1z_2\dots z_k|^{2-\frac{2}{n}}} \sum\limits_{\sigma \in S_k}sgn(\sigma)\sum\limits_{j_1,j_2,\dots,j_k=0}^{k-1}\dfrac{n^{(j_1+j_2+\dots+j_k)}}{j_1!j_2!\dots j_k!}\prod\limits_{\ell=1}^{k}(r_\ell r_{\sigma(\ell)}\alpha_\ell\overline{\alpha_{\sigma(\ell)}})^{\frac{j_\ell}{n}}\omega_n^{j_\ell(i_\ell-i_{\sigma(\ell)})}\label{eqn:perm1} \end{align} Let $\sigma=C_1C_2\dots C_\ell$ where $C_1,C_2,\dots ,C_\ell$ are disjoint cycles of $\sigma$. Then \ref{eqn:perm1} will be reduced to\\ ***************************************\\ need more explanation\\ ***************************************\\ \begin{align} \sum\limits_{i_1,i_2,\dots,i_k=1}^{n}\dfrac{e^{-(r_1^\frac{2}{n}+\dots+r_k^\frac{2}{n})}}{n^{k}|z_1z_2\dots z_k|^{2-\frac{2}{n}}}\sum\limits_{\sigma \in S_n}sgn(\sigma)\sum\limits_{j_{k_1},j_{k_2},\dots,j_{k_\ell}=0}^{k-1}\prod\limits_{v=1}^{\ell}\prod\limits_{u\in C_i}(r_u r_{\sigma(u)})^{j_{k_v}}\frac{n^{j_{k_v}}}{j_{k_v}!} \end{align} \section{Result} \chapter{Proofs of Theorems \ref{thm1} and \ref{thm2}.} \label{ch:proofs5} \section{Outline of proofs.} The proofs here are adapted from the proof of Kabluchko's theorem as presented in \cite{kabluchko}. The proofs involve in analysing the function $L_n(z)$. In case of Theorem \ref{thm1} define $L_n(z)=\frac{P_n'(z)}{P_n(z)} = \sum\limits_{k=1}^{n}\frac{1}{z-\xi_k}$. We shall prove the theorems by showing that the hypotheses of the Theorems \ref{thm1} and \ref{thm2} imply the following three statements. \begin{align} &\text{For Lebesgue a.e. $z \in \mathbb{C}$ } \text{ and for every }\epsilon>0, \lim\limits_{n\rightarrow \infty}\Pr\left(\frac{1}{n}\log|L_n(z)|>\epsilon\right)=0. \tag{A1} \label{A1}\\ &\text{For Lebesgue a.e. $z \in \mathbb{C}$ } \text{ and for every }\epsilon>0, \lim\limits_{n\rightarrow \infty}\Pr\left(\frac{1}{n}\log|L_n(z)|<-\epsilon\right)=0. \tag{A2} \label{A2}\\ &\text{For any }r>0, \text{the sequence }\left\{\int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log^2|L_n(z)|\right\}_{n\geq 1} \text{ is tight.} \tag{A3} \label{A3} \end{align} Statements \eqref{A1} and \eqref{A2} assert that $\frac{1}{n}\log|L_n(z)|$ converge to $0$ in probability. Statement \eqref{A3} assert that the sequence $\{\int\limits_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log^2|L_n(z)|\}_{n \geq 1}$ is tight. A lemma of Tao and Vu links the above two facts to yield that $\{\int\limits_{\text{$\mathbb{D}$}_r}\frac{1}{n}\log|L_n(z)|\}_{n \geq 1}$ converge to $0$ in probability. We state this lemma below. \begin{lemma}[Lemma~3.1 in~\cite{taovu}]\label{lem:tao_vu} Let $(X,\mathcal{A},\nu)$ be a finite measure space and $f_n:X\to \mathbb{R}$, $n\geq 1$ random functions which are defined over a probability space $(\Omega, \mathcal{B}, \mathbb{P})$ and are jointly measurable with respect to $\mathcal{A}\otimes \mathcal{B}$. Assume that: \begin{enumerate} \item For $\nu$-a.e.\ $x\in X$ we have $f_n(x)\to 0$ in probability, as $n\to\infty$. \item For some $\delta>0$, the sequence $\int_X |f_n(x)|^{1+\delta} d\nu(x)$ is tight. \end{enumerate} Then, $\int_X f_n(x)d\nu(x)$ converge in probability to $0$. \end{lemma} Thus it follows from the above assertions \eqref{A1}, \eqref{A2}, \eqref{A3} and Lemma \ref{lem:tao_vu}, that $\int\limits_{\text{$\mathbb{D}$}_r}\frac{1}{n}\log|L_n(z)|dm(z) \rightarrow 0$ in probability for any $r>0$. Choose any $f\in C_c^{\infty}(\mathbb{C})$, assume that $\mbox{support}(f) \subseteq \text{$\mathbb{D}$}_r$ and define $f_n(z)=\frac{1}{n}\left(\log|L_n(z)|\right)\Delta f(z)$. Because $f$ is a bounded function and $\frac{1}{n}\log|L_n(z)|$ satisfy the hypothesis of Lemma \ref{lem:tao_vu}, the functions $f_n$ also satisfy the hypothesis of Lemma \ref{lem:tao_vu}. Therefore we get that $\int\limits_{\text{$\mathbb{D}$}_r}f_n(z)dm(z)\rightarrow 0$ in probability. Applying Green's theorem twice we have the identity, \[\int_{\text{$\mathbb{D}$}_r}^{}f(z)\Delta\frac{1}{n}\log|L_n(z)| = \int_{\text{$\mathbb{D}$}_r}^{}\frac{1}{n}\log|L_n(z)|\Delta f(z)dm(z). \] The left hand side of the above integral is defined in the sense of distributions. Therefore it follows that $\int_{\text{$\mathbb{D}$}_r}^{} f(z)\frac{1}{n}\Delta\log|L_n(z)|\rightarrow 0 $ in probability. This suffices for Theorem \ref{thm2}. We complete the proof of Theorem \ref{thm1} by the following arguments. In the sense of distributions we have \begin{equation}\label{eqn:distribution} \int_{\text{$\mathbb{D}$}_r}^{} f(z)\frac{1}{n}\Delta\log|L_n(z)| = \frac{1}{n}\sum\limits_{k=1}^{n}f(\xi_k)-\frac{1}{n}\sum\limits_{k=1}^{n-1}f(\eta_k^{(n)}) \end{equation} From Lemma \ref{random_seq_limit} it follows that the sequence $\{\xi_n\}_{n\geq1}$ is $\mu$-distributed. Hence \linebreak $\frac{1}{n}\sum_{k=1}^{n}f(\xi_k) \rightarrow \int\limits_{\text{$\mathbb{D}$}_r}f(z)d\mu(z)$ almost surely. Therefore from \eqref{eqn:distribution} we get, \begin{equation} \frac{1}{n}\sum_{k=1}^{n-1}f(\eta_k^{(n)}) \rightarrow \int\limits_{\text{$\mathbb{D}$}_r}f(z)d\mu(z) \text{ in probability.} \label{proof_1} \end{equation} Because for any $f \in C_c^\infty(\text{$\mathbb{C}$})$ and $\epsilon>0$, the sets of the form $\{\mu:|\int\limits_{\text{$\mathbb{C}$}}f(z)d\mu(z)|<\epsilon\}$ form an open base at origin, from Definition \ref{modes of convergence} and \eqref{proof_1} it follows that $\frac{1}{n-1}\sum_{k=1}^{n-1}\delta_{\eta_i^{(n)}} \xrightarrow{w} \mu$ in probability. We show \eqref{A1}, by obtaining moment bounds for $L_n(z)$. To show \eqref{A2} we will use a concentration bound for the function $L_n(z)$. In either of the Theorems \ref{thm1} and \ref{thm2}, observe that $L_n(z)$ is a sum of independent random variables. Kolmogorov-Rogozin inequality gives the concentration bounds for sums of independent random variables. A version of Kolmogorov-Rogozin inequality which will be used later in the proofs is stated below. \paragraph{Kolmogorov-Rogozin inequality (multi-dimensional version)}\label{KR-Inequality} [Corollary 1. of Theorem 6.1 in \cite{KR1}.] Let $X_1,X_2, \dots $ be independent random vectors in $\mathbb{R}^\text{$n$}$. Define the concentration function, \begin{equation} Q(X,\delta) := \sup_{a\in \mathbb{R}^\text{$n$}}\Pr(X \in B(a,\delta)). \end{equation} Let $\delta_i \leq \delta$ for each $i$, then \begin{equation} Q(X_1+\dots + X_n,\delta) \leq \frac{C\delta}{\sqrt{\sum_{i=1}^{n}\delta_i^2(1-Q(X_i,\delta_i))}}.\label{kol-rog-ineq} \end{equation} It remains to show that the hypotheses of Theorems \ref{thm1} and \ref{thm2} imply \eqref{A1}, \eqref{A2} and \eqref{A3}. We show this in the subsequent sections. \section{Proof of Theorem \ref{thm1}} In the following lemma we show that the hypothesis of the Theorem \ref{thm1} imply \eqref{A1}. \begin{lemma}\label{momentbound} Let $L_n(z)=\sum\limits_{k=1}^{n}\frac{1}{z-\xi_k}$ where $\xi_ks$ are as in the Theorem \ref{thm1}. Then for any $\epsilon > 0$, and for Lebesgue a.e. $z \in \mathbb{C}$ we have $\lim\limits_{n\rightarrow \infty}\Pr(\frac{1}{n}\log|L_n(z)|\geq \epsilon)= 0$. \end{lemma} \begin{proof} Define $A_n^{\epsilon}=\bigcup\limits_{k=1}^{n}\{z:|z-a_k|<e^{-n\epsilon} \text{ or } |z-b_k|<e^{-n\epsilon}\}$ and $F^{\epsilon}=\limsup\limits_{n \rightarrow \infty} A_n^{\epsilon}$, then $F^\epsilon$ are decreasing sets in $\epsilon$. For these sets we have $\sum\limits_{n=1}^{\infty}m(A_n^{\epsilon}) \leq \sum\limits_{n=1}^{\infty}2\pi ne^{-2n\epsilon} < \infty$, where $m$ is Lebesgue measure on complex plane. Applying Borel-Cantelli lemma to the sequence $\{A_n^{\epsilon}\}_{n\geq 1}$ we get $m(F^{\epsilon})=0$. Because $F^\epsilon$ are decreasing sets in $\epsilon$, we have that if $F=\bigcup\limits_{\epsilon>0}F^{\epsilon}$, then $m(F)=0$. Choose $z \in F^c$, there is $N_z^{\epsilon}$ such that for any $n>N_z^{\epsilon}$ we have $z \notin A_n^{\epsilon}$. Therefore $\frac{1}{|z-\xi_n|}>e^{n\epsilon}$ is satisfied only for finitely many $n$. Hence we have $|L_n(z)|<M+ne^{n\epsilon}$, where $M$ is the finite random number obtained from the terms for which the inequality $\frac{1}{|z-\xi_n|}>e^{n\epsilon}$ is violated. It follows from here $\limsup\limits_{n\rightarrow \infty}\frac{1}{n}\log|L_n(z)|<\epsilon$ almost surely. Therefore for $z\notin \mathbb{C}$, we have $\lim\limits_{n\rightarrow \infty}\Pr(\frac{1}{n}\log|L_n(z)|\geq \epsilon)= 0$. \end{proof} \begin{remark} In proof of Lemma \ref{momentbound}, we have proved a stronger statement that for Lebesgue almost every $z$, $\limsup\limits_{n\rightarrow \infty}\frac{1}{n}\log|L_n(z)|=0$ almost surely. \end{remark} We will use the null set $F$ defined in the proof of above Lemma \ref{momentbound} in the proofs of subsequent lemmas. In the forthcoming lemma we establish \eqref{A2}. \begin{lemma}\label{kolmogorov-rogozin} Let $L_n(z)=\sum\limits_{k=1}^{n}\frac{1}{z-\xi_k}$ where $\xi_ks$ are as in the Theorem \ref{thm1}. Then for any $\epsilon > 0$, and almost every $z$ we have $\lim\limits_{n\rightarrow \infty}\Pr(\frac{1}{n}\log|L_n(z)|\leq -\epsilon)= 0$. \end{lemma} \begin{proof} Fix $z \in F^c$, where $F$ is as defined in proof of lemma \ref{momentbound}. From Kolmogorov-Rogozin inequality \eqref{kol-rog-ineq} and taking $\delta_i=\delta=e^{-n\epsilon}$ we have, \begin{equation}\label{kreq} \Pr\left(\bigg|\sum\limits_{k=1}^{n}\frac{1}{z-\xi_k}\bigg|<e^{-n\epsilon}\right) \leq \frac{C}{\sqrt{\sum_{k=1}^{n}(1-Q(\frac{1}{z-\xi_k},e^{-n\epsilon}))}}. \end{equation} We shall show that $\sum_{k=1}^{n}(1-Q(\frac{1}{z-\xi_k},e^{-n\epsilon}))$ goes to $\infty$. Observe that, \begin{align} Q\left(\frac{1}{z-\xi_k},e^{-n\epsilon}\right) =& \sup\limits_{\alpha \in \mathbb{C}}\Pr\left( \bigg|\frac{1}{z-\xi_k}-\alpha\bigg|<e^{-n\epsilon} \right) \leq \frac{1}{2}, \end{align} whenever $|\frac{1}{z-a_k}-\frac{1}{z-b_k}| > 2e^{-n\epsilon}$. Define $S_n=\{k\leq n:|\frac{1}{z-a_k}-\frac{1}{z-b_k}| > 2e^{-n\epsilon}\}$. Notice that if $a_k \neq b_k$, then there is $N_k$ such that whenever $n>N_k$, we have $k\in S_n$. Because $a_k\neq b_k$ for infinitely many $k$, $|S_n|$ increases to infinity as $n \rightarrow \infty$. The denominator on the right hand side of \eqref{kreq} is at least $\sqrt{\frac{|S_n|}{2}}$. Therefore $\Pr\left(\bigg|\sum\limits_{k=1}^{n}\frac{1}{z-\xi_k}\bigg|<e^{-n\epsilon}\right)\leq \frac{C\sqrt{2}}{\sqrt{|S_n|}}\rightarrow 0$, as $n \rightarrow \infty$. This completes the proof of the lemma. \end{proof} It remains to prove the tightness for the sequence $\{\int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log^2|L_n(z)|dm(z)\}_{n\geq1}$ and will be proved in the following lemma. \begin{lemma}\label{tight} Let $L_n(z):=\sum\limits_{k=1}^{n}\frac{1}{z-\xi_k}$, where $\xi_ks$ are as in the Theorem \ref{thm1}. Then, for any $r>0$, the sequence $\{\int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log^2|L_n(z)|dm(z)\}_{n\geq1}$ is tight. \end{lemma} \begin{proof} We will first decompose $\log|L_n(z)|$ into its positive and negative parts and analyze them separately. Let $\log|L_n(z)|=\log_+|L_n(z)|-\log_-|L_n(z)|$. Then, \[\int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log^2|L_n(z)|dm(z)= \int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log_+^2|L_n(z)|dm(z)+\int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log^2_-|L_n(z)|dm(z).\] Using \eqref{logplussum}, we get, \begin{align} \int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log_+^2|L_n(z)|dm(z) & = \int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log_+^2\bigg|\sum\limits_{k=1}^{n}\frac{1}{z-\xi_k}\bigg|dm(z),\\ & \leq \int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\left(\sum\limits_{k=1}^{n}\log_+\bigg|\frac{1}{z-\xi_k}\bigg|+\log(n)\right)^2dm(z). \end{align} Using the Cauchy-Schwarz inequality $(a_1+a_2+\dots+a_n)^2\leq n(a_1^2+a_2^2+\dots+a_n^2)$ for the above, we get, \begin{align} \int_{r\text{$\mathbb{D}$}}\frac{1}{n^2}\log_+^2|L_n(z)|dm(z) & \leq \int_{\text{$\mathbb{D}$}_r}\frac{n+1}{n^2}\left(\sum\limits_{k=1}^{n}\log_+^2\bigg|\frac{1}{z-\xi_k}\bigg|+\log^2(n)\right)dm(z),\\ & = \frac{n+1}{n^2}\sum\limits_{k=1}^{n}\int_{\text{$\mathbb{D}$}_r}\log_-^2|z-\xi_k|dm(z) + \frac{n+1}{n^2}\log^2(n)\pi r^2.\label{eqn:lemma:tight:1} \end{align} Because Lebesgue measure on complex plane is translation invariant, we have $$\int_{\text{$\mathbb{D}$}_r}\log_-^2|z-\xi|dm(z)=\int_{\xi+\text{$\mathbb{D}$}_r}\log_-^2|z|dm(z)\leq\int_{\text{$\mathbb{D}$}_1}\log_-^2|z|dm(z)<\infty.$$ Therefore $\sup\limits_{\xi\in \text{$\mathbb{C}$}}\int_{K}\log^2|z-\xi|dm(z)<\infty$ for any compact set $K \subset \mathbb{C}$ it can be seen that each of the terms in the final expression \eqref{eqn:lemma:tight:1} are bounded. Hence the sequence $\{\int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log_+^2|L_n(z)|dm(z)\}_{n\geq1}$ is bounded. We will now show that the sequence $\{\int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log_-^2|L_n(z)|dm(z)\}_{n\geq1}$ is bounded. Let $P_n(z)=\prod\limits_{k=1}^{n}(z-\xi_k)$ and $P_n'(z)=n\prod\limits_{k=1}^{n-1}(z-\eta_{k}^{(n)})$. Applying inequality \eqref{logminusprod} and Cauchy-Schwarz inequality we get, \begin{align} \int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log_-^2|L_n(z)|dm(z) & = \int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log_-^2\bigg|\frac{P_n'(z)}{P_n(z)}\bigg|dm(z),\\ & \leq \int_{\text{$\mathbb{D}$}_r}\frac{2}{n^2}\log_-^2|P_n'(z)|dm(z)+ \int_{\text{$\mathbb{D}$}_r}\frac{2}{n^2}\log_-^2\bigg|\frac{1}{P_n(z)}\bigg|dm(z). \end{align} Again applying inequalities \eqref{logminusprod}, \eqref{logplusprod}, \eqref{logplussum} and Cauchy-Schwarz inequality to the above we obtain, \begin{align} \int_{\text{$\mathbb{D}$}_r}\frac{1}{n^2}\log_-^2 & |L_n(z)|dm(z)\\ & \leq \int_{\text{$\mathbb{D}$}_r}\frac{2}{n^2}\left(\sum\limits_{k=1}^{n-1}\log_-|z-\eta_{k}^{(n)}|\right)^2dm(z) + \int_{\text{$\mathbb{D}$}_r}\frac{2}{n^2}\left(\sum_{k=1}^{n}\log_+|z-\xi_k|\right)^2dm(z),\\ & \leq \frac{2}{n}\sum\limits_{k=1}^{n-1}\int_{\text{$\mathbb{D}$}_r}\log_-^2|z-\eta_{k}^{(n)}|dm(z) \label{eqn:lemma:tight1}\\&+ 2\int_{\text{$\mathbb{D}$}_r}\left(\log(2)+\log_+|z|+\frac{1}{n}\sum\limits_{k=1}^{n}\log_+|\xi_k|\right)^2dm(z). \label{eqn:lemma:tight2} \end{align} From the hypothesis, we have that both the sequences $\{a_n\}_{n\geq1}$ and $\{b_n\}_{n\geq 1}$ are log-Ces\'{a}ro bounded, which in turn implies that $\{\xi_n\}_{n\geq1}$ is also log-Ces\'{a}ro bounded, almost surely. Hence the integrand in \eqref{eqn:lemma:tight2} is bounded uniformly in $n$. Therefore \eqref{eqn:lemma:tight2} is bounded uniformly in $n$. Using the fact that $\sup\limits_{\xi\in \text{$\mathbb{C}$}}\int_{K}\log^2|z-\xi|dm(z)<\infty$ , we get \eqref{eqn:lemma:tight1} is bounded uniformly in $n$. From the above facts we get that the sequence $\left\{\frac{1}{n^2}\int_{\text{$\mathbb{D}$}_r}\log^2|L_n(z)|\right\}_{n\geq1}$ is tight. \end{proof} Lemmas \ref{momentbound}, \ref{kolmogorov-rogozin}, \ref{tight} show that the statements \eqref{A1}, \eqref{A2} and \eqref{A3} are satisfied. Hence the Theorem \ref{thm1} is proved. \section{Proof of Theorem \ref{thm2}} We will prove the theorem when $\omega=0$ i.e, $0$ is not a limit point of the sequence $\{z_n\}_{n\geq1}$. For other cases we can translate all the points by $\omega$ and apply the theorem. We will first prove a general lemma for sequences of numbers which will later be used in proving the subsequent lemmas. \begin{lemma}\label{liminf} Given any sequence $\{z_k\}_{k\geq1}$, where $z_k \in \mathbb{C}$, $\liminf\limits_{n \rightarrow \infty}\left(\inf\limits_{|z|=r}|z-z_n|^{\frac{1}{n}}\right) \geq 1$ for Lebesgue a.e. $r \in \mathbb{R^+}$ w.r.t Lebesgue measure. \end{lemma} \begin{proof} Fix $\epsilon > 0$ and let $A_n = \{r>0:\inf\limits_{|z|=r}|z-z_n|< (1-\epsilon)^{n} \}$. Let $m$ denote the Lebesgue measure on the complex plane. Then, \begin{align} m\left(\left\{r>0 :\liminf\limits_{n \rightarrow \infty}(\inf\limits_{|z|=r}|z-z_n|^\frac{1}{n}) \leq (1-\epsilon)\right\}\right) & = m\left(\limsup\limits_{n \rightarrow \infty}A_n\right) \\ & \leq \lim\limits_{k \rightarrow \infty}m\left(\mathop{\cup}_{n \geq k} A_n\right) \end{align} If $r \in A_k, $ then from the definition of $A_k$ we have that $r \in [|z_k|-(1-\epsilon)^k,|z_k|+(1-\epsilon)^k] $. Hence we get, \begin{align} m&\left(\left\{r>0 :\liminf\limits_{n \rightarrow \infty}(\inf\limits_{|z|=r}|z-z_n|^\frac{1}{n}) \leq (1-\epsilon)\right\}\right)\\ & \leq \lim\limits_{k\rightarrow \infty } \sum_{n=k}^{\infty}m\left(\left\{r:|z_n|-(1-\epsilon)^{n} \leq r \leq |z_n|+(1-\epsilon)^n\right\}\right)\\ & \leq \lim\limits_{k \rightarrow \infty } \sum_{n=k}^{\infty} 2(1-\epsilon)^n = 0 \end{align} The above is true for every $\epsilon >0$, therefore $\liminf\limits_{n \rightarrow \infty}\left(\inf\limits_{|z|=r}|z-z_n|^{\frac{1}{n}}\right) \geq 1$ outside an exceptional set $E\subset\mathbb{R}^\text{$+$}$ whose Lebesgue measure is $0$. \end{proof} Define the set $F=\{z:\liminf\limits_{n \rightarrow \infty}|z-z_n|^{\frac{1}{n}} < 1\}$. Because $0$ is not a limit point of $\{z_n\}_{n\geq1}$, we have $\liminf\limits_{n\rightarrow\infty}|z_n|^\frac{1}{n}\geq1$. Hence $0\notin F$. For $|z|=r$, we have \[\liminf\limits_{n \rightarrow \infty}|z-z_n|^\frac{1}{n}\geq\liminf\limits_{n \rightarrow \infty}\left(\inf\limits_{|z|=r}|z-z_n|^{\frac{1}{n}}\right).\] Hence $F\subseteq \{z:|z|=r, r\in E\}$ and by invoking Fubini's theorem we get $m( \{z:|z|=r, r\in E\})=0$. From the above two observations it follows that $m(F)=0$. The following lemma shows that the hypothesis of the Theorem \ref{thm2} implies \eqref{A1}. \begin{lemma}\label{momentthm2} Let $L_n(z)$ be as in the Theorem \ref{thm2}. Then for any $\epsilon>0$, and Lebesgue a.e. $z \in \mathbb{C}$, $$\limsup\limits_{n \rightarrow \infty}\frac{1}{n}\log|L_n(z)|< \epsilon$$ almost surely. \end{lemma} \begin{proof} From the hypothesis, Lemma \ref{liminf} and Using Markov's inequality we get $$\sum_{n=1}^{\infty}\Pr\left(\sup\limits_{|z|=r}\big|\frac{a_n}{z-z_n}\big|>e^{n\epsilon}\right) \leq \sum_{n=1}^{\infty}\sup\limits_{|z|=r}\frac{\ee{\text{$|a_n|$}}}{|z-z_n|e^{n\epsilon}}.$$ Denoting $t_n(r)=\sup\limits_{|z|=r}\bigg|\frac{1}{z-z_n}\bigg|$ we have \begin{equation} \sum_{n=1}^{\infty}\sup\limits_{|z|=r}\frac{\ee{\text{$|a_n|$}}}{|z-z_n|e^{n\epsilon}}=\sum_{n=1}^{\infty}\frac{\ee{\text{$|a_n|$}}}{e^{n\epsilon}}t_n(r). \label{eqn:power_series} \end{equation} Because $a_n$s are i.i.d. random variables, $\ee{\text{$|a_n|$}}=\ee{\text{$|a_1|$}}$. Using the root test for the convergence of sequences and the Lemma \ref{liminf}, it follows that the right hand side of \eqref{eqn:power_series} is convergent for Lebesgue a.e. $r \in (0,\infty)$. Invoking Borel-Cantelli lemma we can say that $\sup\limits_{|z|=r} \frac{|a_n|}{|z-z_n|}>e^{n\epsilon}$ only for finitely many times. From here we get $|L_n(z)| \leq M_\epsilon + ne^{n\epsilon}$, where $M_\epsilon$ is a finite random number which is obtained by bounding the finite number of terms for which $\sup\limits_{|z|=r} \frac{|a_n|}{|z-z_n|}>e^{n\epsilon}$ is satisfied. Therefore we get that $\limsup\limits_{n \rightarrow \infty}\frac{1}{n}\log|L_n(z)|< \epsilon$ almost surely. \end{proof} Notice that we have proved a stronger version of the Lemma \ref{momentthm2}. We will state this as a remark which will be used further lemmas. \begin{remark}\label{one} Define $M_n(R):=\sup\limits_{|z|=R}|L_n(z)|$. Then for any $\epsilon>0$, we have $$\limsup\limits_{n \rightarrow \infty}\frac{1}{n}\log M_n(R)< \epsilon$$ for almost every $R>0$. \end{remark} For proving a similar result for the lower bound of $\log|L_n(z)|$ and establish \eqref{A2}, we need the Kolmogorov-Rogozin inequality \ref{KR-Inequality} which was stated at the beginning of this chapter. \begin{lemma}\label{krthm2} Let $L_n(z)$ be as in Theorem \ref{thm2}. Then for any $\epsilon>0$, and Lebesgue a.e. $z \in \mathbb{C}$ $$\lim\limits_{n \rightarrow \infty}\Pr\left(\frac{1}{n}\log|L_n(z)|<-\epsilon\right)=0.$$ \end{lemma} \begin{proof} Fix $z \in \mathbb{C}$ which is not in the exceptional set $F$. Let $z_{i_1},z_{i_2},\dots z_{i_{l_n}}$ be the points in $K_z$ from the set $\{z_1,z_2,\dots,z_n\}$. From the definition of concentration function and the fact that the concentration function $ Q(X_1+X_2+\dots+X_n,\delta)$ is decreasing in $n$ we get, \begin{align} \Pr\left(|L_n(z)|\leq e^{-n\epsilon}\right) & \leq Q\left(\sum_{i=1}^{n}\frac{a_i}{z-z_i},e^{-n\epsilon}\right), \\ &\leq Q\left(\sum_{k=1}^{l_n}\frac{a_{i_k}}{z-z_{i_k}},e^{-n\epsilon}\right). \end{align} The random variables $\frac{a_{i_k}}{z-z_{i_k}}$s are independent. Hence we can apply Kolmogorov-Rogozin inequality to get, \begin{align} \Pr\left(|L_n(z)|\leq e^{-n\epsilon}\right) & \leq C_\epsilon\left\{\sum_{i=1}^{l_n}\left(1-Q\bigg(\frac{a_{i_k}}{z-z_{i_k}},e^{-n\epsilon}\bigg)\right)\right\}^{-\frac{1}{2}}. \end{align} Because $|z-z_{i_k}|\leq d(z,K_z)+diam(K_z)$, from above we get, \begin{align} \Pr\left(|L_n(z)|\leq e^{-n\epsilon}\right) & \leq C_\epsilon \left\{\sum_{i=1}^{l_n}\left(1-Q\left(a_{i_k},\left(d(z,K_z)+diam(K_z)\right)e^{-n\epsilon}\right)\right) \right\}^{-\frac{1}{2}}\label{eqn:lemma5.4.8:1} \end{align} Because $a_{i_k}$s are non-degenerate i.i.d random variables and $l_n\rightarrow\infty$, the right hand side of \eqref{eqn:lemma5.4.8:1} converges to $0$ as $n\rightarrow\infty$. Hence the lemma is proved. \end{proof} It remains to show that the hypothesis of Theorem \ref{thm2} implies \eqref{A3}. Fix $R>r$. The idea here is to write the function $\log|L_n(z)|$ for $z \in \text{$\mathbb{D}$}_r$ as an integral on the boundary of a larger disk $\text{$\mathbb{D}$}_R$ and bound the integral uniformly on the disk $\text{$\mathbb{D}$}_r.$ This is facilitated by Poisson-Jensen's formula for meromorphic functions. The Poisson-Jensen's formula is stated below. Let $\alpha_1, \alpha_2, \dots \alpha_k$ and $\beta_1,\beta_2, \dots \beta_\ell$ be the zeros and poles of a meromorphic function $f$ in $\text{$\mathbb{D}$}_R$. Then \begin{align} \log|f(z)| = \frac{1}{2\pi}\int_{0}^{2\pi}\Re\bigg(\frac{Re^{i\theta}+z}{Re^{i\theta}-z}\bigg)\log|f(Re^{i\theta})|d\theta - \sum_{m=1}^{k}\log\bigg|\frac{R^{2}-\overline{\alpha}_jz}{R(z-\alpha_j)}\bigg|\\ + \sum_{m=1}^{l}\log\bigg|\frac{R^{2}-\overline{\beta}_jz}{R(z-\beta_j)}\bigg| \end{align} The following lemma \ref{tighttwo} gives an estimate of the boundary integral obtained in the Poisson-Jensen's formula when applied for the function $\log|L_n(z)|$ at $z=0$. Define \[\text{$\mathcal{I}$}\text{$_n(z,R)$}:=\frac{1}{2\pi}\int_{0}^{2\pi}\Re\bigg(\dfrac{Re^{i\theta}+z}{Re^{i\theta}-z}\bigg)\log|L_n(Re^{i\theta})|d\theta.\] \begin{lemma}\label{tighttwo} There is a constant $c_2>0$ such that \begin{align} \lim\limits_{n \rightarrow \infty}\Pr\left( \dfrac{1}{n}\text{$\mathcal{I}$}(0,R)\leq-c_2\right) = 0. \end{align} \end{lemma} \begin{proof} From Poisson-Jensen's formula at $0$ we get, \begin{equation} \dfrac{1}{n}\text{$\mathcal{I}$}\text{$_n(0;R) = \dfrac{1}{n}\log|L_n(0)| +\dfrac{1}{n}\sum_{m=1}^{k}\log\bigg|\dfrac{z_{i_m}}{R}\bigg| - \dfrac{1}{n}\sum_{m=1}^{l}\log\bigg|\dfrac{\alpha_{i_m}}{R}\bigg|$},\label{eqn:lemma5.4.9:1} \end{equation} where $z_{i_m}s$ and $\alpha_{i_m}s$ are zeros and critical points respectively of $P_n(z)$ in the disk $\text{$\mathbb{D}$}_R$. Because $0$ is not a limit point of $\{z_1,z_2, \dots\}$, $\left\{\dfrac{1}{n}\sum_{m=1}^{k}\log\big|\dfrac{z_{i_m}}{R}\big|\right\}_{n\geq1}$ is a sequence of negative numbers bounded from below. $\left\{\dfrac{1}{n}\sum_{m=1}^{l}\log\bigg|\dfrac{\alpha_{i_m}}{R}\bigg|\right\}_{n\geq1}$ is also a sequence of negative numbers. Therefore the last two terms in the right hand side of \eqref{eqn:lemma5.4.9:1} are bounded below. Because $0$ is not in exceptional set $F$, from Lemma \ref{krthm2} we have that the sequence $\lim\limits_{n\rightarrow\infty}\Pr\left(\frac{1}{n}\log|L_n(z)|<-1\right)=0$ is bounded from below. Therefore there exists $C_1$ such that $$\lim\limits_{n\rightarrow\infty}\Pr\left(\frac{1}{n}\log|L_n(z)|<-1 \text{ and }\dfrac{1}{n}\sum_{m=1}^{k}\log\bigg|\dfrac{z_{i_m}}{R}\bigg| - \dfrac{1}{n}\sum_{m=1}^{l}\log\bigg|\dfrac{\alpha_{i_m}}{R} <-C_1\right)=0.$$ Choosing $c_2=C_1+1$ the statement of lemma is established. \end{proof} Using above lemma \ref{tight} and exploiting formula of Poisson kernel for disk we will now obtain an uniform bound for the corresponding integral $\text{$\mathcal{I}$}_n(z,R)$. \begin{lemma}\label{tight3} There is a constant $b>0$ such that for any $z \in \text{$\mathbb{D}$}\text{$_r$}$ \begin{equation} \lim\limits_{n\rightarrow \infty}\Pr\left(\dfrac{1}{n}\text{$\mathcal{I}$}_n(z,R)\leq -b\right)=0. \end{equation} \end{lemma} \begin{proof} We will decompose the function $\log|L_n(z)|$ into its positive and negative components. Let $\log|L_n(z)|=\log_+|L_n(z)|-\log_-|L_n(z)|$, where $\log_+|L_n(z)|$ and $\log_-|L_n(z)|$ are positive. Using this we can write, \begin{align} 2\pi\text{$\mathcal{I}$}\text{$_n(z)$} = & \int_{0}^{2\pi}\log|L_n(Re^{i\theta})|\Re\bigg(\dfrac{Re^{i\theta}+z}{Re^{i\theta}-z}\bigg)d\theta, \\ = & \int_{0}^{2\pi}\log_+|L_n(Re^{i\theta})|\Re\bigg(\dfrac{Re^{i\theta}+z}{Re^{i\theta}-z}\bigg)d\theta -\int_{0}^{2\pi}\log_-|L_n(Re^{i\theta})|\Re\bigg(\dfrac{Re^{i\theta}+z}{Re^{i\theta}-z}\bigg)d\theta. \end{align} We can find constants $C_3$ and $C_4$ such that for any $z \in \text{$\mathbb{D}$}\text{$_r$}$, $0 < C_3 \leq \Re\bigg(\dfrac{Re^{i\theta}+z}{Re^{i\theta}-z}\bigg) \leq C_4 < \infty$ is satisfied. Therefore, \begin{align} 2\pi\text{$\mathcal{I}$}\text{$_n(z)$} \geq & C_3\int_{0}^{2\pi}\log_+|L_n(Re^{i\theta})|d\theta - C_4\int_{0}^{2\pi}\log_-|L_n(Re^{i\theta})|^-d\theta,\\ \geq & 2\pi C_3\text{$\mathcal{I}$}\textbf{$_n(0)-2\pi(C_4-C_3) M_n(R)$}.\label{eqn:lemma5.4.10:1} \end{align} From the Remark \ref{one} and Lemma \ref{tighttwo} we get \begin{equation} \lim\limits_{n \rightarrow \infty}\Pr\left( \dfrac{1}{n}\text{$\mathcal{I}$}\text{$_n(0)\leq -c \text{ or } \dfrac{1}{n}M_n(R) > 1 $}\right) = \textbf{$0$} \label{eqn:lemma5.4.10:2} \end{equation} The proof is completed from above \eqref{eqn:lemma5.4.10:2} and \eqref{eqn:lemma5.4.10:1} and by choosing $b=2\pi(cC_3+C_4-C_3)$. \end{proof} To complete the argument we now need to control the other terms in Poisson-Jensen's formula. It is shown in the forthcoming expressions. Let $\xi_{i_m}s$ and $\beta_{i_m}s$ be the poles and zeros of $L_n(z)$ in $\text{$\mathbb{D}$}_R$ and $k,l(\leq n)$ are the number of zeros and poles of $L_n(z)$ respectively in $\text{$\mathbb{D}$}_R$. Now applying Poisson-Jensen's formula to $L_n(z)$ we have, \begin{align} \frac{1}{n^{2}}\int_{\text{$\mathbb{D}$}\text{$_r$}}^{}&\log^{2}|L_n(z)|dm(z) \\ &=\frac{1}{n^{2}}\int_{\text{$\mathbb{D}$}\text{$_r$}}\left(\text{$\mathcal{I}$}\textbf{$_n(z)+\sum\limits_{m=1}^{k}\log\biggl|\frac{R(z-\beta_{i_m})}{R^2-\overline{\beta}_{i_m}z}\biggr|+\sum\limits_{m=1}^{l}\log\biggl|\frac{R(z-\xi_{i_m})}{R^2-\overline{\xi}_{i_m}z}\biggr|$}\right)^2dm(z) \end{align} Invoking a case of Cauchy-Schwarz inequality $(a_1+a_2+\dots+a_n)^2\leq n(a_1^2+a_2^2+\dots+a_n^2)$ repeatedly we get, \begin{align} \int_{\text{$\mathbb{D}$}\text{$_r$}}^{}\dfrac{1}{n^{2}}&\log^{2}|L_n(z)|dm(z) \\ & \leq \dfrac{3}{n^{2}}\int_{\text{$\mathbb{D}$}\text{$_r$}}^{}|\text{$\mathcal{I}$}_n(z)|^2dm(z) + \dfrac{3}{n^{2}}\int_{\text{ $\mathbb{D}$}_r}^{}\left(\sum_{m=1}^{k}\log\bigg|\dfrac{R(z-\beta_{i_m})}{R^{2}-\overline{\beta}_{i_m}z}\bigg|\right)^{2}dm(z)\\ &+\dfrac{3}{n^{2}}\int_{\mathbb{D}\text{$_r$}}^{}\left(\sum_{m=1}\log\bigg|\dfrac{R(z-\xi_{i_m})}{R^{2}-\overline{\xi}_{i_m}z}\bigg|\right)^{2}dm(z),\\ &\leq \int_{\mathbb{D}\text{$_r$}}^{}\dfrac{3}{n^{2}}|\text{$\mathcal{I}$}_n(z)|^2dm(z) + \dfrac{3k}{n^{2}}\sum_{m=1}^{k}\int_{\mathbb{D}\text{$_r$}}^{}\log^{2}\bigg|\dfrac{R(z-\beta_{i_m})}{R^{2}-\overline{\beta}_{i_m}z}\bigg|dm(z)\\ &+\dfrac{3l}{n^{2}}\sum_{m=1}^{l}\int_{\mathbb{D}\text{$_r$}}^{}\log^{2}\bigg|\dfrac{R(z-\xi_{i_m})}{R^{2}-\overline{\xi}_{i_m}z}\bigg|dm(z). \end{align} For $z \in\text{ $\mathbb{D}$}_r$, we have $|R^2-\overline{\beta}_{i_m}z|\geq R(R-r)$. Applying this inequality in the above we get, \begin{align} \int_{\mathbb{D}\text{$_r$}}^{}\dfrac{1}{n^{2}}\log^{2}|L_n(z)|dm(z) & \leq \int_{\mathbb{D}\text{$_r$}}^{}\dfrac{3}{n^{2}}|\text{$\mathcal{I}$}_n(z)|^2dm(z) + \dfrac{3k}{n^{2}}\sum_{m=1}^{k}\int_{\mathbb{D}\text{$_r$}}^{}\log^{2}\bigg|\dfrac{z-\beta_{i_m}}{R-r}\bigg|dm(z)\\ &+\dfrac{3l}{n^{2}}\sum_{m=1}^{l}\int_{\mathbb{D}\text{$_r$}}^{}\log^{2}\bigg|\dfrac{z-\xi_{i_m}}{R-r}\bigg|dm(z). \label{eqn:poisson-jensen:1} \end{align} From the Lemmas \ref{tighttwo} and \ref{tight3}, the corresponding sequence $\frac{3}{n^{2}}\int_{\mathbb{D}\text{$_r$}}^{}|\text{$\mathcal{I}$}$$_n(z)|^2dm(z)$ is tight. The function $\log^{2}|z|$ is an integrable function on any bounded set in $\mathbb{C}$. Combining these facts and above inequality \eqref{eqn:poisson-jensen:1} we have that the sequences \linebreak $\left\{\int_{\mathbb{D}\text{$_r$}}^{}\frac{1}{n^{2}}\log^{2}|L_n(z)|dm(z)\right\}_{n\geq1}$ are tight. Hence the hypothesis of the Theorem \ref{thm2} implies \eqref{A3}. Therefore the proof of the theorem is complete. \chapter{Probability that products of real random matrices have all eigenvalues real tend to $1$.} \label{ch:realeigenvalues3} \section{Background} In this chapter we consider products of real random matrices with fixed size. In \cite{arul}, Arul Lakshminarayan observed an interesting phenomenon in products of Ginibre matrices. He considered products of i.i.d Ginibre matrices with real Gaussian entries. Let $p_n^{(k)}$ be the probability that product of $n$ such matrices have all real eigenvalues. Using numerical simulations he computed $p_n^{(k)}$. Based on the observations, he conjectured that $p^{(k)}_n$ increases to $1$ with the size of the product. Peter Forrester, in \cite{forrester}, considered the case of $k \times k$ Ginibre matrices with real Gaussian entries. He gave a formula for $p_n^{(k)}$. From that formula he deduced that this probability increases to $1$ exponentially. We state a generalization of the conjecture stated in \cite{arul}. \begin{conjecture}\label{con1}Let $X_1,X_2, \dots X_n$ be i.i.d. matrices of size $k \times k$, whose entries are i.i.d. real random variables distributed according to probability measure $\mu$ and $A_n=X_1X_2\dots X_n$. Then, $$\lim\limits_{n\rightarrow \infty}\Pr(A_n\text{ has all real eigenvalues})=1.$$ \end{conjecture} \section{Results} We prove the conjecture for the special case when the probability measure $\mu$ has an atom i.e., there is $x \in \mathbb{R}$ such that $\mu(\{x\})>0$. The proof for this case is based on a simple observation that rank of product of matrices is at most the minimum of the ranks of the individual matrices. In the given scenario, each individual matrix will be of rank at most $1$ with non zero probability. If a real matrix has rank at most $1$, then it has all real eigenvalues (they are $0$ and the trace of the matrix). The following theorem formalizes the result. \begin{theorem} Let $X_1,X_2, \dots X_n$ be i.i.d. matrices of size $k \times k$, whose entries are i.i.d. real random variables distributed according to an atomic probability measure $\mu$ and $A_n=X_1X_2\dots X_n$. Then, \[ \lim\limits_{n\rightarrow\infty}\Pr(A_n \text{ has all real real eigenvalues})=1. \] \end{theorem} \begin{proof} Let $x$ be an atom of measure $\mu$. Then $X_j$ has rank at most $1$, with probability at least $\mu(\{x\})^{k^2}$. All the matrices $X_j$s are independent of each other. Therefore, \begin{align} \Pr(A_n \text{ has rank at most } 1 ) & \geq \Pr(\text{at least one of }X_1,X_2,\dots,X_n \text{ has rank at most }1),\\ & \geq 1-(1-\mu(\{x\})^{k^2})^n.\label{eqn:chapter4:lemma1:1} \end{align} We know that real matrices with rank at most $1$, have all eigenvalues real. Hence, $$ \Pr (A_n \text{ has all real eigenvalues}) \geq \Pr(A_n \text{ has rank at most }1). $$ Hence from above and \eqref{eqn:chapter4:lemma1:1} we have, $\lim\limits_{n\rightarrow\infty}\Pr(A_n \text{ has all real eigenvalues})=1.$ \end{proof} We make the following observation about $2 \times 2$ real matrices. It says that if the rows or columns of a real random matrix are exchangeable then that matrix has both real eigenvalues with probability at least $\frac{1}{2}$. Later this will be applied to product of $2 \times 2$ real Ginibre matrices, which gives us the probability that the product of $2 \times 2$ real Ginibre matrices is at least $\frac{1}{2}$. \begin{lemma}\label{halfbound} Let $M = \left[\begin{smallmatrix} a&b\\ c&d \end{smallmatrix}\right]$, where $(a,b)$ and $(c,d)$ are real exchangeable random variables. Then, $$ \Pr(M \text{ has both real eigenvalues}) \geq \frac{1}{2}. $$ \end{lemma} \begin{proof} The characteristic polynomial of the matrix $M$ is $P_M(x)=x^2-(a+d)x+(ad-bc)$. The matrix $M$ has all real eigenvalues if the discriminant of the characteristic polynomial, $$(a+d)^2-4(ad-bc) \geq 0.$$ Hence the probability that $M$ has both real eigen values is, $$\Pr((a+d)^2-4(ad-bc) \geq 0).$$ Because $(a,c)$ and $(b,d)$ are exchangeable we have, $$ \Pr((a+d)^2-4(ad-bc) \geq 0)=\Pr((b+c)^2-4(bc-bd) \geq 0). $$ Therefore, \begin{align} \Pr(M \text{ has both} &\text{ real eigenvalues})\\ &= \frac{1}{2}(\Pr((a+d)^2-4(ad-bc) \geq 0)+\Pr((b+c)^2-4(bc-bd) \geq 0)),\\ & \geq \frac{1}{2}\Pr((a+d)^2-4(ad-bc) \geq 0 \text{ or }(b+c)^2-4(bc-bd) \geq 0). \label{eqn:chapter4:lemma2:1} \end{align} Because $(a+d)^2-4(ad-bc)+(b+c)^2-4(bc-bd)\geq 0$, at least one of $(a+d)^2-4(ad-bc)$ and $(b+c)^2-4(bc-bd)$ is non-negative. Therefore, $$ \Pr((a+d)^2-4(ad-bc) \geq 0 \text{ or }(b+c)^2-4(bc-bd) \geq 0)=1. $$ Combining above and \eqref{eqn:chapter4:lemma2:1} we have that, $\Pr(M \text{ has both real eigenvalues}) \geq \frac{1}{2}.$ \end{proof} Using Lemma \ref{halfbound}, in the case of $2 \times 2$ Ginibre matrices we can obtain that the probability of the products having all real eigenvalues to be at least $\frac{1}{2}$. \begin{corollary} Let $X_1,X_2, \dots X_n$ are i.i.d. matrices of size $2 \times 2$ whose entries are i.i.d real random variables distributed according to probability measure $\mu$ and $A_n=X_1X_2\dots X_n$. Then, $$ \Pr(A_n\text{ has all real eigenvalues})\geq \frac{1}{2}. $$ \end{corollary} \begin{proof} To satisfy the hypothesis of Lemma \ref{halfbound}, it is enough to show that the rows of the matrix $A_n$ are exchangeable. It can be noticed that the matrices $X_1$ and $\left[\begin{smallmatrix} 0 & 1\\ 1 & 0 \end{smallmatrix}\right]X_1$ are identically distributed, so are $A_n$ and $\left[\begin{smallmatrix} 0 & 1\\ 1 & 0 \end{smallmatrix}\right]A_n$. Hence the rows of $A_n$ are exchangeable.\end{proof} \chapter{Declaration} \vspace{0.5in} \noindent I hereby declare that the work reported in this thesis is entirely original and has been carried out by me under the supervision of Prof.~Manjunath Krishnapur at the Department of Mathematics, Indian Institute of Science, Bangalore. I further declare that this work has not been the basis for the award of any degree, diploma, fellowship, associateship or similar title of any University or Institution.\\ \vspace*{1in} \noindent $\begin{array}{lcr} \textrm{Tulasi Ram Reddy A} & \hspace*{1.95in} &~ \\ \textrm{S. R. No. 6910-110-101-08085} & \hspace*{1.95in} &~ \\ \textrm{Indian Institute of Science} & \hspace*{1.95in} & ~ \\ \textrm{Bangalore} & \hspace*{1.95in} & ~ \\ ~ & \hspace*{1.95in} & ~ \\ ~ & \hspace*{1.95in} & ~ \\ ~ & \hspace*{1.95in} & ~ \\ ~& \hspace*{1.95in} & \textrm{Prof.~Manjunath Krishnapur} \\ ~& \hspace*{1.95in} & \textrm{(Research advisor)} \end{array}$ \frontmatter \clearpage \thispagestyle{empty} \par\vspace*{.35\textheight}{\centering TO \\[2em] \large\it My Parents\\ and\\ \large\it Teachers \par} \chapter{Acknowledgements} For an outsider a doctoral thesis might appear as a solitary endeavour. But several people (including virtual communities) have contributed in various forms to this thesis. I make an attempt here to thank the people who have contributed to the same. The set of people I have acknowledged here is a subset of the many people who have helped me in this endeavour and is far from complete. I am deeply indebted to my thesis advisor Manjunath Krishnapur. It is a privilege to be his student. To say the least he has played several roles from being my thesis advisor to a great companion. I believe that I have also acquired some of his inexhaustible enthusiasm towards mathematics, which is indeed very contagious. My interest in the study of random polynomials surged after discussions with Zakhar Kabluchko during the Trondheim spring school 2013, along with several conversations with Manjunath. I have immensely benefited by the courses offered in the Mathematics department at IISc. The many thought provoking talks and seminars held at IISc, ISI, ICTS and TIFR-CAM have also been helpful. Probabilists in Bangalore deserve a special mention for their efforts in organising various academic activities in the city. They, along with many probabilists visiting Bangalore, have certainly made this city very lively and thereby stimulating my research activity at various levels. I was introduced to mathematics at the Indian Statistical Institute, Kolkata. Courses taught by BV Rao, late SC Bagchi, Arup Bose, Gopal Basak, Arnab Chakraborthy and Probal Chaudhuri among others created first impressions on mathematics and probability in me. Arni Srinivasa Rao encouraged me to pursue research in mathematics and has been great support since. From my high school teachers I recall K Subba Rao and M Radhakrishna who constantly encouraged me to do mathematics. I take this opportunity to thank various organizations for providing their support generously. I thank NBHM for providing me travel grant to attend the School on Random matrices and Growth models in 2013 in Trieste, Italy. Apart from this, I was also supported by NBHM during my first year of Ph.D program. I thank CSIR for supporting me with the SPM fellowship. I am grateful to KVPY which provided me a fellowship during my undergraduate days. These fellowships have helped me make choices without any confusion at various junctures in my life. My friends at IISc have bestowed me with several moments which can be cherished forever. This wouldn't have been possible without people like Arpan, Divakaran, Jaikrishnan, Kartick, Nanda, Pranav, Prathamesh, Rajeev, Sayani and Vikram. Kartick, Manjunath and Nanda read through this thesis carefully before pointing out many mistakes. They also suggested many invaluable changes. A special thanks to them. I would also like to thank them along with Koushik Saha for having good discussions in the subject. I thank the office staff in the Mathematics department and the many workers at IISc for ensuring an environment which facilitated a very smooth stay for me in the past five years. Many of my endeavours have posed several challenges for my parents at times. Regardless, they have endorsed my decisions and backed me unflinchingly all along. No less was the patience and love unveiled by my sister and brother-in-law. My grandfather, who probably is my first friend ever, used to present creative answers to my questions during our rounds in the fields. He indeed deserves a special mention here. My aunts, uncles and cousins have made my visits to home very eventful and something to crave for. Friends were constantly in touch with me, despite no efforts from my side in reaching to them. I am sure all of them will be happy seeing this thesis. \chapter{Abstract} In the first part of this thesis, we study critical points of random polynomials. We choose two deterministic sequences of complex numbers, whose empirical measures converge to the same probability measure in complex plane. We make a sequence of polynomials whose zeros are chosen from either of sequences at random. We show that the limiting empirical measure of zeros and critical points agree for these polynomials. As a consequence we show that when we randomly perturb the zeros of a deterministic sequence of polynomials, the limiting empirical measures of zeros and critical points agree. This result can be interpreted as an extension of earlier results where randomness is reduced. Pemantle and Rivin initiated the study of critical points of random polynomials. Kabluchko proved the result considering the zeros to be i.i.d. random variables.\\ In the second part we deal with the spectrum of products of Ginibre matrices. Exact eigenvalue density is known for a very few matrix ensembles. For the known ones they often lead to determinantal point process. Let $X_1,X_2,\dots, X_k$ be i.i.d Ginibre matrices of size $n \times n$ whose entries are standard complex Gaussian random variables. We derive eigenvalue density for matrices of the form $X_1^{\epsilon_1}X_2^{\epsilon_2}\dots X_k^{\epsilon_k}$, where $\epsilon_i=\pm1$ for $i=1,2,\dots,k$. We show that the eigenvalues form a determinantal point process. The case where $k=2$, $\epsilon_1+\epsilon_2=0$ was derived earlier by Krishnapur. In the case where $\epsilon_i=1$ for $i=1,2,\dots,n$ was derived by Akemann and Burda. These two known cases can be obtained as special cases of our result. \tableofcontents \mainmatter \include{introduction1} \include{criticalpoints5} \include{proofs6} \include{matching7} \include{ginibreproducts2} \include{realeigenvalues4} \backmatter
{ "timestamp": "2016-02-18T02:05:51", "yymm": "1602", "arxiv_id": "1602.05298", "language": "en", "url": "https://arxiv.org/abs/1602.05298", "abstract": "In the first part we study critical points of random polynomials. We choose two deterministic sequences of complex numbers,whose empirical measures converge to the same probability measure in complex plane. We make a sequence of polynomials whose zeros are chosen from either of sequences at random. We show that the limiting empirical measure of zeros and critical points agree for these polynomials. As a consequence we show that when we randomly perturb the zeros of a deterministic sequence of polynomials, the limiting empirical measures of zeros and critical points agree. This result can be interpreted as an extension of earlier results where randomness is reduced. Pemantle and Rivin initiated the study of critical points of random polynomials. Kabluchko proved the result considering the zeros to be i.i.d. random variables.In the second part we deal with the spectrum of products of Ginibre matrices. Exact eigenvalue density is known for a very few matrix ensembles. For the known ones they often lead to determinantal point process. Let $X_1,X_2,...,X_k$ be i.i.d matrices of size $n \\times n$ whose entries are independent complex Gaussian random variables. We derive the eigenvalue density for matrices of the form $Y_1.Y_2....Y_n$, where each $Y_i = X_i$ or $X_i^{-1}$. We show that the eigenvalues form a determinantal point process. The case where $k=2$, $Y_1=X_1,Y_2=X_2^{-1}$ was derived earlier by Krishnapur. The case where $Y_i =X_i$ for all $i=1,2,...,n$, was derived by Akemann and Burda. These two known cases can be obtained as special cases of our result.", "subjects": "Probability (math.PR); Mathematical Physics (math-ph); Complex Variables (math.CV)", "title": "On critical points of random polynomials and spectrum of certain products of random matrices", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9850429116504952, "lm_q2_score": 0.8128673246376009, "lm_q1q2_score": 0.8007091962465707 }
https://arxiv.org/abs/1603.08651
Parkable convex sets and finite-dimensional Hilbert spaces
A subset of a convex body $B$ containing the origin in a Euclidean space is {\it parkable in $B$} if it can be translated inside $B$ in such a manner that the translate the origin. We provide characterizations of ellipsoids and of centrally symmetric convex bodies in Euclidean spaces of dimension $\ge 3$ based on the notion of parkability, answering several questions posed by G. Bergman.The techniques used, which are based on characterizations of Hilbert spaces among finite-dimensional Banach spaces in terms of their lattices of subspaces and algebras of endomorphisms, also apply to improve a result of W. Blaschke characterizing ellipsoids in terms of boundaries of illumination.
\section{#1}\setcounter{lemma}{0}} \title{Parkable convex sets and finite-dimensional Hilbert spaces} \author{Alexandru Chirvasitu\footnote{University of Washington, \url{chirva@uw.edu}}} \begin{document} \date{} \maketitle \begin{abstract} A subset of a convex body $B$ containing the origin in a Euclidean space is {\it parkable in $B$} if it can be translated inside $B$ in such a manner that the translate the origin. We provide characterizations of ellipsoids and of centrally symmetric convex bodies in Euclidean spaces of dimension $\ge 3$ based on the notion of parkability, answering several questions posed by G. Bergman. The techniques used, which are based on characterizations of Hilbert spaces among finite-dimensional Banach spaces in terms of their lattices of subspaces and algebras of endomorphisms, also apply to improve a result of W. Blaschke characterizing ellipsoids in terms of boundaries of illumination. \end{abstract} \noindent {\em Key words: convex body, Blaschke, centrally symmetric, ellipsoid, parkable set, Hilbert space, Banach space} \vspace{.5cm} \noindent{MSC 2010: 52A20, 52A21, 47L10, 46C15} \section*{Introduction} The present paper answers several questions raised in \cite{berg} regarding convex bodies in euclidean spaces. The setup is based on the following notion; it is introduced by G. Bergman in the process of studying ``efficient'' embeddings of metric spaces into other metric spaces. All convex sets are understood to be subsets of some ambient Euclidean space $\bR^n$. \begin{definition'}\label{def.parkable} Let $C\subseteq B$ be convex sets with $0\in B$. $C$ is \define{parkable in} $B$ if some translate of $C$ is still contained in $B$ and contains $0$. \end{definition'} The questions in \cite{berg} referred to above have to do with characterizing particularly nice convex bodies in $\bR^n$ (i.e. centrally symmetric or ellipsoids) by means of parkability. First recall \begin{definition'} Let $C\subseteq \bR^n$ be a convex subset. A \define{center of symmetry} for $C$ is a point $p\in C$ such that \begin{equation*} C=2p-C:=\{2p-x\ |\ x\in C\}. \end{equation*} $C$ \define{has a center of symmetry} if a center of symmetry exists. $C$ is \define{centrally symmetric} if $0\in C$ is a center of symmetry for it. \end{definition'} \cite[Question 32]{berg} then reads as follows. \begin{question'}\label{qu.sym} Let $C$ be a compact convex subset of $\bR^n$ for some $n>2$, with the property that for every centrally symmetric compact convex subset $B\subset \bR^n$ containing some translate of $C$, that translate is parkable in $B$. Is it true that $C$ must have a center of symmetry? \end{question'} One of the main results is \begin{theorem'}\label{th.sym} The answer to \Cref{qu.sym} is affirmative. \mbox{}\hfill\ensuremath{\blacksquare} \end{theorem'} A more elaborate result has to do with recognizing ellipsoids by means of parkability. Recall \begin{definition'} A \define{convex body} is a compact convex set with non-empty interior which contains $0$. \end{definition'} The starting point for the discussion that follows is the following result from \cite{berg}. \begin{proposition'}\label{pr.ell} Let $B\subset \bR^n$ be a convex body for some $n>2$. Then, the following properties are successively weaker. \begin{enumerate} \renewcommand{\labelenumi}{(\roman{enumi})} \item $B$ is an ellipsoid centered at $0$. \item $B$ is centrally symmetric and its intersection with every hyperplane has a center of symmetry (provided it is not empty). \item Every closed convex subset of $B$ is parkable in $B$. \end{enumerate} \end{proposition'} It is then natural to ask \cite[Question 31]{berg}: \begin{question'}\label{qu.ell} Is either of the implications from \Cref{pr.ell} reversible? \end{question'} In this context, the second main result is \begin{theorem'}\label{th.ell} The three properties in \Cref{pr.ell} are equivalent to one another. \mbox{}\hfill\ensuremath{\blacksquare} \end{theorem'} The rest of this introduction contains a few remarks about the proofs. \Cref{th.sym,th.ell} do not directly imply one another as such, but they are nevertheless interlinked through the methods used in their proofs. The same auxiliary lemma about plane compact convex sets, for instance, leads both to \Cref{th.sym} and to the fact that condition (iii) in \Cref{pr.ell} implies that $B$ is centrally symmetric (in other words, the partial implication (iii) $\Rightarrow$ (ii)). Once we show that (iii) implies central symmetry, we can assume $B$ to be centrally symmetric throughout the rest of the proof of \Cref{th.ell}. As such, it is the unit ball of a unique Banach space structure $(\bR^n,\|\cdot\|)$ on $\bR^n$, and now functional-analytic techniques and results can be brought to bear. More specifically, using the same auxiliary result referred to in passing above we first prove \begin{proposition'}\label{pr.aux} Let $B$ be a convex centrally symmetric body $B$ satisfying condition (iii) from \Cref{pr.ell}. Then, for every linear hyperplane $H\subset \bR^n$, the non-empty intersections \begin{equation*} (H+x)\cap B,\quad x\in \bR^n \end{equation*} have centers of symmetry. Moreover, these centers are collinear. \mbox{}\hfill\ensuremath{\blacksquare} \end{proposition'} Associating to $L$ the unique line through the origin containing the centers of symmetry from the statement of \Cref{pr.aux} extends, it turns out, to an inclusion-reversing involution on the lattice of subspaces of the Banach space $(\bR^n,\|\cdot\|)$ referred to above. This, together with a lattice-theoretic characterization of Hilbert spaces among Banach spaces due to Kakutani et al. leads to the conclusion that $(\bR^n,\|\cdot\|)$ is a Hilbert space. But this is equivalent to its unit ball being an ellipsoid, and the conclusion follows. There are other results that might be of independent interest, such as an improvement of a result of Blaschke on characterizations of ellipsoids by means of light rays (\cite{bla}). \section{Preliminaries}\label{se.prel} \subsection{Convex geometry}\label{subse.conv} Our main reference on the topic will be \cite{tomo}, to which we refer the reader for basic terminology. We say that two convex bodies in a Euclidean space $\bR^n$ are {\it mutual translates} if one is the image of the other through some translation of $\bR^n$. The following result will make an appearance several times in the sequel; the reader can consult e.g. \cite{Rya15} and the references therein for background on the result (which is also \cite[Theorem 3.1.3]{tomo}). \begin{theorem}\label{th.transl} Let $K$ and $L$ be two compact convex subsets of $\bR^n$ for $n\ge 3$, and $2\le m\le {n-1}$ a positive integer. If the projections of $K$ and $L$ on every $m$-dimensional linear subspace of $\bR^n$ are mutual translates, then so are $K$ and $L$. \mbox{}\hfill\ensuremath{\blacksquare} \end{theorem} As a consequence, we get (\cite[Corollary 3.1.5]{tomo}) \begin{corollary}\label{cor.centr} Let $m$ and $n$ be positive integers as in \Cref{th.transl}, and $B\subset \bR^n$ a compact convex subset. Then, $B$ has a center of symmetry if and only if its projection on every $m$-dimensional linear subspace of $\bR^n$ does. \end{corollary} \begin{proof} Having a center of symmetry is equivalent to $B$ and $-B$ being translates; we can now simply apply \Cref{th.transl} to this pair of convex sets. \end{proof} \subsection{Banach and Hilbert spaces}\label{subse.ban} We will assume some basics on Banach and Hilbert spaces and bounded operators thereon, as covered e.g. in the introductory sections of \cite[Chapters 1 and 12]{rud} or in the first three chapters of \cite{con}. Given a centrally symmetric convex body $B=-B$ in $\bR^n$, we can associate to it the unique Banach space structure on $\bR^n$ making $B$ the unit ball: the norm of $x\in \bR^n$ is defined to be \begin{equation*} \|x\| = \|x\|_B = \inf\{r\ge 0\ |\ x\in rB\}. \end{equation*} The correspondence $B\mapsto \|\cdot\|_B$ is a bijection between centrally symmetric convex bodies and Banach space structures on $\bR^n$; the inverse map associates to a Banach space structure $(\bR^n,\|\cdot\|)$ its unit ball. As \Cref{th.ell} above suggests, one of our main goals will be characterizing ellipsoids among convex bodies. In terms of the bijection $B\leftrightarrow \|\cdot\|_B$ the ellipsoids correspond to the {\it Hilbert space} structures on $\bR^n$, i.e. those Banach space structures whose underlying norm $\|\cdot\|$ arises from an inner product $\langle-,-\rangle$ via the usual formula \begin{equation*} \|x\|^2 = \langle x,x\rangle,\ \forall x\in \bR^n. \end{equation*} For this reason, it will be important to have at our disposal results that allow for the recognition of Hilbert spaces among Banach spaces. One such tool is \cite[Theorem 1.1]{istr} (or rather a variant thereof, with real Banach spaces instead of complex ones): \begin{theorem}\label{th.istr} Let $(\bR^n,\|\cdot\|)$ be a Banach space. The norm is induced by an inner product if and only if there exists an operation $T\mapsto T^*$ on the space $M_n(\bR)$ of endomorphisms of $\bR^n$ such that \begin{enumerate} \item $(T+S)^* = T^*+S^*$; \item $(T^*)^* = T$; \item $(TS)^* = S^*T^*$; \item $\|P\|\le 1$ if $P^2=P=P^*$, where $\|P\|$ is the norm on $M_n(\bR)$ induced by that on $\bR^n$. \end{enumerate} \mbox{}\hfill\ensuremath{\blacksquare} \end{theorem} In other words, we have a recognition criterion for Hilbert space norms in terms of involutions $T\mapsto T^*$ on their algebras of endomorphisms. \section{Universally parkable sets}\label{se.univ} The aim of this section is to prove \Cref{th.sym} above, answering \Cref{qu.sym} in the affirmative. We recall the statement, after introducing the following notion relevant to the setup of the theorem. \begin{definition}\label{def.univ_park} A compact convex set $C\subset \bR^n$ is {\it universally parkable} if for any centrally symmetric convex subset $B\subset \bR^n$ containing a translate of $C$, that translate is parkable in $B$ in the sense of \Cref{def.parkable}. \end{definition} \begin{theorem}\label{th.sym_bis} For a positive integer $n\ge 3$, every universally parkable compact convex subset $C\subset \bR^n$ has a center of symmetry. \end{theorem} We proceed through a series of auxiliary results. First off, we narrow down the class of centrally symmetric convex sets that witness the parkability of $C\subset \bR^n$. For this purpose as well as for use in the sequel, denote by $C_u$ the translate $C+u$ for $u\in \bR^n$. \begin{lemma}\label{le.Cu} A compact convex subset $C\subset \bR^n$ is universally parkable if and only if for every $u\in \bR^n$ the translate $C_u$ is parkable in the convex hull $\co(-C_u\bigcup C_u)$. \end{lemma} \begin{proof} On the one hand, if $C$ is universally parkable, every $\co(-C_u\bigcup C_u)$ is certainly a centrally symmetric convex subset of $\bR^n$ containing a translate $C_u$, and hence the latter must be parkable in the former. Conversely, every centrally symmetric compact convex $B\subset \bR^n$ containing a translate $C_u$ contains $\co(-C_u\bigcup C_u)$, so that if $C_u$ is parkable in the latter then it is parkable in the former as well. \end{proof} As a consequence, we obtain \begin{corollary}\label{cor.Cu} If $C\subset \bR^n$ is universally parkable, then so are its projections on linear subspaces of $\bR^n$. \end{corollary} \begin{proof} This follows from the fact that the property from the statement of \Cref{le.Cu} is clearly invariant under taking orthogonal projections. \end{proof} In order to fix ideas, we now specialize \Cref{th.sym_bis} to $n=3$. \begin{lemma} If \Cref{th.sym_bis} holds for $n=3$ then it holds in general. \end{lemma} \begin{proof} Starting with a universally parkable $C\subset \bR^n$ for some $n\ge 3$, \Cref{cor.Cu} ensures that its projections on all $3$-dimensional linear subspaces of $\bR^n$ are universally parkable. If \Cref{th.sym_bis} holds for $\bR^3$ we can conclude that these projections all have centers of symmetry, which according to \Cref{cor.centr} means that so does $C$. \end{proof} This now allows us to focus on the case $n=3$. Let $\bS\subset \bR^3$ be an origin-centered sphere whose radius is large enough to ensure that no translates $C_u$ (and hence no reflections $-C_u$) contain $0$ as $u$ ranges over $\bS$. The following lemma shows that when $u\in \bS$ there is a certain rigidity in how one may translate $C_u$ inside $\co(-C_u\bigcup C_u)$ so as to engulf $0$. \begin{lemma}\label{le.unique} Let $C\subset \bR^3$ be as in the statement of \Cref{th.sym_bis} For any $u\in \bS$, the direction of a vector $0\ne v\in \bR^3$ such that \begin{equation*} 0\in C_u+v\subset \co(-C_u\bigcup C_u) \end{equation*} is uniquely determined. \end{lemma} \begin{proof} Let $u\in \bS$ and $0\ne v\in \bR^3$ be vectors as in the statement, and let $H\subset \bR^3$ be a $2$-dimensional linear such that the projection $C_u|H$ does not contain $0\in H$ and is not collinear with the origin (i.e. is not a segment whose supporting line contains $0$). Note that such planes $H$ form an open subset of the Grassmannian $\cG(3,2)$ (because their defining property is open), and by our choice of radius for $\bS$ this subset of $\cG(3,2)$ is non-empty. In the plane $H$ the part of the boundary $\partial \co(-C_u|H\bigcup C_u|H)$ that is not contained in the boundaries of $C_u|H$ and $-C_u|H$ consists of two open segments forming two opposite edges of a parallelogram centered at $0$: \begin{equation*} \begin{tikzpicture}[auto,baseline=(current bounding box.center),scale=1] \node (c) at (2.2,.8) {$\scriptstyle C_u|H$}; \node (-c) at (-2.3,-.8) {$\scriptstyle -C_u|H$}; \node (aux) at (-1.125,-.425) {}; \draw (0,0) node[circle, inner sep=1pt, fill=black, label={right:{$\scriptstyle 0$}}] (0) {}; \draw[-] (2,1.5) .. controls (2.5,1.7) and (3.3,1.2) .. (3,.7); \draw[-] (3,.7) .. controls (2.7,.2) and (2.625,.25) .. (2.5,.2); \draw[-] (2.5,.2) to[bend left=9] (1.5,.4); \draw[-] (1.5,.4) .. controls (1,.6) and (1.75,1.4) .. (2,1.5); \draw[-] (-2,-1.5) .. controls (-2.5,-1.7) and (-3.3,-1.2) .. (-3,-.7); \draw[-] (-3,-.7) .. controls (-2.7,-.2) and (-2.625,-.25) .. (-2.5,-.2); \draw[-] (-2.5,-.2) to[bend left=9] (-1.5,-.4); \draw[-] (-1.5,-.4) .. controls (-1,-.6) and (-1.75,-1.4) .. (-2,-1.5); \draw[-] (2,1.5) -- (-2.5,-.2); \draw[-] (-2,-1.5) -- (2.5,.2); \draw[->] (0) to node[pos=.3,auto,swap] {$\scriptstyle w$} (aux); \draw (-.25,.65) node[circle, inner sep=1pt, fill=black] () {}; \draw (.25,-.65) node[circle, inner sep=1pt, fill=black] () {}; \draw[-] (.5,-1.3) -- (-.5,1.3); \draw (-.5,1.3) node[circle, inner sep=0pt, fill=black, label={right:{$\scriptstyle L$}}] () {}; \end{tikzpicture} \end{equation*} More formally, as the diagram illustrates, the two segments in question are the tangents to the boundary $\partial\co(-C_u|H\bigcup C_u|H)$ at the points where a hyperplane (i.e. line) $L$ that separates $C_u|H$ and $-C_u|H$ intersects said boundary. Denoting by $w$ the orthogonal projection of $v$ on $H$, it follows from the above remark about the boundary of $\co(-C_u|H\bigcup C_u|H)$ that $w$ must be parallel to the two segments and point away from $C_u|H$. This determines the direction of $w\in H$ uniquely, and since $w$ was the orthogonal projection of $v$ on any one of the elements of an open subset of $\cG(3,2)$, we deduce the desired uniqueness of the direction of $v$. \end{proof} \begin{remark}\label{re.unique} Note that the proof of \Cref{le.unique} actually shows more than the statement claims: it shows that the unique direction in which $C_u$ can be translated without leaving $\co(-C_u\bigcup C_u)$ is that of $\varphi(u)$. \end{remark} This gives us, for every $u\in \bS$, a unique unit vector $\varphi(u)=\frac v{\|v\|}\in \bS_1$ (unit sphere centered at $0$) such that translation in its direction will position $C_u$ so that it contains $0$. The map $\varphi$ is in fact very well-behaved. Recall that an {\it odd} map between two spheres is one that sends antipodes to antipodes. With this in hand, we have \begin{lemma}\label{le.cont} The map $\varphi:\bS\to \bS_1$ deduced from \Cref{le.unique} as in the discussion above is continuous and odd. \end{lemma} \begin{proof} The fact that $\varphi$ is odd is immediate: if $-\varphi(u)$ belongs to $C_u$ then $\varphi(u)$ belongs to $-C_u$, meaning that translation of $-C_u$ by $-\varphi(u)$ will position the former so that it contains $0$. As for the continuity of $\varphi$, note first that because of our choice of $\bS$ (such that no $C_u$ contains the origin) there is a positive lower bound on the length of vectors $v$ such that $0\in C_u+v$ as $u$ ranges over $\bS$. In conclusion, there is some $t>0$ such that for all $u\in \bS$ we have \begin{equation}\label{eq:cont} C_u+t \varphi(u)\in \co(-C_u\bigcup C_u). \end{equation} The condition \begin{equation*} C_u+tv\in \co(-C_u\bigcup C_u) \end{equation*} is closed in $(u,v)\in \bS\times \bS_1$, and together with \Cref{re.unique} above, \Cref{eq:cont} shows that the set of pairs $(u,v)$ that satisfy it is precisely the graph of $\varphi:\bS\to \bS_1$. Since we now know that the graph of the map $\varphi$ between compact spaces is closed, we can conclude that it is continuous. \end{proof} \begin{corollary}\label{cor.onto} The map $\varphi:\bS\to \bS_1$ from \Cref{le.cont} is onto. \end{corollary} \begin{proof} Indeed, \Cref{le.cont} shows that it is a continuous odd map between $2$-spheres. By one version of the Borsuk-Ulam theorem (e.g. the main result of \cite{dold}) it follows that the map is not nullhomotopic and hence, because the complement of a point in the $2$-sphere is contractible, must be onto. \end{proof} \begin{lemma}\label{le.tough} For every $u\in \bS$, the orthogonal projection of $C_u$ on $H=\varphi(u)^\perp$ is centrally symmetric. \end{lemma} \begin{proof} We have to show that the projections of $C_u$ and $-C_u$ on $H$ coincide. If they do not, then there is a point $x$ on the boundary $\partial(C_u|H)$ that does not belong to $-C_u|H$. The line in $\bR^3$ through $x$ and orthogonal to $H$ (and hence parallel to $\varphi(u)$) intersects $-C_u\bigcup C_u$ along a segment (possibly degenerate, i.e. a single point) contained in the boundary $\partial C_u$. \begin{equation*} \begin{tikzpicture}[auto,baseline=(current bounding box.center),scale=1] \node (c) at (-.7,1.4) {$\scriptstyle C_u$}; \node (-c) at (.8,-1.5) {$\scriptstyle -C_u$}; \draw (0,0) node[circle, inner sep=1pt, fill=black, label={45:{$\scriptstyle 0$}}] () {}; \draw[-] (-2,2) -- (-2,1.5); \draw[-] (-2,2) .. controls (-2,3) and (.5,2) .. (.5,1); \draw[-] (-2,1.5) .. controls (-2,1) and (.5,0) .. (.5,1); \draw[-] (2,-2) -- (2,-1.5); \draw[-] (2,-2) .. controls (2,-3) and (-.5,-2) .. (-.5,-1); \draw[-] (2,-1.5) .. controls (2,-1) and (-.5,0) .. (-.5,-1); \draw[-] (-3,0) -- (3,0); \node (H) at (2.8,.2) {$\scriptstyle H$}; \draw[->] (2,2) to node[pos=.5] {$\scriptstyle \phi(u)$} (2,1); \draw[-] (-2,2.5) -- (-2,-.3); \draw (-2,0) node[circle, inner sep=1pt, fill=black, label={135:{$\scriptstyle x$}}] () {}; \end{tikzpicture} \end{equation*} Translation by $\varphi(u)$ will move one of the endpoints of the segment outside of $\co(-C_u\bigcup C_u)$, contradicting \begin{equation*} C_u+\varphi(u)\subset \co(-C_u\bigcup C_u). \end{equation*} It follows from the contradiction that $-C_u|H=C_u|H$, as desired. \end{proof} We are now ready to complete the proof of the main result of this section. \begin{proof_of_symbis} \Cref{le.tough,cor.onto} show that the orthogonal projection of $C$ on every hyperplane in $\bR^3$ has a center of symmetry and hence, according to \Cref{cor.centr}, so does $C$. \end{proof_of_symbis} \section{Parkability and finite-dimensional Banach spaces}\label{se.park} Here, we prove \Cref{th.ell}. Recall the statement: \begin{theorem}\label{th.ell_bis} For a convex body $B\subset \bR^n$, $n\ge 3$ the following properties are equivalent. \begin{enumerate} \renewcommand{\labelenumi}{(\roman{enumi})} \item $B$ is an ellipsoid centered at $0$. \item $B$ is centrally symmetric and its intersection with every hyperplane has a center of symmetry (provided it is not empty). \item Every closed convex subset of $B$ is parkable in $B$. \end{enumerate} \end{theorem} We already know from \Cref{pr.ell} that (i) implies (ii), which in turn implies (iii); hence, it suffices to go backwards. We once again specialize to $n=3$ in order to simplify some of the proofs and language. \begin{proposition}\label{pr.n=3} If \Cref{th.ell_bis} holds for $n=3$ then it holds in general. \end{proposition} \begin{proof} Property (iii) is clearly preserved by passing to orthogonal projections from $\bR^n$ down to its three-dimensional subspaces. If the implication (iii) $\Rightarrow$ (i) holds in $\bR^3$, then we know that all orthogonal projections of $B$ on such subspaces are ellipsoids. This implies that $B$ itself is an ellipsoid by \cite[Theorem 3.1.7]{tomo}. \end{proof} As explained in the introduction, the first priority will be to prove that (iii) implies the property of being centrally symmetric so as to be able to think of the convex body as the unit ball of a Banach space structure on $\bR^n$. The following lemma provides the first step in this direction. \begin{lemma}\label{le.2impln} If the implication (iii) $\Rightarrow$ (centrally symmetric) holds in $\bR^2$, then it holds for all $\bR^n$, $n\ge 3$. \end{lemma} \begin{proof} Let $B\subset \bR^n$, $n\ge 3$ be a convex body satisfying condition (iii) of the theorem, and suppose (iii) implies central symmetry in the plane. Clearly, (iii) is preserved upon projecting down to any plane, and hence, by assumption, all projections of $B$ on planes are centrally symmetric. The conclusion that $B$ itself is then follows from \Cref{cor.centr} (or rather a variant thereof whereby the center of symmetry is in fact the origin). \end{proof} \begin{remark}\label{re.hyp} As seen in \cite[Lemma 29]{berg}, condition (iii) of \Cref{th.ell_bis} is equivalent to the fact that every hyperplane section of $B$ is parkable. We will henceforth use this equivalence without further comment when convenient. \end{remark} \begin{proposition}\label{pr.2} Let $B\subset \bR^2$ be a convex body such that every intersection of $B$ with a line is parkable in $B$. Then, $B$ is centrally symmetric. \end{proposition} \begin{proof} Let $v\in \bR^2$ be a non-zero vector, and $H_1$ and $H_2$ the two $B$-supporting lines that are orthogonal to $v$. {\bf Claim: The convex hull of the segments $H_1\cap B$ and $H_2\cap B$ contains the origin.} Indeed, if $p_1\in H_1\cap B$ and $p_2\in H_2\cap B$ are two points on the boundary of $B$, then by our hypothesis some translate $\overline{p_1p_2}+w\subset B$ contains the origin. Because $H_i$ are supporting lines the translates $p_1+w$ and $p_2+w$ belong to $H_1$ and $H_2$ respectively (i.e. if it is non-zero, then $w$ is parallel to $H_1$ and $H_2$); this finishes the proof of the claim. \begin{equation*} \begin{tikzpicture}[auto,baseline=(current bounding box.center),scale=.8] \draw (0,-1) node[circle, inner sep=0pt, fill=black, label={225:{$\scriptstyle H_1$}}] () {}; \draw (-1.2,.2) node[circle, inner sep=1pt, fill=black, label={225:{$\scriptstyle p_1$}}] (p1) {}; \draw (4,-2) node[circle, inner sep=0pt, fill=black, label={45:{$\scriptstyle H_2$}}] () {}; \draw (2.5,-.5) node[circle, inner sep=1pt, fill=black, label={45:{$\scriptstyle p_2$}}] (p2) {}; \draw (.3,.3) node[circle, inner sep=1pt, fill=black, label={45:{$\scriptstyle 0$}}] () {}; \draw[-] (-2,1) -- (-1,0); \draw[-] (1,1) -- (3,-1); \draw[-] (-2,1) -- (1,1); \draw[-] (-1,0) -- (3,-1); \draw[-] (-2,1) .. controls (-2.5,1.5) and (.5,1.5) .. (1,1); \draw[-] (-1,0) .. controls (-.5,-.5) and (3.5,-1.5) .. (3,-1); \draw[-] (-2.5,1.5) -- (0,-1); \draw[-] (0,2) -- (4,-2); \draw[->] (4,.5) to node[pos=.5,auto,swap] {$\scriptstyle w$} (3.5,1); \end{tikzpicture} \end{equation*} Given the claim, the conclusion now follows from the technical \Cref{le.supports} below. \end{proof} \begin{lemma}\label{le.supports} Let $B\subset \bR^2$ be a convex body. Suppose that for every non-zero $v\in \bR^2$ the two supporting lines $L_1=L_1(v)$ and $L_2=L_2(v)$ of $B$ that are orthogonal to $v$ satisfy \begin{equation*} 0\in \co((L_1\cap B)\cup(L_2\cap B)). \end{equation*} Then, $B$ is centrally symmetric. \end{lemma} \begin{proof} We have to show that $B$ and $-B$ coincide. Assuming they do not, the condition on $B$ at least ensures that the union $-B\cup B$ is convex. Indeed, otherwise for some $p\in \partial B\cap \partial(-B)$ some $B$-supporting line $L$ through $p$ would not be $-B$-supporting, and hence $v\perp L$ would violate the hypothesis. In conclusion, $B$ and $-B$ admit common parallel supporting lines $L$ and $-L$ at two points $p$ and $-p$ respectively in the intersection $\partial B\cap \partial(-B)$. We will moreover choose coordinates in $\bR^2$ so that the segment $\overline{(-p)p}$ is horizontal and $L$ is vertical. We will derive a contradiction from the assumption that the portions of $\partial B$ and $\partial(-B)$ lying in the lower half plane are distinct. \begin{equation*} \begin{tikzpicture}[auto,baseline=(current bounding box.center),scale=1] \draw (0,0) node[circle, inner sep=1pt, fill=black, label={90:{$\scriptstyle 0$}}] () {}; \draw (-1,0) node[circle, inner sep=1pt, fill=black, label={135:{$\scriptstyle -p$}}] () {}; \draw (1,0) node[circle, inner sep=1pt, fill=black, label={45:{$\scriptstyle p$}}] () {}; \draw (-1,-2) node[circle, inner sep=0pt, fill=black, label={225:{$\scriptstyle -L$}}] () {}; \draw (1,-2) node[circle, inner sep=0pt, fill=black, label={315:{$\scriptstyle L$}}] () {}; \draw (.5,-1) node[circle, inner sep=0pt, fill=black, label={100:{$\scriptstyle B$}}] () {}; \draw (-.5,-2) node[circle, inner sep=0pt, fill=black, label={85:{$\scriptstyle -B$}}] () {}; \draw[-] (-1,1) -- (-1,-2); \draw[-] (1,1) -- (1,-2); \draw[-] (-2,0) -- (2,0); \draw[-] (-1,0) .. controls (-1,-.5) and (.1,-1) .. (.5,-1); \draw[-] (.5,-1) .. controls (.9,-1) and (1,-.5) .. (1,0); \draw[-] (-1,0) .. controls (-1,-.5) and (-.9,-2) .. (-.5,-2); \draw[-] (-.5,-2) .. controls (-.1,-2) and (1,-2) .. (1,0); \end{tikzpicture} \end{equation*} We can now identify $\overline{(-p)p}$ with a closed interval $[-a,a]$, $a>0$ and the portions of $\partial B$ and $\partial(-B)$ depicted above with graphs of convex functions $f<0$ and $g<0$ respectively defined on $(-a,a)$. Moreover, the assumption $B\ne -B$ translates without loss of generality to $g(x)<f(x)$ for all $x\in (-a,a)$. In this setup, the meaning of the hypothesis is that whenever a line through the origin intersects the graphs of $f$ and $g$ at smooth points $(x,f(x))$ and $(y,f(y))$ respectively, the tangents to the two graphs at those points are parallel; equivalently, $f'(x)=g'(y)$. If, in the previous paragraph, we choose $x$ to be negative, then $y<x$. Because $g$ is convex, its derivative is non-decreasing, and hence \begin{equation*} f'(x)=g'(y)\le g'(x) \end{equation*} (always assuming we are working with points where the derivatives exist, which is the case for all but countably many of the elements in the domain $(-a,a)$). In conclusion, whenever both $f$ and $g$ are differentiable at $x\in (-a,0]$ we have $f'(x)\le g'(x)$. This, together with \begin{equation*} f(0)=\int_{-a}^0f'(x)\ \mathrm{d}x\le g(0)=\int_{-a}^0g'(x)\ \mathrm{d}x \end{equation*} contradicts our assumption that $g(0)<f(0)$. \end{proof} Using \Cref{pr.2}, we obtain \begin{corollary}\label{cor.iii_implies_sym} Let $B\subset \bR^n$, $n\ge 3$ be a convex body satisfying condition (iii) of \Cref{th.ell_bis}. Then, $B$ is centrally symmetric. \end{corollary} \begin{proof} This follows immediately from \Cref{le.2impln} via \Cref{pr.2}. \end{proof} We now know that all convex bodies satisfying the weakest condition (iii) of \Cref{th.ell_bis} are centered at the origin, and as a consequence we henceforth specialize the discussion to such bodies. Next, our goal will be to prove an $n=3$ version of \Cref{pr.aux}, announced above. \begin{proposition}\label{pr.aux_bis} Let $B$ be a convex centrally symmetric body $B\subset \bR^3$ satisfying condition (iii) from \Cref{pr.ell}. Then, for every linear hyperplane $H\subset \bR^3$, the non-empty intersections \begin{equation*} (H+x)\cap B,\quad x\in \bR^3 \end{equation*} have centers of symmetry. Moreover, these centers are collinear. \end{proposition} Before embarking on a proof, we need some preparation. As explained in the discussion following \cite[Question 31]{berg}, convex bodies $B$ with parkable hyperplane sections have the following property: for any linear hyperplane $H\subset \bR^3$ there is a line $L\subset \bR^3$ such that $B$ has a supporting translate of $L$ at every point of $H\cap \partial B$ (here, `supporting' is the natural extension to lines of the term `supporting hyperplane'; it simply means that the line intersects only the boundary of $B$). It is also explained in loc. cit. how this property is in a sense dual to one considered by Blaschke \cite[pp. 157-159]{bla} in relation to illuminating convex bodies. In this latter reference, (smooth) convex bodies are considered with the property that when illuminated with parallel light rays, the boundary curve of the illuminated region is planar. In other words, given a line $L$, the intersections of the $B$-supporting translates of $L$ with $B$ form a planar curve in $\partial B$. This is the precise opposite of the situation in the previous paragraph. Motivated by these considerations, we label the two dual properties described above. \begin{definition}\label{def.bla} Let $B\subset \bR^3$ be a convex body containing $0$ in its interior. $B$ has {\it the Blaschke property} (or {\it is Blaschke}, for short) if for any line $L$ there exists a linear hyperplane $H\subset \bR^3$ such that all points in $H\cap \partial B$ lie on some $B$-supporting translate of $L$. $B$ has {\it the dual Blaschke property} (or {\it is dual Blaschke}) if for any linear hyperplane $H\subset \bR^3$ there is a line $L\in \bR^3$ such that all points in $H\cap \partial B$ lie on some $B$-supporting translate of $L$. \end{definition} With these terms in place, the discussion preceding the definition amounts to (\cite[p. 257]{berg}) \begin{proposition}\label{pr.co_bla} A centrally symmetric convex body in $\bR^3$ satisfying condition (iii) of \Cref{th.ell_bis} is dual Blaschke. \mbox{}\hfill\ensuremath{\blacksquare} \end{proposition} Our next aim is to prove the dual version of this result. \begin{proposition}\label{pr.bla} A centrally symmetric convex body in $\bR^3$ satisfying condition (iii) of \Cref{th.ell_bis} is Blaschke. \end{proposition} We now introduce some tools necessary in the proof of \Cref{pr.bla}. Consider a convex body $B$ as in \Cref{pr.co_bla}. For every unit vector $v\in \bR^3$, denote by $H_v$ the plane orthogonal to $v$. We define the subset $\psi(v)=\psi_B(v)$ of the unit sphere to consist of all unit vectors $w$ such that \begin{equation*} w\cdot v>0 \text{ and the line } \{p+tw\ |\ t\in \bR\} \text{ is }B-\text{supporting for every }p\in \partial B\cap H_v. \end{equation*} In other words, it is the set of unit vectors that make acute angles with $v$ and pointing along the directions of those lines $L$ associated to the plane $H_v$ as in \Cref{pr.co_bla,def.bla}. \Cref{pr.co_bla} is what makes $\psi(v)$ non-empty. Note moreover that since the lines supporting $B$ at points in $\partial B\cap H$ cannot be contained in $H$ (for any plane $H$), $\psi(v)$ is closed in the unit sphere $\bS^2$. Finally, it is also {\it convex} as a subset of the sphere, in the sense that for $w,w'\in \psi(v)$ the rescaled convex combinations \begin{equation*} \frac{tw+(1-t)w'}{\|tw+(1-t)w'\|},\ t\in (0,1) \end{equation*} all belong to $\psi(v)$. All in all, we have defined a map \begin{equation*} \psi=\psi_B:\bS^2\to\text{ closed convex subsets of }\bS^2. \end{equation*} The map $\psi$ has a number of other properties that are easy to check: \begin{lemma}\label{le.odd} For a convex body $B$ as in \Cref{pr.co_bla}, the map $\psi=\psi_B$ is odd in the sense that \begin{equation*} \psi(-v)=-\psi(v) \end{equation*} and upper semicontinuous in the sense that its graph \begin{equation*} \{(v,w)\in \bS^2\times \bS^2\ |\ w\in \psi(v)\} \end{equation*} is closed in $\bS^2\times \bS^2$. \end{lemma} \begin{proof} The property of being odd is immediate from the definition of $\psi$: $v$ and $-v$ define the same plane $H_v=H_{-v}$, and hence the same set of lines supporting $B$ at the points of $\partial B\cap H_v$. The upper semicontinuity claim is similarly routine, using the observation that the set of $B$-supporting affine lines is closed in the set of all affine lines of $\bR^3$. \end{proof} \begin{lemma}\label{le.psi_onto} An odd, upper semicontinuous map \begin{equation*} \psi=:\bS^2\to\text{ closed convex subsets of }\bS^2 \end{equation*} is onto in the sense that every $w\in \bS^2$ belongs to some set $\psi(v)$, $v\in \bS^2$. \end{lemma} \begin{proof} Assume the contrary, with, say, $w\in \bS^2$ lying outside all of the sets $\psi(v)$ as $v$ ranges over $\bS^2$. Flowing along great circles through $w$ down to the equator in the plane orthogonal to $w$, we can deform $\psi$ into an odd, upper semicontinuous map \begin{equation*} \psi=:\bS^2\to\text{ closed convex subsets of }\bS^1. \end{equation*} The existence of such a map contradicts the multivalued Borsuk-Ulam theorem as stated e.g. in \cite[Corollary 2.4]{multi_BU}. \end{proof} We have now effectively proven \Cref{pr.bla} above. \begin{proof_of_bla} Using the notation introduced above, the statement amounts to showing that $\psi_B$ is onto; this is precisely what \Cref{le.psi_onto} does. \end{proof_of_bla} This provides us with the necessary tools to attack \Cref{pr.aux_bis}. \begin{proof_of_aux_bis} The statement stipulates a condition that is closed among planes in $\cG(3,2)$, so it suffices to prove that this condition holds for a dense set of planes. Specifically, we will prove it for planes $H$ with the following property: {\it The two supporting planes of $B$ that are parallel to $H$ each intersect $B$ at a single point.} Now fix such a plane $H$, and let $p$ and $q=-p$ be the two points where the two $B$-supporting translates of $H$ intersect $B$. Let also $K$ be a planar section $(H+x)\cap B$ with non-empty relative interior. Finally, fix a line $L\in H$. In the affine plane $H+x$, the convex body $K$ has two supporting lines parallel to $L$; denote these by $L_1$ and $L_2$. {\bf Claim: The convex hull of $L_i\cap K$ contains the point $r=(H+x)\cap \overline{pq}$.} Given the claim, we can finish the proof of the proposition as follows. Since the claim holds for any line $L\in H$, we can apply \Cref{le.supports} inside the affine plane $H+x$, with the point $r$ regarded as the origin, to conclude that $K$ has this point as its center of symmetry. In conclusion, we obtained the desired result: all planar sections of $B$ that are parallel to $H$ are centrally symmetric along points on the line containing $p$ and $q$. It thus remains to prove the claim; this goal will take up the rest of the proof. According to \Cref{pr.bla}, there is a plane $H'\subset \bR^3$ with the property that every point in $H'\cap \partial B$ is contained in a $B$-supported line parallel to $L$. By our choice of $H$ (so that its two $B$-supporting translates meet $B$ at single points $p$ and $q$ respectively), we have $p,q\in H'$. As a consequence of this, the endpoints of the segment $H'\cap K$ are contained in the $B$-supporting translates $L_1$ and $L_2$; this concludes the proof. \end{proof_of_aux_bis} \begin{remark}\label{re.32to31} Note that given the plane $H\subset \bR^3$, the line containing the centers of symmetry for the non-empty planar sections $(H+x)\cap B$ is uniquely determined; this provides a map from the Grassmannian of planes $\cG(3,2)$ to the Grassmannian $\cG(3,1)$ of lines in $\bR^3$. \end{remark} We now take the first step towards constructing a map $\cG(3,1)\to \cG(3,2)$ in the opposite direction to that of \Cref{re.32to31}. \begin{lemma}\label{le.aux_bis_dual} Suppose we are under the assumptions of \Cref{pr.aux_bis}, and $L$ is the line containing the centers of symmetry of $(H+x)\cap B$ for $x$ ranging over $\bR^3$. Then, the midpoint of every non-empty intersection $(L+y)\cap B$ belongs to $H$. \end{lemma} \begin{proof} \Cref{pr.aux_bis} can be restated as saying that the linear involution $T$ of $\bR^3$ that acts as the identity along $L$ and as $x\mapsto -x$ on $H$ preserves the convex body $B$. The involution $-T$ also preserves $B$, acts as the identity on $H$, and as $x\mapsto -x$ along $L$. In other words, $-T$ preserves $H$ and reverses every segment $(L+y)\cap B$; this implies the conclusion. \end{proof} We can now prove the dual version of \Cref{pr.aux_bis}. \begin{proposition}\label{pr.aux_bis_dual} Suppose we are under the hypotheses of \Cref{pr.aux_bis}. For any line $L\subset \bR^3$, the midpoints of the non-empty intersections $(L+y)\cap B$ are coplanar. \end{proposition} \begin{proof} \Cref{le.aux_bis_dual} above already proves the statement for those lines $L$ arising as images of planes $H\in \cG(3,2)$ through the map $\cG(3,2)\to \cG(3,1)$ from \Cref{re.32to31}. It is thus sufficient to show that this map is onto. In other words, we have to prove that {\it every} line $L$ contains the symmetry centers of the non-empty planar sections $(H+x)\cap B$ for some $H\in \cG(3,2)$. Consider the map $\phi=\phi_B:\bS^2\to \bS^2$ defined as follows. For a unit vector $v\in \bS^2$, let $H=H_v$ be the plane orthogonal to $v$, and take $\phi(v)$ be the unit vector making an acute angle with $v$ and pointing along the line $L$ that contains the centers of symmetry of $(H+x)\cap B$, $x\in \bR^3$. Clearly, $\phi$ is continuous and odd, in the sense that $\phi(-v)=-\phi(v)$. We can now apply \Cref{le.psi_onto} in the simpler case of ordinary (rather than multivalued) maps to conclude via Borsuk-Ulam that $\phi$ is onto. \end{proof} \Cref{pr.aux_bis,pr.aux_bis_dual} allow us to introduce the following \begin{notation}\label{not.prime} For a linear subspace $H\subset \bR^3$ define $H'$ to be \begin{itemize} \item the zero subspace $\{0\}$ if $H=\bR^3$; \item the line containing the centers of symmetry of $(H+x)\cap B$ if $H$ is a plane; \item the plane containing the midpoints of $(H+y)\cap B$ if $H$ is a line; \item $\bR^3$ if $H=\{0\}$. \end{itemize} \end{notation} The operation $H\mapsto H'$ on the subspace of $\bR^3$ has the following properties. \begin{lemma}\label{le.inv} \begin{enumerate} \renewcommand{\labelenumi}{(\arabic{enumi})} \item For any two subspaces $L,H$ of $\bR^3$ we have $L\subseteq H\Rightarrow L'\supseteq H'$; \item For every subspace $H\subseteq \bR^3$ we have $H''=H$; \item For every subspace $H\subseteq \bR^3$ we have $H\cap H'=\{0\}$. \end{enumerate} \end{lemma} \begin{proof} In all three statements, the interesting cases are those where the subspaces in question are 1- or 2-dimensional. For this reason, we treat only these cases in the proof. {\bf (1)} Here, we may as well assume that $L$ is a line contained in the plane $H$. $H'$ is by definition the line containing the symmetry centers of the sections $(H+x)\cap B$. Every such center is the midpoint of the segment $(L+x)\cap B$, and is thus on $L'$ by the definition of the latter. {\bf (2)} The identity $H=H''$ for planes is simply a restatement of \Cref{le.aux_bis_dual} above. Applying the operation $\bullet\mapsto \bullet'$ once more we get $H'=H'''$ for planes $H$, and hence $L=L''$ for lines $L$ follows from the observation that every line arises as $H'$ for some plane $H$ (this latter claim is essentially the surjectivity of the map $\phi_B$ from the proof of \Cref{pr.aux_bis_dual}). {\bf (3)} This is immediate: when $H$ is a plane, for instance, there are non-empty sections $(H+x)\cap B$ for non-zero $x\in \bR^3$, and hence there are points of the line $H'$ that are not contained in $H$. A similar argument applies when $H$ is a line (or alternatively, by part (2), in that case $H=H''$ and we can apply the previous argument to the plane $H'$). \end{proof} Properties (1), (2) and (3) in \Cref{le.inv} are, according to \cite[Theorem 1]{KM44}, precisely what is necessary in order to ensure that there is a Hilbert space structure on $\bR^3$ for which $H\mapsto H'$ is the orthogonal complement operation. Such a Hilbert space structure is given by an inner product of the form \begin{equation*} \langle v,w\rangle = v\cdot Aw \end{equation*} where $A\in M_3=M_3(\bR)$ is a positive operator and $\cdot$ is the usual dot product. Since the image of an ellipsoid through a positive operator $A:\bR^3\to \bR^3$ is again an ellipsoid, we may as well assume (for the purposes of proving \Cref{th.ell_bis}) that $A$ is the identity matrix and hence $K\mapsto K'$ is the usual orthogonal complement operation in $\bR^3$; we will do this throughout the rest of the section, in order to simplify the discussion. Now consider the usual transposition operation $T\mapsto T^t$ on $M_3$. It satisfies conditions (1), (2) and (3) of \Cref{th.istr} above. Our goal will be to apply that result to the Banach space $(\bR^3,\|\cdot\|_B)$ induced by the centrally symmetric convex body $B$ as explained in \Cref{subse.ban}. In preparation for that, note that the symmetric idempotents $P\in M_3$ (i.e. those satisfying $P^2=P=P^t$) are precisely those whose range and kernel are orthogonal. \begin{lemma}\label{le.istr_applies} If $\|\cdot\|$ denotes the norm induced on $M_3$ by $\|\cdot\|_B$, then $\|P\|^2\le 1$ for all symmetric idempotents $P\in M_3$ \end{lemma} \begin{proof} If the range of $P$ is three- or zero-dimensional then $P$ is the identity or the zero operator respecticely, so there is nothing to prove. It thus remains to prove the claim when $\dim(\mathrm{Im}(P))$ is $2$ or $1$. We tackle the former case; the latter is entirely analogous. If $H\in \cG(3,2)$ is the range of $P$, then $P$ is the orthogonal projection on $H$. Since, as explained in the discussion following \Cref{le.inv}, we are assuming that $H'=H^\perp$, the convex body $B$ (i.e. the unit ball of $(\bR^3,\|\cdot\|_B)$) is contained in the orthogonal cylinder based on $H\cap B$. But this means that if $\|x\|_B=1$ (i.e. $x\in \partial B$) then the orthogonal projection $Px$ on $H$ is contained in $B\cap H$, and hence $\|Px\|_B\le 1$. In conclusion, \begin{equation*} \|P\| = \sup_{\|x\|_B\le 1}\|Px\|_B\le 1. \end{equation*} This finishes the proof. \end{proof} This concludes the preparatory material necessary for the proof of the main result of this section. \begin{proof_of_ellbis} As mentioned after the statement of the theorem, it suffices to prove that (iii) implies that $B$ is an ellipsoid. Moreover, \Cref{pr.n=3} shows that it is enough to work with $n=3$. \Cref{cor.iii_implies_sym} ensues that $B$ is centrally symmetric, and hence can be regarded as the unit ball of a Banach space $(\bR^3,\|\cdot\|_B)$. \Cref{le.istr_applies,th.istr} now ensure that the norm $\|\cdot\|_B$ is induced by an inner product on $\bR^3$, and hence the unit ball of $\|\cdot\|_B$ is an ellipsoid, as explained in \Cref{subse.ban}. \end{proof_of_ellbis} \section{Illuminated bodies and subspace lattice involutions}\label{se.illum} The techniques employed in \Cref{se.park} will also help in improving on a characterization of ellipsoids via illumination by parallel rays due to Blaschke (\cite[pp. 157-159]{bla}). For the purpose of stating the result briefly, we introduce one more piece of terminology. \begin{definition}\label{def.wbla} A convex body $B$ in $\bR^3$ has {\it the weak Blaschke property} (or {\it is weakly Blaschke}) if for any line $L$ there exists an affine plane $H\subset \bR^3$ such that $H$ intersects the interior of $B$ and all points in $H\cap \partial B$ lie on some $B$-supporting translate of $L$. $B$ has {\it the weak dual Blaschke property} (or {\it is weakly dual Blaschke}) if for any linear plane $H\subset \bR^3$ there is a translate $H'$ of $H$ and a line $L\in \bR^3$ such that $H'$ intersects the interior of $B$ and all points in $H'\cap \partial B$ lie on some $B$-supporting translate of $L$. \end{definition} \begin{remark}\label{re.wbla} Note the difference to \Cref{def.bla}: in \Cref{def.wbla} we do not require the planes to pass through a given, fixed point (such as the origin in the case of the former definition). This justifies the adjective `weak' in the definition. In fact, it is easy to see that centrally symmetric weakly (dual) Blaschke convex bodies are automatically (dual) Blaschke. \end{remark} \begin{remark} The weak Blaschke property can be interpreted as saying that if the body is illuminated with parallel rays, then the boundary of the shaded region contains a coplanar curve; hence the title of this section. \end{remark} With this language in place, Blaschke's result referred to above states that a smooth, strictly convex convex body in $\bR^3$ that is weakly Blaschke must be an ellipsoid. In this section we remove the smoothness and strict convexity requirements: \begin{theorem}\label{th.illum} A convex $B\subset \bR^3$ that is weakly Blaschke is an ellipsoid. \end{theorem} We once more prepare the ground before giving the proof proper. First, we show that the weak Blaschke property entails its dual. \begin{lemma}\label{le.self-dual} A weakly Blaschke convex body is weakly dual Blaschke. \end{lemma} \begin{proof} This is very similar in spirit to the proof of \Cref{pr.bla} from \Cref{pr.co_bla} via \Cref{le.psi_onto}, using the same multivalued Borsuk-Ulam theorem. To any unit vector $v\in \bS^2$ associate the subset of the unit $2$-sphere consisting of those vectors $w$ with the property that $w\cdot v>0$ and $w$ is orthogonal to those planes associated as in the definition of the weak Blaschke property to the line containing $v$. This gives rise, as in the discussion following \Cref{pr.bla}, to a map \begin{equation*} \psi=:\bS^2\to\text{ closed convex subsets of }\bS^2. \end{equation*} $\psi$ can be shown to be upper semicontinuous and odd as in \Cref{le.odd}, and hence is onto according to \Cref{le.psi_onto}. This concludes the proof: we have just shown that {\it every} plane in $\bR^3$ is parallel to some affine plane associated to a line as in the definition of the weak Blaschke property. \end{proof} Next, we prove the existence of a center of symmetry. \begin{lemma}\label{le.wbla_symm} A weakly Blaschke convex body $B\subset \bR^3$ has a center of symmetry. \end{lemma} \begin{proof} It suffices to show that the group of affine symmetries of $B$ is infinite. For this purpose, let $H\subset \bR^3$ be a plane with the property that the two $B$-supporting translates of $H$ intersect $B$ at single point $p$ and $q$. We now proceed very much as in the proof of \Cref{pr.aux_bis}. We denote by $K$ a planar section $(H+x)\cap B$ of $B$ that has non-empty relative interior. Then choose a line $L\subset H$. Just as in the cited proof, we can show using the weak Blaschke property that if $L_1$ and $L_2$ are the $K$-supporting translates of $L$, then the convex hull of $L_i\cap K$ contains the point $\overline{pq}\cap K$. \Cref{le.supports} then ensures that $K$ has a center of symmetry at $\overline{pq}\cap K$. Since the $H$-parallel section $K$ of $B$ was arbitrary, we conclude that the affine involution of $\bR^3$ that preserves the points on the line $pq$ and acts as reflection across $pq\cap (H+x)$ along the plane $H+x$ preserves $B$. The above construction provides us with infinitely many affine symmetries of $B$, one for each choice of plane $H$ with the generic property specified at the beginning. \end{proof} \begin{proof_of_illum} We now know from \Cref{le.self-dual} that $B$ is both weakly Blaschke and weakly dual Blaschke. Since moreover it has a center of symmetry by \Cref{le.wbla_symm}, we can assume that $B$ is centered at $0$ and hence is both Blaschke and dual Blaschke (cf. \Cref{re.wbla}). Running through the proof of \Cref{th.ell_bis}, this is sufficient to conclude that $B$ is an ellipsoid. \end{proof_of_illum}
{ "timestamp": "2016-03-30T02:06:11", "yymm": "1603", "arxiv_id": "1603.08651", "language": "en", "url": "https://arxiv.org/abs/1603.08651", "abstract": "A subset of a convex body $B$ containing the origin in a Euclidean space is {\\it parkable in $B$} if it can be translated inside $B$ in such a manner that the translate the origin. We provide characterizations of ellipsoids and of centrally symmetric convex bodies in Euclidean spaces of dimension $\\ge 3$ based on the notion of parkability, answering several questions posed by G. Bergman.The techniques used, which are based on characterizations of Hilbert spaces among finite-dimensional Banach spaces in terms of their lattices of subspaces and algebras of endomorphisms, also apply to improve a result of W. Blaschke characterizing ellipsoids in terms of boundaries of illumination.", "subjects": "Metric Geometry (math.MG); Functional Analysis (math.FA)", "title": "Parkable convex sets and finite-dimensional Hilbert spaces", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9850429129677614, "lm_q2_score": 0.8128673155708975, "lm_q1q2_score": 0.8007091883862414 }
https://arxiv.org/abs/1806.03774
Counting subgroups of fixed order in finite abelian groups
We use recurrence relations to derive explicit formulas for counting the number of subgroups of given order (or index) in rank 3 finite abelian p-groups and use these to derive similar formulas in few cases for rank 4. As a consequence, we answer some questions by M. T$\ddot{a}$rn$\ddot{a}$uceanu in \cite{MT} and L. T$\dot{\acute{o}}$th in \cite{LT}. We also use other methods such as the method of fundamental group lattices introduced in \cite{MT} to derive a similar counting function in a special case of arbitrary rank finite abelian p-groups.
\section{Introduction} The subject of counting various kinds of subgroups of finite abelian groups has a long and rich history. For instance, in the early 1900's, Miller\cite{MI04} determined the number of cyclic subgroups of prime power order in a finite abelian p-group $G$, where p is a prime number. In about the same time, Hilton\cite{HH} found a necessary and sufficient condition for the existence of subgroups of a given order and then gave a general procedure for counting subgroups of order $p^r$ in a finite abelian p-group of order $p^n$, for positive integers $r\leq n.$ For recent work on counting subgroups of finite abelian p-groups, see \cite{GB96}, \cite{GBJW}, \cite[Sections 6 \& 7]{GLP}, \cite{HT}, \cite{AI}, \cite{MI39}, \cite{MT} and \cite{LT}. As an example, Hilton determined the total number of normal subgroups of index $p^2$ in any finite abelian p-group and noted that giving a general formula for the number of subgroups of order $p^r$ for every value of $r$ would be somewhat complicated. Progress has been made since then and there are now recursive formulas based on works of Stehling\cite{TS} and Birkhoff\cite{GB} that can compute the numbers for any $r.$ It can also be deduced using Hall's polynomials\cite{IGM} that the function that counts the number of subgroups is a polynomial in powers of $p.$ In addition, Shokuev\cite{VNS} finds an exact expression for the number of subgroups of any (possible) order of an arbitrary finite p-group. However, this requires subgroup structure of smaller rank subgroups in order to determine the expression. So far as we know, the explicit form of these counting polynomials has not been determined beyond that for rank 2 finite abelian p-groups. In \cite[Theorem 3]{GBJW01}, Bhowmik and Wu get a unimodality result on the coefficients and determine the leading coefficient of the polynomials in any rank. In this paper, we determine the explicit form of the counting polynomials in rank 3 and for some special values of $r$ in any rank. However, finding the general explicit polynomials in rank 4 or higher will be a tedious and difficult computation and we will illustrate this with examples. By the fundamental theorem on the structure of finite abelian groups, any finite abelian group is isomorphic to a direct product of prime power order cyclic groups. Therefore, the number of subgroups in the finite abelian group is the product of the numbers of the subgroups contained in these cyclic groups\cite{RS}. This reduces the problem of counting subgroups in finite abelian groups to counting subgroups in finite abelian p-groups. Let $\mathbb{Z}_{n}$ denote the cyclic group of order n. Given partitions $\lambda,\mu,\nu$ and $G$ an abelian $p$-group of type $\lambda = (a_d,\dots,a_1),$ i.e., $G \cong \mathbb{Z}/p^{a_1}\times \cdots \times \mathbb{Z}/p^{a_d},$ for some positive integers $a_1,\dots,a_d$ where $1\leq a_1\leq \dots \leq a_d,$ G has $g^\lambda_{\mu\nu}(p)$ subgroups $K$ such that $K$ has type $\mu$ and $G/K$ has type $\nu,$ where $g^\lambda_{\mu\nu}(X)\in \mathbb{Z}[X]$ is a Hall polynomial \cite{IGM}. Let \begin{equation} \label{no_of_subgrps} h_b^\lambda (p) = \sum_{|\nu| = b} \sum_{\mu}g^\lambda_{\mu\nu} (p) \end{equation} \noindent where $\nu = (\nu_l,\dots,\nu_1)$ with $\nu_1\leq \nu_2\leq \dots \leq \nu_l$ and $|\nu| = \nu_1+\dots +\nu_l.$ Then $h_b^\lambda(p) = h_b^{(a_d,\dots,a_1)}(p)$ is the number of subgroups of index $p^b$ (or equivalently of order $p^{m-b}$ where $m=\sum_{i=1}^d a_i$) in the rank $d$ abelian $p$-group $\mathbb{Z}/p^{a_1}\times \cdots \times \mathbb{Z}/p^{a_d}.$ Moreover, $h_b^\lambda(p)$ is a polynomial as it is a sum of Hall polynomials. When the finite abelian $p$-group has rank 2, the number of subgroups of order $p^b$ in $\mathbb{Z}/p^{a_1}\times \mathbb{Z}/p^{a_2}$ is given by \cite[Theorem 3.3]{MT} as follows (for an alternate proof, see the end of Section \ref{sec:convolution}): \begin{equation} \label{rank2ratfns} h_b^{(a_2,a_1)}(p)=\begin{cases} \frac{p^{b+1}-1}{p-1} & \text{if } 0\le b \leq a_1\\ \frac{p^{a_1+1}-1}{p-1} & \text{if } a_1\le b \leq a_2\\ \frac{p^{a_1+a_2-b+1}-1}{p-1} & \text{if } a_2\le b \leq a_1+a_2.\\ \end{cases} \end{equation} \noindent Except for rank 2 finite abelian p-groups and some other specific cases, there was no general direct closed formula (i.e. explicit polynomial depending on the type $\lambda$ and $b$; expressed above as a rational function in each case) that computes $h_b^\lambda(p)$ given $\lambda$ and $b.$ We determine in the next section the polynomials that give the number of subgroups of given order in rank 3 and few cases of rank 4 and one particular case for arbitrary rank finite abelian $p$-group. This answers part of a problem posed by M. T$\ddot{a}$rn$\ddot{a}$uceanu\cite[Problem 5.1]{MT}. As an application of the rational function formulas, we will prove a conjecture posed by Laszlo T${\acute{o}}$th in \cite{LT}. \section{Rank 3} G. Birkhoff shows in \cite{GB} that $h_b^{(a_d,\dots,a_1)}$ is symmetric in $b$, that is, $h_b^{(a_d,\dots,a_1)} = h_{a_1+\dots+a_d-b}^{(a_d,\dots,a_1)}.$ So, for instance, in the three cases in equation (\ref{rank2ratfns}) for rank 2, the first case is symmetric with the third case while the second case is symmetric to itself. For example, when $d=2,$ the case $0\le b \leq a_1$ is symmetric to $a_2\le b \leq a_1+a_2$ because $$0\le b \leq a_1$$ $$-a_1\le -b \leq 0$$ $$a_2\leq a_1+a_2-b \leq a_1+a_2.$$ \noindent Therefore, in order to determine the formula in the case $a_2\le b \leq a_1+a_2$ from the case when $0\le b \leq a_1$, we replace $b$ in the formula for $0\le b \leq a_1$ by $a_1+a_2-b.$ So, for instance in rank 2, it suffices to determine the compact form (as rational functions) of the polynomials in only two cases (first and second cases) and then use symmetry to deduce the remaining case (the last case). Let $\lambda=(a_3,a_2, a_1)$ and $\lambda^{'}=(a_2, a_1)$. The partition of the interval $[0,a_1+a_2+a_3]$ as in Hironaka's paper \cite{YH} leads to 10 cases for rank 3. For each of these cases, there is a polynomial $h_b^{(a_1,a_2,a_3)}(p)$ (easily described as a rational function) enumerating the number of subgroups of index $p^b$ (or order $p^{a_1+a_2+a_3-b}$) in $\mathbb{Z}/p^{a_1}\times \mathbb{Z}/p^{a_2}\times \mathbb{Z}/p^{a_3}.$ There are at least four methods of deducing the formulas. The first method is to use a lemma \cite[Lemma 2.3]{YH} that Y. Hironaka deduces from a recurrence relation of T. Stehling in \cite{TS}. This is the method used in the proof of the theorem below as her lemma shows how to express the formulas in higher rank in terms of those of lower rank which we already know and this requires the least amount of computation. The second method was suggested to the first author by Gautam Chinta (the first author's doctoral advisor) while working on a related problem on subgroup growth as such subgroup counting formulas help compute smaller rank cotype zeta functions as in \cite{CKK}. This method involves using a series of contour integrals, the residue theorem and convolution of generating series to arrive at the formulas. The third method is based on an extension of results in the second author's previous paper in \cite{PKASSS}. This will be explicated another time. The last method, called the method of fundamental group lattices, was introduced by M. T$\ddot{a}$rn$\ddot{a}$uceanu in \cite{MT} and we have been able to apply it to derive a compact formula for counting subgroups in case 1 of all ranks. We demo the second method of proof in a special case at the end of the paper. \begin{theorem} \label{rank3} For every $0\leq b \leq a_1+a_2+a_3$, the number $h_b^{(a_3,a_2,a_1)}(p)$ of all subgroups of order $p^b$ (equivalently of index $p^{a_1+a_2+a_3-b}$) in the finite abelian $p$-group $\mathbb{Z}/p^{a_1}\times \mathbb{Z}/p^{a_2}\times \mathbb{Z}/p^{a_3}$ where $1\leq a_1\leq a_2 \leq a_3,$ is given by one of the following polynomials expressed as rational functions:\\ Case 1: $0\le b \leq a_1$ \begin{equation*} h_b^{(a_3,a_2,a_1)}(p) = \frac{p^{2b+3} - p^{b+2} - p^{b+1} +1}{ (p-1)(p^2-1)}. \end{equation*} Case 2: $a_1\le b \leq a_2$ \begin{equation*} h_b^{(a_3,a_2,a_1)}(p) = \frac{p^{b+a_1+3}+p^{b+a_1+2}-p^{2a_1+2}-p^{b+2}-p^{b+1}+1}{ (p-1)(p^2-1)}. \end{equation*} Case 3: $a_2 < b \le a_3 \leq a_1 + a_2$ \begin{equation*} h_b^{(a_3,a_2,a_1)}(p) =\frac{(b-a_{2}+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-(b-a_2)p^{a_1+a_2+1}-p^{2a_1+2}-p^{b+2}-p^{b+1}+1}{(p-1)(p^{2}-1)}. \end{equation*} Case 4: $a_2 < b \le a_1 + a_2 \leq a_3$ $$h_b^{(a_3,a_2,a_1)}(p) = \frac{(b-a_{2}+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-(b-a_2)p^{a_1+a_2+1}-p^{2a_1+2}-p^{b+2}-p^{b+1}+1}{(p-1)(p^{2}-1)}.$$ Case 5: $a_1 + a_2 \le b \le a_3$ $$h_b^{(a_3,a_2,a_1)}(p) = \frac{(a_{1}+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-a_{1}p^{a_1+a_2+1}-p^{2a_1+2}-p^{a_{1}+a_{2}+2}-p^{a_{1}+a_{2}+1}+1}{(p-1)(p^{2}-1)}.$$ Case 6: $a_3 < b \leq a_{1}+a_{2} $ \begin{multline*} h_b^{(a_3,a_2,a_1)}(p) = \frac{(a_3-a_2+1)p^{a_2+a_1+3}+ 2p^{a_2+a_1+2}-(a_3-a_2-1)p^{a_2+a_1+1}-p^{a_3+a_2+a_1-b+2}}{(p-1)(p^2-1)}\\ + \frac{-p^{a_3+a_2+a_1-b+1} -p^{2a_1+2}-p^{b+2}-p^{b+1}+1}{(p-1)(p^2-1)}. \end{multline*} The remaining cases are symmetric to one of the cases above, i.e., they can be obtained when we replace $b$ by $a_1+a_2+a_3-b.$\\ Case 7: $a_1 + a_2 < a_3 < b \le a_1 + a_3$ \begin{multline*} h_b^{(a_3,a_2,a_1)}(p) = \frac{(a_1+a_3-b+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-(a_1+a_3-b)p^{a_1+a_2+1}-p^{2a_1+2}}{(p-1)(p^{2}-1)}\\ + \frac{-p^{a_1+a_2+a_3-b+2}-p^{a_1+a_2+a_3-b+1}+1}{(p-1)(p^{2}-1)}. \end{multline*} Case 8: $ a_3 < a_1 + a_2 < b \le a_1 + a_3$ \begin{multline*} h_b^{(a_3,a_2,a_1)}(p) = \frac{(a_1+a_3-b+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-(a_1+a_3-b)p^{a_1+a_2+1}-p^{2a_1+2}}{(p-1)(p^{2}-1)}\\ + \frac{-p^{a_1+a_2+a_3-b+2}-p^{a_1+a_2+a_3-b+1}+1}{(p-1)(p^{2}-1)}. \end{multline*} Case 9: $a_1 + a_3 \le b \le a_2 + a_3$ $$h_b^{(a_3,a_2,a_1)}(p) = \frac{p^{2a_1+a_2+a_3-b+3}+p^{2a_1+a_2+a_3-b+2}-p^{2a_1+2}-p^{a_1+a_2+a_3-b+2}-p^{a_1+a_2+a_3-b+1}+1}{(p-1)(p^{2}-1)}.$$ Case 10: $a_2 + a_3 \le b \le a_1 + a_2 + a_3$ $$h_b^{(a_3,a_2,a_1)}(p) = \frac{p^{2a_1+2a_2+2a_3-2b+3}-p^{a_1+a_2+a_3-b+2}-p^{a_1+a_2+a_3-b+1}+1}{(p-1)(p^{2}-1)}.$$ \end{theorem} \begin{proofof}{Theorem \ref{rank3}} To simplify notation, $h_b^{(a_3,a_2,a_1)}(p)$ will be denoted by $N_{b}(\lambda)$ as in the notation used in \cite{YH}. By \cite[Lemma 2.3]{YH}, we have the recursive formula \begin{equation} N_{b}(\lambda)=\sum_{i=0}^{b} p^{i}N_{i}(\lambda^{'}) - \sum_{|\lambda| +1-k}^{|\lambda^{'}|} p^{i}N_{i}(\lambda^{'}), 0\leq b\le |\lambda| \end{equation} \noindent where $\lambda = (a_d,\dots, a_1), \lambda^{'} = (a_{d-1},\dots, a_1), |\lambda| = a_d+\dots+a_1$ and the second summation appears only when $b > a_d.$ Thus, we compute each of the cases as follows: \\ Case 1: $0\le b \leq a_1$ \\ $N_{b}(\lambda)=\sum_{i=0}^{b} p^{i}N_{i}(\lambda^{'})=\sum_{i=0}^{b} p^{i} \frac{p^{i+1}-1}{p-1}=\sum_{i=0}^{b} \frac{p^{2i+1}-p^{i}}{p-1}=\frac{p^{2b+3}-p^{b+2}-p^{b+1}+1}{(p-1)(p^{2}-1)}.$\\ Case 2: $a_1\le b \leq a_2$ \\ $N_{b}(\lambda)=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{1}+1}^{b} p^{i}N_{i}(\lambda^{'})=\frac{p^{2a_1+3}-p^{a_1+2}-p^{a_1+1}+1}{(p-1)(p^{2}-1)}+\sum_{i=a_{1}+1}^{b} p^{i} \frac{p^{a_{1}+1}-1}{p-1}$\\ $=\frac{p^{b+a_1+3}+p^{b+a_{1}+2}-p^{2a_1+2}-p^{b+2}-p^{b+1}+1}{(p-1)(p^{2}-1)}.$\\ Case 3: $a_2 < b \le a_3 \leq a_1 + a_2$\\ $N_{b}(\lambda)=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{1}+1}^{a_{2}} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{2}+1}^{b} p^{i}N_{i}(\lambda^{'})$\\ $=\frac{p^{a_{2}+a_1+3}+p^{a_{2}+a_{1}+2}-p^{2a_1+2}-p^{a_{2}+2}-p^{a_{2}+1}+1}{(p-1)(p^{2}-1)}+\sum_{i=a_{2}+1}^{b} p^{i}\frac{p^{a_{1}+a_{2}+1-i}-1}{p-1}$\\ $=\frac{(b-a_{2}+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-(b-a_2)p^{a_1+a_2+1}-p^{2a_1+2}-p^{b+2}-p^{b+1}+1}{(p-1)(p^{2}-1)}.$\\ Case 4: $a_2 < b \le a_1 + a_2 \leq a_3$ \\ $N_{b}(\lambda)=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{1}+1}^{a_{2}} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{2}+1}^{b} p^{i}N_{i}(\lambda^{'})$\\ $=\frac{p^{a_{2}+a_1+3}+p^{a_{2}+a_{1}+2}-p^{2a_1+2}-p^{a_{2}+2}-p^{a_{2}+1}+1}{(p-1)(p^{2}-1)}+\sum_{i=a_{2}+1}^{b} p^{i}\frac{p^{a_{1}+a_{2}+1-i}-1}{p-1}$\\ $=\frac{(b-a_{2}+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-(b-a_2)p^{a_1+a_2+1}-p^{2a_1+2}-p^{b+2}-p^{b+1}+1}{(p-1)(p^{2}-1)}.$\\ Case 5: $a_1 + a_2 \le b \le a_3$ \\ $N_{b}(\lambda)=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{1}+1}^{a_{2}} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{2}+1}^{a_{1}+a_{2}} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{1}+a_{2}+1}^{b} p^{i}N_{i}(\lambda^{'})\\ - \sum_{i=a_{1}+a_{2}+a_{3}+1-b}^{a_{1}+a_{2}}p^{i}N_{i}(\lambda^{'})$\\ $=\frac{(a_{1}+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-a_{1}p^{a_1+a_2+1}-p^{2a_1+2}-p^{a_{1}+a_{2}+2}-p^{a_{1}+a_{2}+1}+1}{(p-1)(p^{2}-1)}.$\\ Case 6: $a_3 < b \leq a_{1}+a_{2} $ \\ $N_{b}(\lambda)=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{1}+1}^{a_{2}} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{2}+1}^{a_{3}} p^{i}N_{i}(\lambda^{'})+\sum_{i=(+1}^{b} p^{i}N_{i}(\lambda^{'})\\ - \sum_{i=a_{1}+a_{2}+a3+1-b}^{a_{1}+a_{2}}p^{i}N_{i}(\lambda^{'})$\\ $=\frac{(a_{3}-a_{2}+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-(a_{3}-a_2)p^{a_1+a_2+1}-p^{2a_1+2}-p^{a_{3}+2}-p^{a_{3}+1}+1}{(p-1)(p^{2}-1)}+\sum_{i=a_{3}+1}^{b} p^{i}\frac{p^{a_{1}+a_{2}+1-i}-1}{p-1}$\\ $-\sum_{i=a_{1}+a_{2}+a_{3}+1-b}^{a_{1}+a_{2}}p^{i}\frac{p^{a_{1}+a_{2}+1-i}-1}{p-1}$\\ $=\frac{(a_3-a_2+1)p^{a_2+a_1+3}+ 2p^{a_2+a_1+2}-(a_3-a_2-1)p^{a_2+a_1+1}-p^{a_3+a_2+a_1-b+2}-p^{a_3+a_2+a_1-b+1}}{(p-1)(p^2-1)} \\ + \frac{-p^{2a_1+2}-p^{b+2}-p^{b+1}+1}{(p-1)(p^2-1)}. $\\ The remaining cases are symmetric to one of the cases above, i.e., they can be obtained when we replace $b$ by $a_1+a_2+a_3-b.$\\ Case 7: $a_1 + a_2 < a_3 < b \le a_1 + a_3$ Replace $b$ in case 4 by $a_1+a_2+a_3-b.$\\ $N_{b}(\lambda)= \frac{(a_1+a_3-b+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-(a_1+a_3-b)p^{a_1+a_2+1}-p^{2a_1+2}-p^{a_1+a_2+a_3-b+2}-p^{a_1+a_2+a_3-b+1}+1}{(p-1)(p^{2}-1)}.$\\ Case 8: $ a_3 < a_1 + a_2 < b \le a_1 + a_3$ Replace $b$ in case 3 by $a_1+a_2+a_3-b.$\\ $N_{b}(\lambda) = \frac{(a_1+a_3-b+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-(a_1+a_3-b)p^{a_1+a_2+1}-p^{2a_1+2}-p^{a_1+a_2+a_3-b+2}-p^{a_1+a_2+a_3-b+1}+1}{(p-1)(p^{2}-1)}.$\\ Case 9: $a_1 + a_3 \le b \le a_2 + a_3$ Replace $b$ in case 2 by $a_1+a_2+a_3-b.$\\ $N_{b}(\lambda) = \frac{p^{2a_1+a_2+a_3-b+3}+p^{2a_1+a_2+a_3-b+2}-p^{2a_1+2}-p^{a_1+a_2+a_3-b+2}-p^{a_1+a_2+a_3-b+1}+1}{(p-1)(p^{2}-1)}.$\\ Case 10: $a_2 + a_3 \le b \le a_1 + a_2 + a_3$ Replace $b$ in case 1 by $a_1+a_2+a_3-b.$\\ $N_{b}(\lambda) = \frac{p^{2a_1+2a_2+2a_3-2b+3}-p^{a_1+a_2+a_3-b+2}-p^{a_1+a_2+a_3-b+1}+1}{(p-1)(p^{2}-1)}.$ \end{proofof} \begin{remark} We can deduce all other cases using only case 6 as follows:\\ Case 1: replace $a_3,a_2$ and $a_1$ in case 6 formula by $b.$\\ Case 2: replace $a_3$ and $a_2$ in case 6 formula by $b.$\\ Cases 3 and 4: replace $a_3$ in case 6 formula by $b.$\\ Case 5: replace $a_3$ and $b$ in case 6 formula by $a_2+a_1$ \\ Cases 7 and 8: replace $a_3$ in case 6 formula by $a_1+a_2+a_3-b.$\\ Case 9: replace $a_3$ and $a_2$ in case 6 formula by $a_1+a_2+a_3-b.$\\ Case 10: replace $a_3,a_2$ and $a_1$ in case 6 formula by $a_1+a_2+a_3-b.$ \end{remark} \section{Rank 4} \subsection{Number of subgroups of given order} While it was a somewhat lengthy but accomplishable task to determine the cases for rank 3, there does not appear to be a systematic way of determining all the cases for rank 4. We were able to list more than 22 cases some of which we list below. These formulas are also proved in exactly the same way as in Theorem \ref{rank3}. Here $\lambda=(a_4,a_3,a_2, a_1)$ and $\lambda^{'}=(a_3,a_2, a_1).$\\ When $0\le b \leq a_1,$ we get \\ $N_{b}(\lambda)=\sum_{i=0}^{b} p^{i}N_{i}(\lambda^{'})=\sum_{i=0}^{b} p^{i} \frac{p^{2i+3}-p^{i+2}-p^{i+1}+1}{(p-1)(p^{2}-1)}=\sum_{i=0}^{b} \frac{p^{3i+3}-p^{2i+2}-p^{2i+1}+p^{i}}{(p-1)(p^{2}-1)}$\\ $=\frac{p^{3b+6}-p^{2b+5}-p^{2b+4}-p^{2b+3}+p^{b+3}+p^{b+2}+p^{b+1}-1}{(p-1)(p^{2}-1)(p^{3}-1)}.$\\ When $a_1\le b \leq a_2,$ we get $N_{b}(\lambda)=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{1}+1}^{b} p^{i}N_{i}(\lambda^{'})$\\ $=\frac{p^{3a_{1}+6}-p^{2a_{1}+5}-p^{2a_{1}+4}-p^{2a_{1}+3}+p^{a_{1}+3}+p^{a_{1}+2}+p^{a_{1}+1}-1}{(p-1)(p^{2}-1)(p^{3}-1)}+\sum_{i=a_{1}+1}^{b} p^{i} \frac{p^{i+a_1+3}+p^{i+a_{1}+2}-p^{2a_1+2}-p^{i+2}-p^{i+1}+1}{(p-1)(p^{2}-1)}$\\ $=\frac{p^{2b+a_{1}+6}+p^{2b+a_{1}+5}+p^{2b+a_{1}+4}-p^{2a_{1}+b+5}-p^{2a_{1}+b+4}-p^{2a_{1}+b+3}+p^{3a_{1}+3}-p^{2b+5}-p^{2b+4}-p^{2b+3}+p^{b+3}+p^{b+2}+p^{b+1}-1}{(p-1)(p^{2}-1)(p^{3}-1)}.$\\ And as a final example, when $a_2\le b \leq min \{a_3, a_{1}+a_2 \},$ we have \\ $N_{b}(\lambda)=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{1}+1}^{a_{2}} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{2}+1}^{b} p^{i}N_{i}(\lambda^{'})-\sum_{i=a_{1}+a_{2}+a_{3}+a_{4}+1-b}^{a_{1}+a_{2}+a_{3}}p^{i}N_{i}(\lambda^{'})$\\ $=\frac{p^{2a_{2}+a_{1}+6}+p^{2a_{2}+a_{1}+5}+p^{2a_{2}+a_{1}+4}-p^{2a_{1}+a_{2}+5}-p^{2a_{1}+a_{2}+4}-p^{2a_{1}+a_{2}+3}+p^{3a_{1}+3}}{(p-1)(p^{2}-1)(p^{3}-1)}\\ +\frac{-p^{2a_{2}+5}-p^{2a_{2}+4}-p^{2a_{2}+3}+p^{a_{2}+3}+p^{a_{2}+2}+p^{a_{2}+1}-1}{(p-1)(p^{2}-1)(p^{3}-1)}\\ +\sum_{i=a_{2}+1}^{b} p^{i} \frac{(i-a_{2}+1)p^{a_2+a_1+3}+p^{a_2+a_{1}+2}-(i-a_2)p^{a_1+a_2+1}-p^{2a_1+2}-p^{i+2}-p^{i+1}+1}{(p-1)(p^{2}-1)}$\\ $=\frac{(b+1-a_2)p^{a_{2}+a_{1}+b+6}+(b+1-a_{2})p^{a_{2}+a_{1}+b+5}+(a_{2}-b-1)p^{a_{2}+a_{1}+b+3}+(a_{2}-b-1)p^{a_{1}+a_{2}+b+2}}{(p-1)(p^{2}-1)(p^{3}-1)}\\ +\frac{p^{2a_{2}+a_{1}+4}+p^{a_{1}+2a_{2}+3}+p^{a_{1}+2a_{2}+2}}{(p-1)(p^{2}-1)(p^{3}-1)}$\\ $+\frac{-p^{2a_{1}+b+5}-p^{2a_{1}+b+4}-p^{2a_{1}+b+3}+p^{b+3}+p^{b+2}+p^{b+1}+p^{3a_{1}+3}-p^{2b+5}-p^{2b+4}-p^{2b+3}-1}{(p-1)(p^{2}-1)(p^{3}-1)}.$ \subsection{Total number of subgroups} Let $\lambda = (a_4,a_3,a_2,a_1)$ and let $N(\lambda)$ denote the total number of subgroups of $\mathbb{Z}/p^{a_1}\times \mathbb{Z}/p^{a_2}\times \mathbb{Z}/p^{a_3}\times \mathbb{Z}/p^{a_4}$ where $1\le a_1\le a_2\le a_3\le a_4,$ and $n= a_1+a_2+a_3+a_4.$ Then \begin{equation*} N(\lambda) = \sum_{b=0}^n N_b(\lambda). \end{equation*} A conjecture of L. T$\acute{o}$th in \cite[Conjecutre 10]{LT} claims that the degree of the polynomial $N(m,m,m,m)$, i.e. when $\lambda = (m,m,m,m),$ is $4m$ and that it has leading coefficient of 1. We prove that this follows from our formulas in Theorem \ref{rank3}. First, let's state a corollary for the special case of a rank 3 abelian p-group of type $(m,m,m).$ The proof of the corollary follows immediately from setting $a_1=a_2=a_3=m$ in the theorem. \begin{corollary}\label{mmm} For every $0\leq b \leq 3m$, the number $h_b^{(m,m,m)}(p)$ of all subgroups of order $p^b$ in the finite abelian $p$-group $\mathbb{Z}/p^{m}\times \mathbb{Z}/p^{m}\times \mathbb{Z}/p^{m}$ where $1\leq m,$ is given by one of the following polynomials expressed as rational functions:\\ Case 1: $0\le b \leq m$ $$h_b^{(m,m,m)}(p) = \frac{p^{2b+3} - p^{b+2} - p^{b+1} +1}{(p-1)(p^2-1)}.$$ Case 2: $m \le b \leq 2m$ $$h_b^{(m,m,m)}(p) = \frac{p^{2m+3}+p^{2m+2}+p^{2m+1}-p^{3m+2-b}-p^{3m+1-b}-p^{b+2}-p^{b+1}+1}{ (p-1)(p^2-1)}.$$ Case 3: $2m \le b \le 3m$ $$h_b^{(m,m,m)}(p) =\frac{p^{6m+3-2b}-p^{3m+2-b}-p^{3m+1-b}+1}{(p-1)(p^{2}-1)}.$$ \end{corollary} \begin{theorem}\label{Conj10} The the total number of subgroups $N(m,m,m,m)$ of the rank 4 finite abelian p-group $\mathbb{Z}/p^{m}\times \mathbb{Z}/p^{m}\times \mathbb{Z}/p^{m}\times \mathbb{Z}/p^{m}$ where $1\le m$ is given by a polynomial expressed as a rational function as below: \begin{multline*} N(m,m,m,m) = \sum_{b=1}^{4m} N_b(m,m,m,m)\\ = \frac{ {\left(p^{2} + p + 1\right)}^{3} {\left(p^{2} + 1\right)} p^{4 \, m + 2} - {\left({\left(2 \, m + 3\right)} p^{3} - 2 \, m - 1\right)} {\left(p^{3} + p\right)} {\left(p + 1\right)}^{3} p^{3 \, m} }{{\left(p^{2} - 1\right)}^{2} {\left(p^3 - 1\right)}^{2}}\\ - \frac{4 \, m p^{4} + 4 \, m p^{3} + 7 \, p^{4} + 9 \, p^{3} - 4 \, m p + 6 \, p^{2} - 4 \, m + p - 1}{{\left(p^{2} - 1\right)}^{2} {\left(p^3 - 1\right)}^{2}}. \end{multline*} In particular, the degree of the polynomial is $4m$ and its leading coefficient is 1. \end{theorem} \begin{proofof}{Theorem \ref{Conj10}} We break the interval $[0,4m]$ into 4 sub-intervals and use \cite[Lemma 3.2]{YH} and the corollary above in each case to derive a polynomial expression for the number of subgroups of order $p^b$ in a rank 4 abelian p-group of type $(m,m,m,m)$. Summing over these polynomials will result in the required polynomial expression.\\ Case 1: $0\le b\le m$ \begin{align*} N_b(m,m,m,m) &= \sum_{i=0}^{b} p^{i}N_{i}(m,m,m)=\sum_{i=0}^{b} p^{i} \frac{p^{2i+3} - p^{i+2} - p^{i+1} +1}{(p-1)(p^2-1)}\\ &= \frac{p^{3 \, b + 6} - p^{2 \, b + 5} - p^{2 \, b + 4} - p^{2 \, b + 3} + p^{b + 3} + p^{b + 2} + p^{b + 1} - 1}{{\left(p - 1\right)} {\left(p^2 - 1\right)} {\left(p^3 - 1\right)}}. \end{align*} Case 2: $m < b\le 2m$ \begin{align*} N_b(m,m,m,m) &= \sum_{i=0}^{m} p^{i}N_{i}(m,m,m) + \sum_{i=m+1}^{b} p^{i}N_{i}(m,m,m) - \sum_{i=4m+1-b}^{3m} p^{i}N_{i}(m,m,m)\\ &= \sum_{i=0}^{m} p^{i}\frac{p^{2i+3} - p^{i+2} - p^{i+1} +1}{(p-1)(p^2-1)}\\ &+ \sum_{i=m+1}^{b} p^{i}\frac{p^{2m+3}+p^{2m+2}+p^{2m+1}-p^{3m+2-i}-p^{3m+1-i}-p^{i+2}-p^{i+1}+1}{ (p-1)(p^2-1)}\\ &- \sum_{i=4m+1-b}^{3m} p^{i}\frac{p^{6m+3-2i}-p^{3m+2-i}-p^{3m+1-i}+1}{(p-1)(p^{2}-1)}\\ &= -\frac{p^{2 \, b + 5} + p^{2 \, b + 4} + p^{2 \, b + 3} - p^{b + 2 \, m + 6} - p^{b + 2 \, m + 5} - 2 \, p^{b + 2 \, m + 4}}{(p-1)(p^2-1)(p^3-1)} \\ &+\frac{ - p^{b + 2 \, m + 3} - p^{b + 2 \, m + 2} - p^{b + 3} - p^{b + 2}}{(p-1)(p^2-1)(p^3-1)} \\ &-\frac{ - p^{b + 1} - p^{-b + 4 \, m + 3} - p^{-b + 4 \, m + 2} - p^{-b + 4 \, m + 1} + p^{3 \, m + 5} + 2 \, p^{3 \, m + 4}}{(p-1)(p^2-1)(p^3-1)}\\ &+ \frac{2 \, p^{3 \, m + 3} + 2 \, p^{3 \, m + 2} + p^{3 \, m + 1} + 1}{(p-1)(p^2-1)(p^3-1)}. \end{align*} Case 3: $2m < b \le 3m$ \begin{align*} N_b(m,m,m,m) &= \sum_{i=0}^{m} p^{i}N_{i}(m,m,m) + \sum_{i=m+1}^{2m} p^{i}N_{i}(m,m,m) + \sum_{i=2m+1}^{b} p^{i}N_{i}(m,m,m)\\ &- \sum_{i=4m+1-b}^{2m} p^{i}N_{i}(m,m,m) -\sum_{i=2m+1}^{3m} p^{i}N_{i}(m,m,m)\\ &= \sum_{i=0}^{m} p^{i}\frac{p^{2i+3} - p^{i+2} - p^{i+1} +1}{(p-1)(p^2-1)}\\ &+ \sum_{i=m+1}^{2m} p^{i}\frac{p^{2m+3}+p^{2m+2}+p^{2m+1}-p^{3m+2-i}-p^{3m+1-i}-p^{i+2}-p^{i+1}+1}{ (p-1)(p^2-1)}\\ &+ \sum_{i=2m+1}^{b} p^{i}\frac{p^{6m+3-2i}-p^{3m+2-i}-p^{3m+1-i}+1}{(p-1)(p^{2}-1)}\\ &- \sum_{i=4m+1-b}^{2m} p^{i}\frac{p^{2m+3}+p^{2m+2}+p^{2m+1}-p^{3m+2-i}-p^{3m+1-i}-p^{i+2}-p^{i+1}+1}{ (p-1)(p^2-1)}\\ &- \sum_{i=2m+1}^{3m} p^{i}\frac{p^{6m+3-2i}-p^{3m+2-i}-p^{3m+1-i}+1}{(p-1)(p^{2}-1)}\\ &=\frac{p^{b + 3} + p^{b + 2} + p^{b + 1} + p^{-b + 6 \, m + 6} + p^{-b + 6 \, m + 5} + 2 \, p^{-b + 6 \, m + 4}}{(p-1)(p^2-1)(p^3-1)}\\ &+ \frac{p^{-b + 6 \, m + 3} + p^{-b + 6 \, m + 2} + p^{-b + 4 \, m + 3}}{(p-1)(p^2-1)(p^3-1)}\\ &+ \frac{p^{-b + 4 \, m + 2} + p^{-b + 4 \, m + 1} - p^{-2 \, b + 8 \, m + 5} - p^{-2 \, b + 8 \, m + 4} - p^{-2 \, b + 8 \, m + 3} - p^{3 \, m + 5} - 2 \, p^{3 \, m + 4}}{(p-1)(p^2-1)(p^3-1)}\\ &+ \frac{- 2 \, p^{3 \, m + 3} - 2 \, p^{3 \, m + 2} - p^{3 \, m + 1} - 1}{(p-1)(p^2-1)(p^3-1)}. \end{align*} Case 4: $3m < b\le 4m$ \begin{align*} N_b(m,m,m,m) &= \sum_{i=0}^{m} p^{i}N_{i}(m,m,m) + \sum_{i=m+1}^{2m} p^{i}N_{i}(m,m,m) + \sum_{i=2m+1}^{3m} p^{i}N_{i}(m,m,m) \\ &- \sum_{i=4m+1-b}^{m} p^{i}N_{i}(m,m,m) -\sum_{i=m+1}^{2m} p^{i}N_{i}(m,m,m) - \sum_{i=2m+1}^{3m} p^{i}N_{i}(m,m,m)\\ &= \sum_{i=0}^{m} p^{i}\frac{p^{2i+3} - p^{i+2} - p^{i+1} +1}{(p-1)(p^2-1)}\\ &+ \sum_{i=m+1}^{b} p^{i}\frac{p^{2m+3}+p^{2m+2}+p^{2m+1}-p^{3m+2-i}-p^{3m+1-i}-p^{i+2}-p^{i+1}+1}{ (p-1)(p^2-1)} \end{align*} \begin{align*} &+ \sum_{i=2m+1}^{3m} p^{i}\frac{p^{6m+3-2i}-p^{3m+2-i}-p^{3m+1-i}+1}{(p-1)(p^{2}-1)}\\ &- \sum_{i=4m+1-b}^{m} p^{i}\frac{p^{2i+3} - p^{i+2} - p^{i+1} +1}{(p-1)(p^2-1)}\\ &- \sum_{i=m+1}^{2m} p^{i}\frac{p^{2m+3}+p^{2m+2}+p^{2m+1}-p^{3m+2-i}-p^{3m+1-i}-p^{i+2}-p^{i+1}+1}{ (p-1)(p^2-1)}\\ &- \sum_{i=2m+1}^{3m} p^{i}\frac{p^{6m+3-2i}-p^{3m+2-i}-p^{3m+1-i}+1}{(p-1)(p^{2}-1)}\\ &=\frac{p^{-b + 4 \, m + 3} + p^{-b + 4 \, m + 2} + p^{-b + 4 \, m + 1} - p^{-2 \, b + 8 \, m + 5}}{(p-1)(p^2-1)(p^3-1)}\\ &+ \frac{- p^{-2 \, b + 8 \, m + 4} - p^{-2 \, b + 8 \, m + 3} + p^{-3 \, b + 12 \, m + 6} - 1}{(p-1)(p^2-1)(p^3-1)}. \end{align*} Now summing over the 4 cases, we get that \begin{align*} N(m,m,m,m)&=\sum_{b=0}^{4m} N_b(m,m,m,m)\\ &= \frac{ {\left(p^{2} + p + 1\right)}^{3} {\left(p^{2} + 1\right)} p^{4 \, m + 2} - {\left({\left(2 \, m + 3\right)} p^{3} - 2 \, m - 1\right)} {\left(p^{3} + p\right)} {\left(p + 1\right)}^{3} p^{3 \, m} }{{\left(p^{2} - 1\right)}^{2} {\left(p^3 - 1\right)}^{2}}\\ &- \frac{4 \, m p^{4} + 4 \, m p^{3} + 7 \, p^{4} + 9 \, p^{3} - 4 \, m p + 6 \, p^{2} - 4 \, m + p - 1}{{\left(p^{2} - 1\right)}^{2} {\left(p^3 - 1\right)}^{2}}. \end{align*} Now, we can see that the degree of the polynomial is $4m$ and the leading coefficient is 1, as conjectured by L. T$\acute{o}$th. \end{proofof} \begin{remark} In an unpublished paper \cite{CCL}, Chew, Chin, and Lim derive an explicit formula for the number of subgroups of a finite abelian p-group of rank 4. Here we will use the formula to reprove the above theorem and also prove one more conjecture by T$\acute{o}$th in \cite[Conjecture 9]{LT}. \begin{theorem} (Chew, Chin, and Lim) Let $1 \le w \le x \le y \le z$. The number of subgroups of $\mathbb{Z}/p^{w}\times \mathbb{Z}/p^{x}\times \mathbb{Z}/p^{y}\times \mathbb{Z}/p^{z}$ is \begin{align*} N(w,x,y,z) &= \sum_{i=0}^{w-1}\sum_{j=0}^{i} [(w+x+y+z-4i+1)(2i-2j+1)p^{3i+j}\\ &+(w+x+y+z-4i-1)(2i-2j+1)p^{3i+j+1}\\ &+ 2(w+x+y+z-4i-2)(i-j+1)p^{3i+j+2}]\\ &+\sum_{i=0}^{w}\sum_{j=w}^{x-1} (w+j-2i+1)[(x+y+z-3j+1)p^{w+2j+i}\\ \end{align*} \begin{align*} &+(x+y+z-3j-1)p^{w+2j+i+1}] \\ &+\sum_{i=0}^{w}\sum_{j=x}^{y} (y+z-2j+1)(w+x+-2i+1)p^{w+x+i+j}. \end{align*} \end{theorem} Since all coefficients are positive, none of the terms will cancel out and the degree of the polynomial in $p$ can be easily determined. By comparing the exponents, we can see that the highest degree appears in the last double sum. Setting $i=w$ and $j=y,$ we get that the leading term of $N(w,x,y,z)$ is $(z-y+1)(x-w+1)p^{2w+x+y},$ with degree $2w+x+y,$ confirming conjecture 9 in \cite{LT}. Moreover, setting $w=x=y=z=m,$ in the above leading term, we get that the degree of the leading term in $N(m,m,m,m)$ is $4m,$ confirming Conjecture 10 in the same paper. \end{remark} \begin{remark} What is new in Theorem \ref{Conj10} is the explicit polynomial. Else, both conjectures 9 and 10 of Toth follow from Theorem 3 in \cite{GBJW01} as they have determined the leading coefficient and degree of the counting polynomial for all ranks. \end{remark} \section{A special case of any rank} The recursive counting of the subgroups used in the proof of Theorem \ref{rank3} based on \cite[Lemma 2.3]{YH} enabled us to deduce a compact rational form for the polynomials in all the cases for ranks 2 and 3. Here we will demo the method of fundamental group lattices in one case for all ranks. We remark that it is difficult to use this method to give an explicit formula in the remaining cases (those that are not symmetric to Case 1) of rank 3 or higher. Below we extend T$\ddot{a}$rn$\ddot{a}$uceanu's result for counting subgroups as in \cite[Theorem 3.3]{MT} for case 1 to any rank. \begin{theorem} \label{funlatt} Let $0\leq b \leq a_1.$ Then \begin{equation} h_b^{(a_k,\dots,a_1)}(p) = \prod_{i=2}^k\frac{p^{b+i-1}-1}{p^{i-1}-1}. \end{equation} Similarly, if $a_2+\dots+a_k\le b \leq a_1+\dots+a_k$, then replace $b$ in the above formula by $a_1+\dots+a_k-b$ to get \begin{equation} h_b^{(a_k,\dots,a_1)}(p) = \prod_{i=2}^k\frac{p^{a_1+\dots+a_k-b+i-1}-1}{p^{i-1}-1}. \end{equation} \end{theorem} \begin{proofof}{Theorem \ref{funlatt}} Following the notation in \cite{MT}, let $A= (a_{ij})$ be a solution of $(\ast)$ below corresponding to a subgroup of order $p^{a_1+\dots+a_k-b}$ in $\mathbb{Z}/p^{a_1}\times \mathbb{Z}/p^{a_2}\times \cdots \times \mathbb{Z}/p^{a_k}.$ \\ $$(\ast) \begin{cases} \text{i) } a_{ij} = 0 \text{ for any } i>j\\ \text{ii) } 0\leq a_{1j},a_{2j},\dots,a_{j-1,j} < a_{jj} \text{ for any } j \in \{1,2,\dots, k\}\\ \text{iii) } \text{1) } a_{11}|p^{a_1}\\ \hspace{0.2in}\text{ 2) } a_{22}|(p^{a_2},p^{a_1}\frac{a_{12}}{a_{11}})\\ \hspace{0.2in}\text{ 3) } a_{33}|(p^{a_3},p^{a_2}\frac{a_{23}}{a_{22}},p^{a_1}\frac{\begin{vmatrix} a_{12} &a_{13} \\ a_{22} & a_{23} \end{vmatrix}}{a_{22}a_{11}})\\ \vspace{0.1in}\\ \hspace{1in} \longvdots{1.5em}\\ \hspace{0.2in}\text{ k) } a_{kk}|(p^{a_k},p^{a_{k-1}}\frac{a_{k-1,k}}{a_{k-1,k-1}},p^{a_{k-2}}\frac{\begin{vmatrix} a_{k-2,k-1} &a_{k-2,k} \\ a_{k-1,k-1} & a_{k-1,k} \end{vmatrix}}{a_{k-1,k-1}a_{k-2,k-2}},\dots,p^{a_1}\frac{\begin{vmatrix} a_{12} & a_{13} & \cdots & a_{1,k} \\ a_{22} & a_{23} & \cdots & a_{2,k} \\ \textbf{.} & \textbf{.} & & \textbf{.} \\ \textbf{.} & \textbf{.} & & \textbf{.} \\ \textbf{.} & \textbf{.} & & \textbf{.} \\ a_{k-1,2} & a_{k-1,3} & \cdots & a_{k-1,k} \end{vmatrix}}{a_{k-1,k-1}a_{k-2,k-2}\dots a_{11}}). \end{cases} $$ Put $a_{11}=p^{i_1}$ where $0\leq i_1\leq a_1.$ By the first remark on pp. 376 of \cite{MT}, the order of a subgroup corresponding to $A= (a_{ij})$ is $$\frac{p^{\sum_{i=1}^k a_i}}{\prod_{i=1}^k a_{ii}} = p^{a_1+\dots+a_k-b}$$ Therefore, $a_{11}\cdots a_{kk}=p^b.$ Since $a_{11}=p^{i_1},$ we have $a_{22}\cdots a_{kk}=p^{b-i_1}.$ Let\\ \hspace{2in} $a_{22} = p^{i_2}$ where $0\leq i_2 \leq b-i_1,$ \\ \hspace{2in} $a_{33}=p^{i_3}$ where $0\leq i_3\leq b-i_1-i_2,$\\ \hspace{2in} $\dots $ \\ \hspace{2in} $a_{k-1,k-1}=p^{i_{k-1}}$ where $0\leq i_{k-1} \leq b-i_1-\dots - i_{k-2}.$ \\ Then $a_{kk}=p^{b-i_1-\dots - i_{k-1}}.$ Now, the conditions become $$p^{i_2}|(p^{a_2},p^{a_1-i_1}a_{12}) = p^{a_1-i_1}(p^{a_2-a_1+i_1},a_{12})$$ $$p^{i_3}|(p^{a_3},p^{a_2-i_2}a_{12}, p^{a_1-i_1-i_2}(a_{12}a_{23}-p^{i_2}a_{13})$$ $$\dots$$ $$p^{b-i_1-\dots - i_{k-1}}|(p^{a_k},p^{a_{k-1}}\frac{a_{k-1,k}}{a_{k-1,k-1}},p^{a_{k-2}}\frac{\begin{vmatrix} a_{k-2,k-1} &a_{k-2,k} \\ a_{k-1,k-1} & a_{k-1,k} \end{vmatrix}}{a_{k-1,k-1}a_{k-2,k-2}},\dots,p^{a_1}\frac{\begin{vmatrix} a_{12} & a_{13} & \cdots & a_{1,k} \\ a_{22} & a_{23} & \cdots & a_{2,k} \\ \textbf{.} & \textbf{.} & & \textbf{.} \\ \textbf{.} & \textbf{.} & & \textbf{.} \\ \textbf{.} & \textbf{.} & & \textbf{.} \\ a_{k-1,2} & a_{k-1,3} & \cdots & a_{k-1,k} \end{vmatrix}}{a_{k-1,k-1}a_{k-2,k-2}\dots a_{11}}).$$ These conditions are satisfied by all $$a_{12}< a_{22}=p^{i_2},$$ $$a_{13},a_{23} < a_{33}=p^{i_3},$$ $$\cdots$$ $$a_{1,k-1},\dots,a_{k-2,k-1}< a_{k-1,k-1}=p^{i_{k-1}}$$ $$a_{1,k},\dots,a_{k-1,k}< a_{k,k}=p^{b-i_1-\cdots-i_{k-1}}.$$ So, we get $$p^{i_2}\cdot (p^{i_3})^2 \cdots (p^{i_{k-1}})^{k-2}\cdot (p^{b-i_1-\cdots-i_{k-1}})^{k-1} = p^{(k-1)b-(k-1)i_1-(k-2)i_2-\cdots -2i_{k-2}-i_{k-1}}$$ distinct solutions of $(\ast)$ for given $a_{11}=p^{i_1},a_{22}=p^{i_2},\cdots,a_{k-1,k-1}=p^{i_{k-1}},$ and $a_{k,k}=p^{b-i_1-\cdots-i_{k-1}}.$ The number of subgroups of index $p^b$ (or order $p^{a_1+\dots+a_k-b}$) in $\mathbb{Z}/p^{a_1}\times \cdots \times \mathbb{Z}/p^{a_k}$ is then \begin{equation} h_b^{(a_k,\dots,a_1)}(p) = \sum_{i_1=0}^{b}\sum_{i_2=0}^{b-i_1}\cdots \sum_{i_{k-1}=0}^{b-i_1-\cdots-i_{k-2}} p^{(k-1)b-(k-1)i_1-(k-2)i_2-\cdots -2i_{k-2}-i_{k-1}}. \end{equation} Now we will use induction to prove that \begin{equation}\label{anyrankclaim} \sum_{i_1=0}^{b}\sum_{i_2=0}^{b-i_1}\cdots \sum_{i_{k-1}=0}^{b-i_1-\cdots-i_{k-2}} p^{(k-1)b-(k-1)i_1-(k-2)i_2-\cdots -2i_{k-2}-i_{k-1}} = \prod_{i=2}^k\frac{p^{b+i-1}-1}{p^{i-1}-1}. \end{equation} By Theorem \ref{rank3} and previous discussion, equation (\ref{anyrankclaim}) is true for when $k=2,3.$ Let us assume that it is true for rank k. Now consider the left hand side of equation (\ref{anyrankclaim}) for rank $k+1,$ that is, $$\sum_{i_1=0}^{b}\sum_{i_2=0}^{b-i_1}\cdots \sum_{i_{k-1}=0}^{b-i_1-\cdots-i_{k-2}} \sum_{i_{k}=0}^{b-i_1-\cdots-i_{k-2}-i_{k-1}}p^{(k)b-(k)i_1-(k-1)i_2-\cdots -3i_{k-2}-2i_{k-1}-i_{k}} $$\\ and put $\alpha=b-i_1$. Then \\ $$\sum_{i_1=0}^{b}p^{b-i_1}\sum_{i_2=0}^{\alpha}\cdots \sum_{i_{k-1}=0}^{\alpha-\cdots-i_{k-2}} \sum_{i_{k}=0}^{\alpha-\cdots-i_{k-2}-i_{k-1}}p^{(k-1)\alpha-(k-1)i_2-\cdots -3i_{k-2}-2i_{k-1}-i_{k} }$$ \begin{align*} &= \sum_{i_1=0}^{b}p^{b-i_1} \prod_{i=3}^{k+1}\frac{p^{b+i-1}-1}{p^{i-1}-1}\\ &= \prod_{i=3}^{k+1}\frac{p^{b+i-1}-1}{p^{i-1}-1}\sum_{i_1=0}^{b}p^{b-i_1}\\ &= \prod_{i=2}^{k+1}\frac{p^{b+i-1}-1}{p^{i-1}-1}. \end{align*} This completes the proof. \end{proofof} \begin{remark} An application of the method used in Theorem \ref{rank3} also shows that we can rederive the above result in a simpler way as follows. Let $\lambda=(a_k,\dots, a_1), \lambda^{'}=(a_{k-1},\dots, a_1),$ and $0\le b \leq a_1.$ Then a repeated application of the recurrence (\ref{Stehrec}) for $N_b(\lambda)$ shows that \begin{align*} N_{b}(\lambda) &= \sum_{i_1=0}^{b} p^{i_1}N_{i_1}(\lambda^{'})\\ &=\sum_{i_1=0}^{b} p^{i_1}\sum_{i_2=0}^{i_1} p^{i_2}N_{i_2}(\lambda^{''})\\ &=\dots \\ &=\sum_{i_1=0}^{b} \sum_{i_2=0}^{i_1} \dots\sum_{i_{k-1}=0}^{i_{k-2}} p^{i_1+\dots +i_{k-1}} N_{i_{k-1}}(a_1)\\ &=\sum_{i_1=0}^{b} \sum_{i_2=0}^{i_1} \dots\sum_{i_{k-1}=0}^{i_{k-2}} p^{i_1+\dots+i_{k-1}}. \end{align*} Now let $j_t = b - \sum_{l=1}^t i_t$ for $1\le t\le k-2.$ Then $i_t = j_{t-1} - j_t$ for $1\le t\le k-2,$ letting $j_0=b.$ As $i_t$ varies between 0 and $j_{t-1},$ $j_t$ will also vary between 0 and $j_{t-1}.$ Also $(k-1)b-(k-1)i_1-(k-2)i_2-\cdots -2i_{k-2}-i_{k-1}= j_1+\dots+j_{k-1}.$ Hence, \begin{align*} \sum_{j_1=0}^{b} \sum_{j_2=0}^{j_1} \dots\sum_{j_{k-1}=0}^{j_{k-2}} p^{j_1+\dots+j_{k-1}} &= \sum_{i_1=0}^{b} \sum_{i_2=0}^{j_1} \dots\sum_{i_{k-1}=0}^{j_{k-2}} p^{j_1+\dots+j_{k-1}} \\ &=\sum_{i_1=0}^{b}\sum_{i_2=0}^{b-i_1}\cdots \sum_{i_{k-1}=0}^{b-i_1-\cdots-i_{k-2}} p^{(k-1)b-(k-1)i_1-(k-2)i_2-\cdots -2i_{k-2}-i_{k-1}}, \end{align*} which is the same as the left hand side of equation (\ref{anyrankclaim}). Therefore, \begin{equation} \sum_{i_1=0}^{b} \sum_{i_2=0}^{i_1} \dots\sum_{i_{k-1}=0}^{i_{k-2}} p^{i_1+\dots+i_{k-1}} = \prod_{i=2}^k\frac{p^{b+i-1}-1}{p^{i-1}-1}. \end{equation} \end{remark} \section{The method of convolution} \label{sec:convolution} Here, we will use convolution of generating series and a recurrence relation of T. Stehling\cite{TS} to prove the rational function formula in case 1 of rank 2. This is only to demo the method and the rest of the cases can be handled similarly. To ease the use of the recurrence relation, let's adopt Stehling's notation for type $\alpha = (\alpha_1,\dots,\alpha_d)$ where $ \alpha_1\ge \cdots \alpha_d\ge 0$ which reverses the order in the notation we used in previous sections. Let $0\le b \leq \alpha_2$ (i.e. case 1 of rank 2), then \begin{equation} \label{c1rk2} h_b^{(\alpha_1,\alpha_2)}(p) = \frac{ p^{b+1} -1}{p-1}. \end{equation} The symmetric case (i.e. $\alpha_1\le b \leq \alpha_1+\alpha_2$) is then (replacing $b$ by $\alpha_1+\alpha_2-b$ in the above rational function): \begin{equation} h_b^{(\alpha_1,\alpha_2)}(p) = \frac{ p^{\alpha_1+\alpha_2-b+1} -1}{p-1}. \end{equation} The idea behind the method is the following. Let $F(x) = \sum a_nx^n$ and $G(x) = \sum b_nx^n$ be two convergent series. Then the series $H(x) = \sum a_nb_nx^n$ can be obtained from $F$ and $G$ as follows. Define the \textbf{convolution} of $F$ and $G$ denoted $F\star G (x)$ as the following contour integral around a small circle about $0$ where $x$ is a complex number very small in magnitude. \begin{align*} F\star G (x) &= \frac{1}{2\pi i} \oint_{y\in B_{\epsilon}(0)} F(y)G(\frac{x}{y})\frac{dy}{y}\\ &= \frac{1}{2\pi i} \oint_{y\in B_{\epsilon}(0)} \sum a_nb_my^n\frac{x^m}{y^m}\frac{dy}{y} \\ &=\sum a_nb_mx^m \frac{1}{2\pi i} \oint_{y\in B_{\epsilon}(0)} y^{n-m-1}dy\\ &= \sum a_nb_nx^n. \end{align*} Thus $H(x) = F\star G (x).$ Now, we will prove equation (\ref{c1rk2}) using this method. Let $F_d(x,y)$ be the generating series \begin{equation} F_d(x,y) = \sum_{\substack{\alpha = (\alpha_1,\dots,\alpha_d) \\ \alpha_1\ge \cdots \alpha_d\ge 0 \\ 0\le r\le \alpha_1+\dots+\alpha_d}} h_r^{(\alpha_1,\dots,\alpha_d)}(p)x^{\alpha}y^r \end{equation} where $x = (x_1,\dots, x_d)$ and $x^\alpha = x_1^{\alpha_1}x_2^{\alpha_2}\cdots x_d^{\alpha_d}.$ Let's denote $h_r^{\alpha}(p)$ by $N_{\alpha}^{(d)}(r)$ or simply $N_{\alpha}(r)$ when the rank is understood. By \cite[Corollary]{TS}, we have the recurrence formula \begin{equation}\label{Stehrec} N_{\alpha}(r) = N_{\tilde{\alpha}}(r-1)+p^rN_{\hat{\alpha}}(r) \end{equation} where $\hat{\alpha} = (\alpha_2,\dots,\alpha_d)$ and $\tilde{\alpha} = \alpha$ with $-1$ added to the $k^{th}$ position where $$k=\begin{cases} 1 & \text{if } \alpha_1>\alpha_2\\ 2 & \text{if } \alpha_1=\alpha_2>\alpha_3 \\ 3 & \text{if } \alpha_1=\alpha_2=\alpha_3>\alpha_4\\ \dots & \dots \end{cases}$$ Note $N_{\alpha}^{(1)}(r) =1$ if $0\le r\le \alpha.$ So, \begin{equation} F_1(x,y) = \frac{1}{(1-x)(1-xy)}. \end{equation} For rank 2, let's break down $F_2(x_1,x_2,y)$ as follows: \begin{align} \label{F2} F_2(x_1,x_2,y) &= \sum_{\substack{ \alpha_1 \ge \alpha_2\ge 0 \\ 0\le r\le \alpha_1+\alpha_2}} N_{\alpha_1,\alpha_2}^{(2)}(r)x_1^{\alpha_1}x_2^{\alpha_2}y^r = F_2^{(0)}+F_2^{(1)} \end{align} where \begin{align} F_2^{(0)}(x_1,x_2,y) &= \sum_{\alpha_1 = \alpha_2\ge 0} N_{\alpha_1,\alpha_2}^{(2)}(r)x_1^{\alpha_1}x_2^{\alpha_2}y^r \\ F_2^{(1)}(x_1,x_2,y) &= \sum_{\alpha_1 > \alpha_2\ge 0} N_{\alpha_1,\alpha_2}^{(2)}(r)x_1^{\alpha_1}x_2^{\alpha_2}y^r \end{align} Let's first compute $F_2^{(0)}(x_1,x_2,y). $ \begin{align*} F_2^{(0)}(x_1,x_2,y) &= \sum_{\alpha=0}^\infty \sum_r N_{\alpha,\alpha}^{(2)}(r)(x_1x_2)^{\alpha}y^r \\ &= \sum_{\alpha} \sum_r N_{\alpha,\alpha-1}^{(2)}(r-1)(x_1x_2)^{\alpha}y^r + \sum_{\alpha} p^rN_{\alpha}^{(1)}(r)(x_1x_2)^{\alpha}y^r\\ &= \sum_{\alpha} \sum_r N_{\alpha-1,\alpha-1}^{(2)}(r-2)(x_1x_2)^{\alpha}y^r + \sum_{\alpha} \sum_r p^{r-1}N_{\alpha-1}^{(1)}(r-1)(x_1x_2)^{\alpha}y^r \\ &+ \sum_{\alpha} p^rN_{\alpha}^{(1)}(r)(x_1x_2)^{\alpha}y^r\\ &= \sum_{\alpha} \sum_r N_{\alpha,\alpha}^{(2)}(r)(x_1x_2)^{\alpha+1}y^{r+2} + \sum_{\alpha} \sum_r N_{\alpha}^{(1)}(r)(x_1x_2)^{\alpha+1}p^{r}y^{r+1}\\ &+ \sum_{\alpha} N_{\alpha}^{(1)}(r)(x_1x_2)^{\alpha}p^ry^r\\ &= x_1x_2y^2F_2^{(0)}(x_1,x_2,y) + x_1x_2yF_1(x_1x_2,py) + F_1(x_1x_2,py). \end{align*} Solving for $F_2^{(0)},$ results in \begin{equation} F_2^{(0)}(x_1,x_2,y) = \frac{1+x_2x_2y}{(1-x_1x_2)(1-px_1x_2y)(1-x_1x_2y^2)}. \end{equation} Similarly, we have \begin{align*} F_2^{(1)}(x_1,x_2,y) &= \sum_{\alpha_1>\alpha_2\ge 0} \sum_r N_{\alpha}^{(2)}(r)x_1^{\alpha_1}x_2^{\alpha_2}y^r \\ &=\sum_{\alpha_1>\alpha_2\ge 0} \left [ N_{\tilde{\alpha}}^{(2)}(r-1)+p^rN_{\hat{\alpha}}^{(1)}(r) \right ]x_1^{\alpha_1}x_2^{\alpha_2}y^r\\ &=\sum_{\alpha_1>\alpha_2\ge 0} \left [ N_{\alpha_1-1,\alpha_2}^{(2)}(r-1)+p^rN_{\alpha_2}^{(1)}(r) \right ]x_1^{\alpha_1}x_2^{\alpha_2}y^r\\ &=\sum_{\alpha_1 \ge \alpha_2\ge 0} N_{\alpha_1,\alpha_2}^{(2)}(r-1)x_1^{\alpha_1+1}x_2^{\alpha_2}y^{r+1} + \sum_{ \substack{ \alpha_1 > \alpha_2\ge 0\\ 0\le r\le \alpha_2}} x_1^{\alpha_1}x_2^{\alpha_2}(py)^r\\ &= x_1y\left [ F_2^{(1)}(x_1,x_2,y) + F_2^{(0)}(x_1,x_2,y) \right ] + \sum_{ \alpha_1 > \alpha_2\ge 0} x_1^{\alpha_1}x_2^{\alpha_2}\frac{1-(py)^{\alpha_2+1}}{1-py}. \end{align*} Solving for $F_2^{(1)}(x_1,x_2,y),$ we have \begin{equation} F_2^{(1)}(x_1,x_2,y) = \frac{1+y-x_1y+x_1^2x_2y^2}{(1-x_1)(1-x_1y)(1-x_1x_2)(1-x_1x_2y)(1-qx_1x_2y)}. \end{equation} Finally, combining the two sums, equation (\ref{F2}) becomes \begin{equation*} F_2(x,y) = F_2(x_1,x_2,y) = \sum_{\substack{\alpha = (\alpha_1,\alpha_2) \\ \alpha_1\ge \alpha_2\ge 0 \\ 0\le r\le \alpha_1+\alpha_2}} h_r^{(\alpha_1,\alpha_2)}(p)x_1^{\alpha_1}x_2^{\alpha_2}y^r\end{equation*} \begin{equation} = \frac{x_1^2x_2y^2 + x_1^2x_2y - x_1x_2y - 1}{(1-x_1)(1-x_1y)(1-x_1x_2)(1-x_1x_2y^2)(px_1x_2y-1)}. \end{equation} We will use the idea described before to pick out the rational function expression for case 1 from the general rational function expression of $F_2$ above. The final step will be extracting a general formula for the coefficients of the general term of the power series expansion of the resulting rational function using products of power series and change of variables. Now consider small circles in the complex plane around 0 where the complex numbers $x_1,x_2,y$ are very small in magnitude. We evaluate the following repeated contour integrals of the convolution of $F_2$ and $G$ by applying Cauchy's residue theorem to get the generating series for case 1. Let \begin{equation} G(x_1,x_2,y) = \sum_{ \alpha_1\ge \alpha_2\ge r\ge 0 } x_1^{\alpha_1}x_2^{\alpha_2}y^r = \frac{1}{(1-x_1)(1-x_1x_2)(1-x_1x_2y)}. \end{equation} Then \begin{align*} F_2\star G (x_1,x_2,y) &=\frac{1}{(2\pi i)^3} \oint_{u_3\in B_{\epsilon_3}(0)} \oint_{u_2\in B_{\epsilon_2}(0)} \oint_{u_1\in B_{\epsilon_1}(0)} F_2(u_1,u_2,v)G(\frac{x_1}{u_1},\frac{x_2}{u_2},\frac{y}{v})\frac{du_1}{u_1}\frac{du_2}{u_2}\frac{dv}{v} \\ &= \sum_{ \alpha_1\ge \alpha_2\ge r\ge 0 } h_r^{(\alpha_1,\alpha_2)}(p)x_1^{\alpha_1}x_2^{\alpha_2}y^r. \end{align*} Using the rational functions for $F_2$ and $G$ and evaluating the integrals, we get that \begin{equation} \sum_{ \alpha_1\ge \alpha_2\ge r\ge 0 } h_r^{(\alpha_1,\alpha_2)}(p)x_1^{\alpha_1}x_2^{\alpha_2}y^r = \frac{1}{(1-x_1)(1-x_1x_2)(1-x_1x_2y)(1-px_1x_2y)}. \end{equation} It remains to extract the coefficient of $x_1^{\alpha_1}x_2^{\alpha_2}y^r$ from the last rational function in order to determine $h_r^{(\alpha_1,\alpha_2)}(p)$ for the case $0\le r \le a_2.$ We do this as follows: \begin{align*} &\frac{1}{(1-x_1)(1-x_1x_2)(1-x_1x_2y)(1-px_1x_2y)}\\ &= \big( \sum_{k_1\ge 0} x_1^{k_1}\big)\big( \sum_{k_2\ge 0} (x_1x_2)^{k_2}\big)\big( \sum_{k_3\ge 0} (x_1x_2y)^{k_3}\big)\big( \sum_{k_4\ge 0} (px_1x_2y)^{k_3}\big)\\ &= \sum_{k_1,k_2,k_3,k_4\ge 0} p^{k_4}x_1^{k_1+k_2+k_3+k_4}x_2^{k_2+k_3+k_4}y^{k_3+k_4}. \end{align*} Let $r = k_3+k_4, a_2=k_2+k_3+k_4=k_2+r, a_1=k_1+k_2+k_3+k_4=k_1+a_2.$ The last sum above becomes \begin{align*} \frac{1}{(1-x_1)(1-x_1x_2)(1-x_1x_2y)(1-px_1x_2y)} &=\sum_{a_1\ge a_2\ge r\ge 0} x_1^{a_1}x_2^{a_2}y^{r}\sum_{k_4=0}^r p^{k_4} \\ &= \sum_{a_1\ge a_2\ge r\ge 0} \frac{p^{r+1}-1}{p-1}x_1^{a_1}x_2^{a_2}y^{r}. \end{align*} Finally, by comparing coefficients, we get the required result: $$h_r^{(\alpha_1,\alpha_2)}(p) = \frac{p^{r+1}-1}{p-1}.$$ \begin{remark} In comparison, we can quickly prove the formulas for all cases of rank 2 using the recurrence used in the proof of Theorem \ref{rank3} as follows. First let's return to the notation of the previous sections, i.e. $a = (a_d, \dots,a_1)$ with $a_1\leq \dots \le a_d.$ Let $\lambda=(a_2, a_1)$ and $\lambda^{'}=( a_1).$\\ \text{Case 1:} $0\le b \leq a_1$ $N_{b}(\lambda)=\sum_{i=0}^{b} p^{i}N_{i}(\lambda^{'})=\sum_{i=0}^{b} p^{i}=\frac{p^{b+1}-1}{p-1}.$\\ Note $N_{i}(\lambda^{'})=1$ or 0 according as $ i \leq a_1$ or $ i> a_1 ,$ respectively.\\ \text{Case 2:} $a_1\le b \leq a_2$ $N_{b}(\lambda)=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{1}+1}^{b} p^{i}N_{i}(\lambda^{'})=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})=\sum_{i=0}^{a_1} p^{i}=\frac{p^{a_1+1}-1}{p-1}.$\\ \text{Case 3:} $a_2\le b \leq a_{1}+a_2$ (This case can also be deduced by replacing $b$ in case 1 by $a_1+a_2-b$) $N_{b}(\lambda)=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{1}+1}^{a_{2}} p^{i}N_{i}(\lambda^{'})+\sum_{i=a_{2}+1}^{b} p^{i}N_{i}(\lambda^{'})-\sum_{i=a_{1}+a_{2}+1-b}^{a_1}p^{i}N_{i}(\lambda^{'})$\\ $=\sum_{i=0}^{a_1} p^{i}N_{i}(\lambda^{'})-\sum_{i=a_{1}+a_{2}+1-b}^{a_1}p^{i}=\frac{p^{a_{1}+a_{2}+1-b}-1}{p-1}.$\\ In summary, we have $$h_b^{(a_2,a_1)}(p)=\begin{cases} \frac{p^{b+1}-1}{p-1} & \text{if } 0\le b \leq a_2\\ \frac{p^{a_2+1}-1}{p-1} & \text{if } a_2\le b \leq a_1\\ \frac{p^{a_1+a_2-b+1}-1}{p-1} & \text{if } a_1\le b \leq a_1+a_2. \end{cases}$$ \end{remark} \textit{Acknowledgement.} The first named author would like to thank the City University of New York HPCC facilities for letting us verify every formula using the SageMath and Mathematica software on their HPC systems based at the College of Staten Island, New York City. \bibliographystyle{hplain}
{ "timestamp": "2018-06-18T02:04:29", "yymm": "1806", "arxiv_id": "1806.03774", "language": "en", "url": "https://arxiv.org/abs/1806.03774", "abstract": "We use recurrence relations to derive explicit formulas for counting the number of subgroups of given order (or index) in rank 3 finite abelian p-groups and use these to derive similar formulas in few cases for rank 4. As a consequence, we answer some questions by M. T$\\ddot{a}$rn$\\ddot{a}$uceanu in \\cite{MT} and L. T$\\dot{\\acute{o}}$th in \\cite{LT}. We also use other methods such as the method of fundamental group lattices introduced in \\cite{MT} to derive a similar counting function in a special case of arbitrary rank finite abelian p-groups.", "subjects": "Group Theory (math.GR); Combinatorics (math.CO)", "title": "Counting subgroups of fixed order in finite abelian groups", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9793540728763411, "lm_q2_score": 0.8175744695262775, "lm_q1q2_score": 0.800694886610274 }
https://arxiv.org/abs/2103.13727
Conway's spiral and a discrete Gömböc with 21 point masses
We show an explicit construction in 3 dimensions for a convex, mono-monostatic polyhedron (i.e., having exactly one stable and one unstable equilibrium) with 21 vertices and 21 faces. This polyhedron is a 0-skeleton, with equal masses located at each vertex. The above construction serves as an upper bound for the minimal number of faces and vertices of mono-monostatic 0-skeletons and complements the recently provided lower bound of 8 vertices. This is the first known construction of a mono-monostatic polyhedral solid. We also show that a similar construction for homogeneous distribution of mass cannot result in a mono-monostatic solid.
\section{Introduction}\label{sec:intro} \subsection{Mono-stability and homogeneous polyhedra} If a rigid body has one single stable position then we call it \emph{mono-stable}, and this property was probably first explored by Archimedes as he developed his famous design for ships \cite{Archimedes}. Mono-stability might also be of advantage for rigid bodies under gravity, supported on a rigid (frictionless) surface, as it facilitates self-righting. Beyond these applications, mono-stable bodies have also attracted considerable mathematical interest. In particular, in case of convex polyhedra with homogeneous mass distribution, it is still unclear what are the minimal numbers $F^S, V^S$ of faces and vertices necessary to achieve mono-stability. Conway and Guy in 1967 \cite{Conway} offered the first upper bound by describing such an object with $F=19$ faces and $V=34$ vertices. The Conway-Guy construction was improved by Bezdek \cite{Bezdek} to $(F,V)=(18,18)$ and later by Reshetov \cite{Reshetov} to $(F,V)=(14,24)$. The mentioned values of $F$ and $V$ define the best known \emph{upper bounds} for a mono-stable polyhedron, so we have $F^S \leq 14 , V^S \leq 18$. Even less is known about the lower bounds: the only known result is due to Conway \cite{Dawson} who proved that a homogeneous tetrahedra have at least two stable equilibria, from which $F^S, V^S \geq 5$ follows. \subsection{Mono-unstable and mono-monostatic homogeneous polyhedra} The natural dual property to being mono-stable is being \emph{mono-unstable}, i.e. to have one single unstable static balance position. The Conway-Guy polyhedron has, beyond the single stable position on one face, 4 unstable equilibria at 4 vertices. The first example for a mono-unstable polyhedron was demonstrated in \cite{balancing}, having $F=18$ faces and $V=18$ vertices and in the same paper it was proven that a homogeneous tetrahedron can not be mono-unstable. Thus, for the minimal numbers $F^U, V^U$ for the faces and vertices that a homogeneous, mono-unstable polyhedron may have, the following bounds apply: $ 5 \leq F^U \leq 18$, $5 \leq V^U \leq 18$. If a rigid body is either mono-stable or mono-unstable then we call it monostatic. If it has both properties, then we call it mono-monostatic. The construction of the first convex, homogeneous, mono-monostatic body called G\"{o}mb\"{o}c \cite{VarkonyiDomokos} in 2006 raised the interest in the subject, because a polyhedral version of the G\"{o}mb\"{o}c is not known. This implies that for the minimal numbers $F^{\star}, V^{\star}$ for the faces and vertices of a mono-monostatic polyhedron the only known bounds are $F^{\star}, V^{\star} \geq 5.$ \subsection{0-skeletons and the main result} Here we highlight a new aspect of this problem: instead of looking at uniform mass distribution, we consider polyhedra with unit masses at the vertices, also called polyhedral 0-skeletons. The latter problem may appear, at first sight, almost `unsportingly' easy. However, the minimal vertex number $V^{\star}_0$ and face number $F^{\star}_0$ to produce a mono-monostatic polyhedral 0-skeleton are not known. Even more curiously, the minimal number of vertices for a mono-monostatic, \emph{polygonal} 0-skeleton (in 2 dimensions) is not known either. The first related results have been reported in \cite{Bozoki} where, for the minimal number of vertices $V^{U}_0$ for a mono-unstable polyhedral 0-skeleton $V^{U}_0 \geq 8$ was proven and this implies the lower bounds $F^{U}_0 \geq 6$ (via the theorem of Steinitz \cite{Steinitz}) and it also implies the bounds $F^{\star}_0 \geq 6, V^{\star}_0 \geq 8$ for mono-monostatic polyhedral 0-skeletons. In this paper we explain the background and show some constructions which may inspire further research. In particular, by providing an explicit construction of a mono-monostatic polyhedral 0-skeleton with 21 faces and 21 vertices, we prove \begin{theorem}\label{th1} $F^{\star}_0, V^{\star}_0 \leq 21.$ \end{theorem} Our example, illustrated in Figure \ref{fig:intro}(c) and defined on line 3 of Table~\ref{tb:monomono}, appears to be the first discrete construction of a mono-monostatic object and it may help to inspire thinking about the bounds $F^{\star}, V^{\star}$ for the homogeneous case. The paper is structured as follows: in Section~\ref{spirals} we explain the geometric idea behind Conway's classical construction and how this idea may be generalized in various directions. In Section~\ref{skeletons}, by relying on an idea by Dawson \cite{Dawson}, we describe the construction for a mono-monostatic 0-skeleton in 2 dimensions, having $V_0=11$ vertices and then we proceed to prove Theorem \ref{th1} by providing the construction of the mono-monostatic 0-skeleton. In Section~\ref{other} we show the connection to other problems, including the mechanical complexity of polyhedra, and also point out why the particular geometry of our constructions may not be applied to the construction of a homogeneous mono-monostatic polyhedron. In Section~\ref{sum} we draw conclusions. \section{The geometry of Conway spirals}\label{spirals} \begin{figure}[ht] \begin{center} \includegraphics[width=\textwidth]{intro.eps} \caption{Construction of symmetric, mono-monostatic discs and polyhedra; a) Geometry of the Conway spiral $P_0 , \ldots, P_n$. $P_0$ is fixed at $z=1$ and each radius $OP_i$ is perpendicular to the corresponding edge $P_{i-1}P_i$. The geometry of the spiral is uniquely described in terms of $n$ angular variables $\alpha_1, \ldots, \alpha_n$; b) 2D mirror-symmetric mono-monostatic polygon with 11 vertices for $n=5$ and $k=2$, see Table~\ref{tb:monomono}, line 6 for numerical data; c) 3D mono-monostatic polyhedron with 5-fold rotational symmetry for $n=4$ and $k=5$, see Table~\ref{tb:monomono}, line 3 for numerical data.} \label{fig:intro} \end{center} \end{figure} \subsection{The classical Conway double spiral and the Conway-Guy monostable polyhedron} The essence of the Conway-Guy polyhedron is a remarkable planar construction to which we will briefly refer as the \emph{Conway spiral}, illustrated in Figure~\ref{fig:intro}(a). In terms of symbols shown in the figure, it can be defined as an open planar polygon $M$ composed of the sequence of points $P_0 , \ldots, P_n$ with $\angle O P_i P_{i-1} = \pi/2$, $i = 1\ldots n$. Without loss of generality, $O$ is considered here as the origin of the coordinate system, all points $P_i$ lie in the plane $xz$ and the coordinates of $P_0$ are fixed at (0,0,1). If we consider double Conway spirals generated by reflection symmetry, for the $x$-coordinate of center of mass $C$ of any double Conway spiral we have $x_C=0$ and due to the special design, the double Conway spiral is monostatic if and only if $z_C<0$. The original Conway-Guy construction is equivalent to Figure~\ref{fig:intro}(a) if all central angles are equal, i.e., we have \begin{equation}\label{eq:conway} \alpha _1 = \alpha _2 =\dots =\alpha _{n+1}, \end{equation} implying that all triangles $P_iP_{i+1}O$ are similar. This case, to which we refer as the \emph{classical Conway spiral} admits a discrete family of shapes, parametrized by the integer $n$, and a corresponding family of double Conway spirals. None of these polygons (interpreted as homogeneous discs rolling along their circumference on a horizontal plane) is monostatic, i.e., we have $z_C>0$ for all values of $n$, since convex monostatic, homogeneous discs do not exist \cite{DomokosRuina}. Still, the Conway spiral may be regarded as a \emph{best shot} at a monostatic polyhedral disc with reflection symmetry. The same intuition suggests that a Conway spiral may need minimal added `bottom weight' to become monostatic. Conway and Guy added this bottom weight by extending the shape into 3D as an oblique prism and they computed the minimal value of $n$ necessary to make this homogeneous oblique prism (with the cross-section of a classical Conway spiral) monostable as $n=8$, resulting in a homogeneous, convex polyhedron with 34 vertices and 19 faces. \subsection{The modified Conway double spiral and Dawson's monostable simplices in higher dimensions} The idea of the Conway spiral may be generalized to bear more fruits. In \cite{Dawson} Dawson, seeking monostatic simplices in higher dimensions, considered the generalized version with \begin{equation}\label{eq:dawson} \alpha_i = c^{i-1}\alpha _1, \quad i=1,2, \dots n \qquad\mbox{and}\qquad \alpha_{n+1}=\alpha_n \end{equation} to which we refer as a \emph{modified Conway spiral}. To describe Dawson's construction we again consider a double spiral, with the mirror images of the vertex $P_i$ defined as $P_{-i}$. In this model the vectors $\mathbf{x}_i=OP_i$, $i=-n, -n+1 \dots n$ are interpreted as the \emph{face vectors} of a simplex ($\mathbf{x}_i$ being orthogonal to the face $f_i$ and having magnitude proportional to the area of $f_i$). To qualify as face vectors, any set of vectors must be balanced \cite{Minkowski}, i.e., we must have \begin{equation}\label{eq:sum} \sum_{i=-n}^{n}\mathbf{x}_i=0. \end{equation} Dawson proved that the condition for the simplex tipping from face $f_i$ to $f_j$ can be written as \begin{equation}\label{dawson1} \lvert\mathbf{x}_i\rvert < \lvert\mathbf{x}_j\rvert\cos\theta_{ij}, \end{equation} where $\theta_{ij}$ is the angle between $\mathbf{x}_i$ and $\mathbf{x}_j$. By using this \emph{tipping condition} he found that for $n=5, c=1.5$ the modified Conway spiral (\ref{eq:dawson}) yields a set of balanced vectors, the small perturbation of which defines a 10-dimensional, homogeneous mono-stable simplex. \section{Mono-monostatic 0-skeletons}\label{skeletons} \subsection{The generalized double Conway spiral and planar 0-skeletons}\label{ss:11gon} If, instead of considering double Conway spirals as homogeneous disks we associate unit masses with the vertices then we obtain objects which may be called \emph{polygonal 0-skeletons}. Since there are relatively many vertices with negative $z$ coordinate and relatively few ones with positive $z$ coordinate, this interpretation appears to be a convenient manner to add `bottom weight' to the geometric double Conway spiral. In this interpretation as planar 0-skeletons, one may ask whether mono-monostable double Conway spirals exist and if yes, what is the minimal number of their vertices necessary to have this property. Since static balance equations for such a skeleton coincide with (\ref{eq:sum}) and the tipping condition (\ref{dawson1}) is equivalent to prohibit an unstable equilibrium at vertex $v_i$ \cite{Bozoki}, it is easy to see that Dawson's geometric construction, interpreted as a 0-skeleton, has $z_C<0$ and it defines a polygon with $V=11$ vertices which is mono-monostatic. One can ask whether this construction is optimal in two ways: whether there exists a smaller value of $n$ which defines a mono-monostatic modified double Conway spiral (interpreted as a 0-skeleton) and whether by keeping $n=5$, one may pick other values for $\alpha_i$ which yield a center of mass with larger negative coordinate. The first question was answered in \cite{Dawson3} in the negative by proving that monostable simplices in $d<9$ dimensions do not exist. This implies that for $n<5$ no mono-monostatic Conway spiral (interpreted as a 0-skeleton) exists, but nothing is known about the existence of mono-monostatic 10-gonal disks as 0-skeletons since they cannot be represented by a symmetric double Conway spiral. The second question may be addressed if we admit \emph{generalized} Conway spirals with arbitrary $\alpha_i$ and optimize this construction to seek the minimum of $z_C$. In any case, to verify monostatic property of a given double Conway spiral, $z_C$ needs to be computed. In terms of coordinates $z_i$, we have from Figure~\ref{fig:intro}(a): \begin{equation} \label{eq:z_C} z_C = \dfrac{1+k \displaystyle\sum_{i=1}^{n} z_i}{1+kn}, \end{equation} where $k$ stands for the multiplicity of Conway spirals; now $k=2$. Furthermore, any $z_i$ can be expressed in terms of angles $\angle P_0 O P_i = \sum_{j=1}^i \alpha_j$ and distances $$r_i = \overline{OP_i} = \overline{OP_0}\cdot\prod_{j=1}^i \cos\alpha_j$$ as follows: \begin{equation} \label{eq:z_i} z_i = \prod_{j=1}^i \cos\alpha_j\cdot\cos\left(\sum_{j=1}^i \alpha_j\right). \end{equation} By merging (\ref{eq:z_C}) and (\ref{eq:z_i}) we get \begin{equation} \label{eq:z_prodsum} z_C(\boldsymbol{\alpha}) = \dfrac{1+k \displaystyle\sum_{i=1}^{n} \prod_{j=1}^i \cos\alpha_j\cdot\cos\left(\sum_{j=1}^i \alpha_j\right)}{1+kn}, \end{equation} or briefly, \begin{equation} \label{eq:z_C_simp} z_C(\boldsymbol{\alpha}) = \dfrac{1+k S_n(\boldsymbol{\alpha})}{1+kn}. \end{equation} We performed an optimization for $\boldsymbol{\alpha} = (\alpha_1 \ldots \alpha_n)$ and found the shape in Figure~\ref{fig:intro}(b) (see Table~\ref{tb:monomono}, line 6 for computed values of $\boldsymbol{\alpha}$). Note that this single result is an alternative proof for the existence of monostable $10$-dimensional simplices given by Dawson \cite{Dawson}. We remark that a similar optimization process of the Conway spiral is discussed in \cite{Minich} for the homogeneous case. \subsection{Proof of Theorem \ref{th1}: Conway $k$-spirals and mono-monostatic 0-skeletons in 3 dimensions}\label{ss:21hedron} \begin{proof} Generalized Conway spirals may be used as the building blocks of mono-monostatic 0-skeletons in 3 dimensions. The key idea is to consider instead of a double Conway spiral \emph{multiple} Conway spirals in a $D_k$-symmetrical arrangement around the $z$-axis, rotated at angles $\beta=2\pi/k$. We call such a construction a Conway $k$-spiral. Planar double spirals correspond to $k=2$, while for higher values of $k$ one may seek to find mono-monostatic 0-skeletons. If for $k=2$ the Conway spiral defines a mono-monostatic planar 0-skeleton then we expect that for higher values of $k$ we will obtain mono-monostatic polyhedral 0-skeletons. The procedure of finding mono-monostatic Conway $k$-spirals (interpreted as 0-skeletons) is as follows: Let us consider a planar polygonal line $M$ as the intersection of a symmetry plane bisecting a sequence of faces, while another polygonal line $N(Q_0, \ldots,Q_n)$ remains on a sequence of edges as before. \begin{figure}[!ht] \begin{center} \includegraphics[scale=0.6]{revol.eps} \caption{Construction of polyhedra with rotational symmetry: side and top views. Polygonal lines $M(P_0, \ldots,P_n)$ (solid line) and $N(Q_0, \ldots,Q_n)$ (dashed line) lie in symmetry planes through faces and edges, respectively. Any optimal construction requires $Q_0Q_1$ instead of $P_0P_1$ to be perpendicular to radius $OQ_1$.} \label{fig:revol} \end{center} \end{figure} Let $e_i$ be an edge $Q_iQ_{i+1}, i=0\ldots n-1$ and face $f_i$ be adjacent to $e_i$. Call face $f_i$ (edge $e_i$) `outwards' if its upper edge (endpoint) is farther from the axis of symmetry than the bottom one, i.e., for a face $f_i$, $\sum_{j=i+2}^{n+1} \alpha_j \leq \pi/2$. Clearly, $e_i$ is outwards if and only if $f_i$ does. By construction, $e_0$ and $f_0$ are never outwards but we assume from now on that any $e_i$, $f_i$ with $i>0$ are outwards edges and faces. For them it is clear that $\angle OQ_{i+1}Q_i > \angle OP_{i+1}P_i$ and if this latter equals $\pi/2$, there will be no equilibrium points inside $f_i$. Non-outwards edges, however, are just on the contrary and therefore an optimal construction for the entire polyhedron requires $\angle OQ_{i+1}Q_i = \pi/2$, causing the top vertex $Q_0$ to be moved up by a positive distance $h$ as shown in Figure~\ref{fig:revol}. It is easy to read from the right triangle $OP_0P_1$ that $z_1 = \cos^2\alpha_1$ and $x_1 = \sin\alpha_1 \cos\alpha_1$. Let the distance between $z$ and $Q_1$ (also between $z$ and $Q'_1$ in the figure) be denoted by $x'_1$. Since $x_1 = x'_1 \cos(\pi/k)$ (see the top view) and $OQ'_1Q_0$ is also a right triangle, for its height of length $x'_1$ the following equality holds: \begin{equation*} \label{eq:height} \cos^2\alpha_1 (\sin^2\alpha_1 + h) = \left(\dfrac{\sin\alpha_1 \cos\alpha_1}{\cos(\pi/k)}\right)^2, \end{equation*} which yields \begin{equation*} \label{eq:h} h = \sin^2\alpha_1 \tan^2\dfrac{\pi}{k}. \end{equation*} Since it affects the vertical position of the top vertex and thus of the centroid, (\ref{eq:z_C_simp}) should be modified as \begin{equation} \label{eq:z_C_ast} z^\ast_C(\boldsymbol{\alpha}) = \dfrac{1+k S^\ast_n(\boldsymbol{\alpha})}{1+kn}, \end{equation} where \begin{equation} \label{eq:S_n_ast} S^\ast_n(\boldsymbol{\alpha}) = S_n(\boldsymbol{\alpha}) + \dfrac{1}{k} \sin^2\alpha_1 \tan^2\dfrac{\pi}{k}. \end{equation} We performed calculations in search of minimum $z^\ast_C$ that lead to different constructions (denoted as $P_{n,k} ), one of these constructions with $n=4 , k=5$ is illustrated in Figure~\ref{fig:intro}(c). Table~\ref{tb:monomono} summarizes the possible mono-monostatic objects with minimum required $k$ found by the above method ($v = kn+1$ stands for the number of vertices or/and faces): \begin{table}[!ht] \begin{center} \begin{tabular}{|c|ccclp{7.7cm}|} \hline no. & $n$ & $k$ & $v$ & $z_C$ & $(\alpha_{n+1},\alpha_n, \ldots, \alpha_1)$ \\ \hline \hline 1 & 2 & 25 & 51 & -0.00051277 & $(49.799, 49.799, 80.402)^\circ$ \\ \hline 2 & 3 & 8 & 25 & -0.0061413 & $(30.273, 30.273, 46.543, 72.912)^\circ$ \\ \hline 3 & 4 & 5 & 21 & -0.015354 & $(19.716, 19.716, 29.875, 44.519, 66.173)^\circ$ \\ \hline 4 & 5 & 4 & 21 & -0.029972 & $(13.494, 13.494, 20.336, 29.781, 43.215, 59.680)^\circ$ \\ \hline 5 & 7 & 3 & 22 & -0.042695 & $( 7.1815, 7.1815, 10.7864, 15.6392, 22.1409,$ $30.9129, 43.0793, 43.0788)^\circ$ \\ \hline 6 & 5 & 2$^\ast$ & 11 & -0.017984 & $( 13.201, 13.201, 19.890, 29.110, 42.172, 62.427)^\circ$ \\ \hline \end{tabular} \vskip 5mm \caption{List of some mono-monostatic 0-skeletons $P_{n,k}$ with $D_k$-symmetry and $v = nk+1$ vertices; $z_C$ can be verified via (\ref{eq:z_prodsum}). $k=2$ marked by `$\ast$' is the two -dimensional case already mentioned at the end of Subsection~\ref{ss:11gon}. The minimum number of vertices for monostatic 3D rotational polyhedra is 21.} \label{tb:monomono} \end{center} \end{table} \end{proof} We believe that this construction is close to a (local) optimum, i.e., we think that this may be the mono-monostatic 0-skeleton defined by multiple generalized Conway spirals which has the least number of vertices. This, however, does not exclude the existence of mono-monostatic 0-skeletons with smaller number of vertices which have less symmetry. Our construction provides 21 as an \emph{upper bound} for the minimal number of vertices and faces of a mono-monostatic 0-skeleton. The lower bound for the number of vertices was given in \cite{Bozoki} as 8, from which a lower bound of 6 for the number of faces follows \cite{Steinitz1}. \section{Connection to related other problems}\label{other} \subsection{Mechanical complexity of polyhedra} It is apparent that constructing monostatic polyhedra is not easy. In \cite{balancing} this general observation was formalized by introducing the \emph{mechanical complexity} $C(P)$ of a polyhedron $P$ as \begin{equation}\label{complexity} C(P) = 2(V(P) + F(P) - S(P) - U(P)), \end{equation} where $V(P),F(P),S(P),U(P)$ stand for the number of vertices, faces, stable and unstable equilibrium points of $P$, respectively. The \emph{equilibrium class} of polyhedra with given numbers $S,U$ of stable and unstable equilibria is denoted by $(S,U)^E$ and the complexity of such class was defined as \begin{equation}\label{complexity1} C(S,U)=min\{C(P): P \in (S,U)^E\}. \end{equation} The only material distribution considered in \cite{balancing} was uniform density. Other types of homogeneous mass distributions, commonly referred to as \emph{h-skeletons} are also possible: 0-skeletons have mass uniformly distributed on their vertices, 1-skeletons have mass uniformly distributed on the edges, 2-skeletons have mass uniformly distributed on the faces. To distinguish between these cases we will apply an upper index to the symbol $C$ of complexity, indicating the type of skeleton (the absence of index indicates classical homogeneity). In the case of uniform density (classical homogeneity), the complexity for all non-monostatic equilibrium classes $(S,U)^E$ for $S,U>1$ has been computed in \cite{balancing}. On the other hand, the complexity has not yet been determined for any of the monostatic classes $(1,U)^E, (S,1)^E$. Lower and upper bounds exist for $C(S,1),C(1,U)$ for $S,U>1$. The most difficult appears to be the mono-monostatic class $(1,1)^E$ for the complexity $C(1,1)$ of which the prize USD $1.000.000/C(1,1)$ has been offered in \cite{balancing}. Not only is $C(1,1)$ unknown, at this point there is no upper bound known either. \subsection{Complexity of some monostable and mono-unstable polyhedral 0-skeletons}\label{ss:complexity} Admittedly, computing upper bounds for 0-skeletons is easier. This is already apparent in the planar case, where monostatic discs with homogeneous mass distribution in the interior do not exist \cite{DomokosRuina} whereas a monostatic 0-skeleton could be constructed with $V=11$ vertices \cite{Dawson}. In 3D, our construction of a 0-skeleton with $F=21$ faces and $V=21$ vertices (see the top left polyhedron in Figure~\ref{fig:compl} and Table~\ref{tb:monomono}, line 3) offers such an upper bound as \begin{equation}\label{bound} C^0(1,1)\leq 2(21+21-1-1) =80 \end{equation} This is the first known such construction and its existence may help to solve the more difficult cases, in particular, the case with uniform density. In Figure~\ref{fig:compl} we provide upper bounds for the complexity of 0-skeletons in some other monostatic equilibrium classes as well. \begin{figure}[ht] \begin{center} \includegraphics[width=\textwidth]{compl.eps} \caption{Complexity of some monostable and mono-unstable polyhedra. Drawn representatives of equilibrium classes $(S,U)$ prove an upper bound for complexity of the respective class, see the bracketed numbers as lower and upper bounds, respectively, in the top left corner of their cells. Since mono-unstable polyhedra with less than 8 vertices (and therefore, by Steinitz's theorem, with less than 6 faces) cannot exist, 24 is a lower bound of complexity of classes $(S,1)$. Complexity of the four non-monostatic classes is exactly known by the existence of simplicial representatives of each class \cite{balancing}. Coordinates of drawn polyhedra, except for the one in class (1,1), are given in Table~\ref{tb:coord}. } \label{fig:compl} \end{center} \end{figure} \subsection{Existence and non-existence of certain types of mono-monostatic 0-skeletons and homogeneous bodies}\label{ss:existence} The following paragraphs illustrate the relative difficulty of constructing mono-monostatic $h$-skeletons from a different point of view. Firstly, it is known from \cite{DomokosRuina} that no homogeneous mono-monostatic two-dimensional objects rolling along their perimeter exist; however, $P_{5,2}$ drawn in Figure~\ref{fig:intro}b is a mono-monostatic 0-skeleton in 2D. A similar property of non-existence of homogeneous mono-monostatic objects will be proven below for Conway $k$-spirals, interpreted as homogeneous solids. \begin{theorem} \label{thm:nonexst} Let $P$ be a convex solid with center of mass at $C$. Let $a$ denote an axis intersecting $P$ and let $h(a)$ be a half-plane the boundary of which is $a$. Let $N$ denote the intersection of $P$ and $h(a)$ and let us describe $N$ as the polar distance $r(\varphi)$, measured from $C$ as origin. If there exists an axis $a$ such that $r(\varphi)$ is strictly monotonic for all possible $h(a)$ then $P$ is not mono-monostatic. \end{theorem} \begin{proof} Let an axis $z$ be directed along $a$ and let a point $Q$ on the surface of $P$ be parametrized as $Q(\theta, \varphi,r)$ where $0 \leq\theta\leq \pi$ is the meridian angle between $CQ$ and $z$, $0\leq\varphi< 2\pi$ is the azimuth angle (with respect to a fixed starting position), $r=\lvert Q-C \rvert$. Since $P$ is convex, $r = r(\theta, \varphi)$ for all surface points is uniquely defined. In this polar system, $C$ can only be the centre of mass of $P$ if \begin{equation} \label{eq:Cheight} \int\limits_{0}^{2\pi} \int\limits_{0}^{\pi} \dfrac{2}{9} r(\theta, \varphi)^4 \sin\theta \cos\theta d\theta d\varphi = 0, \end{equation} once $(1/3)r^3 \sin\theta d\theta d\varphi$ is the volume of an elementary pyramid with its apex at $C$ and $(2/3)r\cos \theta$ measures the $z$ coordinate for the centre of mass of an elementary pyramid. From the condition of the theorem, it follows that $r$ is strictly monotonic in $\theta$: assume now that $\theta_1 < \theta_2 \iff r_1 > r_2$ for all $Q_1(\theta_1, \varphi, r_1), Q_2(\theta_2, \varphi, r_2)$ and rewrite (\ref{eq:Cheight}) as follows: \begin{equation*} \dfrac{2}{9} \int\limits_{0}^{2\pi} \int\limits_{0}^{\pi/2} \left(r(\theta, \varphi)^4 \sin\theta \cos\theta + r(\pi-\theta, \varphi)^4 \sin(\pi-\theta) \cos(\pi-\theta) \right) d\theta d\varphi = 0 \end{equation*} \begin{equation} \dfrac{1}{9} \int\limits_{0}^{2\pi} \int\limits_{0}^{\pi/2} \left(r(\theta, \varphi)^4 - r(\pi-\theta, \varphi)^4 \right) \sin 2\theta d\theta d\varphi = 0. \end{equation} Here both terms of the product in the integrand are positive, so the definite integral cannot evaluate to zero. \end{proof} \begin{cor} Conway $k$-spirals, interpreted as homogeneous solids, are never mono-monostatic. \end{cor} \begin{proof} We prove the Corollary by showing that a Conway $k$-spiral satisfies the monotonicity condition of the theorem. Consider $a$ to be aligned with axis $z$ again. Since we consider polyhedral solids, the `level sets' for $r$ are concentric circles on all faces. By construction, perpendicular projection of $C$ on the base $k$-gon is incident to $a$, so $r$ increases monotonically within that $k$-gon along any $h$. For all other faces, assume that there is a plane $h$ intersecting or being tangent to a level set, but it would immediately imply that a non-horizontal edge (surely contained by some plane $h$) of the same face carries an equilibrium point which contradicts the mono-monostatic property. As a consequence, any $h$ intersects all set levels without being even tangent to any of them, which is a necessary and sufficient condition for $r$ being strictly monotonic along any line $N$ started and ended at the axis $a$. \end{proof} We note that Theorem~\ref{thm:nonexst} also implies that any homogeneous smooth solid of revolution cannot be mono-monostatic. \section{Concluding comments}\label{sum} In this paper, by relying on the geometric idea of Conway spirals, we demonstrated the existence of mono-monostatic 0-skeletons in two and three dimensions. In the former case, by drawing on an earlier result of Dawson \cite{Dawson} we showed that mono-monostatic planar 0-skeletons with $V=11$ vertices exist. It follows from another result of Dawson \cite{Dawson3} that for $V=9$ such constructions can not exist The $V=10$ case is not known. In three dimensions we showed an explicit construction with $V=21$ vertices, thus providing an upper bound for the minimal number of vertices. The lower bound is $V=8$ \cite{Bozoki} and other results are not known. We hope that these constructions will motivate further research to find the minimal number of $V$ for a mono-monostatic 0-skeleton, both in two and in 3 dimensions. \begin{table}[ht] \begin{center} \begin{scriptsize} \begin{tabular}{|r|r|r|} \hline \multicolumn{3}{|c|}{$(S,U) = (1,2)$}\\ $x$ & $y$ & $z$ \\ \hline 0 & 374 & 0 \\ 154 & 80 & 0 \\ 124 & -32 & 0 \\ 81 & -78 & 0 \\ 47 & -95 & 0 \\ 24 & -100 & 0 \\ -24 & -100 & 0 \\ -47 & -95 & 0 \\ -81 & -78 & 0 \\ -124 & -32 & 0 \\ -154 & 80 & 0 \\ 0 & -1200 & 5000 \\ \hline \end{tabular} \begin{tabular}{|r|r|r|} \hline \multicolumn{3}{|c|}{$(S,U) = (1,3)$}\\ $x$ & $y$ & $z$ \\ \hline 0 & 466 & 0 \\ 166 & 70 & 0 \\ 121 & -47 & 0 \\ 71 & -87 & 0 \\ 35 & -100 & 0 \\ -35 & -100 & 0 \\ -71 & -87 & 0 \\ -121 & -47 & 0 \\ -166 & 70 & 0 \\ 0 & -100 & -900 \\ 0 & -100 & 900 \\ \hline \multicolumn{3}{c}{\phantom{0} } \end{tabular} \vskip 5mm \begin{tabular}{|r|r|r|} \hline \multicolumn{3}{|c|}{$(S,U) = (2,1)$}\\ $x$ & $y$ & $z$ \\ \hline 0 & 374.328 & 0 \\ 153.589 & 80.2023 & 20 \\ 124.268 & -32.3675 & 14.9819 \\ 81.1006 & -77.5258 & 8.45141 \\ 46.9121 & -94.4981 & 3.41302 \\ 23.4562 & -100 & 0 \\ -23.4562 & -100 & 0 \\ -46.9121 & -94.4981 & 3.41302 \\ -81.1006 & -77.5258 & 8.45141 \\ -124.268 & -32.3675 & 14.9819 \\ -153.589 & 80.2023 & 20 \\ \hline \multicolumn{3}{c}{\phantom{0} } \end{tabular} \begin{tabular}{|r|r|r|} \hline \multicolumn{3}{|c|}{$(S,U) = (3,1)$}\\ $x$ & $y$ & $z$ \\ \hline 0 & 334.907 & 0 \\ 145.019 & 83.7267 & 10 \\ 145.019 & 0 & 9.6018 \\ 94.9161 & -68.9606 & 5.40618 \\ 53.5898 & -92.8203 & 2.10256 \\ 26.7949 & -100 & 0 \\ -26.7949 & -100 & 0 \\ -53.5898 & -92.8203 & 2.10256 \\ -94.9161 & -68.9606 & 5.40618 \\ -145.019 & 0 & 9.6018 \\ -145.019 & 83.7267 & 10 \\ \hline \multicolumn{3}{c}{\phantom{0} } \end{tabular} \end{scriptsize} \caption{Coordinates of some polyhedra shown in Figure~\ref{fig:compl}. Monostable objects are provided with integer coordinates which would be difficult for mono-unstable ones due to oblique polygonal faces.} \label{tb:coord} \end{center} \end{table}
{ "timestamp": "2021-10-14T02:23:03", "yymm": "2103", "arxiv_id": "2103.13727", "language": "en", "url": "https://arxiv.org/abs/2103.13727", "abstract": "We show an explicit construction in 3 dimensions for a convex, mono-monostatic polyhedron (i.e., having exactly one stable and one unstable equilibrium) with 21 vertices and 21 faces. This polyhedron is a 0-skeleton, with equal masses located at each vertex. The above construction serves as an upper bound for the minimal number of faces and vertices of mono-monostatic 0-skeletons and complements the recently provided lower bound of 8 vertices. This is the first known construction of a mono-monostatic polyhedral solid. We also show that a similar construction for homogeneous distribution of mass cannot result in a mono-monostatic solid.", "subjects": "Metric Geometry (math.MG); Combinatorics (math.CO)", "title": "Conway's spiral and a discrete Gömböc with 21 point masses", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9793540692607816, "lm_q2_score": 0.817574471748733, "lm_q1q2_score": 0.8006948858308556 }
https://arxiv.org/abs/1401.6619
On cycles in intersection graph of rings
Let $R$ be a commutative ring with non-zero identity. We describe all $C_3$- and $C_4$-free intersection graph of non-trivial ideals of $R$ as well as $C_n$-free intersection graph when $R$ is a reduced ring. Also, we shall describe all complete, regular and $n$-claw-free intersection graphs. Finally, we shall prove that almost all Artin rings $R$ have Hamiltonian intersection graphs. We show that such graphs are indeed pancyclic.
\section{Introduction} If $S=\{S_1,\ldots,S_n\}$ is a family of sets, then the intersection graph of $S$, is the graph having $S$ as its vertex set with $S_i$ adjacent to $S_j$ if $i\neq j$ and $S_i\cap S_j\neq\emptyset$. A well-know theorem due to Marczewski \cite{tam-frm} states that all graphs are intersection graph. An interesting case of intersection graphs is when the members of $S$ have an algebraic structure. Bosak \cite{jb} was the first who studied graphs arising from semigroups. Cs\'{a}k\'{e}any and Poll\'{a}k \cite{bc-gp} defined and studied the intersection graphs of nontrivial proper subgroups of groups. Zelinka \cite{bz} continued the work of Cs\'{a}k\'{e}any and Poll\'{a}k on intersection graphs of subgroups of finite abelian groups, and later Shen \cite{rs} studies such graphs and classifies all finite groups whose intersection graphs of nontrivial subgroups are disconnected. Herzog, Longobardi and Maj \cite{mh-pl-mm} study the intersection graphs of maximal subgroups of finite groups and among other results classify all finite groups with disconnected graph. The same as for groups, the intersection graphs of ideals of rings and subspaces of vector spaces have been discussed in \cite{ic-sg-tkm-mks,shj-njr-1,shj-njr-2}. Let $R$ be a commutative ring with a non-zero identity. The intersection graph of $R$, denoted by $\Gamma(R)$, is a graph whose vertices are the nontrivial ideals of $R$ and two distinct vertices are joined by an edge if the corresponding ideals of $R$ have a non-zero intersection. In this paper, we study the cycle structure of intersection graphs. First we classify all Artin rings with a regular (hence complete) intersection graph. Next we shall investigate all rings $R$ whose intersection graphs $\Gamma(R)$ do not have an induced cycle of length $3$ or $4$. Also, we show that if $R$ is a reduce ring, then $\Gamma(R)$ is $C_n$-free $(n\geq5)$ if and only if $R$ has no ideal which is the direct sum of $n$ non-zero ideals. The same result is also established for $n$-claws instead of $n$-cycles. In the last section, we shall prove that except few cases all other Artin rings have Hamiltonian intersection graphs. Using simple modifications of the given Hamiltonian cycle, we show that $\Gamma(R)$ is pancyclic whenever it is Hamiltonian. Recall that an $n$-claw (a claw) is the star graph $K_{1,n}$ ($K_{1,3}$). Also a graph is called pancyclic if it contains cycles of possible arbitrary sizes $\geq3$. The following theorem will be used without further reference. \begin{recalltheorem}[{\cite[Theorem VI.2]{brm}}] Let $R$ be an Artin commutative ring with a non-zero identity. Then \[R=R_1\oplus\cdots\oplus R_n,\] where $R_1,\ldots,R_n$ are local rings. \end{recalltheorem} If $R$ is a ring, then the ideals $\a_1,\ldots,\a_n$ are called \textit{independent} if \[\a_i\cap(\a_1+\cdots+\a_{i-1}+\a_{i+1}+\cdots+\a_n)=0\] for $i=1,\ldots,n$. In other words, $(\a_1,\ldots,\a_n)=\a_1\oplus\cdots\oplus\a_n$ is the direct sum of $\a_1,\ldots,\a_n$. All rings in this paper are commutative rings with a non-zero identity. \section{$C_n$-free intersection graphs} As a most simple property we may investigate on intersection graph $\Gamma(R)$ of ideals of a ring, is whether $\Gamma(R)$ is a complete graph. We show that the class of Artin rings with a complete intersection graph coincides with the class of Artin rings with a regular intersection graph and then characterize all such rings. \begin{theorem} Let $R$ be an Artin ring, which is not a direct sum of two fields. If $\Gamma(R)$ is regular, then it is complete. \end{theorem} \begin{proof} First we show that $R$ has no direct factor, which is a field. If $R=S\oplus F$, where $F$ is a field and $S$ is not a field, then $N_{\Gamma(R)}(F)=\{\a\oplus F:0\neq\a\vartriangleleft S\}$ and $N_{\Gamma(R)}(S)=\{\a,\a\oplus F:0\neq\a\vartriangleleft S\}$. Hence, $\deg_{\Gamma(R)}S>\deg_{\Gamma(R)}F$, which is a contradiction. Therefore, each maximal ideal of $R$ is adjacent to all other vertices of $\Gamma(R)$, from which it follows that $\Gamma(R)$ is a complete graph. \end{proof} \begin{theorem} If $R$ is an Artin ring, then the graph $\Gamma(R)$ is complete if and only if there exists a sequence of rings $R_1,\ldots,R_n$, in which $R=R_1$, $(R_i,R_{i+1})$ is a local ring for all $i=1,\ldots,n-1$ and $R_n$ is a field. \end{theorem} \begin{proof} If $R$ is not a local ring, then $R=S\oplus T$ for some non-zero rings $S$ and $T$. But then $S\cap T=0$, which is a contradiction. Thus $R=(R,\mathfrak{m})$ is a local ring. Continuing this way for $\mathfrak{m}$ instead of $R$ the result follows. The converse is obvious. \end{proof} In the following two theorems, we shall consider conditions under which the intersection graph of a ring is a star graph, which also results in a characterization of rings with a bipartite intersection graph. \begin{theorem}\label{pendant} Let $R$ be a ring, which is neither a direct sum of two fields nor a direct sum of a field with a local ring $(S,\mathfrak{m})$ such that $\mathfrak{m}$ is a field. If $\Gamma(R)$ has a pendant, then $\Gamma(R)$ is a star graph. \end{theorem} \begin{proof} Let $\a\in V(\Gamma(R))$ be a pendant. If $\a$ is a maximal ideal, then it is easy to see that $R=(R,\a)$ is a local ring and $\a=(x)$ is a principal ideal. Let $\b$ be the ideal of $R$ adjacent to $\a$. Then $\b=(x^2)$ and $(x^3)=0$, hence $\Gamma(R)$ is an edge. If $\a$ is not a maximal ideal, then there exists a unique maximal ideal $\mathfrak{m}$ of $R$ containing $\a$. Clearly, $\a=(x)$ is principal and $(x^2)=0$. If $R$ is not a local ring, then there exists a maximal ideal $\mathfrak{n}$ such that $\a\cap\mathfrak{n}=0$. Thus $R=\a\oplus\mathfrak{n}$ and $\a$ is a field. Then $\mathfrak{m}=\a+\b$ for some ideal $\b$ of $\mathfrak{n}$. But then $(\mathfrak{n},\b)$ is a local ring such that $\b$ is a field, which is a contradiction. Therefore $(R,\mathfrak{m})$ is a local ring. Clearly, $\mathfrak{m}=(x,y)$ for some $y\in R$. If $\mathfrak{m}$ is principal, then we may assume that $\mathfrak{m}=(y)$. Thus $(y^2)=(x)$ and $(y^3)=0$, which implies that $\Gamma(R)$ is an edge. If $\mathfrak{m}$ is not principal, then $(x)\cap(y)=0$ and consequently $xy=0$. Since $(x)\subseteq (x)+(y^2)\subseteq (x)+(y)=\mathfrak{m}$, it follows that $(y^2)=0$. Hence $\mathfrak{m}^2=0$ so that $\mathfrak{m}$ is a vector space over the field $F=R/\mathfrak{m}$, where the multiplication is defined by $(r+\mathfrak{m})\cdot m=rm$ for all $r\in R$ and $m\in\mathfrak{m}$. Clearly, there is a one to one correspondence between ideals of $R$ contained in $\mathfrak{m}$ and subspaces of $(\mathfrak{m},F)$. Hence $\dim_F\mathfrak{m}=2$ so that $\Gamma(R)$ is a star graph. \end{proof} \begin{theorem}\label{triangle} \label{main} If $\Gamma(R)$ is triangle-free, then $\Gamma(R)$ is star or two isolated vertices. \end{theorem} \begin{proof} If $R$ is not a local ring, then there exist two distinct maximal ideals $\mathfrak{m}_1$ and $\mathfrak{m}_2$ in $R$. Since $\Gamma(R)$ is triangle-free, we should have $\mathfrak{m}_1\cap\mathfrak{m}_2=0$. Hence $R=\mathfrak{m}_1\oplus\mathfrak{m}_2$. Let $F_1=R/\mathfrak{m}_1$ and $F_2=R/\mathfrak{m}_2$. Then $R\cong F_1\oplus F_2$ and $\Gamma(R)$ is the union of two isolated vertices. Now, suppose that $(R,\mathfrak{m})$ is a local ring. We have two cases for $\mathfrak{m}$. Case 1: $\mathfrak{m}$ is not a principal ideal. First we show that $\mathfrak{m}^2=0$. If $x\in\mathfrak{m}$ and $y\in \mathfrak{m}\setminus xR$, then $ xR\cap yR=0$. Thus $xy=0$ so that $(\mathfrak{m}\setminus xR)x=0$. On the other hand, if $r\in R$, then $xr=y+(xr- y)$ so that $xrx =0$. Thus $xRx=0$ and consequently $\mathfrak{m} x=0$. Hence $\mathfrak{m}^2=0$. Let $F=R/\mathfrak{m}$. The same as in the proof of Theorem \ref{pendant}, $(\mathfrak{m},F)$ is a vector space. If $\dim_F\mathfrak{m}\geq3$ and $\lbrace x, y, z\rbrace$ is an independent set in $(\mathfrak{m},F)$, then the set of ideals $\{(x),(x,y),(x,y,z)\}$ induces a triangle in $\Gamma(R)$, which is a contradiction. Thus $\dim_F\mathfrak{m}\leq2$. If $\dim_F\mathfrak{m}=2$, then every two distinct non-trivial ideals of $R$ different from $\mathfrak{m}$ are disjoint. Thus $\Gamma(R)$ is a star graph with $\mathfrak{m}$ at the center. If $\dim_F\mathfrak{m}=1$, then $\Gamma(R)$ is a single vertex and we are done. Case 2: $\mathfrak{m}=xR$ is a principal ideal. Let $\a$ be a non-zero ideal of $R$. Then $ \a\subseteq\mathfrak{m}$. If $y\in\a$, then $y=rx$ for some $r \in R$. If $r$ is a unit, then $x=yr^{-1} \in \a$ and hence $ \a= \mathfrak{m} $. If $\a\neq\mathfrak{m}$, then $r$ is not unit and so $r=sx$, for some $s\in R$. Thus $y=sx^2$ and subsequently $\a\subseteq\mathfrak{m}^2\subseteq\mathfrak{m} $. Since $\Gamma(R)$ is triangle free, it follows that $\a= \mathfrak{m}^2$. Therefore $\Gamma(R)$ is either a single vertex when $\mathfrak{m}=\mathfrak{m}^2$ or it is an edge when $\mathfrak{m}\neq\mathfrak{m}^2$. \end{proof} The following corollary is a direct consequence of the preceding two theorems. \begin{corollary} Let $R$ be a ring, which is neither a direct sum of two fields nor a direct sum of a field with a local ring $(S,\mathfrak{m})$ such that $\mathfrak{m}$ is a field. Then the following conditions are equivalent: \begin{itemize} \item[(1)]$\Gamma(R)$ is triangle-free, \item[(2)]$\Gamma(R)$ has a pendant, \item[(3)]$\Gamma(R)$ is bipartite. \item[(4)]$\Gamma(R)$ is star. \end{itemize} \end{corollary} In what follows, we shall concentrate on cycle structure of intersection graphs and give a characterization of almost all intersection graphs under investigation that do not have an induced cycle of length greater than $3$. \begin{theorem} The graph $\Gamma(R)$ is $C_{4}$-free if and only if $R$ has no set of four non-zero independent ideals. \end{theorem} \begin{proof} First suppose that $R$ has an ideal which is a direct sum of four non-zero ideals, namely $\a_1,\a_2,\a_3$ and $\a_4$. Then $\a_1\oplus\a_2$, $\a_2\oplus\a_3$, $\a_3\oplus\a_4 ,\a_4\oplus\a_1$ induces a cycle of length $4$ in $\Gamma(R)$. Conversely, suppose that $R$ has an induced $4$-cycle with vertices $\a_1,\a_2,\a_3$ and $\a_4$. Then $\a_1\cap\a_3=\a_2\cap\a_4=0$. Since $\a_2\cap\a_3+\a_3\cap\a_4\subseteq\a_3$, we have \begin{align*} (\a_1\cap\a_2)\cap(\a_2\cap\a_3+ \a_3\cap\a_4+\a_4\cap\a_1)&\subseteq(\a_1\cap\a_2)\cap(\a_3+(\a_4\cap\a_1))\\ &=(\a_1\cap\a_2)\cap (\a_3\oplus\a_4\cap\a_1). \end{align*} If $a+b\in (\a_1\cap\a_2)\cap(\a_3\oplus\a_4\cap\a_1)$, where $a\in\a_3$ and $b\in\a_4\cap\a_1$, then $a\in\a_1$, which implies that $a=0$. Then $b\in\a_2$ and similarly $b=0$. Hence \[(\a_1\cap\a_2)\cap(\a_2\cap\a_3+\a_3\cap\a_4+\a_4\cap\a_1)=0.\] Similar arguments show that $(\a_1\cap\a_2),(\a_2\cap\a_3),(\a_3\cap\a_4)$ and $(\a_4\cap\a_1)$ are non-zero independent ideals and the proof is complete. \end{proof} Recall that a ring is reduced if it has no non-zero nilpotent element. \begin{theorem} Let $R$ be a reduced ring. Then $\Gamma(R)$ is $C_n$-free ($n\geq5$) if and only if $R$ has no set of $n$ independent of ideals. \end{theorem} \begin{proof} First suppose $\Gamma(R)$ is $C_n$-free. If $R$ has $n$ non-zero independent ideals $\a_1,\ldots\a_n$, then $\a_1\oplus\a_2,\a_2\oplus\a_3,\ldots,\a_n\oplus\a_1$ induces a cycle of length $n$ in $\Gamma(R)$, which is a contradiction. Now, suppose that $R$ has no set of $n$ non-zero independent ideals and the ideals $\a_1,\ldots,\a_n$ induce a cycle of length $n$. Let $\b_n=\a_n\cap\a_1$ and $\b_i =\a_i\cap\a_{i+1}$ for all $i=1,\ldots,n-1$. Then for all distinct $1\leq i,j\leq n$, we have $\b_i\b_j= 0$. Let $\b_i^*=\b_1+\cdots+\b_{i-1}+\b_{i+1}+\cdots+\b_n$. Then $\b_i\b_i^*=0$. Thus $(\b_i\cap\b_i^*)^2\subseteq\b_i\b_i^*=0$, for all $i=1,\ldots,n$. Since $R$ is reduced, it follows that $\b_i\cap\b_i^*=0$, from which it follows that $\{\b_1,\ldots,\b_n\}$ is a set of non-zero independent ideals of $R$, which is a contradiction. \end{proof} In the sequel, we give another approaches to induced cycles in intersection graphs. The following lemma is straightforward. \begin{lemma} Suppose $\a_1,\ldots,\a_n$ induce a cycle of length $n$ in $\Gamma(R)$. Then there exist $t$ independent ideals $\a_{i_1},\ldots,\a_{i_t}$ such that $2\leq t\leq\lfloor\frac{n}{2}\rfloor$ and $\a_{i_1}\oplus\cdots\oplus\a_{i_t}$ is adjacent to $\a_i$ for all $i=1,\ldots,n$. \end{lemma} \begin{theorem} Suppose $\a_1,\ldots,\a_n$ induce a cycle of length $n$ in $\Gamma(R)$ and the number $t$ introduced in the previous lemma takes it maximum value $\lfloor\frac{n}{2}\rfloor$. Then $R$ has a set of $n$ non-zero independent ideals if $n$ is even and it has a set of $n-1$ non-zero independent ideals if $n$ is odd. \end{theorem} \begin{proof} Without loss of generality we may assume that $\a_1,\a_3,\ldots,\a_{2\lfloor\frac{n}2\rfloor-1}$ are independent. A simple verification shows that \[\{\a_1\cap\a_2,\a_2\cap\a_3,\ldots,\a_{n-1}\cap\a_n,\a_n\cap\a_1\}\] when $n$ is even, \[\{\a_1\cap\a_2,\a_2\cap\a_3,\ldots,\a_{2\lfloor\frac{n}2\rfloor-1}\cap\a_{n -1},\a_n\cap\a_1\}\] when $n$ is odd are sets of non-zero independent ideas of $R$, as required. \end{proof} \begin{theorem} Suppose $\a_1,\ldots,\a_n$ $(n\geq3)$ are independent ideals of $R$. Let $\b_i=\a_{i_1}\oplus\cdots\oplus\a_{i_{n_i}}$, for $i=1,\ldots,n$. Then $\b_1,\ldots,\b_n$ induce a cycle of length $n$ if and only if there exist a permutation $\pi\in S_n$ such that $\b_i=\a_{\pi(i)}\oplus\a_{\pi(i+1)}$. \end{theorem} \begin{proof} If $n=3$ then the result is obvious. If there exist $\pi \in S_n$ such that $\b_i=\a_{\pi(i)}\oplus\a_{\pi(i +1)}$, for all $i=1,\ldots,n$, then there is nothing to prove. Hence we may assume that $\b_1,\ldots,\b_n$ are vertices of an induced cycle with length $n\geq4$. Then $n_i\geq 2$, for all $i=1,\ldots,n$, otherwise $\b_j=\a_{j_1}$ for some $j$. But then $\a_{j_1}$ is adjacent to $\b_{j-1}$ and $\b_{j+1}$, which implies that $\b_{j-1}$ and $\b_{j+1}$ are adjacent, a contradiction. Hence $2n\leq \sum_{i=1}^{n}n_i$. On the other hand, the number of $\b_j$ containing $\a_i$ is at most two for all $i=1,\ldots,n$, which implies that $\sum_{i=1}^{n}n_i\leq2n$. Therefore $\sum_{i=1}^{n}n_i=2n$ and hence $n_i=2$, for all $i=1,\ldots,n$. Now the result is straightforward. \end{proof} Utilizing the same method used before, we may prove the following result for $n$-claws instead of $n$-cycles. \begin{theorem} Let $R$ be a reduced ring. Then the ideals $\a_1,\ldots,\a_n$ of $R$ are independent and $\a_1\oplus\cdots\oplus\a_n$ is a proper ideal of $R$ if and only if there exist an induced $n$-claw in $\Gamma(R)$. \end{theorem} \begin{proof} If $\a_1,\ldots,\a_n$ are independent ideals of $R$ such that $\a_1\oplus\cdots\oplus\a_n$ is a proper ideal of $R$, then clearly $\{\a_1,\ldots,\a_n,\a_1\oplus\cdots\oplus\a_n\}$ induces an $n$-claw in $\Gamma(R)$. Now, suppose that the ideals $\a_1,\ldots,\a_n$ and $\a$ are pendants and the center of an induced $n$-claw, respectively. Let \[\a_i^*=\a_1+\cdots+\a_{i-1}+\a_{i +1}+\cdots+\a_n,\] for all $i=1,\ldots,n$. Then \[(\a_i\cap\a_i^*)^2\subseteq\a_i\a_i^*=\sum_{j \neq i}\a_i\a_j\subseteq\sum_{j\neq i}\a_i\cap\a_j=0,\] for all $i=1,\ldots,n$, which implies that $\a_1,\ldots,\a_n$ are independent. If $R\neq\a_1\oplus\cdots\oplus\a_n$, then we are done. Now, suppose that $R=\a_1\oplus\cdots\oplus\a_n$. If $\a_i$ is not a field for some $1\leq i\leq n$, then by replacing $\a_i$ by one of its non-zero proper ideals, we may assume that $R\neq\a_1\oplus\cdots\oplus\a_n$, as required. Otherwise $\a_1,\ldots,\a_n$ are all fields. But then $\a_i\subseteq\a$, for all $i=1,\ldots,n$, which implies that $\a=R$, a contradiction. \end{proof} \section{Hamilton cycles} The aim of this section is to show that except few cases all intersection graphs are Hamiltonian. Indeed, we shall prove the stronger result that such graphs are pancyclic. A simple verification shows that if $R=S\oplus F$, where $F$ is a field and $\Gamma(S)$ has a Hamiltonian path, then $\Gamma(R)$ has a Hamiltonian cycle. This fact enables us to prove the following result. In what follows, the set of all ideals of a ring $R$ is denoted by $\mathbf{I}(R)$. \begin{theorem}\label{hamiltonian} Let $R$ be an Artin ring. Then $\Gamma(R)$ is Hamiltonian if and only if $R$ is not isomorphic to the following rings: \begin{itemize} \item[(1)]$F$ or $E\oplus F$, \item[(2)]$S$ or $E\oplus S$ such that $(S,F)$ is a local ring, \item[(3)]$S$ such that $(S,T)$ is a local ring and $(T,F)$ is a local ring, \end{itemize} where $E$ and $F$ are fields. \end{theorem} \begin{proof} If $R$ is isomorphic to one of the rings in parts (1), (2) or (3), then clearly $\Gamma(R)$ is not Hamiltonian. Now, suppose that $R$ is a ring such that $\Gamma(R)$ is not Hamiltonian. We proceed in some steps: Case 1: $R=R_1\oplus R_2$ such that $|\mathbf{I}(R_1)|,|\mathbf{I}(R_2)|\geq4$. Let \[\mathbf{I}(R_1)=\{0=\a_0,\a_1,\ldots,\a_m=R_1\}\] and \[\mathbf{I}(R_2)=\{0=\b_0,\b_1,\ldots,\b_n=R_2\}.\] Clearly an arbitrary ideal of $R$ can be expressed as $\a_i\oplus\b_j$ for some $1\leq i\leq m$ and $1\leq j\leq n$. Consider an $(m\times n)$-grid and put $\a_i\oplus\b_j$ on the $(i,j)$-th coordinate. By Figures 1, 2 and 3, the subgraph induced by ideals $\a_i\oplus\b_j$ in which $\a_i,\b_j\neq0$ is Hamiltonian with a Hamiltonian cycle in which there exists at least one edge on every row as well as one edge on every column. If $\{\a_i\oplus\b_j,\a_{i+1}\oplus\b_j\}$ is an edge such that $i,j>0$, then by removing this edge and adding two edges $\{\a_i\oplus\b_j,\b_j\}$ and $\{\b_j,\a_{i+1}\oplus\b_j\}$ we reach to a new cycle including the vertex $\b_j$. Similarly, we may enlarge the resulting cycle in which the new cyclic contains an arbitrary $\a_i\neq0$. Continuing this way, we reach to a Hamiltonian cycle for $\Gamma(R)$, a contradiction. Case 2: $R=R_1\oplus R_2$ such that $\mathbf{I}(R_1)\geq3$ and $|\mathbf{I}(R_2)|=3$. The same as in case 1, we may present ideals of $R$ on the grids as it is shown in Figures 4 and 5, which gives rise to a Hamiltonian cycle for $\Gamma(R)$. Hence $\Gamma(R)$ is Hamiltonian, which is a contradiction. Case 3: $R=S\oplus F$, where $F$ is a field. If $S=S_1\oplus S_2$, where either $S_1$ or $S_2$, say $S_1$ is not a field, then $R=S_1\oplus(S_2\oplus F)$ and by case 1, $\Gamma(R)$ is Hamiltonian. Now, suppose that $S_1$ and $S_2$ are both fields. Then \[S_1\sim S_1\oplus S_2\sim S_2\sim S_2\oplus F\sim F\sim S_1\oplus F\sim S_1\] is a Hamiltonian cycle for $\Gamma(R)$. Hence $\Gamma(R)$ is Hamiltonian, a contradiction. Case 4: If $R$ is a field or it is a direct sum of two fields, then we are done. If not, by cases 1, 2 and 3, there exists a sequence $\{(S_i,R_i)\}_{i=1}^n$ of local rings and a sequence $\{F_i\}_{i=1}^n$ of fields such that $R=R_0=S_1$ or $S_1\oplus F_1$ and $R_i=S_{i+1}$ or $S_{i+1}\oplus F_{i+1}$ for all $1\leq i<n$. Moreover, $R_n$ is a field. If $n=1$, then either $\Gamma(R)$ is a single vertex or it is a path of length three. If $n=2$, then $R=S_1, R_1=S_2$ and $\Gamma(R)$ is an edge. If $n\geq3$, then since $\Gamma(R_{n-2})$ is a path, $\Gamma(R_{n-3})$ and hence $\Gamma(R)$ is Hamiltonian, which is a contradiction. The proof is complete. \end{proof} \begin{center} \begin{tikzpicture}[scale=0.6,rotate=90] \draw [dotted] (0,0) grid (4,4); \draw [dotted] (5,0) grid (9,4); \draw [dotted] (0,5) grid (4,9); \draw [dotted] (5,5) grid (9,9); \draw [color=white] (0,8)--(0,9)--(1,9); \draw [color=white] (8,0)--(9,0)--(9,1); \draw [thick] (3,4)--(3,0)--(1,0)--(1,1)--(2,1)--(2,2)--(1,2)--(1,3)--(2,3)--(2,4)--(1,4); \draw [thick] (2,5)--(1,5)--(1,6)--(2,6)--(2,7)--(1,7)--(1,8)--(2,8); \draw [thick] (3,5)--(3,8)--(4,8)--(4,5); \draw [thick] (4,4)--(4,0); \draw [thick] (5,0)--(5,4); \draw [thick] (6,4)--(6,0)--(7,0)--(7,4); \draw [thick] (8,2)--(8,4); \draw [thick] (8,0)--(8,1)--(9,1)--(9,4); \draw [thick] (5,5)--(5,8)--(6,8)--(6,5); \draw [thick] (7,5)--(7,8)--(8,8)--(8,5); \draw [thick] (9,5)--(9,8); \draw [thick] (8,0) to [out=77, in=-77] (8,2); \draw [thick] (2,8) to [out=13, in=167] (9,8); \draw [loosely dotted,thick] (4.2,0.5)--(4.8,0.5); \draw [loosely dotted,thick] (4.2,1.5)--(4.8,1.5); \draw [loosely dotted,thick] (4.2,2.5)--(4.8,2.5); \draw [loosely dotted,thick] (4.2,3.5)--(4.8,3.5); \draw [loosely dotted,thick] (4.2,5.5)--(4.8,5.5); \draw [loosely dotted,thick] (4.2,6.5)--(4.8,6.5); \draw [loosely dotted,thick] (4.2,7.5)--(4.8,7.5); \draw [loosely dotted,thick] (4.2,8.5)--(4.8,8.5); \draw [loosely dotted,thick] (0.5,4.2)--(0.5,4.8); \draw [loosely dotted,thick] (1.5,4.2)--(1.5,4.8); \draw [loosely dotted,thick] (2.5,4.2)--(2.5,4.8); \draw [loosely dotted,thick] (3.5,4.2)--(3.5,4.8); \draw [loosely dotted,thick] (5.5,4.2)--(5.5,4.8); \draw [loosely dotted,thick] (6.5,4.2)--(6.5,4.8); \draw [loosely dotted,thick] (7.5,4.2)--(7.5,4.8); \draw [loosely dotted,thick] (8.5,4.2)--(8.5,4.8); \draw [loosely dotted,thick] (4.2,4.2)--(4.82,4.82); \draw [loosely dotted,thick] (4.82,4.2)--(4.2,4.82); \end{tikzpicture}\\ Figure 1. $(|\mathbf{I}(R_1)|,|\mathbf{I}(R_2)|)=(\mbox{odd}>3,\mbox{even}>3)$ \end{center} \begin{center} \begin{tikzpicture}[scale=0.6,rotate=90] \draw [dotted] (0,0) grid (4,4); \draw [dotted] (5,0) grid (9,4); \draw [dotted] (0,5) grid (4,9); \draw [dotted] (5,5) grid (9,9); \draw [color=white] (0,8)--(0,9)--(1,9); \draw [color=white] (8,0)--(9,0)--(9,1); \draw [thick] (4,0)--(4,4); \draw [thick](3,4)--(3,0)--(1,0)--(1,1)--(2,1)--(2,2)--(1,2)--(1,3)--(2,3)--(2,4)--(1,4); \draw [thick] (2,8)--(1,8)--(1,7)--(2,7)--(2,6)--(1,6)--(1,5)--(2,5); \draw [thick] (3,5)--(3,8)--(4,8)--(4,5); \draw [thick] (5,4)--(5,0)--(6,0)--(6,4); \draw [thick] (7,4)--(7,0)--(8,0); \draw [thick] (8,4)--(8,1)--(9,1)--(9,4); \draw [thick] (5,8)--(5,5); \draw [thick] (6,5)--(6,8)--(7,8)--(7,5); \draw [thick] (8,8)--(8,5); \draw [thick] (9,8)--(9,5); \draw [thick] (2,8) to [out=13, in=167] (9,8); \draw [thick] (2,8) to [out=13, in=167] (9,8); \draw [thick] (8,0) to [out=77, in=-77] (8,8); \draw [loosely dotted,thick] (4.2,0.5)--(4.8,0.5); \draw [loosely dotted,thick] (4.2,1.5)--(4.8,1.5); \draw [loosely dotted,thick] (4.2,2.5)--(4.8,2.5); \draw [loosely dotted,thick] (4.2,3.5)--(4.8,3.5); \draw [loosely dotted,thick] (4.2,5.5)--(4.8,5.5); \draw [loosely dotted,thick] (4.2,6.5)--(4.8,6.5); \draw [loosely dotted,thick] (4.2,7.5)--(4.8,7.5); \draw [loosely dotted,thick] (4.2,8.5)--(4.8,8.5); \draw [loosely dotted,thick] (0.5,4.2)--(0.5,4.8); \draw [loosely dotted,thick] (1.5,4.2)--(1.5,4.8); \draw [loosely dotted,thick] (2.5,4.2)--(2.5,4.8); \draw [loosely dotted,thick] (3.5,4.2)--(3.5,4.8); \draw [loosely dotted,thick] (5.5,4.2)--(5.5,4.8); \draw [loosely dotted,thick] (6.5,4.2)--(6.5,4.8); \draw [loosely dotted,thick] (7.5,4.2)--(7.5,4.8); \draw [loosely dotted,thick] (8.5,4.2)--(8.5,4.8); \draw [loosely dotted,thick] (4.2,4.2)--(4.82,4.82); \draw [loosely dotted,thick] (4.82,4.2)--(4.2,4.82); \end{tikzpicture}\\ Figure 2. $(|\mathbf{I}(R_1)|,|\mathbf{I}(R_2)|)=(\mbox{odd}>3,\mbox{odd}>3)$ \end{center} \begin{center} \begin{tikzpicture}[scale=0.6,rotate=90] \draw [dotted] (0,0) grid (4,4); \draw [dotted] (5,0) grid (9,4); \draw [dotted] (0,5) grid (4,9); \draw [dotted] (5,5) grid (9,9); \draw [color=white] (0,8)--(0,9)--(1,9); \draw [color=white] (8,0)--(9,0)--(9,1); \draw [thick] (1,0)--(2,0)--(2,1)--(1,1)--(1,2)--(2,2)--(2,3)--(1,3)--(1,4)--(2,4); \draw [thick] (2,5)--(1,5)--(1,6)--(2,6)--(2,7)--(1,7)--(1,8)--(2,8); \draw [thick] (3,5)--(3,8)--(4,8)--(4,5); \draw [thick] (3,4)--(3,0); \draw [thick] (4,4)--(4,0); \draw [thick] (5,0)--(5,4); \draw [thick] (6,4)--(6,0)--(7,0)--(7,4); \draw [thick] (8,2)--(8,4); \draw [thick] (8,0)--(8,1)--(9,1)--(9,4); \draw [thick] (5,5)--(5,8)--(6,8)--(6,5); \draw [thick] (7,5)--(7,8)--(8,8)--(8,5); \draw [thick] (9,5)--(9,8); \draw [thick] (8,0) to [out=77, in=-77] (8,2); \draw [thick] (2,8) to [out=13, in=167] (9,8); \draw [thick] (1,0) to [out=13, in=167] (3,0); \draw [loosely dotted,thick] (4.2,0.5)--(4.8,0.5); \draw [loosely dotted,thick] (4.2,1.5)--(4.8,1.5); \draw [loosely dotted,thick] (4.2,2.5)--(4.8,2.5); \draw [loosely dotted,thick] (4.2,3.5)--(4.8,3.5); \draw [loosely dotted,thick] (4.2,5.5)--(4.8,5.5); \draw [loosely dotted,thick] (4.2,6.5)--(4.8,6.5); \draw [loosely dotted,thick] (4.2,7.5)--(4.8,7.5); \draw [loosely dotted,thick] (4.2,8.5)--(4.8,8.5); \draw [loosely dotted,thick] (0.5,4.2)--(0.5,4.8); \draw [loosely dotted,thick] (1.5,4.2)--(1.5,4.8); \draw [loosely dotted,thick] (2.5,4.2)--(2.5,4.8); \draw [loosely dotted,thick] (3.5,4.2)--(3.5,4.8); \draw [loosely dotted,thick] (5.5,4.2)--(5.5,4.8); \draw [loosely dotted,thick] (6.5,4.2)--(6.5,4.8); \draw [loosely dotted,thick] (7.5,4.2)--(7.5,4.8); \draw [loosely dotted,thick] (8.5,4.2)--(8.5,4.8); \draw [loosely dotted,thick] (4.2,4.2)--(4.82,4.82); \draw [loosely dotted,thick] (4.82,4.2)--(4.2,4.82); \end{tikzpicture}\\ Figure 3. $(|\mathbf{I}(R_1)|,|\mathbf{I}(R_2)|)=(\mbox{even}>3,\mbox{even}>3)$ \end{center} \begin{center} \begin{tikzpicture}[scale=0.6] \draw [dotted] (0,0) grid (4,2); \draw [dotted] (5,0) grid (9,2); \draw [color=white] (0,1)--(0,0)--(1,0); \draw [color=white] (8,2)--(9,2)--(9,1); \draw [thick] (0,1)--(0,2)--(1,2)--(1,1)--(2,1)--(2,2)--(3,2)--(3,1)--(4,1)--(4,2); \draw [thick] (5,2)--(5,1)--(6,1)--(6,2)--(7,2)--(7,1)--(8,1)--(8,2); \draw [thick] (8,0)--(9,0)--(9,1); \draw [thick] (8,2) to [out=-60, in=60] (8,0); \draw [thick] (0,1) to [out=-13, in=-167] (9,1); \draw [loosely dotted,thick] (4.2,.5)--(4.82,.5); \draw [loosely dotted,thick] (4.2,1.5)--(4.82,1.5); \end{tikzpicture}\\ Figure 4. $(|\mathbf{I}(R_1)|,|\mathbf{I}(R_2)|)=(3,\mbox{even}\geq3)$ \end{center} \begin{center} \begin{tikzpicture}[scale=0.6] \draw [dotted] (0,0) grid (4,2); \draw [dotted] (5,0) grid (9,2); \draw [color=white] (0,1)--(0,0)--(1,0); \draw [color=white] (8,2)--(9,2)--(9,1); \draw [thick] (0,1)--(0,2)--(1,2)--(1,1)--(2,1)--(2,2)--(3,2)--(3,1)--(4,1)--(4,2); \draw [thick] (5,1)--(5,2)--(6,2)--(6,1)--(7,1)--(7,2)--(8,2)--(8,0); \draw [thick] (8,0)--(9,0)--(9,1); \draw [thick] (0,1) to [out=-13, in=-167] (9,1); \draw [thick , color=white] (0,0)--(1,0); \draw [loosely dotted,thick] (4.2,.5)--(4.82,.5); \draw [loosely dotted,thick] (4.2,1.5)--(4.82,1.5); \end{tikzpicture}\\ Figure 5. $(|\mathbf{I}(R_1)|,|\mathbf{I}(R_2)|)=(3,\mbox{odd}\geq3)$ \end{center} \begin{theorem}\label{pancyclic} Let $R$ be an Artin ring. Then $\Gamma(R)$ is Hamiltonian if and only if it is pancyclic. \end{theorem} \begin{proof} If $\Gamma(R)$ is pancyclic, then clearly $\Gamma(R)$ is Hamiltonian. Now, we show that the converse is also true. Suppose on the contrary that there is an Artin ring $R$ such that $\Gamma(R)$ is a non-pancyclic Hamiltonian graph and that $R$ is minimal with this property. If $R$ is neither a local ring nor a direct sum of a local ring with a field, then by applying the following transformations on the Hamiltonian cycles constructed in cases 1 and 2 of Theorem \ref{hamiltonian}, along with replacing horizontal or vertical paths of length two to a path of length one, by joining its end vertices, we would reach to cycles with possible arbitrary length $\geq4$. \begin{center} \begin{tikzpicture}[scale=0.6] \draw [dotted] (0,0) grid (1,1); \draw [dotted] (3,0) grid (4,1); \draw [dotted] (7,0) grid (8,1); \draw [dotted] (10,0) grid (11,1); \draw [dotted] (0,2) grid (1,3); \draw [dotted] (3,2) grid (4,3); \draw [dotted] (7,2) grid (8,3); \draw [dotted] (10,2) grid (11,3); \draw [thick] (0,1)--(0,0)--(1,0)--(1,1); \draw [->] (1.5,0.5)--(2.5,0.5); \draw [thick] (3,1)--(4,1); \draw [thick] (7,1)--(8,1)--(8,0)--(7,0); \draw [->] (8.5,0.5)--(9.5,0.5); \draw [thick] (10,0)--(10,1); \draw [thick] (1,2)--(0,2)--(0,3)--(1,3); \draw [->] (1.5,2.5)--(2.5,2.5); \draw [thick] (4,2)--(4,3); \draw [thick] (7,2)--(7,3)--(8,3)--(8,2); \draw [->] (8.5,2.5)--(9.5,2.5); \draw [thick] (10,2)--(11,2); \end{tikzpicture} \end{center} On the other hand, by Theorem \ref{triangle}, the graphs under consideration contain triangles, which implies that $\Gamma(R)$ is pancyclic, a contradiction. Hence either $R=S$ or $R=S\times F$, where $(S,\mathfrak{m})$ is a local ring and $F$ is a field. If $\Gamma(\mathfrak{m})$ is Hamiltonian, then either $\Gamma(\mathfrak{m})$ is pancyclic, which implies that $\Gamma(R)$ is pancyclic too, contradicting the hypothesis, or $\Gamma(\mathfrak{m})$ is not pancyclic which contradicts the minimality of $R$. Thus $\Gamma(\mathfrak{m})$ is not Hamiltonian and $\mathfrak{m}$ is isomorphic to one of the five rings given in Theorem \ref{hamiltonian}. Now, a simple verification shows that in each case either $\Gamma(\mathfrak{m})$ is not Hamiltonian or it is pancyclic, which is a contradiction. The proof is complete. \end{proof}
{ "timestamp": "2014-01-28T02:08:50", "yymm": "1401", "arxiv_id": "1401.6619", "language": "en", "url": "https://arxiv.org/abs/1401.6619", "abstract": "Let $R$ be a commutative ring with non-zero identity. We describe all $C_3$- and $C_4$-free intersection graph of non-trivial ideals of $R$ as well as $C_n$-free intersection graph when $R$ is a reduced ring. Also, we shall describe all complete, regular and $n$-claw-free intersection graphs. Finally, we shall prove that almost all Artin rings $R$ have Hamiltonian intersection graphs. We show that such graphs are indeed pancyclic.", "subjects": "Commutative Algebra (math.AC)", "title": "On cycles in intersection graph of rings", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9908743604549526, "lm_q2_score": 0.8080672089305841, "lm_q1q2_score": 0.8006930788537111 }
https://arxiv.org/abs/1911.12336
Synchronization of Kuramoto Oscillators in Dense Networks
We study synchronization properties of systems of Kuramoto oscillators. The problem can also be understood as a question about the properties of an energy landscape created by a graph. More formally, let $G=(V,E)$ be a connected graph and $(a_{ij})_{i,j=1}^{n}$ denotes its adjacency matrix. Let the function $f:\mathbb{T}^n \rightarrow \mathbb{R}$ be given by $$ f(\theta_1, \dots, \theta_n) = \sum_{i,j=1}^{n}{ a_{ij} \cos{(\theta_i - \theta_j)}}.$$ This function has a global maximum when $\theta_i = \theta$ for all $1\leq i \leq n$. It is known that if every vertex is connected to at least $\mu(n-1)$ other vertices for $\mu$ sufficiently large, then every local maximum is global. Taylor proved this for $\mu \geq 0.9395$ and Ling, Xu \& Bandeira improved this to $\mu \geq 0.7929$. We give a slight improvement to $\mu \geq 0.7889$. Townsend, Stillman \& Strogatz suggested that the critical value might be $\mu_c = 0.75$.
\section{Introduction} We study a simple problem that can be understood from a variety of perspectives. Perhaps its simplest formulation is as follows: let $G=(V,E)$ be a connected graph and $(a_{ij})_{i,j=1}^{n}$ denotes its adjacency matrix. We assume the graph is simple, and thus $a_{ii} = 0$ for $i = 1, \ldots, n$. We are then interested in the behavior of the energy functional $f:\mathbb{T}^n \cong [0,2\pi]^n \rightarrow \mathbb{R}$ given by \begin{equation} \label{main} f(\theta_1, \dots, \theta_n) = \sum_{i,j=1}^{n}{ a_{ij} \cos{(\theta_i - \theta_j)}}. \end{equation} Ling, Xu \& Bandeira \cite{ling} ask the following very interesting \begin{quote} \textbf{Question.} What is the relationship between the existence of local maxima and the topology of the network? \end{quote} $f$ assumes its global maximum when $\theta_i \equiv \theta$ is constant and this is the unique global maximum up to rotation. Factoring out the rotation symmetry, there are at least $2^n$ critical points of the form $\theta_i \in \left\{0, \pi\right\}$. The main question is under which condition we can exclude the existence of local maxima that are not global maxima.\\ This is related to the Kuramoto model as follows: suppose we consider the system of ordinary differential equations given by $$ \frac{d \theta_i}{dt} = -\sum_{j=1}^{n}{a_{ij} \sin{(\theta_i - \theta_j)}}.$$ We can interpret this system of ODEs as a gradient flow with respect to the energy $$ E(\theta_1, \dots, \theta_n) = - \sum_{i,j=1}^{n}{ a_{ij} \cos{(\theta_i - \theta_j)}}.$$ In this case, local maxima that are not global correspond to stable local minima of the gradient flow. In light of this model, particles on the circle that are connected by springs, it is natural to assume that no spurious local minimizer of this energy exist if there are enough springs. This motivated an existing line of research: Taylor \cite{taylor} proved that if each vertex is connected to at least $\mu(n-1)$ vertices for $\mu \geq 0.9395$, then \eqref{main} does not have local maxima that are not global. Ling, Xu \& Bandeira \cite{ling} improved this to $\mu \geq 0.7929$. They also showed the existence of a configuration coming from the family of Wiley-Strogatz-Girvan networks \cite{wiley} where each vertex is connected to $0.68n$ other vertices that indeed has local maxima that are not global. Townsend, Stillman \& Strogatz \cite{townsend} suggest that the critical value might be $\mu_c = 0.75$ and identify networks with $\mu = 0.75$ having interesting spectral properties. \\ The problem itself arises in a variety of settings. We refer to the surveys \cite{review1, review2, review3} for an overview regarding synchronization problems, to \cite{chaos} for insights into complexities of the Kuramoto model and to \cite{lopes, lopes2} for random Kuramoto models. There is also recent interest in the landscape of non-convex loss functionals for which this problem is a natural test case, we refer to \cite{loss1, loss2, loss3, loss4}. \begin{theorem} If $G=(V,E)$ is a connected graph such that the degree of every vertex is at least $0.7889(n-1)$, then $$ f(\theta_1, \dots, \theta_n) = \sum_{i,j=1}^{n}{ a_{ij} \cos{(\theta_i - \theta_j)}}$$ does not have local maxima that are not global. \end{theorem} The main idea behind the argument is a refinement of the approach of Ling, Xu \& Bandeira \cite{ling} in a certain parameter range using a new decomposition of the points. We consider the problem a natural benchmark for testing our understanding of the geometry of energy landscapes. We conclude by reiterating the original question from \cite{ling}: which kind of assumption on the network (this paper, for example, is only dealing with edge-density assumptions) implies synchronization? \section{Proof} \subsection{Ingredients.} The purpose of this section is to sketch several of the tools that go into the argument which is a variation on the argument given by Ling, Xu \& Bandeira \cite{ling}. We first recall their argument. They start by introducing a useful Proposition (precursors of which can be found in Taylor \cite{taylor}). \begin{proposition}[Ling, Xu, Bandeira \cite{ling}] Let $(\theta_1, \dots, \theta_n) \in \mathbb{T}^n$ be a strict local maximizer of \eqref{main}. If there exists an angle $\theta_r$ such that $$ \forall~1 \leq i \leq n: \qquad |\sin{(\theta_i - \theta_r)}| < \frac{1}{\sqrt{2}},$$ then all the $\theta_i$ have the same value. \end{proposition} The idea behind this argument is as follows: if $(\theta_1, \dots, \theta_n) \in \mathbb{T}^n$ is a local maximizer, then the quadratic form (corresponding to the negative Hessian) is positive semi-definite. In other words, a necessary condition for being a local maximum (derived in \cite{ling}) is that for all vectors $w \in \mathbb{R}^n$ $$ \sum_{i,j=1}^{n}{ a_{ij} \cos{(\theta_i - \theta_j)} (w_i - w_j)^2} \geq 0.$$ We can derive a contradiction by defining a vector $w \in \left\{-1, 1\right\}^n$ depending on which of the two `cones' the variable $\theta_i$ is in. Then the summation only ranges over pairs that are in opposite sides of the cone. The cosine is negative for those values and since the graph is connected, there is at least one connection leading to a contradiction.\\ A second important ingredient is the Kuramoto parameter \cite{kuramoto} $$ r = \| r\| e^{i \theta_r}:= \sum_{j=1}^{n}{ e^{i \theta_j}}.$$ The second part in the argument \cite{ling} is based on showing that \begin{equation} \label{length} \begin{aligned} \|r\|^2 &\geq \frac{n^2}{2} - \sum_{i \neq j}{ (1-a_{ij}) \Abs{ \cos{(\theta_i - \theta_j)} - \cos^2{(\theta_i - \theta_j)} }} \\ &\geq \left(2 \mu - \frac{3}{2} \right)n^2 + 2(1-\mu)n, \end{aligned} \end{equation} where the second inequality follows from $$ \Abs{\cos{(\theta_i - \theta_j)} - \cos^2{(\theta_i - \theta_j)}} \leq 2$$ and $$ \sum_{i \neq j}{(1-a_{ij})} = \sum_{i=1}^{n} \sum_{j \neq i}{(1-a_{ij})} \leq \sum_{i=1}^{n}{ (1-\mu)(n-1)} = (1-\mu) (n-1)n.$$ The argument in \cite{ling} proceeds by writing $$ \| r\| e^{-i (\theta_i - \theta_r)} = r e^{- i\theta_i} = \sum_{j=1}^{n}{ e^{-i(\theta_i - \theta_j)}}$$ and taking imaginary parts to obtain $$ \|r\| \sin{(\theta_i - \theta_r)} = \sum_{j=1}^{n}{ \sin{(\theta_i - \theta_j)}}.$$ However, the first order condition in a maximum implies $$ \sum_{j=1}^{n}{a_{ij} \sin{(\theta_i - \theta_j)}} = 0$$ and thus we obtain $$ \|r\| \sin{(\theta_i - \theta_r)} = \sum_{j=1}^{n}{(1-a_{ij}) \sin{(\theta_i - \theta_j)}}.$$ As a consequence, we have \begin{equation} \label{bound} | \sin{(\theta_i - \theta_r)} | \leq \frac{1}{\|r\|} \Biggl\lvert \sum_{j=1}^{n}{(1-a_{ij}) \sin{(\theta_i - \theta_j)}} \Biggr\rvert \leq \frac{(1-\mu)n}{\|r\|}. \end{equation} The Proposition together with \eqref{length} and \eqref{bound} then imply the result. \subsection{The Proof} Our proof is motivated by the following simple observation: inequality \eqref{length} is only sharp when $$ \Abs{\cos{(\theta_i - \theta_j)} - \cos^2{(\theta_i - \theta_j)}} = 2$$ for all pairs of angles $\theta_i, \theta_j$ for which $a_{ij} = 0$. This only happens when $\theta_i - \theta_j = \pi$. So the only extreme case of equality is when the points that are not connected are concentrated on two different sides of the torus. Then, however, we can dramatically improve inequality \eqref{bound} since $\sin{(\pi)} = 0$. The proof will decouple into two steps: either inequality \eqref{length} is very far from sharp, in that case the origin Ling-Xu-Bandeira argument will result in the desired improvement, or the inequality is close to being sharp in which case we can try to extract more information from it.\\ Our proof makes use of several new parameters. As will come to no surprise to the reader, we obtain them by working with unspecified coefficients in the beginning and then solving the arising optimization problem to obtain the optimal selection of parameters. For readers who prefer explicit values to have an idea of scales, we will later set $$ \varepsilon = 0.5 \qquad \mbox{and} \qquad \delta = 0.88.$$ \begin{proof} The proof decouples into several steps.\\ \textbf{Step 1: Finding Good Vertices.} The first part of our proof emulates the argument from \cite{ling} with one slight modification. We assume that we are working with a slightly improved bound in \eqref{length}. Specifically, we first assume that we have the identity $$ \sum_{i \neq j} (1 - a_{ij}) \Abs{\cos(\theta_i - \theta_j) - \cos(\theta_i-\theta_j)^2} = (2-\alpha) (1-\mu) (n-1)n $$ for some value of $\alpha > 0.0537$. In that case, running the original Ling-Xu-Bandeira argument again shows that we have $$ | \sin{(\theta_i - \theta_r)} | < \frac{1}{\sqrt{2}}$$ as long as $\mu \geq 0.788897$. It thus remains to obtain a similar bound in the case where this assumption does not hold. We may thus additionally assume \begin{equation} \label{standingassumption} \sum_{i \neq j} (1 - a_{ij}) \Abs{\cos(\theta_i - \theta_j) - \cos(\theta_i-\theta_j)^2} = (2-\alpha) (1-\mu) (n-1)n, \end{equation} where $\alpha$ is defined by the equation and satisfies $0 \leq \alpha \leq 0.0537$. The first assumption that we derive from this assumption is that there are many `good' vertices, a term that we will use repeatedly and that we now formally define. \begin{definition} A vertex $i$ is `$\varepsilon-$good' if \begin{equation} \label{firstcond} \sum_{j=1,\, j \neq i}^{n} (1 - a_{ij})\Abs{\cos(\theta_i - \theta_j) - \cos(\theta_i-\theta_j)^2} \geq (2-\varepsilon) (1-\mu)(n-1). \end{equation} \end{definition} In practice, we will assume that $\varepsilon$ is fixed and to be optimized over later and we will refer to them as `good vertices' or `good points'. The nomenclature is motivated by the fact that the double sum has to be large (see equation \eqref{standingassumption}) and `good points' contribute to achieving this goal. It will also be reflected in the fact that we will derive some improved bounds for good points early on in the paper. The first step in our argument is to conclude that, for every given $\varepsilon \in [\alpha, 2]$, there are at least $(1-\frac{\alpha}{\varepsilon})n$ vertices that are $\varepsilon-$good. Suppose this is false, then the double sum in \eqref{standingassumption} could be bounded from above by \begin{align*} \text{LHS of }\eqref{standingassumption}&< \frac{\alpha}{\varepsilon} n (2-\varepsilon) (1-\mu)(n-1) + \Bigl(1-\frac{\alpha}{\varepsilon}\Bigr)n 2 (1-\mu)(n-1) \\ &= (1-\mu)(n-1)n \Bigl( \frac{\alpha}{\varepsilon}(2-\varepsilon) + 2-2\frac{\alpha}{\varepsilon}\Bigr)\\ &= (2-\alpha)(1-\mu)(n-1)n \end{align*} which is a contradiction to \eqref{standingassumption}. \\ Let us now take a $\varepsilon-$good vertex $i$. We argue that, for every given $ \delta \in (\varepsilon, 2)$ there are many non-neighbors, indices $j$ such that $a_{ij} = 0$, for which $$ \Abs{\cos(\theta_i - \theta_j) - \cos(\theta_i-\theta_j)^2} \geq 2 - \delta.$$ Let us assume their number is $(1-\mu - c)(n-1)$. Then we can bound, using the fact that the total number of non-neighbors is at most $(1-\mu)(n-1)$, that $$ \sum_{j=1}^{n} (1 - a_{ij})\Abs{\cos(\theta_i - \theta_j) - \cos(\theta_i-\theta_j)^2} \leq (1-\mu-c)(n-1) 2 + c(n-1)(2-\delta).$$ We require, using \eqref{firstcond}, that $$ (2-\varepsilon) (1-\mu)\leq (1-\mu-c) 2 + (2-\delta)c$$ and thus $$ c \leq (1-\mu)\frac{\varepsilon}{\delta}.$$ This shows that for any $\varepsilon-$good vertex, the number of non-neighbors for which the cosine quantity exceeds $2-\delta$ is at least $$(1-\mu - c)(n-1) \geq (1-\mu)(n-1)\left(1- \frac{\varepsilon}{\delta} \right).$$ We summarize our arguments up to this step. \begin{enumerate} \item Let us consider the value $\alpha$ as defined in \eqref{standingassumption}. If $\alpha > 0.0537$, then we get the desired result directly from the argument of Ling, Xu \& Bandeira. It thus remains to study the cases where $0 \leq \alpha \leq 0.0537$. \item In this case, for each $\varepsilon \geq \alpha$, there are at least $(1-\frac{\alpha}{\varepsilon})n$ vertices $i$ (`the $\varepsilon$-good vertices') for which we have the inequality $$ \sum_{j=1,j\neq i}^{n} (1 - a_{ij})\Abs{\cos(\theta_i - \theta_j) - \cos(\theta_i-\theta_j)^2} \geq (2-\varepsilon) (1-\mu)(n-1). $$ \item For each $\varepsilon < \delta < 2$, each of these $(1-\frac{\alpha}{\varepsilon})n$ good points has at least $(1-\mu) (n-1)(1 - \frac{\varepsilon}{\delta})$ non-neighbors, $a_{ij} = 0$, for which $$ \Abs{\cos(\theta_i - \theta_j) - \cos(\theta_i-\theta_j)^2} \geq 2- \delta.$$ \end{enumerate} \textbf{Step 2: Improved Bounds for Good Vertices.} It is an elementary trigonometric fact that if $0 \leq \delta \leq 1$ (in fact the inequality below holds for $0 \leq \delta < 7/4$), then $$ \Abs{\cos(x) - \cos(x)^2} \geq 2 - \delta,\quad \text{implies} \quad \left|\sin{(x)}\right| \leq \frac{1}{\sqrt{2}} \sqrt{ \sqrt{9 - 4 \delta} - 3 + 2\delta} =: s_{\delta},$$ where we introduced the shorthand $s_{\delta}$ for simplicity of exposition. Combining these facts, we can show that for all the good points, of which there are at least $(1-\frac{\alpha}{\varepsilon})n$, we have \begin{align*} \|r\| \cdot \left|\sin{(\theta_i - \theta_r)}\right| &= \left| \sum_{j=1}^{n}{ (1-a_{ij}) \sin{(\theta_i - \theta_j)}} \right|\\ &\leq (1-\mu)(n-1) \Bigl(1-\frac{\varepsilon}{\delta}\Bigr) s_{\delta} \\ &\qquad + (1-\mu)(n-1)\frac{\varepsilon}{\delta}\\ &= (1-\mu)(n-1)\Bigl( s_{\delta} + \frac{\varepsilon}{\delta} (1-s_{\delta})\Bigr). \end{align*} This, in turn, implies that any good point $i$ satisfies $$ \left|\sin{(\theta_i - \theta_r)}\right|^2 \leq \frac{ \left( s_{\delta} + \frac{\varepsilon}{\delta} (1-s_{\delta})\right)^2 (1-\mu)^2 n^2}{\|r\|^2}$$ Note that we also have \eqref{length} and \eqref{standingassumption} implying \begin{align*} \|r\|^2 \geq \left(\frac{1}{2} - (2-\alpha)(1-\mu) \right)n^2 \end{align*} Therefore, if $i$ is a good point, then $$ \left|\sin{(\theta_i - \theta_r)}\right|^2 \leq \frac{ \left( s_{\delta} + \frac{\varepsilon}{\delta} (1-s_{\delta})\right)^2 (1-\mu)^2}{ \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)}.$$ Recall also that there are at most $\frac{\alpha}{\varepsilon} n$ `bad' points (points which are not good, we will also call then `outliers'). We define $$\varphi = \varphi(\mu, \varepsilon, \delta, \alpha)$$ as the positive angle satisfying $$ \left|\sin{(\varphi)}\right|^2 = \frac{ \left( s_{\delta} + \frac{\varepsilon}{\delta} (1-s_{\delta})\right)^2 (1-\mu)^2}{ \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)}.$$ This is the bound we get on the angle that good points have with $\theta_r$. There is a balancing act: by choosing the parameters in a more restricted fashion, we can get a better upper bound for this angle but there will be less good points. Choosing the parameters in the other direction will result in a bigger number of good points but less control on their geometric distribution. We will require, further along in the argument, that $\varepsilon$ and $\delta$ are chosen in such a way that \begin{equation} \label{sincond} \left|\sin{(\varphi)}\right|^2 = \frac{ \left( s_{\delta} + \frac{\varepsilon}{\delta} (1-s_{\delta})\right)^2 (1-\mu)^2}{ \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)} < \frac{1}{2}. \end{equation} \textbf{Step 3: The Distribution of Good Points.} We now introduce a bit of orientation and assume, without loss of generality after possibly rotating all the points, that $$ r = \| r\| e^{i \theta_r} = \sum_{j=1}^{n}{ e^{i \theta_j}} \qquad \mbox{is a positive real number.}$$ We know that the good points, of which there are at least $(1-\frac{\alpha}{\varepsilon})n$ many, satisfy \eqref{sincond} and are thus contained in one of two cones. We assume that we have $\gamma_1 n$ good points in the left cone and $\gamma_2n$ good points in the right cone (see Figure~\ref{fig:cones} for a sketch of this). Without loss of generality, we can reflect the picture if necessary and assume that $\gamma_2 \geq \gamma_1$. Recalling our lower bound on the number of good points, we have \begin{equation}\label{eq:gamma12} \gamma_1 + \gamma_2 \geq 1 - \frac{\alpha}{\varepsilon}. \end{equation} \begin{center} \begin{figure}[h!] \begin{tikzpicture} \draw [thick] (0,0) circle (2cm); \draw [->, thick] (0,0) -- (3,0); \node at (3, -0.3) {$r$}; \draw [thick, dashed] (-1.6, -1.2) -- (1.6, 1.2); \draw [thick, dashed] (-1.6, 1.2) -- (1.6, -1.2); \node at (-4,0) {$\gamma_1 n$ points here}; \draw[->] (-2.7,0) -- (-2.2,0); \node at (4,1) {$\gamma_2 n$ points here}; \draw [->] (2.8, 0.8) -- (2.2, 0.5); \node at (0.6, 0.2) {$\varphi$}; \node at (3,-1) {$\sin^2{(\varphi)} < 1/2$}; \end{tikzpicture} \caption{Introducing orientation: $r$ being a positive real forces all the good points to be in two cones. The outliers can be anywhere (and could also be in the cone).\label{fig:cones}} \end{figure} \end{center} We note that the outliers, of which there are (at most) $\frac{\alpha}{\varepsilon} n$, might also be in the left or the right cone, we do not make any statement about their actual location and will always assume that they are working against us. The inequality $$ \|r\|^2 \geq \left(\frac{1}{2} - (2-\alpha)(1-\mu) \right)n^2$$ forces some restrictions on $\gamma_1$ and $\gamma_2$: in particular, if all the good points and all the outliers were distributed somewhat evenly, then $\|r\|$ would actually be quite small. However, we do have a lower bound on $\|r\|$ and this forces some restrictions which we will now explore. Assuming the worst case (where all the outliers are actually working in our favor and contribute to making $r$ as big as possible), we have $$ \|r\| \leq \left(\gamma_2 - \cos{(\varphi)} \gamma_1 + \frac{\alpha}{\varepsilon}\right) n$$ and therefore $$ \gamma_2 - \cos{(\varphi)} \gamma_1 + \frac{\alpha}{\varepsilon} \geq \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)^{1/2}$$ which implies, using $-\gamma_1 \leq \gamma_2 -1 + \frac{\alpha}{\varepsilon}$, $$ \bigl(1 + \cos(\varphi)\bigr) \Bigl(\gamma_2 + \frac{\alpha}{\varepsilon}\Bigr) - \cos(\varphi) \geq \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)^{1/2}.$$ This inequality implies there cannot be too few points inside the right cone for otherwise $r$ could not attain the size it does. More precisely, this forces \begin{equation}\label{eq:lowerboundgamma2} \gamma_2 \geq \frac{\cos(\varphi) + \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)^{1/2}}{1 + \cos{(\varphi)}} - \frac{\alpha}{\varepsilon}. \end{equation} \textbf{Step 4: Using the Hessian.} So far, we have obtained fairly precise information about the number of good points and their approximate location. We have not yet made strong use of the fact that the configuration is a local maximum (we did use it implicitly when appealing to results of Ling-Xu-Bandeira). In this section, we will explicitly use the fact that the Hessian has to be negative-semidefinite in a maximum to derive a criterion that has to be satisfied for all local maxima. We will then contrast this criterion with the precise structure we have derived to conclude that for some parameters this condition is violated thus excluding them. This will prove the result. We conclude the argument by using that the configuration of points considered in Step 3 is a local maximum: this means that the Hessian is definite which in our case implies that for any $(w_1, \dots, w_n) \in \mathbb{R}^n$, we have \begin{equation} \label{positivity} \sum_{i,j}{ a_{ij} \cos{(\theta_i - \theta_j)} (w_i - w_j)^2} \geq 0. \end{equation} We are free to choose the constants $w_i$ and will choose them in a way that makes the quadratic form as small as possible: this means we want to pick $w_i, w_j$ to be quite different when $i,j$ are on different sides of the cone. We pick these numbers as follows: for a constant $v \in \mathbb{R}$ to be determined $$ w = \begin{cases} 1 &\qquad \mbox{for the ones on the left cone}\\ v &\qquad \mbox{for the outliers}\\ -1 &\qquad \mbox{for the ones on the right cone}. \end{cases}$$ We will bound the expression from above using the information we have about $\gamma_1$ and $\gamma_2$: we know that $\gamma_1 + \gamma_2$ is not too small, that $\gamma_1 \leq \gamma_2$ and we have a lower bound on $\gamma_2$. The quadratic form, which we know to be positive, is bounded from above \begin{align*} \frac{1}{2} \sum_{i,j}{ a_{ij} \cos{(\theta_i - \theta_j)} (w_i - w_j)^2} &= \sum_{i~\tiny \mbox{left}}~ \sum_{j~ \tiny \mbox{right}}{ a_{ij} \cos{(\theta_i - \theta_j)} (w_i - w_j)^2}\\ &+ \sum_{i~\tiny \mbox{outlier}} ~ \sum_{j~\tiny \mbox{right/left cone}}{ a_{ij} \cos{(\theta_i - \theta_j)} (w_i - w_j)^2} \end{align*} Now, since $|\sin(\varphi)|^2 \leq 1/2$ by assumption, we have that the cosine is negative for any pair of points where one is contained in the left cone and one is contained in the right cone. Indeed, we see that for any pair $i,j$ of good points on different sides of the cone, we have $$ \cos{(\theta_i - \theta_j)} \leq \cos{(\pi - (2\pi))},$$ where $\phi$ is the angle introduced above. We further bound the quadratic form from above by assuming that the number of connections running across good points on different sides of the cone is minimized which leads to an upper bound since we know that each such connection contributed a nonpositive number. To simplify this step in the argument, we will make one more assumption: $$ \gamma_2 \geq 1 - \mu.$$ Using this assumption, it becomes possible to determine the minimal configuration: each good point on the left-hand side is connected to all points except the $\gamma_2$ points in the right cone as much as possible. Of course, once $\gamma_2 > 1-\mu$, each good point on the left has to connect to at least $(\gamma_2 - (1-\mu))n$ good points on the right side of the cone since a point can only be not connected to at most $(1-\mu)n$ other points. This results in \begin{align} \sum_{i~\tiny \mbox{left}} ~~\sum_{j~\tiny \mbox{right}}{ a_{ij} \cos{(\theta_i - \theta_j)} (w_i - w_j)^2} & \leq \gamma_1 (\gamma_2 - (1-\mu)) \cos{(\pi - 2\varphi)} (1 -(-1))^2 n^2 \notag \\ & = 4 \gamma_1 (\gamma_2 - (1-\mu)) \cos{(\pi - 2\varphi)} n^2. \label{eq:sumleftright} \end{align} As for outliers, we have no control on where they are. Let $i$ be an outlier and consider the quantity that we need to bound $$ S = \sum_{j~\tiny \mbox{left/right cone}}{ a_{ij} \cos{(\theta_i - \theta_j)} (w_i - w_j)^2}.$$ We can increase this contribution by assuming that the outlier is somewhere in $\varphi \leq \theta_i \leq \pi - \varphi$ and that all the good points in the left and right cone are located at angles $\varphi$ and $\pi - \varphi$: this means that the good points in the cone are as close to each other as they are allowed to be from the cone condition which simplifies getting large interactions with both of them from the monotonicity of the cosine in that range. If $i$ is an outlier located at angle $\theta$, this leads to the upper bound $$ S \leq \gamma_1 n (1-v)^2 \cos{(\theta - (\pi - \varphi))} + \gamma_2n (1+v)^2 \cos{(\theta- \varphi)}.$$ We bound expressions of this type via \begin{align*} A \cos{(\theta - (\pi - \phi))} + B \cos{(\theta- \varphi)} &= -A \cos{(\theta + \varphi)} + B \cos{(\theta - \varphi)} \\ &= \mbox{Re} \Bigl(-A e^{i (\theta + \varphi)} + B e^{i(\theta - \varphi)}\Bigr)\\ &= \mbox{Re}~ e^{i \theta} \left( -A e^{i \varphi} + B e^{-i \varphi} \right)\\ &\leq \left| -A e^{i \varphi} + B e^{-i \varphi} \right|. \\ &= \left| -A e^{ 2i \varphi} + B \right|. \end{align*} We note that, by assumption, $|\sin{\varphi}|^2 < 1/2$ and thus, since $A,B$ are positive reals, $$ \left| -A e^{ 2i \varphi} + B \right| \leq \sqrt{A^2+B^2}.$$ This allows us to bound $$ S \leq n\sqrt{ \gamma_1^2 (1-v)^4 + \gamma_2^2 (1+v)^4}.$$ We now choose the value of $v$ that minimizes this expression. This value is $$ v = \frac{\gamma_1 - \gamma_2}{\gamma_1 + \gamma_2}$$ and we obtain $$ S \leq n \frac{4 \gamma_1 \gamma_2 (\gamma_1^2 + \gamma_2^2)^{1/2}}{(\gamma_1 + \gamma_2)^2} = n \frac{4\gamma_1 \gamma_2}{\gamma_1 + \gamma_2} \frac{\sqrt{\gamma_1^2 + \gamma_2^2}}{\gamma_1 + \gamma_2} \leq n \frac{4\gamma_1 \gamma_2}{\gamma_1 + \gamma_2}. $$ Altogether, summing over all the outliers shows $$ \sum_{i~\tiny \mbox{outlier}} \sum_{j~\tiny \mbox{left/right}}{ a_{ij} \cos{(\theta_i - \theta_j)} (w_i - w_j)^2} \leq \frac{\alpha}{\varepsilon} \frac{4\gamma_1 \gamma_2}{\gamma_1 + \gamma_2} n^2.$$ We note that \eqref{eq:gamma12} implies an upper bound on the number of outliers and $$ \frac{1}{\varepsilon} \frac{1}{\gamma_1 + \gamma_2} \leq \frac{1}{\varepsilon}\frac{1}{1 - \frac{\alpha}{\varepsilon}} = \frac{1}{\varepsilon - \alpha}.$$ Combining this with \eqref{eq:sumleftright}, we reach a contradiction to \eqref{positivity} if \begin{equation} \label{condition} \gamma_1(\gamma_2 - (1-\mu)) \cos{(\pi - 2\varphi)} + \frac{\gamma_1 \gamma_2 \alpha}{\varepsilon- \alpha} < 0. \end{equation} Indeed, if this inequality is satisfied, then the quadratic form corresponding to the Hessian is not definite implying that the configuration we are in does not correspond to a local maximum. We will now analyze the condition.\\ \textbf{Step 5: Analyzing the Condition.} We will now try to understand under what conditions \eqref{condition} holds. It clearly requires $\gamma_1 > 0$. We first show that if $\gamma_1 = 0$, then the only stable configuration that can arise is actually the one where all points are in the same spot. Then we deal with the more elaborate case that arises when $\gamma_1 > 0$. \\ \textit{Part 1: $\gamma_1 > 0$.} We will show that if $\gamma_1 = 0$, then $\|r\|$ has to be quite big and this will allow us to immediately deduce the desired statement via the Ling-Xu-Bandeira framework. We now discuss this in greater detail. If $\gamma_1 = 0$, then there are at least $(1- \frac{\alpha}{\varepsilon})n$ points in the right cone since all the good points are in one of the two cones, none of them in the left cone and there are at least that many good points in total. Since the opening angle is less than $45^{\circ}$ (recall that this is an assumption and will be the case for all the parameter we will consider below), we know that the $x-$coordinate of $e^{i \theta_j}$ for each good point is at least $1/\sqrt{2}$. The $x-$coordinate of an outlier is, trivially, at most $-1$. Therefore $$ \|r\| = \left| \sum_{j=1}^{n}{ e^{i \theta_j}} \right| \geq \mbox{Re} \sum_{j=1}^{n}{ e^{i \theta_j}} \geq \frac{1}{\sqrt{2}} \left( 1 - \frac{\alpha}{\varepsilon}\right)n - \frac{\alpha}{\varepsilon}n.$$ Recalling \eqref{bound}, we have, for any $i$, that $$ \left| \sin{(\theta_i - \theta_r)} \right| \leq \frac{(1-\mu)n}{\|r\|} \leq \sqrt{2} \frac{1-\mu}{1 - (1+\sqrt{2})\frac{\alpha}{\varepsilon}}.$$ Recalling our regime of interest, $0 \leq \alpha \leq 0.0537$ and $\mu \geq 0.78$ as well as our parameter selection $\varepsilon =0.5$, we see that $$ \left| \sin{(\theta_i - \theta_r)} \right| \leq \frac{1}{\sqrt{2}} \frac{1-\mu}{1 - 2\frac{\alpha}{\varepsilon}} \leq 0.45 < \frac{1}{\sqrt{2}}.$$ This case thus reduces to the Proposition of Ling-Xu-Bandeira discussed above and we see that the only possible case is that all the points are in the same spot.\\ \textit{Part 2: Using $\gamma_1 > 0$.} We can thus assume that $\gamma_1 > 0$ in which case the configuration is clearly not the one where all the points are in one spot. We obtain a contradiction to \eqref{positivity} if $$ (\gamma_2 - (1-\mu)) \cos{(\pi - 2\varphi)} + \frac{ \gamma_2 \alpha}{\varepsilon- \alpha} < 0.$$ Note that $\cos{(\pi - 2\varphi)} = - \cos(2 \varphi) = -1 + 2 \sin(\varphi)^2$. By assumption, we know that $\sin(\varphi)^2 < 1/2$ and thus the cosine contribution in the above inequality is negative. We reach a contradiction if $$ \gamma_2 \left( \cos{(\pi - 2\varphi)} + \frac{\alpha}{\varepsilon - \alpha} \right) < (1-\mu) \cos{(\pi - 2\varphi)}.$$ We would like to extract a bound for $\gamma_2$ from this but we cannot simply divide by a term without knowing its sign which we now determine. The right-hand side is negative; this means that in order to reach a contradiction, we certainly would need to require the quantity in the parentheses to be negative, i.e. \begin{equation} \label{condition2} \cos{(\pi - 2\varphi)} + \frac{\alpha}{\varepsilon - \alpha} < 0. \end{equation} Once this is the case, we can divide by that term and deduce that we reach a global contradiction if \begin{equation} \label{condition3} \gamma_2 \geq \frac{ (1-\mu) \cos{(\pi - 2\varphi)} }{ \cos{(\pi - 2\varphi)} + \frac{\alpha}{\varepsilon - \alpha} }. \end{equation} Notice that this is condition implies the weaker condition $\gamma_2 \geq (1 - \mu)$ that we assumed above, and thus the latter will be dropped. Summarizing, we have derived a lower bound on $\gamma_2$ that leads to a global contradiction. At this point, we recall that we did already derive a lower bound on $\gamma_2$ in \eqref{eq:lowerboundgamma2} stating that \begin{align*} \gamma_2 & \geq \frac{\cos(\varphi) + \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)^{1/2}}{1 + \cos{(\varphi)}} - \frac{\alpha}{\varepsilon} \\ & = 1 - \frac{\alpha}{\varepsilon} - \frac{1 - \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)^{1/2}}{1 + \cos{(\varphi)}}. \end{align*} It is not clear that if our already established lower bound \eqref{eq:lowerboundgamma2} is larger than the lower bound leading to a contradiction, i.e. if $$ 1 - \frac{\alpha}{\varepsilon} - \frac{1 - \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)^{1/2}}{1 + \cos{(\varphi)}} \geq \frac{ (1-\mu) \cos{(\pi - 2\varphi)} }{ \cos{(\pi - 2\varphi)} + \frac{\alpha}{\varepsilon - \alpha} }, $$ then we have obtained a contradiction. Any collection of points with these parameters must necessarily give rise to a Hessian with a negative definite eigenvalue and thus cannot be a local maximum.\\ \textbf{Summary.} In order to obtain a contradiction, it suffices to find, for each $\alpha \leq 0.0537$ two variables $\varepsilon$ and $\delta$ $$ \alpha < \varepsilon < \delta < 1$$ such that, abbreviating once again, $$ s_{\delta} = \frac{1}{\sqrt{2}} \sqrt{ \sqrt{9 - 4 \delta} - 3 + 2\delta},$$ the following properties hold \begin{enumerate} \item we have $$\frac{ \left( s_{\delta} + \frac{\varepsilon}{\delta} (1-s_{\delta})\right)^2 (1-\mu)^2}{ \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)} < \frac{1}{2}.$$ This says that the parameters define a critical angle $\varphi = \varphi(\alpha, \mu, \varepsilon, \delta)$ corresponding to an angle less than $45^{\circ}$ which is required for our argument. \item we require that this angle satisfies $$ \cos{(\pi - 2\varphi)} + \frac{\alpha}{\varepsilon - \alpha} < 0$$ which was required when dividing by it and flipping the sign in Step 5, Part 2. \item we also require the angle $\varphi$ satisfies $$ 1 - \frac{\alpha}{\varepsilon} - \frac{1 - \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)^{1/2}}{1 + \cos{(\varphi)}} > \frac{ (1-\mu) \cos{(\pi - 2\varphi)} }{ \cos{(\pi - 2\varphi)} + \frac{\alpha}{\varepsilon - \alpha} }.$$ This shows that the lower bound \eqref{eq:lowerboundgamma2} we derived for $\gamma_2$ is big enough to imply a contradiction, we refer to Step 5, Part 2.\\ \end{enumerate} \textbf{Conclusion.} We distinguish the cases $\alpha \leq 0.0537$ and $\alpha > 0.0537$. If $\alpha > 0.0537$, then we immediately obtain a contradiction if $$ \mu \geq 0.788897.$$ This follows from rerunning the Ling-Xu-Bandeira argument, described in \S 2.1. verbatim and using the definition of $\alpha$, and the value of $\mu$ comes from \eqref{length} and \eqref{bound}. Let us now assume that $\alpha \leq 0.0537$. We set $$ \varepsilon = 0.5 \qquad \mbox{and} \qquad \delta = 0.88.$$ An easy (mathematica) check shows that we obtain a contradiction for the entire range $0 \leq \alpha \leq 0.0537$ and $0.788897 \leq \mu \leq 0.794$, where $0.794$ is the bound proved in \cite{ling}. More precisely, the inequalities are true with room to spare: we have, over this entire parameter range $$\frac{ \left( s_{\delta} + \frac{\varepsilon}{\delta} (1-s_{\delta})\right)^2 (1-\mu)^2}{ \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)} < 0.46 < \frac{1}{2}$$ $$ \cos{(\pi - 2\varphi)} + \frac{\alpha}{\varepsilon - \alpha} < -0.05 < 0$$ and $$ 1 - \frac{\alpha}{\varepsilon} - \frac{1 - \left( \frac{1}{2} - (2-\alpha)(1-\mu)\right)^{1/2}}{1 + \cos{(\varphi)}} > \frac{ (1-\mu) \cos{(\pi - 2\varphi)} }{ \cos{(\pi - 2\varphi)} + \frac{\alpha}{\varepsilon - \alpha} } + 0.004 .$$ It is the last expression which is almost satisfied and does not allow an extension to smaller parameter ranges (the critical values occur when $\alpha \sim 0.0537$ and $\mu \sim 0.7889$), the inequality is satisfied with a much bigger gap away from these parameters. \end{proof} \textbf{Acknowledgment.} We are grateful to the anonymous referees for the substantial and detailed reports resulting in a greatly improved manuscript.
{ "timestamp": "2020-04-20T02:14:12", "yymm": "1911", "arxiv_id": "1911.12336", "language": "en", "url": "https://arxiv.org/abs/1911.12336", "abstract": "We study synchronization properties of systems of Kuramoto oscillators. The problem can also be understood as a question about the properties of an energy landscape created by a graph. More formally, let $G=(V,E)$ be a connected graph and $(a_{ij})_{i,j=1}^{n}$ denotes its adjacency matrix. Let the function $f:\\mathbb{T}^n \\rightarrow \\mathbb{R}$ be given by $$ f(\\theta_1, \\dots, \\theta_n) = \\sum_{i,j=1}^{n}{ a_{ij} \\cos{(\\theta_i - \\theta_j)}}.$$ This function has a global maximum when $\\theta_i = \\theta$ for all $1\\leq i \\leq n$. It is known that if every vertex is connected to at least $\\mu(n-1)$ other vertices for $\\mu$ sufficiently large, then every local maximum is global. Taylor proved this for $\\mu \\geq 0.9395$ and Ling, Xu \\& Bandeira improved this to $\\mu \\geq 0.7929$. We give a slight improvement to $\\mu \\geq 0.7889$. Townsend, Stillman \\& Strogatz suggested that the critical value might be $\\mu_c = 0.75$.", "subjects": "Optimization and Control (math.OC); Dynamical Systems (math.DS); Adaptation and Self-Organizing Systems (nlin.AO)", "title": "Synchronization of Kuramoto Oscillators in Dense Networks", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9907319885346244, "lm_q2_score": 0.8080672204860316, "lm_q1q2_score": 0.8005780442217729 }
https://arxiv.org/abs/2103.15097
Compound matrices in systems and control theory
The multiplicative and additive compounds of a matrix play an important role in several fields of mathematics including geometry, multi-linear algebra, combinatorics, and the analysis of nonlinear time-varying dynamical systems. There is a growing interest in applications of these compounds, and their generalizations, in systems and control theory. This tutorial paper provides a gentle introduction to these topics with an emphasis on the geometric interpretation of the compounds, and surveys some of their recent applications.
\section{Introduction} Let~$A\in\R^{n\times n}$. Fix~$k\in\{1,\dots,n\}$. The $k$-multiplicative and $k$-additive compounds of~$A$ play an important role in several fields of applied mathematics. Recently, there is a growing interest in the applications of these compounds, and their generalizations, in systems and control theory (see, e.g.~\cite{cheng_diag_stab,kordercont,rami_osci,rola_spect,CTPDS,margaliot2019revisiting, pines2021,Eyal_k_posi,DT_K_POSI,9107214, gruss1,gruss2,gruss3,gruss4,gruss5}). This tutorial paper reviews the~$k$-compounds, focusing on their geometric interpretations, and surveys some of their recent applications in systems and control theory, including~$k$-positive systems, $k$-cooperative systems, and~$k$-contracting systems. The results described here are known, albeit never collected in a single paper. The only exception is the new generalization principle described in Section~\ref{sec:k_generalizations}. We use standard notation. For a set~$S$, $\operatorname{int}(S)$ is the interior of~$S$. For scalars~$\lambda_i$, $i\in\{1,\dots, n\}$, $\operatorname{diag}(\lambda_1,\dots,\lambda_n)$ is the $n\times n$ diagonal matrix with diagonal entries~$\lambda_i$. Column vectors [matrices] are denoted by small [capital] letters. For a matrix~$A$, $A^T$ is the transpose of~$A$. For a square matrix~$B$, $\operatorname{tr}(B)$ [$\det(B)$] is the trace [determinant] of~$B$. $B $ is called Metzler if all its off-diagonal entries are non-negative. The positive orthant in~$\R^n$ is~$\R^n_+ = \{x\in\R^n: x_i\geq 0,\; i=1,\dots,n\}$. \section{Geometric motivation} $k$-compound matrices provide information on the evolution of $k$-dimensional polygons subject to a linear time-varying dynamics. To explain this in a simple setting, consider the LTI system: \begin{equation}\label{eq:ltia} \dot x =\operatorname{diag}(\lambda_1,\lambda_2,\lambda_3)x , \end{equation} with~$\lambda_i\in\R$ and~$x:\R_+ \to\R^3$. Let~$e^i$, $i=1,2,3$, denote the~$i$th canonical vector in~$\R^3$. For~$x(0)=e^i$ we have~$x(t)=\exp(\lambda_i t)x(0)$. Thus, $\exp(\lambda_i t)$ describes the rate of evolution of the line~$e^i$ subject to~\eqref{eq:ltia}. What about 2D areas? Let~$S\subset\R^3 $ denote the square generated by~$e^i$ and~$e^j$, with~$i\not =j$. Then~$S(t):= x(t,S )$ is the rectangle generated by~$\exp(\lambda_i t) e^i$ and~$\exp(\lambda_j t) e^j$, so the area of~$S(t)$ is~$\exp( (\lambda_i + \lambda_j) t)$. Similarly, if~$B \subset\R^3 $ is the 3D box generated by~$e^1,e^2$, and~$e^3$ then the volume of~$B(t):=x(t,B)$ is~$\exp( (\lambda_1+\lambda_2+\lambda_3)t)$ (see Fig.~\ref{fig:GeometricalEvolution}). Since~$\exp(At)=\operatorname{diag} ( \exp(\lambda_1 t), \exp(\lambda_2 t), \exp(\lambda_3 t) )$, this discussion suggests that it may be useful to have a~$3\times 3$ matrix whose eigenvalues are the sums of any two eigenvalues of~$\exp(At)$, and a~$1\times 1$ matrix whose eigenvalue is the sum of the three eigenvalues of~$\exp(At)$. With this geometric motivation in mind, we turn to recall the notions of the multiplicative and additive compounds of a matrix. For more details and proofs, see e.g.~\cite[Ch.~6]{fiedler_book}\cite{schwarz1970}. \begin{figure} \begin{center} \includegraphics[scale=0.33]{GeometricalEvolution.eps} \caption{The evolution of lines, areas, and volumes under the the LTI~\eqref{eq:ltia}.} \label{fig:GeometricalEvolution} \end{center} \end{figure} \section{Compound Matrices} Let~$A\in\C^{n\times m }$. Fix~$k\in\{1,\dots,\min\{n,m\} \}$. Let~$Q(k,n)$ denote the set of increasing sequences of~$k$ integers in~$\{1,\dots,n\}$, ordered lexicographically. For example, \[ Q(2,3)=\{ \{1,2\} ,\{1,3\},\{2,3\} \}. \] For~$\alpha \in Q(k,n),\beta \in Q(k,m)$, let~$A[\alpha|\beta]$ denote the $ k\times k$ submtarix obtained by taking the entries of~$A$ in the rows indexed by~$\alpha$ and the columns indexed by~$\beta$. For example \[ A[ \{1,2\}|\{2,3\}]=\begin{bmatrix} a_{12} & a_{13}\\ a_{22} &a_{23} \end{bmatrix} . \] The minor of~$A $ corresponding to~$\alpha , \beta$ is~$A(\alpha|\beta):=\det (A[\alpha|\beta]) $. For example,~$Q(n,n)$ includes the single set~$\alpha=\{1,\dots,n\}$ and~$A(\alpha|\alpha)=\det(A)$. \begin{Definition}\label{def:multi} Let $A\in C^{n\times m}$ and fix $k \in \{ 1,\dots,\min \{ n,m \} \}$. The \emph{$k$-multiplicative compound} of~$A$, denoted~$A^{(k)}$, is the~$\binom{n}{k}\times \binom{m}{k}$ matrix that contains all the~$k$-order minors of~$A$ ordered lexicographically. \end{Definition} For example, if~$n=m=3$ and~$k=2$ then \[ A^{(2)}= \begin{bmatrix} A(\{12\}|\{12\}) & A(\{12\}|\{13\}) & A(\{12\}|\{23\}) \\ A(\{13\}|\{12\}) & A(\{13\}|\{13\}) & A(\{13\}|\{23\}) \\ A(\{23\}|\{12\}) & A(\{23\}|\{13\}) & A(\{23\}|\{23\}) \end{bmatrix}. \] In particular, Definition~\ref{def:multi} implies that~$A^{(1)}=A$, and if~$n=m$ then~$A^{(n)}=\det(A)$. Let~$A\in\mathbb{C}^{n\times m},\; B\in\mathbb{C}^{m\times p}$. The Cauchy-Binet Formula (see e.g.~\cite{notes_comb_algebra}) asserts that for any $k\in \{1,\dots,\min \{n,m,p\} \}$, \begin{align}\label{eq:AB_MultComp} (AB)^{(k)} = A^{(k)} B^{(k)}. \end{align} Hence the term multiplicative compound. Note that for~$n=m=p$, Eq.~\eqref{eq:AB_MultComp} with~$k=n$ reduces to the familiar formula~$\det(AB)= \det (A) \det (B)$. Let~$I_s$ denote the~$s\times s$ identity matrix. Definition~\ref{def:multi} implies that~$I_n^{(k)} = I_r$, where~$r:=\binom{n}{k}$. Hence, if~$A\in\R^{n\times n}$ is non-singular then~$(AA^{-1})^{(k)}=I_r$ and combining this with~\eqref{eq:AB_MultComp} yields~$(A^{-1})^{(k)} = (A^{(k)})^{-1}$. In particular, if~$A$ is non-singular then so is~$A^{(k)}$. The~$k$-multiplicative compound has an important spectral property. For~$A\in\C^{n\times n}$, let~$\lambda_i$,~$i=1,\dots,n$, denote the eigenvalues of~$A$. Then the eigenvalues of~$A^{(k)}$ are all the products \begin{equation}\label{eq:prodeig} \prod_{\ell=1}^k \lambda_{i_\ell}, \text{ with } 1\leq i_1< i_2<\dots< i_k \leq n. \end{equation} For example, suppose that~$n=3$ and $ A=\begin{bmatrix} a_{11} & a_{12} & a_{13}\\ 0 & a_{22} & a_{23}\\ 0 & 0 & a_{ 33} \end{bmatrix} . $ Then a calculation gives \[ A^{(2)}= \begin{bmatrix} a_{11} a_{22} & a_{11} a_{23}& a_{12} a_{23}-a_{13} a_{22} \\ 0& a_{11} a_{33}& a_{12} a_{33}\\ 0& 0& a_{22} a_{33} \end{bmatrix} , \] so, clearly, the eigenvalues of~$A^{(2)}$ are of the form~\eqref{eq:prodeig}. \begin{Definition} Let~$A\in\mathbb{C}^{n\times n}$. The \emph{$k$-additive compound} matrix of~$A$ is defined by: \begin{align*} A^{[k]} := \frac{d}{d\epsilon} (I+\epsilon A)^{(k)} |_{\epsilon=0} . \end{align*} \end{Definition} Note that this implies that~$ A^{[k]} = \frac{d}{d\epsilon} (\exp(A\epsilon ))^{(k)} |_{\epsilon=0}$, and also that \begin{align}\label{eq:(I+epsilonA)^k} (I+\epsilon A)^{(k)} = I + \epsilon A^{[k]} + o(\epsilon), \end{align} In other words,~$A^{[k]}$ is the first-order term in the Taylor expansion of~$(I+\epsilon A)^{(k)}$. Let~$\lambda_i$,~$i=1,\dots,n$, denote the eigenvalues of~$A$. Then the eigenvalues of~$(I+\epsilon A)^{(k)}$ are the products~$\prod_{\ell=1}^k (1+\epsilon \lambda_{i_\ell}) $, and~\eqref{eq:(I+epsilonA)^k} implies that the eigenvalues of~$A^{[k]}$ are all the sums \[ \sum_{\ell=1}^k \lambda_{i_\ell}, \text{ with } 1\leq i_1< i_2<\dots< i_k \leq n. \] Another important implication of the definitions above is that for any~$A,B\in\C^{n\times n}$ we have \[ (A+B)^{[k]}=A^{[k]}+B^{[k]}. \] This justifies the term additive compound. Moreover, the mapping~$A\to A^{[k]}$ is linear. The following result gives a useful explicit formula for~$A^{[k]}$ in terms of the entries~$a_{ij}$ of~$A$. Recall that any entry of~$A^{(k)}$ is a minor~$A(\alpha|\beta)$. Thus, it is natural to index the entries of~$A^{(k)}$ and~$A^{[k]}$ using~$\alpha,\beta \in Q(k,n)$. \begin{Proposition}\label{prop:Explicit_A_k} Fix~$\alpha,\beta \in Q(k,n)$ and let~$\alpha=\{i_1,\dots,i_k\}$ and~$\beta=\{j_1,\dots,j_k\}$. Then the entry of~$A^{[k]}$ corresponding to~$(\alpha,\beta)$ is equal to: \begin{enumerate} \item $\sum_{\ell=1}^{k} a_{ i_{\ell} i_{\ell} }$, if $i_{\ell} = j_{\ell}$ for all $\ell \in \{ 1,\hdots,k \}$; % \item $(-1)^{\ell +m} a_{i_{\ell} j_{m}} $, if all the indices in $ \alpha $ and $ \beta$ agree, except for a single index $ i_{\ell} \ne j_m$; and % \item $0$, otherwise. \end{enumerate} \end{Proposition} \begin{Example}\label{ex:A^[3]} For~$A\in\R^{4\times 4}$ and~$k=3$, Prop.~\ref{prop:Explicit_A_k} yields \[ A^{[3]} = \left[ \scalemath{0.67}{ \begin{array}{ccccccc} a_{11}+a_{22}+a_{33} & a_{34} & -a_{24} & a_{14} \\ a_{43} & a_{11}+a_{22}+a_{44} & a_{23} & -a_{13} \\ -a_{42} & a_{32} & a_{11}+a_{33}+a_{44} & a_{12} \\ a_{41} & -a_{31} & a_{21} & a_{22}+a_{33}+a_{44} \end{array}} \right]. \] The entry in the second row and fourth column of $A^{[3]}$ corresponds to $(\alpha | \beta) = ( \{ 1,2,4 \} | \{ 2,3,4 \} )$. As $\alpha$ and $\beta$ agree in all indices except for the index $i_{ 1} =1$ and $j_{ 2} =3$, this entry is equal to~$ (-1)^{1+2} a_{13} = -a_{13}$. \end{Example} Note that Prop.~\ref{prop:Explicit_A_k} implies in particular that~$A^{[n]}=\operatorname{tr}(A)$. The next section describes applications of compound matrices for dynamical systems described by~ODEs. For more details and proofs, see~\cite{schwarz1970,muldowney1990}. \section{Compound Matrices and ODEs} Fix an interval~$[a,b] \in \R$. Let~$A:[a,b] \to \R^{n\times n} $ be a continuous matrix function, and consider the LTV system: \begin{equation}\label{Eq:ltv} \dot x(t)=A(t)x(t),\quad x(a)=x_0. \end{equation} The solution is given by~$x(t)=\Phi(t ,a)x(a)$, where~$\Phi(t,a )$ is the solution at time~$t$ of the matrix differential equation \begin{equation}\label{eq:phidot} \dot \Phi(s)=A(s) \Phi(s), \quad \Phi(a)=I_n. \end{equation} Fix~$k\in\{1,\dots,n\}$ and let~$r:=\binom{n}{k}$. A natural question is: how do the~$k$-order minors of~$\Phi(t)$ evolve in time? The next result provides a beautiful formula for the evolution of~$\Phi^{(k)}(t):= (\Phi(t))^{(k)}$. \begin{Proposition}\label{prop:odeforphik} If~$\Phi$ satisfies~\eqref{eq:phidot} then \begin{equation}\label{eq:expat} \frac{d}{dt} \Phi^{(k)}(t)=A^{[k]}(t) \Phi^{(k)}(t),\quad \Phi^{(k)}(a)=I_r, \end{equation} where~$A^{[k]}(t):= (A(t)) ^ {[k]} $. \end{Proposition} Thus, the~$k\times k$ minors of~$\Phi$ also satisfy an~LTV. In particular, if~$A(t)\equiv A$ and~$a=0$ then~$\Phi(t)=\exp( At) $ so~$\Phi^{(k)}(t)=(\exp(At))^{(k)}$ and~\eqref{eq:expat} gives \[ (\exp(At))^{(k)} = \exp(A ^{[k]} t). \] Note also that for~$k=n$, Prop.~\ref{prop:odeforphik} yields \[ \frac{d}{dt} \det(\Phi(t)) = \operatorname{tr}(A(t)) \det(\Phi(t)), \] which is the Abel-Jacobi-Liouville identity. Roughly speaking, Prop.~\ref{prop:odeforphik} implies that under the LTV dynamics~\eqref{Eq:ltv}, $k$-dimensional polygons evolve according to the dynamics~\eqref{eq:expat}. We now turn to consider the nonlinear system: \begin{equation}\label{eq:nonlin} \dot x (t) = f(t,x). \end{equation} For the sake of simplicity, we assume that the initial time is zero, and that the system admits a convex and compact state-space~$\Omega$. We also assume that~$f\in C^1$. The Jacobian of the vector field~$f$ is~$J(t,x):=\frac{\partial}{\partial x}f(t,x)$. Compound matrices can be used to analyze~\eqref{eq:nonlin} by using an LTV called the variational equation associated with~\eqref{eq:nonlin}. To define it, fix~$a,b \in \Omega$. Let~$z(t):=x(t,a)-x(t,b)$, and for~$s\in[0,1]$, let~$\gamma(s):=s x(t,a)+(1-s)x(t,b)$, i.e. the line connecting~$x(t,a)$ and~$x(t,b)$. Then \begin{align*} \dot z(t)&=f(t,x(t,a))-f(t,x(t,b)) \\ &=\int_0^1 \frac{\partial }{\partial s} f(t,\gamma(s)) \dif s , \end{align*} and this gives the variational equation: \begin{equation}\label{eq:var_eqn} \dot z(t)=A^{ab}(t)z(t), \end{equation} where \begin{equation}\label{eq:at_int} A^{ab}(t):=\int_0^1 J(t,\gamma(s)) \dif s. \end{equation} \begin{comment} Let~$W(t,x_0):=\frac{\partial}{\partial x_0}x(t,x_0)$, that is, the change in the solution at time~$t$ due to a small change in the initial condition~$x_0$. Then \begin{equation}\label{eq:var_eqn} \dot W(t,x_0)= J(t,x(t,x_0)) W(t,x_0),\quad W(0,x_0)=I_n. \end{equation} This LTV is the \emph{variational equation along the solution~$x(t,x_0)$}. Prop.~\ref{prop:odeforphik} implies that \begin{equation}\label{eq:varewithk} \frac{d}{dt} W^{(k)}(t,x_0)= J^{[k]}(t,x(t,x_0)) W^{(k)}(t,x_0), \end{equation} with~$W^{(k)}(0,x_0)=I_r$. \end{comment} We can use the results above to describe a powerful approach for deriving useful ``$k$-generalizations'' of important classes of dynamical systems including cooperative systems~\cite{hlsmith}, contracting systems~\cite{sontag_cotraction_tutorial,LOHMILLER1998683}, and diagonally stable systems~\cite{diag_Stab_book}. \section{$k$-generalizations of dynamical systems}\label{sec:k_generalizations} Consider the LTV~\eqref{Eq:ltv}. Suppose that~$A(t)$ satisfies a specific \emph{property}, e.g. \emph{property} may be that~$A(t)$ is Metzler for all~$t$ (so the LTV is positive) or that~$\mu(A(t))\leq-\eta<0$ for all~$t$, where~$\mu:\R^{n\times n}\to\R$ is a matrix measure (so the~LTV is contracting). Fix~$k\in\{1,\dots,n\}$. We say that the LTV satisfies~\emph{$k$-property} if~$A^{[k]}$ (rather than~$A$) satisfies \emph{property}. For example, the LTV is~\emph{$k$-positive} if $A^{[k]}(t)$ is Metzler for all~$t$; the LTV is~\emph{$k$-contracting} if~$\mu(A^{[k]}(t))\leq-\eta<0$ for all~$t$, and so on. This generalization approach makes sense for two reasons. First, when~$k=1$,~$A^{[k]}$ reduces to~$A^{[1]}=A$, so \emph{$k$-property} is clearly a generalization of \emph{property}. Second, we know that~$A^{[k]}$ has a clear geometric meaning, as it describes the evolution of~$k$-dimensional polygons along the dynamics. The same idea can be applied to the nonlinear system~\eqref{eq:nonlin} using the variational equation~\eqref{eq:var_eqn}. For example, if~$\mu(J(t,x))\leq -\eta<0$ for all~$t\geq 0$ and~$x\in\Omega$ then~\eqref{eq:nonlin} is contracting: the distance between any two solutions (in the norm that induced~$\mu$) decays at an exponential rate. If we replace this by the condition~$\mu(J^{[k]}(t,x))\leq -\eta<0$ for all~$t\geq 0$ and~$x\in\Omega$ then~\eqref{eq:nonlin} is called~$k$-contracting. Roughly speaking, this means that the volume of $k$-dimensional polygons decays to zero exponentially along the flow of the nonlinear system. We now turn to describe several such~$k$-generalizations. \section{$k$-contracting systems} $k$-contracting systems were introduced in~\cite{kordercont} (see also the unpublished preprint~\cite{weak_manchester} for some preliminary ideas). For~$k=1$ these reduce to contracting systems. This generalization was motivated in part by the seminal work of Muldowney~\cite{muldowney1990} who considered nonlinear systems that, in the new terminology, are~$2$-contracting. He derived several interesting results for time-invariant $2$-contracting systems. For example, every bounded trajectory of a time-invariant, nonlinear, $2$-contracting system converges to an equilibrium (but, unlike in the case of contracting systems, the equilibrium is not necessarily unique). For the sake of simplicity, we introduce the ideas in the context of an LTV system. The analysis of nonlinear systems is based on assuming that their variational equation is a~$k$-contracting LTV. Recall that a vector norm $|\cdot |:\mathbb{R}^{n}\to\mathbb{R}_{+}$ induces a matrix norm $||\cdot ||:\mathbb{R}^{n\times n}\to\mathbb{R}_{+}$ by: \begin{align*} ||A||:=\max_{|x|=1} |Ax|, \end{align*} and a matrix measure $\mu (\cdot ):\mathbb{R}^{n\times n}\to\mathbb{R}$ by: \begin{align*} \mu (A) :=\lim_{\epsilon \downarrow 0} \frac{||I+\epsilon A||-1}{\epsilon}. \end{align*} For~$p\in\{1,2,\infty\}$, let~$|\cdot|_p:\R^n\to\R_+$ denote the~$L_p$ vector norm, that is, $|x|_1:=\sum_{i=1}^n|x_i|$, $|x|_2:=\sqrt{x^T x}$, and~$|x|_\infty:=\max_{ i\in\{1,\dots,n\} } |x_i|$. The induced matrix measures are~\cite{vid}: \begin{equation}\label{eq:matirx_measures_12infty} \begin{aligned} \mu_1(A) &= \max_{j\in\{1,\dots,n\}} (a_{jj}+ \sum_{\substack{i=1 \\ i \neq j}}^n |a_{ij}| ) , \\ \mu_2(A) &= \lambda_1 ( {A + A^T} )/2 , \\ \mu_{\infty}(A) &= \max_{i\in\{1,\dots,n\}} (a_{ii}+ \sum_{\substack{j=1 \\ j \neq i}}^n |a_{ij}| ) , \end{aligned} \end{equation} where~$\lambda_i (S)$ denotes the $i$-th largest eigenvalue of the symmetric matrix~$S$, that is, \[ \lambda_1(S) \geq \lambda_2(S) \geq \cdots \geq \lambda_n(S). \] \begin{Definition} The LTV~\eqref{Eq:ltv} is called~\emph{$k$-contracting} with respect to (w.r.t.) the norm~$|\cdot| $ if \begin{equation}\label{eq:kconinfi} \mu(A^{[k]}(t))\leq-\eta<0 \text{ for all } t \geq 0, \end{equation} where~$\mu$ is the matrix measure induced by~$|\cdot|$. \end{Definition} Note that for~$k=1$ condition~\eqref{eq:kconinfi} reduces to the standard infinitesimal condition for contraction~\cite{sontag_cotraction_tutorial}. For the~$L_p$ norms, with~$p\in\{1,2,\infty\}$, this condition is easy to check using the explicit expressions for~$\mu_1$, $\mu_2$, and~$\mu_\infty$. This carries over to~$k$-contraction, as combining Prop.~\ref{prop:Explicit_A_k} with~\eqref{eq:matirx_measures_12infty} gives~\cite{muldowney1990}: \begin{align*} \mu_1(A^{[k]}) &= \max_{\alpha \in Q(k,n) } ( \sum_{p=1}^k a_{\alpha_p,\alpha_p} + \sum_{\substack{j \notin \alpha }}(|a_{j,\alpha_1}| + \cdots + |a_{j,\alpha_k}|) ) ,\nonumber \\ \mu_2(A^{[k]}) &= \sum_{i=1}^k \lambda_i ( {A + A^T} )/2 , \\ \mu_{\infty}(A^{[k]}) &= \max_{\alpha \in Q(k,n) } ( \sum_{p=1}^k a_{\alpha_p,\alpha_p} + \sum_{\substack{j \notin\alpha }}(|a_{\alpha_1,j}| + \cdots + |a_{\alpha_k,j}|) ) . \end{align*} For~$k=n$, $A^{[n]}$ is the scalar~$\operatorname{tr}(A)$, so condition~\eqref{eq:kconinfi} becomes~$\operatorname{tr}(A(t))\leq -\eta<0 $ for all $t \geq 0$. Combining Coppel's inequality~\cite{coppel1965stability} with~\eqref{eq:expat} yields the following result. \begin{Proposition} If the LTV~\eqref{Eq:ltv} is~$k$-contracting then \[ ||\Phi^{(k)}(t) || \leq \exp(-\eta t) || \Phi^{(k)}(0)|| = \exp(-\eta t) \] for all~$t\geq 0$. \end{Proposition} Geometrically, this means that under the LTV dynamics the volume of $k$-dimensional polygons converges to zero exponentially. The next example illustrates this. \begin{Example}\label{exa:squares} Consider the LTV~\eqref{Eq:ltv} with~$n=2$ and $ A(t) = \begin{bmatrix} -1 & 0 \\ -2\cos( t) &0 \end{bmatrix}. $ The corresponding transition matrix is: $ \Phi(t)= \begin{bmatrix} \exp(-t)&0 \\ -1+\exp(-t)(\cos(t)-\sin(t)) &1\end{bmatrix}. $ This implies that the LTV is uniformly stable, and that for any~$x(0)\in\R^2$ we have \[ \lim _{t\to \infty} x(t,x(0))=\begin{bmatrix} 0 \\ x_2(0)-x_1(0) \end{bmatrix} . \] The LTV is not contracting, w.r.t. any norm, as there is more than a single equilibrium. However,~$ A^{[2]}(t) =\operatorname{tr}(A(t)) \equiv -1$, so the system is~$2$-contracting. Let~$S\subset\R^2$ denote the unit square, and let~$S(t):=x(t,S )$, that is, the evolution at time~$t$ of the unit square under the dynamics. Fig.~\ref{fig:squares} depicts~$S(t) +2t$ for several values of~$t$, where the shift by~$2t$ is for the sake of clarity. It may be seen that the area of~$S(t)$ decays with~$t$, and that~$S(t)$ converges to a line. \end{Example} \begin{figure} \begin{center} \includegraphics[scale=0.5]{squares.eps} \caption{Evolution of the unit square in Example~\ref{exa:squares}.} \label{fig:squares} \end{center} \end{figure} As noted above, time-invariant $2$-contracting systems have a ``well-odered'' asymptotic behaviour~\cite{muldowney1990,li1995}, and this has been used to derive a global analysis of important models from epidemiology (see, e.g.~\cite{SEIR_LI_MULD1995}). A recent paper~\cite{searial12contracting} extended some of these results to systems that are not necessarily~$2$-contracting, but can be represented as the serial interconnections of~$k$-contracting systems, with~$k\in\{1,2\}$. \section{$\alpha$-compounds and $\alpha$-contracting systems} A recent paper~\cite{pines2021} defined a generalizations called the~$\alpha$-multiplicative compound and~$\alpha$-additive compound of a matrix, where~$\alpha$ is a \emph{real} number. Let~$A\in\C^{n\times n} $ be non-singular. If~$\alpha=k+s$, where~$k\in\{1,\dots,n-1\}$ and~$s\in(0,1)$ then the~$\alpha$-multiplicative of~$A$ is defined by: \[ A^{(\alpha)} : = (A^{(k)})^{1-s} \otimes (A^{(k+1)})^{s}, \] where~$\otimes$ denotes the Kronecker product. This is a kind of ``multiplicative interpolation'' between~$ A^{(k)} $ and~$A^{(k+1)} $. For example,~$A^{(2.1)} = (A^{(2)})^{0.9} \otimes (A^{(3)})^{0.1}$. The~$\alpha$-additive compound is defined just like the~$k$-additive compound, that is, \begin{align*} A^{[\alpha]} := \frac{d}{d\epsilon} (I+\epsilon A)^{(\alpha)} |_{\epsilon=0} , \end{align*} and it was shown in~\cite{pines2021} that this yields \[ A^{[\alpha]} = ((1-s)A^{[k]}) \oplus (sA^{[k+1]}), \] where~$\oplus $ denotes the Kronecker sum. The system~\eqref{eq:nonlin} is called \emph{$\alpha$-contracting} w.r.t. the norm~$|\cdot| $ if \begin{equation} \label{eq:alphacon} \mu(J^{[\alpha]}(t,x))\leq -\eta<0 , \end{equation} for all~$t\geq 0$ and~all~$x$ in the state space~\cite{pines2021}. Using this notion, it is possible to restate the seminal results of Douady and Oesterl\'{e}~\cite{Douady1980} as follows. \begin{Theorem}\cite{pines2021} Suppose that~\eqref{eq:nonlin} is~$\alpha$-contracting for some~$\alpha\in [1,n]$. Then any compact and strongly invariant set of the dynamics has a Hausdorff dimension smaller than~$\alpha$. \end{Theorem} Roughly speaking, the dynamics contracts sets with a larger Hausdorff dimension. The next example, adapted from~\cite{pines2021}, shows how these notions can be used to ``de-chaotify'' a nonlinear dynamical system by feedback. \begin{Example} Thomas' cyclically symmetric attractor~\cite{thomas99,chaos_survey} is a popular example for a chaotic system. It is described by: \begin{align} \label{eq:thom} \dot x_1 =& \sin(x_2)-bx_1 , \nonumber \\ \dot x_2 =& \sin(x_3) - bx_2, \\ \dot x_3 =& \sin(x_1) - bx_3, \nonumber \end{align} where $b>0$ is the dissipation constant. The convex and compact set~$D: = \{x\in\R^3: b |x|_\infty \leq 1 \}$ is an invariant set of the dynamics. Fig.~\ref{fig:chaos} depicts the solution of the system emanating from~$\begin{bmatrix} 1& -2&1\end{bmatrix}^T$ for $ b=0.1$. Note the symmetric strange attractor. The Jacobian $J_f$ of the vector field in~\eqref{eq:thom} is \begin{align*} J_f (x)=\begin{bmatrix}-b&\cos(x_2)&0 \\ 0&-b&\cos(x_3) \\ \cos(x_1)&0&-b\end{bmatrix}, \end{align*} and thus \begin{align*} J_f ^{[2]}(x)=\begin{bmatrix}-2b&\cos(x_3)&0 \\ 0&-2b&\cos(x_2) \\ -\cos(x_1)&0&-2b\end{bmatrix}, \end{align*} and~$ J_f ^{[3]}=\operatorname{trace}(J (x))=-3b$. Since~$b>0$, this implies that the system is $3$-contracting w.r.t. any norm. Let~$\alpha = 2+s$, with~$s\in(0,1)$. Then \begin{align*} &J_f ^{[\alpha]}(x) =(1-s)J_f^{[2]}(x)\oplus sJ_f^{[3]}(x)\\ &= \begin{bmatrix} -(2+s)b & (1-s)\cos(x_3) & 0 \\ 0 & -(2+s)b & (1-s)\cos(x_2) \\ -(1-s)\cos(x_1) & 0 & -(2+s)b \\ \end{bmatrix}. \end{align*} This implies that \begin{equation}\nonumber \mu_{1 }(J_f ^{[\alpha]}(x)) \leq 1-2b-s(b+1), \text{ for all } x\in D. \end{equation} We conclude that for any~$b\in(0,1/2) $ the system is~$(2+s)$-contracting for any $ s>\frac{1-2b}{1+b} $. We now show how $\alpha$-contarction can be used to design a partial-state controller for the system guaranteeing that the closed-loop system has a ``well-ordered'' behaviour. Suppose that the closed-loop system is: \[ \dot x = f(x) + g(x), \] where~$g $ is the controller. Let~$\alpha = 2 + s$, with~$s \in(0,1)$. The Jacobian of the closed-loop system is~$J_{cl}:=J_f+J_g$, so \begin{align*} \mu_1(J_{cl}^{[\alpha]})& = \mu_1(J_f^{[\alpha]}+J_g^{[\alpha]}) \\& \leq \mu_1(J_f^{[\alpha]})+\mu_1(J_g^{[\alpha]}) \\ &\leq 1-2b-s(b+1)+\mu_1(J_g^{[\alpha]}). \end{align*} This implies that the closed-loop system is~$\alpha$-contracting if \begin{align}\label{eq:cond_cont} \mu_1(J^{[\alpha]}_g (x) ) < s(b+1)+2b-1 \text{ for all } x\in D . \end{align} Consider, for example, the controller $ g(x_1,x_2)=c \operatorname{diag}(1,1,0) x $, with gain~$c<0$. Then $ J_g^{[\alpha]} = c\operatorname{diag}( 2 ,1+ s,1+ s ) $ and for any~$c<0$ condition~\eqref{eq:cond_cont} becomes \begin{align}\label{eq:cond_cont1} (1+ s)c < s(b+1)+2b-1. \end{align} This provides a simple recipe for determining the gain~$c$ so that the closed-loop system is~$(2+s)$-contracting. For example, when~$s \to 0$, Eq.~\eqref{eq:cond_cont1} yields $ c<2b-1 $, and this guarantees that the closed-loop system is~$2$-contracting. Recall that in a~$2$-contracting system every nonempty omega limit set is a single equilibrium, thus ruling out chaotic attractors and even non-trivial limit cycles~\cite{li1995}. Fig.~\ref{fig:chaos_closed} depicts the behaviour of the closed-loop system with~$b=0.1$ and~$c=2b-1.1$. The closed-loop system is thus $2$-contracting, and as expected every solution converges to an equilibrium. \end{Example} \begin{figure}[t] \begin{center} \includegraphics[width=8cm,height=6cm]{THOMAS.eps} \caption{A trajectory emanating from~$x(0)=\begin{bmatrix} 1& -2&1 \end{bmatrix}^T$. }\label{fig:chaos} \end{center} \end{figure} \begin{figure}[t] \begin{center} \includegraphics[width=8cm,height=6cm]{CLOSED_LOOP_THOMAS.eps} \caption{Several trajectories of the closed-loop system. The circles denote the initial conditions of the trajectories. }\label{fig:chaos_closed} \end{center} \end{figure} \section{$k$-positive systems} Ref.~\cite{Eyal_k_posi} introduced the notions of~$k$-positive and~$k$-cooperative systems. The LTV~\eqref{Eq:ltv} is called~$k$-positive if~$A^{[k]}(t)$ is Metzler for all~$t$. For~$k=1$ this reduces to requiring that~$A(t)$ is Metzler for all~$t$. In this case the system is positive i.e. the flow maps~$\R^n_+$ to~$\R^n_+$ (and also~$\R^n_-:=-\R^n_+$ to~$\R^n_-$)~\cite{farina2000}. In other words, the flow maps the set of vectors with zero sign variations to itself. $k$-positive systems map the set of vectors with up to~$k-1$ sign variations to itself. To explain this, we recall some definitions and results from the theory of totally positive~(TP) matrices, that is, matrices whose minors are all positive~\cite{total_book,pinkus}. For a vector~$x\in\R^n\setminus\{0\}$, let~$s^-(x) $ denote the number of sign variations in~$x$ after deleting all its zero entries. For example,~$s^-(\begin{bmatrix}-1&0&0&2&-3\end{bmatrix}^T)=2$. We define~$s^-(0):=0$. For a vector~$x\in\R^n $, let~$s^+(x)$ denote the maximal possible number of sign variations in~$x$ after setting every zero entry in~$x$ to either~$-1$ or~$+1$. For example,~$s^+(\begin{bmatrix}-1&0&0&2&-3\end{bmatrix}^T)=4$. These definitions imply that $ 0\leq s^-(x)\leq s^+(x)\leq n-1$, for all $ x\in\R^n. $ For any $k\in \{1,\dots,n\}$, define the sets \begin{align}\label{eq:Pksets} P_{-}^{k} &:= \{ z\in\mathbb{R}^{n}: \; s^{-} (z) \le k-1 \},\nonumber \\ P_{+}^{k} &:= \{ z\in \mathbb{R}^{n}: \; s^{+} (z) \le k-1 \}. \end{align} In other words, these are the sets of all vectors with up to~$k-1$ sign variations. Then~$P_{-}^{k}$ is closed, and it can be shown that~$P_{+}^{k}=\operatorname{int}(P_{-}^{k})$. For example, \begin{align*} P_{-}^{1} = \mathbb{R}_{+}^{n} \cup \mathbb{R}_{-}^{n}, \;\;\; P_{+}^{1} = \operatorname{int}(\mathbb{R}_{+}^{n}) \cup \operatorname{int}( \mathbb{R}_{-}^{n}). \end{align*} \begin{Definition} The LTV~\eqref{Eq:ltv} is called \emph{$k$-positive} on an interval~$[a,b]$ if for any~$a<t_0 < b$, \[ x(t_0)\in P^k_- \implies x(t,x(t_0)) \in P^k_- \text{ for all } t_0 \leq t <b , \] and is called \emph{strongly $k$-positive} if \[ x(t_0)\in P^k_- \implies x(t,x(t_0)) \in P^k_+ \text{ for all } t_0 < t < b. \] \end{Definition} In other words, the sets of up to~$k-1$ sign variations are invariant sets of the dynamics. An important property of TP matrices is their sign variation diminishing property: if~$A\in\R^{n\times n}$ is~TP and~$x\in\R^n\setminus\{0\}$ then~$s^+(Ax)\leq s^-(x)$. In other words, multiplying a vector by a TP matrix can only decrease the number of sign variations. For our purposes, we need a more specialized result. Recall that~$A\in\R^{n\times n}$ is called \emph{sign-regular of order~$k$} if its minors of order~$k$ are all non-positive or all non-negative, and \emph{strictly sign-regular of order~$k$} if its minors of order~$k$ are all positive or all negative \begin{Proposition}\label{thm:BenAvraham_SSRk} \cite{CTPDS} Let $A\in\mathbb{R}^{n\times n}$ be a non-singular matrix. Pick $k\in \{1,\dots,n\}$. Then the following two conditions are equivalent: \begin{enumerate} \item For any $x\in\mathbb{R}^{n} $ with $s^{-}(x) \le k-1$, we have~$s^{-} (Ax) \le k-1$. \item $A$ is sign-regular of order~$k$. \end{enumerate} Also, the following two conditions are equivalent: \begin{enumerate}[I.] \item For any $x\in\mathbb{R}^{n} \setminus \{ 0 \}$ with $s^{-}(x) \le k-1$, we have~$s^{+} (Ax) \le k-1$. \item $A$ is strictly sign-regular of order~$k$. \end{enumerate} \end{Proposition} Using these tools allows to characterize the behaviour of~$k$-positive LTVs. \begin{Theorem}\label{thm:k_pois_ltv} The LTV~\eqref{Eq:ltv} is~$k$-positive on~$[a,b]$ iff~$A^{[k]}(t)$ is Metzler for all~$t\in (a,b)$. It is strongly $k$-positive on~$[a,b]$ iff~$A^{[k]}(t)$ is Metzler for all~$t\in(a,b)$, and~$A^{[k]}(t)$ is irreducible for all~$t \in(a,b)$ except, perhaps, at isolated time points. \end{Theorem} The proof is simple. Consider for example the second assertion in the theorem. The Metzler and irreducibility assumptions imply that the matrix differential system~\eqref{eq:expat} is a positive linear system, and furthermore, that all the entries of~$\Phi^{(k)}(t,t_0)$ are positive for all~$t>t_0$. Thus,~$\Phi(t,t_0)$ is strictly sign-regular of order~$k$ for all~$t>t_0$. Since~$x(t,x(t_0))=\Phi(t,t_0)x(t_0)$, applying Prop.~\ref{thm:BenAvraham_SSRk} completes the proof. This line of reasoning demonstrates a general and useful principle, namely, given conditions on~$A^{[k]}$ we can apply standard tools from dynamical systems theory to the ``$k$-compound dynamics''~\eqref{eq:expat}, and deduce results on the behaviour of the solution~$x(t)$ of~\eqref{Eq:ltv}. A natural question is: when is~$A^{[k]}$ a Metzler matrix? This can be answered using Prop.~\ref{prop:Explicit_A_k} in terms of sign pattern conditions on the entries~$a_{ij}$ of~$A$. This is useful as in fields like chemistry and systems biology, exact values of various parameters are typically unknown, but their signs may be inferred from various properties of the system~\cite{sontag_near_2007}. \begin{Proposition}\label{prop:sign_pattern_for_k_posi} Let~$A \in\R^{n\times n}$ with~$n\geq 3$. Then \begin{enumerate} \item \label{case:nminus1} $A^{[ n-1]}$ is Metzler iff $a_{ij}\geq 0$ for all~$i,j$ with~$i-j$ odd, and~$a_{ij}\leq 0$ for all~$i,j$ with~$i\not =j$ and~$i-j$ even; \item \label{case:kodd} for any odd~$k $ in the range~$1<k<n-1$, $A^{[k]}$ is Metzler iff $a_{1n},a_{n1}\geq 0$, $a_{ij}\geq 0$ for all~$|i-j|=1$, and~$a_{ij}=0$ for all~$1<|i-j|<n-1$; \item \label{case:keven} for any even~$k $ in the range~$1<k<n-1$, $A^{[k]}$ is Metzler iff $a_{1n},a_{n1}\leq 0$, $a_{ij}\geq 0$ for all~$|i-j|=1$, and~$a_{ij}=0$ for all~$1<|i-j|<n-1$. \end{enumerate} \end{Proposition} In Case~\ref{case:nminus1}) there exists a non-singular matrix~$T$ such~$-TAT^{-1}$ is Metzler. In other words, there exists a coordinate transformation such that in the new coordinates the dynamics is competitive. Thus, $k$-positive systems, with~$k\in\{1,\dots,n-1\}$, may be viewed as a kind of interpolation from cooperative to competitive systems. In Case~\ref{case:kodd}),~$A$ is in particular Metzler. Case~\ref{case:keven}) is illustrated in the next example. \begin{Example} Consider the case~$n=3$ and~$A=\begin{bmatrix} -1& 1& -2\\ 0& 1& 0.1 \\ -3& 0 &1 \end{bmatrix} $. Note that~$A$ is not Metzler, yet $ A^{[2]}=\begin{bmatrix} 0& 0.1& 2\\ 0& 0& 1 \\ 3& 0 &2 \end{bmatrix} $ is Metzler (and irreducible). Thm.~\ref{thm:k_pois_ltv} guarantees that for any~$x(0)$ with~$s^-(x(0))\leq 1$, we have \begin{equation}\label{eq:boundsmin} s^-(x(t,x(0)))\leq 1\text{ for all }t\geq 0. \end{equation} Fig.~\ref{fig:signs} depicts~$s^-(x(t,x(0)))=s^-(\exp(At)x(0))$ for~$x(0)=\begin{bmatrix} 4&-21&-1 \end{bmatrix}^T$. Note that~$s^-(x(0))=1$. It may be seen that~$s^-(x(t,x(0)))$ decreases and then increases, but always satisfies the bound~\eqref{eq:boundsmin}. \end{Example} \begin{figure} \begin{center} \includegraphics[scale=0.5]{sign_changes.eps} \caption{ $s^-(x(t,x(0)))$ as a function of~$t$.} \label{fig:signs} \end{center} \end{figure} \subsection{Totally positive differential systems} A matrix~$A\in\R^{n\times n} $ is called a \emph{Jacobi matrix} if~$A$ is tri-diagonal with positive entries on the super- and sub-diagonals. An immediate implication of Prop.~\ref{prop:sign_pattern_for_k_posi} is that~$A^{[k]}$ is Metzler and irreducible for all~$k\in\{1,\dots,n-1\}$ iff~$A$ is Jacobi. It then follows that for any~$t>0$ the matrices~$(\exp (At) )^{(k)} $, $k=1,\dots,n$, are positive, that is,~$\exp(At)$ is TP for all~$t>0$. Combining this with Thm.~\ref{thm:k_pois_ltv} yields the following. \begin{Proposition}\cite{schwarz1970}\label{thm:TP} The following two conditions are equivalent. \begin{enumerate} \item $A$ is Jacobi. \item for any~$x_0\in\R^n \setminus\{0\}$ the solution of the LTI~$\dot x(t)=Ax(t)$, $x(0)=x_0$, satisfies \[ s^+(x(t, x_0) )\leq s^-(x_0) \text { for all } t>0. \] \end{enumerate} \end{Proposition} In other words,~$s^-(x(t,x_0))$ and also~$s^+(x(t,x_0))$ are non-increasing functions of~$t$, and may thus be considered as piece-wise constant Lyapunov functions for the dynamics. Prop.~\ref{thm:TP} was proved by Schwarz~\cite{schwarz1970}, yet he only considered linear systems. It was recently shown~\cite{margaliot2019revisiting} that important results on the asymptotic behaviour of time-invariant and periodic time-varying nonlinear systems with a Jacobian that is a Jacobi matrix for all~$t,x$~\cite{smillie,periodic_tridi_smith} follow from the fact that the associated variational equation is a totally positive~LTV. \section{$k$-cooperative systems} We now review the applications of~$k$-positivity to the time-invariant nonlinear system: \begin{equation}\label{eq:time_invariant_non_linear} \dot x=f(x), \end{equation} with~$f\in C^1$. Let~$J(x):=\frac{\partial }{\partial x}f(x)$. We assume that the trajectories of~\eqref{eq:time_invariant_non_linear} evolve on a convex and compact state-space~$\Omega\subseteq\R^n$. Recall that~\eqref{eq:time_invariant_non_linear} is called \emph{cooperative} if~$J(x)$ is Metzler for all~$x\in \Omega$. In other words, the variational equation associated with~\eqref{eq:time_invariant_non_linear} is positive. The slightly stronger condition of strong cooperativity has far reaching implications. By Hirsch's quasi-convergence theorem~\cite{hlsmith}, almost every bounded trajectory converges to the set of equilibria. It is natural to define~$k$-cooperativity by requiring that the variational equation associated with~\eqref{eq:time_invariant_non_linear} is~$k$-positive. \begin{Definition}\label{def:k-coop}\cite{Eyal_k_posi} The nonlinear system~\eqref{eq:time_invariant_non_linear} is called \emph{[strongly] $k$-cooperative} if the associated LTV~\eqref{eq:var_eqn} is [strongly]~$k$-positive for any~$a,b\in\Omega$. \end{Definition} Note that for~$k=1$ this reduces to the definition of a cooperative [strongly coopertive] dynamical system. One immediate implication of Definition~\ref{def:k-coop} is the existence of certain invariant sets of the dynamics. \begin{Proposition} \label{prop:inhgt} Suppose that~\eqref{eq:time_invariant_non_linear} is~$k$-cooperative. Pick~$a,b\in\Omega$. Then \[ a-b\in P^k_- \implies x(t,a)-x(t,b) \in P^k_- \text{ for all } t \geq 0. \] If, furthermore,~$0 \in \Omega $ and~$0$ is an equilibrium point of~\eqref{eq:time_invariant_non_linear}, i.e.~$f(0)=0$ then \[ a \in P^k_- \implies x(t,a) \in P^k_- \text{ for all } t \geq 0. \] \end{Proposition} The sign pattern conditions in Prop.~\ref{prop:sign_pattern_for_k_posi} can be used to provide simple to verify sufficient conditions for [strong] $k$-cooperativity of~\eqref{eq:time_invariant_non_linear}. Indeed, if~$J(x)$ satisfies a sign pattern condition for all~$x\in \Omega$ then the integral of~$J$ in the variational equation~\eqref{eq:var_eqn} satisfies the same sign pattern, and thus so does~$A^{ab}$. The next example, adapted from~\cite{Eyal_k_posi}, illustrates this. \begin{Example} Ref.~\cite{Elkhader1992} studied the nonlinear system \begin{align}\label{eq:alexsys} \dot x_1&=f_1(x_1,x_n),\nonumber\\ \dot x_i &= f_i(x_{i-1},x_i,x_{i+1}), \quad i=2,\dots,n-1,\nonumber\\ \dot x_n&=f_n(x_{n-1},x_n), \end{align} with the following assumptions: the state-space~$\Omega\subseteq\R^n$ is convex, $f_i\in C^{n-1}$, $i=1,\dots,n$, and there exist~$\delta_i\in\{-1,1\}$, $i=1,\dots,n$, such that \begin{align*} \delta_1\frac{\partial }{\partial x_n}f_1(x) &>0,\\ \delta_2\frac{\partial }{\partial x_1}f_2(x) , \delta_3\frac{\partial }{\partial x_3}f_2(x)&>0,\\ &\vdots\\ \delta_{n-1} \frac{\partial }{\partial x_{n-2}}f_{n-1}(x), \delta_n \frac{\partial }{\partial x_{n}}f_{n-1}(x) &>0,\\ \delta_n \frac{\partial }{\partial x_{n-1}}f_n(x)&>0, \end{align*} for all~$x\in\Omega$. This is a generalization of the monotone cyclic feedback system analyzed in~\cite{poin_cyclic}. As noted in~\cite{Elkhader1992}, we may assume without loss of generality that~$\delta_2=\delta_3=\dots=\delta_n=1$ and~$\delta_1 \in \{-1,1\}$. Then the Jacobian of~\eqref{eq:alexsys} has the form \[ J(x)=\begin{bmatrix} *& 0 &0 &0 &\dots & 0 & 0 & \operatorname{{\mathrm sgn}}(\delta_1) \\ >0 & * &>0 &0 &\dots & 0& 0 & 0 \\ 0& >0 & * &>0 &\dots &0& 0 & 0 \\ &&&\vdots\\ 0& 0 & 0 & 0 &\dots & 0 &>0& * \\ \end{bmatrix} , \] for all~$x\in \Omega$. Here~$*$ denotes ``don't care''. Note that~$J(x)$ is irreducible for all~$x\in \Omega$. If~$\delta_1=1$ then~$J(x)$ is Metzler, so the system is strongly~$1$-cooperative. If~$\delta_1=-1$ then~$J(x)$ satisfies the sign pattern in Case~\ref{case:keven} in Prop.~\ref{prop:sign_pattern_for_k_posi}, so the system is strongly $2$-cooperative. (If~$n$ is odd then~$J(x)$ also satisfies the sign pattern in Case~\ref{case:nminus1}, so there is a coordinate transformation for which the system is also strongly competitive.) \end{Example} The main result in~\cite{Eyal_k_posi} is that strongly~$2$-cooperative systems satisfy a strong Poincar\'{e}-Bendixson property. \begin{Theorem} \label{thm:2dim} Suppose that~\eqref{eq:time_invariant_non_linear} is strongly~$2$-cooperative. Pick~$a\in \Omega$. If the omega limit set~$\omega(a)$ does not include an equilibrium then it is a closed orbit. \end{Theorem} The proof of this result is based on the seminal results of Sanchez~\cite{sanchez2009cones}. Yet, it is considerably stronger than the main result in~\cite{sanchez2009cones}, as it applies to \emph{any} trajectory emanating from~$\Omega$ and not only to so called \emph{pseudo-ordered} trajectories (see the definition in~\cite{sanchez2009cones}). The Poincar\'{e}-Bendixon property is useful because often it can be combined with a local analysis near the equilibrium points to provide a global picture of the dynamics. For a recent application of Thm.~\ref{thm:2dim} to a model from systems biology, see~\cite{Margaliot868000}. \section{Conclusion} $k$-compound matrices describe the evolution of~$k$-dimensional polygons along an LTV dynamics. This geometric property has important consequences in systems and control theory. This holds for both LTVs and also time-varying nonlinear systems, as their variational equation is an LTV. Due to space limitations, we considered here only a partial list of applications. Another application, for example, is based on generalizing diagonal stability for the LTI~$\dot x= Ax$ to~\emph{$k$-diagonal stability} by requiring that there exists a diagonal and positive-definite matrix~$D$ such that $ D A^{[k]}+(A^{[k]})^T D $ is negative-definite~\cite{cheng_diag_stab}. Another interesting line of research is based on analyzing systems with inputs and outputs. A SISO system is called \emph{externally~$k$-positive} if any input with up to~$k$ sign variations induces an output with up to~$k$ sign variations~\cite{gruss1,gruss2,gruss3,gruss4,gruss5}. For LTIs with a zero initial condition the input-output mapping is described by a convolution with the impulse response and then external~$k$-positivity is related to interesting results in statistics~\cite{ibragimov1956} and the theory of infinite-dimensional linear operators~\cite{karlin_tp}. \begin{comment} \section{Positive Systems} \begin{Definition} A matrix $A\in\mathbb{R}^{n\times n}$ is called \emph{totally nonnegative} [\emph{totally positive}] if every sub-minor of $A$ is \emph{nonnegative} [\emph{positive}]. \end{Definition} \begin{Example} Let $E_{i,j}\in\mathbb{R}^{n\times n}$ denote the matrix with a single entry $(i,j)$ equal one, and all the other entries equal zero. Define \begin{align}\label{eq:L_U} L_{i} (p) &:= I + p E_{i,i-1}, \\ U_{i} (p) &:= I + p E_{i-1,i}. \end{align} where $p\in\mathbb{R}$ and $i\in \{ 2,\hdots,n\}$. These matrices are called elemntary bidiagonal ($EB$) matrices, and generalized elementary bidiagonal ($GEB$) if a diagonal matrix $D$ is added to the matrices in~\eqref{eq:L_U}. Note that $EB$ matrices are $TN$ for $p \ge 0$, and $GEB$ matrices are $TN$ for $p \ge 0$ and every entry of $D$ is componentwise nonnegative~\cite{Eyal_Smiliie}. \end{Example} \begin{Example}\label{ex:dominance_condition} Consider the tridiagonal matrix \begin{align*} \begin{bmatrix} a_1 & b_1 & 0 & \hdots & 0 \\ c_1 & a_2 & \ddots & \hdots & \vdots \\ 0 & \ddots & \ddots & \hdots & \vdots \\ \vdots & \ddots & \ddots & \hdots & b_{n-1} \\ 0 & \hdots & \hdots & c_{n-1} & a_{n} \end{bmatrix} \end{align*} where $b_{i}, c_{i} \ge 0 \; \forall i\in [1,n-1]$. Therefore, $A$ is $TN$~\cite[Chapter~0]{total_book}~\cite{Eyal_Smiliie} if the dominance condition holds: \begin{align*} a_{i} \ge b_{i} + c_{i-1} \; \forall i\in [1,n], \end{align*} where $c_{0} = b_{n} := 0$. \end{Example} \subsection{Totally Positive Differential Systems (TPDS)} \begin{Definition} Consider the LTV system in~\eqref{Eq:ltv}. If every minor of $\Phi (t,t_0)$ is positive ($\forall t>0)$, i.e., $\Phi(t,t_0)$ is a \sl{totally positive} matrix (TP) [\sl{totally nonnegative}] ($\forall t>0)$, then~\eqref{Eq:ltv} is a~\sl{totally positive differential system} (TPDS) [\sl{totally nonnegative differential system (TNDS)}]~\cite{schwarz1970,Eyal_k_posi}. \end{Definition} \subsubsection{Coordinate Transformation of the Compound Matrix} Consider the system in~\eqref{Eq:ltv}. Applying the following coordinate transformation $Z(t) := TY(t)$, where $T$ is a non-singular square matrix $T\in\mathbb{R}^{n\times n}$. Thus, the $k$th compound of~\eqref{eq:phidot} is: \begin{align*} (TAT^{-1})^{[k]} = T^{(k)} A^{[k]} ( T^{-1} )^{(k)}. \end{align*} See~\cite{Eyal_k_posi} for the proof. \begin{Theorem} \cite{Eyal_Smiliie,schwarz1970} Consider the matrix differential system $\dot \Phi = A (t) \Phi, \; \Phi (t_0) = I$, where $A(t)$ is a continuous matrix for $t\in (a,b)$. The system is $TNDS$ iff $A(t)\in\mathbb{M}$ for all $t\in (a,b)$. It is $TPDS$ iff $A(t)\in\mathbb{M}$ for all $t\in (a,b)$, and every entry on the super- and sub-diagonals of $A(t)$ is not zero on a time interval. \end{Theorem} \subsubsection{Metzler Matrices} \begin{Definition} Let $M_{2}^{n}$ denote the set of matrices $A\in\mathbb{R}^{n\times n}$ satisfying: \begin{enumerate} \item $a_{1n},a_{n1} \le 0$, \item $a_{ij} \ge 0\; \forall i,j$ with $|i-j|=1$, \item $a_{ij} = 0 \; \forall i,j$ with $1< |i-j|<n-1$. \end{enumerate} \end{Definition} \begin{Example} Consider the $5$-dimension case. Then, the $M_{2}^{5}$ matrices consist of the following sign pattern: \begin{align*} \begin{bmatrix} * & \ge 0 & 0 & 0 & \le 0 \\ \ge 0 & * & \ge 0 & 0 & 0 \\ 0 & \ge 0 & * & \ge 0 & 0 \\ 0 & 0 & \ge 0 & * & \ge 0 \\ \le 0 & 0 & 0 & \ge 0 & * \end{bmatrix}, \end{align*} where $*$ denotes "don't care". \end{Example} \begin{Lemma} Let $A\in\mathbb{R}^{n\times n}$ with $n > 2$. Then $A^{[2]}$ is Metzler iff $A\in M_{2}^{n}.$ \end{Lemma} \begin{Example} Consider the $4$-dimensional case ($n=4$) as in Example~\eqref{ex:A^[3]}. Then, $A^{[2]}$ is Metzler iff $a_{12},a_{23},a_{34},a_{21},a_{32},a_{43} \ge 0$, $a_{14},a_{41} \le 0$, and $a_{13} = a_{24} = a_{31} = a_{42} = 0$, i.e., iff $A\in M_{2}^{4}$. \end{Example} This is a generalization of Smillie's theorem~\cite{smillie}. Indeed, Smillie assumed that $J(x)\in\mathbb{M}^{+}$ for all $x\in\Omega$. Note that $J(x)\in\mathbb{M}$ means that $J(x)$ is tridiagonal and Metzler, so~\eqref{eq:nonlin} is a tridiagonal cooperative system in the sense of Hirsch (\cite{hlsmith}). \begin{Remark} Thus, this is a generalization of the theory of $TNDS$s developed by Schwarz, Smith and Smillies. This is a generalization of the stability and entrainment results under a weaker condition, namely that the Jacobian is tridiagonal but may have nonnegative (rather than positive) entries on its super- and sub-diagonals, along with a suitable observability-type condition~\cite{9107214}. \end{Remark} \subsection{Oscillatory Matrices} \begin{Definition} A matrix $A\in\mathbb{R}^{n\times n}$ is called oscillatory if $A$ is $TN$ and there exists an integer $k>0$ such that $A^{k}$ is $TP$. A $TN$ matrix $A$ is oscillatory iff it is non-singular and irreducible~\cite[Chapter~2]{total_book}, and in this case $A^{n-1}$ is $TP$. \end{Definition} \begin{Example} Consider the matrix $A = \begin{bmatrix} 1 & \epsilon & 0 \\ \epsilon & 1 & \epsilon \\ 0 & \epsilon & 1 \end{bmatrix}$, with $\epsilon\in (0,1/2)$. This matrix is non-singular (as $\det (A) = 1 - 2\epsilon^{2} \ne 0 )$, $TN$ (by the result in Example~\eqref{ex:dominance_condition} ), and irreducible, so it is an oscillatory matrix. Here $A^{n-1} = A^{2} = \begin{bmatrix} 1+\epsilon^{2} & 2\epsilon & \epsilon^{2} \\ 2\epsilon & 1+2\epsilon^{2} & 2\epsilon \\ \epsilon^{2} & 2\epsilon & 1+\epsilon^{2} \end{bmatrix}$, and it is $TP$. \end{Example} \begin{Definition} \cite{gk_book} A square matrix $A\in \mathbb{R}^{n\times n}$ is called oscillatory if it is $TN$ and there exists an integer $k\ge 1$ such that $A^{k}$ is $TP$. \end{Definition} The following Theorem analyzes the spectral structure of oscillatory matrices. \begin{Theorem}\label{thm:gantmacher_and_Krein_OscillatoryMatrices} \cite{gk_book, pinkus} If $A\in\mathbb{R}^{n\times n}$ is an oscillatory matrix then its eigenvalues are all real, positive, and distinct. Thus, the eigenvalues can be order in the following manner: \begin{align*} \lambda_{1} > \lambda_{2} > \hdots \lambda_{n} > 0, \end{align*} Let $u^{k} \in \mathbb{R}^{n}$ denote the corresponding eigenvector to $\lambda_{k}$. Then, for any $1\le i \le j \le n$, and for any real scalars $c_{i},\hdots,c_{j}$ that are not all zero, \begin{align*} i-1 \le s^{-} \left ( \sum_{k=i}^{j} c_{k} u^{k} \right ) \le s^{+} \left ( \sum_{k=i}^{j} c_{k} u^{k} \right ) \le j-1. \end{align*} \begin{Remark} In particular, Theorem~\eqref{thm:gantmacher_and_Krein_OscillatoryMatrices} implies that $s^{-} (u^{i}) = s^{+} (u^{i}) = i-1 \; \forall i\in [1,n]$. \end{Remark} \end{Theorem} \begin{Example}\cite{Eyal_k_posi} Consider the following oscillatory matrix \begin{align*} \begin{bmatrix} 2 & 1 & 0 \\ 1 & 3 & 1 \\ 0 & 1 & 2 \end{bmatrix} \end{align*} Its eigenvalues are $\lambda_{1} = 4,\; \lambda_{2} = 2$ and $\lambda_{3} = 1$. The corresponding eigenvectors are $u^{1} = \begin{bmatrix} 1 & -1 & 1 \end{bmatrix}^{T},\; u^{2} = \begin{bmatrix} -1 & 0 & 1 \end{bmatrix}^{T}$ and $u^{3} = \begin{bmatrix} 1 & -1 & 1 \end{bmatrix}^{T}$. Notice that \begin{align*} s^{-} (u^{k}) = s^{+} (u^{k}) = k-1 \; \forall k\in [1,3]. \end{align*} \end{Example} \subsection{$k$-positive linear systems} For any $k\in [1,n]$, define the sets $P_{-}^{k} := \{ z\in\mathbb{R}^{n}: \; s^{-} (z) \le k-1 \}$, and $P_{+}^{k} := \{ z\in \mathbb{R}^{n}: \; s^{+} (z) \le k-1 \}$. \\ $P_{-}^{k}$ is closed, and $P_{+}^{k}$ is open. \\ Notice that, \begin{align*} P_{-}^{1} = \mathbb{R}_{+}^{n} \bigcup \mathbb{R}_{-}^{n}, \;\;\; P_{+}^{1} = \operatorname{int}\mathbb{R}_{+}^{n} \bigcup \operatorname{int} \mathbb{R}_{-}^{n}, \end{align*} and that \begin{align*} P_{+}^{k} &= \operatorname{int} ( P_{-}^{k} ) \; \forall k\in [1,n-1], \\ P_{-}^{1} &\subset P_{-}^{2} \subset \hdots \subset P_{-}^{n} = \mathbb{R}^{n}, \\ P_{+}^{1} &\subset P_{+}^{2} \subset \hdots \subset P_{+}^{n} = \mathbb{R}^{n}. \end{align*} \subsubsection{odd- and even-positivity} \begin{Definition} ~\cite{Eyal_k_posi,9107214} Choose $n\ge 4$, and $k\in [2,n-2]$. Consider the sign matrix $\overline{A}_{k}^{n}\in \{ *,-,0,+\}^{n\times n}$ , where \begin{enumerate} \item $\overline{a}_{ii} = * \;\; \forall i$, \item If $k$ is odd [even] then $\overline{a}_{1n},\overline{a}_{n1} = + \; [\overline{a}_{1n},\overline{a}_{n1} = -]$, \item $\overline{a}_{ij} = + \; \forall i,j$ such that $|i-j| = 1$, \item $\overline{a}_{ij} = 0 \; \forall i,j$ such that $1<|i-j|<n-1$. \end{enumerate} \end{Definition} \begin{Example}~\cite{9107214} Consider \begin{align*} \overline{A}_{2}^{4} = \begin{bmatrix} * & + & 0 & - \\ + & * & + & 0 \\ 0 & + & * & + \\ - & 0 & + & * \end{bmatrix} \end{align*} If $k$ is odd, then $\overline{A}_{2}^{4}$ is Metzler, whereas if $k$ is even, then $\overline{A}_{k}^{n}$ is not necessarily Metzler. \end{Example} \begin{Remark}~\cite{9107214} \begin{enumerate} \item $\overline{A}_{2}^{n} = \overline{A}_{4}^{n} = \hdots = \overline{A}_{2k}^{n}$ $\forall k$ satisfying $2k \le n-2$. \item $\overline{A}_{3}^{n} = \overline{A}_{5}^{n} = \hdots = \overline{A}_{2k+1}^{n}$ $\forall k$ satisfying $2k+1 \le n-2$. \item $A(t)$ satisfies the sign pattern of both $\overline{A}_{k}^{n}$ and $\overline{A}_{k+1}^{n}$ iff $A(t)$ is tridiagonal with non-negative entries on the super- and sub-diagonals for almost all $t$. Such a system is called a~\sl{totally positive differential system} (TPDS)~\cite{schwarz1970}. \end{enumerate} \end{Remark} \begin{Theorem}\label{thm:timeinterval_positivity}~\cite{Eyal_k_posi,9107214} Choose $n\ge 4$ and $k\in [2,n-2]$. The LTV system~\eqref{Eq:ltv} is $k$-positive on the time interval $(t_0,t_1)$ iff $A(t)$ admits the sign structure $\overline{A}_{k}^{n}$ for almost all $(t_0,t_1)$ \end{Theorem} \begin{Remark}~\cite{9107214} $k$-positivity, for $k\in [2,n-2]$, can be classified to two cases only: \begin{enumerate} \item \sl{odd-positivity} - the flow of~\eqref{Eq:ltv} maps $P^{k}$ to itself for ll odd $k\in [2,n-2]$. % \item \sl{even-positivity} - the flow of~\eqref{Eq:ltv} maps $P^{k}$ to itself for all even $k\in [2,n-2]$. \end{enumerate} Note that $k$-positivity as described in Theorem~\eqref{thm:timeinterval_positivity} is not invariant to coordinate transformations. \end{Remark} \begin{Example} Consider \begin{align*} \overline{A} = \begin{bmatrix} * & 0 & 0 & + \\ + & * & + & 0 \\ 0 & 0 & * & - \\ + & 0 & - & * \end{bmatrix} \end{align*} Note that $\overline{A} \notin M_{2}^{4}$, and therefore in general not even-positive. However, applying a permutation matrix $P$ may transform this to a system which is even-positive. Such a permutation matrix $P$ is for example $ P:= \begin{bmatrix} 0 & 0 & 1 & 0 \\ 0 & 1 & 0 & 0 \\ 1 & 0 & 0 & 0 \\ 0 & 0 & 0 & 1 \end{bmatrix}$. Thus, $\overline{B} := P\overline{A} P^{T} = \begin{bmatrix} * & 0 & 0 & - \\ + & * & + & 0 \\ 0 & 0 & * & + \\ - & 0 & + & * \end{bmatrix}$, where indeed $\overline{B}\in M_{2}^{4}$, and therefore, the system $x^{p} (t):= P x(t)$ is even-positive. \end{Example} \subsubsection{Influence Graphs} \subsubsection{$k$-positive up to permutations} \end{comment} \begin{comment} MATLAB PROGRAMS - DO NOT DELETE hold off; A=[-1 1 -2; 0 1 0.1 ; -3 0 1 ]; alp=-7;beta=5; x0=[-1 0 1]'+alp*[0 3 1 ]'+beta*[1 0 1]'; val=[]; oldnum=1; for t=0:0.001:1 xt=expm(A*t)*x0; num=0;if xt(1)*xt(2)<0 num=1; end if xt(2)*xt(3)<0 num=num+1; end val=[val ; num]; if oldnum ~= num [t oldnum num ] end oldnum=num; plot(t,num,'.k'); hold on; end grid on; xlabel('$t$','fontsize', 16 ,'interpreter','latex'); hold off; for t=0:.5:2 mat=[exp(-t) 0 ; -1+exp(-t)*(cos(t)-sin(t)) 1]; p1=mat*[0;0]; p2=mat*[1;0]; p3=mat*[1;1]; p4=mat*[0;1]; x=[ p1(1) p2(1) p3(1) p4(1) p1(1)]+2*t; y=[ p1(2) p2(2) p3(2) p4(2) p1(2)]+2*t; h=fill(x,y,'k-'); hold on; end text(1/8,1.2,'$t=0$','interpreter','latex','fontsize',14); text(0.7,2.1,'$t=0.5$','interpreter','latex','fontsize',14); text(1.7,3.1,'$t=1$','interpreter','latex','fontsize',14); text(2.7,4.1,'$t=1.5$','interpreter','latex','fontsize',14); text(3.7,5.1,'$t=2$','interpreter','latex','fontsize',14); grid on; xlabel('$x_1$','fontsize', 16 ,'interpreter','latex'); ylabel('$x_2$','fontsize', 16 ,'interpreter','latex'); axis([0 5.5 0 5.5 ] );axis('square'); hold off; timelen=2000; options = odeset('RelTol',1e-10,'AbsTol',1e-10); [t,x]=ode45(@RAZ_eq,[0 timelen], [ 1; -2;1],options); plot3(x(:,1),x(:,2),x(:,3),'k');hold on; h=xlabel('$x_1$','interpreter','latex','FontSize',16); ylabel('$x_2$','interpreter','latex','FontSize',16); zlabel('$x_3$','interpreter','latex','FontSize',16); grid on; box on; function ret=RAZ_eq(t,x) b = 0.1 ; ret=[ sin(x(2))-b*x(1) ; sin(x(3))-b*x(2) ; sin(x(1))-b*x(3) ]; function ret=RAZ_CL_eq(t,x) b = 0.1 ; c=2*b-1.1 ; ret=[ sin(x(2))-b*x(1)+c*x(1); sin(x(3))-b*x(2)+c*x(2); sin(x(1))-b*x(3) ]; hold off; timelen=25; options = odeset('RelTol',1e-10,'AbsTol',1e-10); [t,x]=ode45(@RAZ_CL_eq,[0 timelen],1/2*[-1; 1;1],options); plot3(x(1,1),x(1,2),x(1,3),'ok');hold on; plot3(x(:,1),x(:,2),x(:,3),'k');hold on; [t,x]=ode45(@RAZ_CL_eq,[0 timelen],[-1; 1;1],options); plot3(x(1,1),x(1,2),x(1,3),'ok');hold on; plot3(x(:,1),x(:,2),x(:,3),'k');hold on; [t,x]=ode45(@RAZ_CL_eq,[0 timelen],[ 1; 1;1],options); plot3(x(1,1),x(1,2),x(1,3),'ok');hold on; plot3(x(:,1),x(:,2),x(:,3),'k');hold on; [t,x]=ode45(@RAZ_CL_eq,[0 timelen],[ 1; -1;1],options); plot3(x(1,1),x(1,2),x(1,3),'ok');hold on; plot3(x(:,1),x(:,2),x(:,3),'k');hold on; [t,x]=ode45(@RAZ_CL_eq,[0 timelen],[ 1; 1;-1],options); plot3(x(1,1),x(1,2),x(1,3),'ok');hold on; plot3(x(:,1),x(:,2),x(:,3),'k');hold on; [t,x]=ode45(@RAZ_CL_eq,[0 timelen],[-1; 1;-1],options); plot3(x(1,1),x(1,2),x(1,3),'ok');hold on; plot3(x(:,1),x(:,2),x(:,3),'k');hold on; [t,x]=ode45(@RAZ_CL_eq,[0 timelen],[1/2;1/4;0 ],options); plot3(x(1,1),x(1,2),x(1,3),'ok');hold on; plot3(x(:,1),x(:,2),x(:,3),'k');hold on; [t,x]=ode45(@RAZ_CL_eq,[0 timelen],[1/20;1/40;0 ],options); plot3(x(1,1),x(1,2),x(1,3),'ok');hold on; plot3(x(:,1),x(:,2),x(:,3),'k');hold on; [t,x]=ode45(@RAZ_CL_eq,[0 timelen],[1/2;-1/2;-2 ],options); plot3(x(1,1),x(1,2),x(1,3),'ok');hold on; plot3(x(:,1),x(:,2),x(:,3),'k');hold on; h=xlabel('$x_1$','interpreter','latex','FontSize',16); ylabel('$x_2$','interpreter','latex','FontSize',16); zlabel('$x_3$','interpreter','latex','FontSize',16); grid on; box on; set(gca,'View',[-28,35]); [az,el] = view \end{comment}
{ "timestamp": "2021-03-30T02:25:43", "yymm": "2103", "arxiv_id": "2103.15097", "language": "en", "url": "https://arxiv.org/abs/2103.15097", "abstract": "The multiplicative and additive compounds of a matrix play an important role in several fields of mathematics including geometry, multi-linear algebra, combinatorics, and the analysis of nonlinear time-varying dynamical systems. There is a growing interest in applications of these compounds, and their generalizations, in systems and control theory. This tutorial paper provides a gentle introduction to these topics with an emphasis on the geometric interpretation of the compounds, and surveys some of their recent applications.", "subjects": "Optimization and Control (math.OC)", "title": "Compound matrices in systems and control theory", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9907319852508435, "lm_q2_score": 0.808067208930584, "lm_q1q2_score": 0.8005780301199057 }
https://arxiv.org/abs/2006.14482
A metric on directed graphs and Markov chains based on hitting probabilities
The shortest-path, commute time, and diffusion distances on undirected graphs have been widely employed in applications such as dimensionality reduction, link prediction, and trip planning. Increasingly, there is interest in using asymmetric structure of data derived from Markov chains and directed graphs, but few metrics are specifically adapted to this task. We introduce a metric on the state space of any ergodic, finite-state, time-homogeneous Markov chain and, in particular, on any Markov chain derived from a directed graph. Our construction is based on hitting probabilities, with nearness in the metric space related to the transfer of random walkers from one node to another at stationarity. Notably, our metric is insensitive to shortest and average walk distances, thus giving new information compared to existing metrics. We use possible degeneracies in the metric to develop an interesting structural theory of directed graphs and explore a related quotienting procedure. Our metric can be computed in $O(n^3)$ time, where $n$ is the number of states, and in examples we scale up to $n=10,000$ nodes and $\approx 38M$ edges on a desktop computer. In several examples, we explore the nature of the metric, compare it to alternative methods, and demonstrate its utility for weak recovery of community structure in dense graphs, visualization, structure recovering, dynamics exploration, and multiscale cluster detection.
\section{Introduction} \subsection{Motivation} Many finite spaces can be endowed with meaningful metrics. For undirected graphs, the geodesic (shortest path), commute time (effective resistance), and diffusion distance~\cite{lafon04diffusion,coifman05geometric,coifman06diffusion} metrics are widely applied~\cite{coifman05geometric,Liben_Nowell_2003,abraham2010highway}. The first two can be naively generalized to directed graphs by summing shortest/average walk length in each direction, whereas the third is specifically undirected. We know of only one graph metric specifically designed for directed graphs, namely the generalized effective resistance distance developed in~\cite{Young_2016b,Young_2016a}. Overlaying a metric onto a directed structure is a challenge since, by definition, the metric is symmetric. A related problem is finding metrics on the state space of a finite-state, discrete time Markov chain. In this case, there is also limited prior work, consisting of mean commute time~\cite{rozinas,chebotarev-2020,choi19resistance} and a constant-curvature metric~\cite{vollering2018}. Metrics fit into the broader category of dissimilarity measures, with the decision whether to impose all metric axioms being application dependent. When a metric is used, this additional structure can enable various algorithmic accelerations, improved guarantees, and useful inductive biases~\cite{elkan2003kmeans,moore00anchors,hamerly10kmeans,boytsov13prune,pitis2020bias}. Furthermore, the metric structure is a key ingredient in proofs of convergence, consistency, and stability. While mostly settled for undirected graphs~\cite{Osting_2017,singer2012vector,singer2017spectral,trillos2018variational,trillos2016consistency}, the development of such theories for directed graphs (digraphs) and Markov chains is an open research problem. The first positive result for digraphs appeared recently~\cite{Yuan2020}. In the present work, we introduce and analyze a new metric for digraphs and Markov chains based on the \emph{hitting probability} from one node to another, by which we mean the probability that a random walker starting at one node will reach the other node before returning to its starting node. By correctly combining these probabilities with the invariant distribution of an irreducible Markov chain, a metric can be constructed. This metric differs from other metrics by being insensitive to walk length, thus measuring information that is, in a sense, orthogonal to commute time, as illustrated in examples. In the special case of undirected graphs and with the scale parameter $\beta=1$ (defined below), the hitting probabilities metric is actually the logarithm of effective resistance/commute time (plus a constant), a striking fact proven in~\cite{doyle2000random}, section 1.3.4. For other values of the scale parameter, the hitting probabilities metric is a new addition to the limited catalogue of undirected graph metrics. We illustrate the utility of our metric in several examples, both analytical and numerical, related to graph symmetrization, clustering, structure detection, data exploration, and geometry detection. \subsection{Our contributions} Let ${(X_t)}_{t\geq 0}$ be a discrete-time Markov chain on the state space $[n] = \{ 1, \ldots , n\}$ with initial distribution $\lambda$ and irreducible transition matrix $P$, {\it i.e.\/}, \[ \mathbf{P}(X_0 = i) = \lambda_i \qquad \textrm{and} \qquad \mathbf{P}(X_{t+1} = j \mid X_t = i) = P_{i,j}\,. \] We emphasize that $X$ is not required to be aperiodic. Let $\phi \in \mathbb{R}_{+}^n$ be the invariant distribution for $P$, {\it i.e.\/}, $P^T \phi = \phi$. The \emph{hitting time} (starting from a random state distributed like $\lambda$) for a state $i\in [n]$ is the random variable given by \[ \tau_i := \inf\{ t \geq 1 \colon X_t = i \}\,. \] For $i,j \in [n]$, let us define \begin{equation} Q_{i,j} := \mathbf{P}_i [\tau_j < \tau_i]\,, \end{equation} which denotes the probability that starting from site $i$ ({\it i.e.\/}, the subscript on $\mathbf{P}_i$ is used to indicate that $\lambda = \delta_{i}$) the hitting time of $j$ is less than the time it takes to return to $i$. We emphasize that we consider $\tau_j < \tau_i$ here for a single walk and take the probability of such an event over all walks starting at $i$ when computing $Q_{i,j}$. An expression for the \emph{hitting probability matrix}, $Q$, in terms of the transition matrix will be given in~\cref{eq:Q}; see \cref{s:CompMeth} on computational methods. \begin{lemma}\label{t:KeyIdentity} The following relationship holds\footnote{\Cref{t:KeyIdentity} was previously (and independently) proven in~\cite{chien04link}, in the context of Markov chain perturbation theory applied to the internet. It was possibly known even earlier.} for $i\ne j$: \begin{equation}% \label{Qpi} Q_{i,j} \phi_i = Q_{j,i} \phi_j\,. \end{equation} \end{lemma} The weighting by the invariant measure is motivated by connections between the invariant measure and random walks as found in \cite[Section 1.7]{Norris_1997}. A proof of \cref{t:KeyIdentity} is given in \cref{s:Proofs}. \begin{rem} \Cref{t:KeyIdentity}~implies that, with appropriate choice of $Q_{ii}$, $\frac1n Q$ is a reversible Markov chain with invariant distribution $\phi$. \end{rem} We define the \emph{normalized hitting probabilities matrix}, $\Ahtb \in \mathbb R^{n \times n}$, by \begin{equation} \label{Aht} \Ahtb_{i,j} := \begin{cases} \dfrac{ \phi_i^{\beta} }{ \phi_j^{1-\beta} } Q_{i,j} & i \ne j \\ 1 & i=j \end{cases} \end{equation} where $\beta \in [\sfrac12, \infty)$. In contexts where the choice of $\beta$ is not important, we simply write $A^{(\mathrm{hp})} = \Ahtb$. Two useful choices for $\beta$ are $1$ and $1/2$. The $Q_{i,j}$ matrix has recently been shown to play a key role in determining the error of a family of stratified Markov chain Monte Carlo methods~\cite{dinner2017stratification,Thiede_2015}. From \cref{t:KeyIdentity}, we immediately have the following Corollary. \begin{corollary}\label{l:symmetric} The matrix $\Ahtb$ defined in~\cref{Aht} is symmetric. In particular, \begin{equation} \Ahtb[\sfrac12]_{i,j} = \sqrt{Q_{i,j}Q_{j,i}}. \label{d2sym} \end{equation} \end{corollary} \begin{proof} We observe \begin{align*} \Ahtb_{i,j} & = \frac{ \phi_i^\beta }{ \phi_j^{1-\beta} } Q_{i,j} = \frac{ \phi_i^{\beta-1} }{ \phi_j^{1-\beta} } \phi_i Q_{i,j} \\ & = \frac{ \phi_i^{\beta-1} }{ \phi_j^{1-\beta} } \phi_j Q_{j,i} = \frac{ \phi_j^{\beta} }{ \phi_i^{1-\beta} } Q_{j,i} = A^{(\mathrm{hp},\beta)}_{j,i}. \end{align*} Hence, $\Ahtb$ is symmetric. To prove \eqref{d2sym}, we observe that ${\left(\Ahtb[\sfrac12]_{i,j}\right)}^2 = \frac{\phi_i}{\phi_j} Q_{i,j}^2 = Q_{i,j} Q_{j,i}$ by~\cref{Qpi}. \end{proof} In some applications, information about relatedness of vertices in a graph will be most immediately encoded in the form of a non-stochastic adjacency matrix $A$. In this case the input adjacency matrix can be transformed into a stochastic matrix $P$ either by a similarity transformation involving the dominant right eigenvector of $A$ or by normalization of the rows of $A$ so that they sum to 1. The resulting stochastic matrix $P$ can then be used as in~\cref{Aht} to construct $A^{(\textrm{hp},\beta)}$, itself a symmetric adjacency matrix on the vertices of the network. In this article we do not address the relative merits of methods to transform an adjacency matrix into a stochastic matrix. We use row normalization unless otherwise stated. Given an irreducible stochastic matrix $P$, we can thus define a distance $d^\beta \colon [n] \times [n] \to \mathbb R$, which we refer to as the \emph{hitting probability metric}, by \begin{equation} \label{e:Dist} d^\beta(i,j) = - \log \left( \Ahtb_{i,j} \right). \end{equation} \begin{theorem}\label{t:Metric} The hitting probability metric, $d^\beta \colon [n] \times [n] \to \mathbb R$, defined in~\cref{e:Dist} is a metric for $\beta \in (\sfrac12, 1]$. For $\beta = \sfrac12$, $d^\beta$ is a pseudo-metric\footnotemark and there exists a quotient graph on which the distance function becomes a metric. \end{theorem} \footnotetext{Recall that a pseudo-metric on $[n]$ is a non-negative real function $f\colon [n] \times [n] \to \mathbb R_{\geq 0}$ satisfying $d(i,i) = 0$, symmetry $d(i,j) = d(j,i)$, and the triangle inequality $d(i,j) \leq d(i,k) + d(k,j)$. A pseudo-metric is a metric if we can identify indiscernible values, {\it i.e.\/}, $d(i,j) = 0 \iff i=j$.} In~\cref{t:12bounds}, we show that there exists a quotient graph on which $d^{\sfrac12}$ is a metric and which preserves many of the metric properties of the original graph.\footnote{While the usual pseudo-metric quotienting procedure could apply here, there is no guarantee that there would be a corresponding subgraph, which is why~\cref{t:12bounds} is needed.} The key observation for the $d^{\sfrac12}$ pseudometric is that in order for two vertices to be distance $0$ from each other, the probability of hitting the other vertex before returning must be $1$ for both. Hence we provide (in~\cref{structure,quotients}) a means of effectively collapsing these vertices to a single vertex, carefully preserving the overall probabilities relative to the remaining vertices. \begin{rem} In light of~\cref{t:KeyIdentity,t:Metric}, $A^{(\mathrm{hp})}$ has two interpretations, first as a symmetrization of $A$, and second as a weighted similarity graph corresponding to $d$, since $A(i,j) = e^{-d(i,j)}$. The practice of associating a finite (subset of a) metric space with a similarity graph in this way is widespread, especially in the manifold learning and graph-based machine-learning communities.\footnote{\cite{self-tuning} cites~\cite{shi-malik,relocalization} as this similarity function's first use specifically for graph-based clustering.} Thus, in our experiments, we favored the use of $A^{(\mathrm{hp})}$ for certain applications where it seemed more natural. \end{rem} Finally, we show how advances from~\cite{Thiede_2015} enable us to compute the distance matrix in $O(n^3)$ operations, allowing us to scale up to $\approx 38M$ edges in examples on a Lenovo ThinkStation P410 desktop with Xeon E5--1620V4 3.5 GHz CPU and 16 GB RAM using {\sc MATLAB} R2019a Update 4 (9.6.0.1150989) 64-bit (glnxa64). We also provide various synthetic examples to help develop an intuition for the metric and its differences from other measures. We conclude with an example using New York City taxi data to illustrate how our metric can aid in data exploration. \subsection{Relationship to other notions of similarity and metrics}\label{s:RelWork} In this section, we discuss some related notions of similarity and metrics on finite state spaces with asymmetric (directional) relationships. Our focus is on symmetric notions of dissimilarity, with an emphasis on metrics. While, in some applications, asymmetric similarity scores may be the right choice (see, {\it e.g.\/}, Tversky's seminal work on features of similarity~\cite{Tversky_1977}), we restrict our scope to symmetric notions. We do, however, wish to mention directed metrics (also called quasi-metrics), which are a natural analogue to metric spaces for relaxations of digraph cut problems~\cite{directed_metric}. From~\cite{chebotarev-2020,kemeny,rozinas}, we know that commute time is a metric on ergodic Markov chains. In~\cite{choi19resistance,Young_2016b,Young_2016a}, generalizations of effective resistance are developed for ergodic Markov chains and directed graphs. Commute time and resistance-based metrics are popular and more robust than shortest-path distances, although they are not informative in certain large-graph limits~\cite{vonluxborg2014}. In~\cref{sec:gluedcycles}, we compare the effective resistance of~\cite{Young_2016b,Young_2016a} to the hitting probability metric on a particular example. In~\cite{vollering2018}, a metric is developed on Markov chains. This metric gives the chain constant curvature, in an appropriate generalized sense. Distance in this metric is then related to the distinguishability after one step of random walks beginning at the two distinct nodes. The metric is constructed jointly with the curvature using a fixed point argument. It is expected to be useful in proving, for example, concentration inequalities for Markov chains. Notions of diffusion distance to a set $B$ on undirected graphs have been explored recently for the connection Laplacian~\cite{singer2012vector} and for the graph Laplacian~\cite{cheng2019diffusion}. The notion of distance is determined by taking $\ell$ steps using the random walk generated by the symmetric graph adjacency matrix $A$ with degree matrix $D$, i.e.\ it counts the number of walks of length $2\ell$ from $i$ to $j$. Diffusion distances from a vertex $i$ to a sub-graph $B$ in~\cite{cheng2019diffusion} is defined as the smallest number of steps for all random walks started at $i$ to reach $B$. The work~\cite{singer2012vector} established that diffusion distances converge to geodesic distances in the high density limit of random graphs on manifolds, and \cite{cheng2019diffusion} explored how eigenvectors relate to this notion of distance. Directed graphs have been represented as magnetic connection Laplacians on undirected graphs through a notion of polarization, see~\cite{fanuel2018magnetic}, after which a version of diffusion distance can be applied. A variety of methods exist in machine learning to compute ``graph representations,'' which are learned embeddings of nodes, subgraphs, or entire (possibly directed) graphs into Euclidean space so that they can be fed into standard machine learning tools~\cite{hamilton17representation}. These can be seen as imposing a metric on directed graphs, with the main drawbacks relative to the hitting probability metric being model complexity, difficulty of interpretation, and difficulty of analysis. In~\cite{malliaros13clustering}, existing symmetrization techniques for directed graphs are surveyed. In particular, we mention \cite{satuluri2011symmetrizations,zhou2005semi,lai2010extracting,chen2008clustering}. In each of these articles, clustering, community detection and/or semi-supervised learning techniques are considered on directed graphs using various symmetrizations, such as that of Fan Chung ({\it e.g.\/}~\cite{satuluri2011symmetrizations}) or using commute times similar to those in the effective resistance metric ({\it e.g.\/}~\cite{chen2008clustering}). Our results use $A^{(\mathrm{hp})}$ as a symmetrization, and we will see that this enables us to perform the tasks just mentioned, although with different and sometimes more helpful results. In~\cite{fitch}, the metric of~\cite{Young_2016a,Young_2016b} is used as the basis for a digraph symmetrization technique. It is guaranteed to preserve effective resistances, possibly relying on negative entries. Rigorous applications to directed cut and graph sparsification are given. \subsection*{Outline} We prove \cref{t:KeyIdentity,l:symmetric,t:Metric} in \cref{s:Proofs}. In \cref{s:CompMeth}, we describe computational methods to compute the normalized hitting probabilities matrix, $\Ahtb$. In \cref{s:Examples}, we give some examples of the computed metric. We conclude in~\cref{s:Disc}. \section{Proofs and discussion of structural properties}\label{s:Proofs} \subsection{Structure of the normalized hitting probabilities matrix} \begin{proof}[Proof of \cref{t:KeyIdentity}] The probability that $X_t$ starts from $i$ and hits $j$ at least $k+1$ times before returning to $i$ can be expressed as \[ \mathbf{P}_i [\tau_j < \tau_i] \mathbf{P}_j {[\tau_j < \tau_i]}^k\,. \] We let $V_i^j$ be the number of times $X_t$ hits $j$ before returning to $i$, $ V_i^j = \sum_{t=1}^{\tau_i} {1}_{X_t = j}\,. $ Then, we have \[ \mathbf{P}_i [\tau_j < \tau_i] \mathbf{P}_j {[\tau_j < \tau_i]}^k = \mathbf{P}_i [V_i^j \geq k+1]\,. \label{r1} \] Now observe that \begin{align} &\label{geo} \sum_{k=0}^\infty \mathbf{P}_i [\tau_j < \tau_i] \mathbf{P}_j {[\tau_j < \tau_i]}^k =\sum_{k=0}^\infty \mathbf{P}_i [ V_i^j \ge k + 1] =\sum_{k=0}^\infty \mathbf{E}_i \left[1_{V_i^j \ge k+1}\right] \\ &\hspace{1cm} = \mathbf{E}_i \left[\sum_{k=0}^\infty {1}_{V_i^j \geq k+1}\right] = \mathbf{E}_i \left[\sum_{k=0}^\infty k 1_{V_i^j=k} \right] = \sum_{k=0}^\infty k \mathbf{P}[V_i^j=k] \notag \\ &\hspace{1cm}= \mathbf{E}_i [ V_i^j ] \label{r3}\,. \end{align} The expectation in \cref{r3} is known to satisfy \begin{equation} \label{gammaid} \mathbf{E}_i [ V_i^j ] = \frac{\phi_j}{\phi_i}\,, \end{equation} which is proved in, for example,~\cite[Theorem 1.7.6]{Norris_1997}. However, we recognize the expression \cref{geo} as a geometric series and hence have \begin{align*} \sum_{k=0}^\infty \mathbf{P}_i [\tau_j < \tau_i] \mathbf{P}_j {[\tau_j < \tau_i]}^k &= \mathbf{P}_i [\tau_j<\tau_i]{\left(1-\mathbf{P}_j[\tau_j<\tau_i]\right)}^{-1} \\ &= \mathbf{P}_i [\tau_j < \tau_i] \mathbf{P}_j {[\tau_i < \tau_j]}^{-1} = Q_{i,j} Q_{j,i}^{-1}\,. \end{align*} Combining this with~\cref{gammaid} we arrive at $Q_{i,j} Q_{j,i}^{-1}=\frac{\phi_j}{\phi_i}$. \end{proof} To prove~\cref{t:Metric}, we will need one more lemma \begin{lemma}% \label{Qlemma} The following inequality hold \begin{equation} \label{Qineq} Q_{i,j} \geq Q_{i,k} Q_{k,j}\,. \end{equation} \end{lemma} \begin{proof} Consider the corresponding auxiliary Markov process restricted to nodes $i$, $j$, and $k$ with $3\times 3$ transition matrix $F$, the elements of which we denote by, {\it e.g.\/}, $F_{i,j} = \mathbf{P}_i[\tau_j < \min \{ \tau_i, \tau_k\}]$ and $F_{i,i} = \mathbf{P}_i[\tau_i < \min \{\tau_j, \tau_k\}]$. That is, $F_{i,j}$ gives the probability of a random walker starting at $i$ eventually reaching $j$ before either reaching $k$ or returning to $i$, while $F_{i,i}$ gives the probability of a random walker starting at $i$ returning to $i$ before reaching either $j$ or $k$. Since $F_{i,i}+F_{i,j}+F_{i,k}=1$, we have \begin{equation*} Q_{i,j} = F_{i,j} + \frac{ F_{i,k}F_{k,j}}{1-F_{k,k}}\,, \quad Q_{i,k} = F_{i,k} + \frac{ F_{i,j}F_{j,k}}{1-F_{j,j}}\,, \quad Q_{k,j} = F_{k,j} + \frac{ F_{k,i}F_{i,j}}{1-F_{i,i}}\,. \end{equation*} Hence, we observe \begin{align*} Q_{i,k} Q_{k,j} & = \left( F_{i,k} + \frac{F_{i,j} F_{j,k}}{1-F_{jj} } \right) \left( F_{k,j} + \frac{F_{k,i} F_{i,j}}{1-F_{ii} } \right) \\ & = F_{i,k}F_{k,j} + F_{i,j} \left( \frac{F_{j,k} F_{k,j}}{1-F_{jj} } + \frac{F_{i,k} F_{k,i}}{1-F_{ii} } + \frac{F_{j,k} F_{k,i} F_{i,j}}{(1-F_{ii}) (1-F_{jj} ) } \right)\,. \end{align*} Using \[ \frac{F_{j,k}}{1-F_{j,j}} = \frac{F_{j,k}}{F_{j,i} + F_{j,k}} < 1\,, \] we then observe \begin{align*} Q_{i,k} Q_{k,j} & \leq F_{i,k}F_{k,j} + F_{i,j} \left( F_{k,j} + \frac{F_{i,k} F_{k,i}}{1-F_{ii} } + \frac{ F_{k,i} F_{i,j}}{1-F_{ii}} \right) \\ &= F_{i,k}F_{k,j} + F_{i,j} \left( F_{k,j} + F_{k,i} \frac{F_{i,k} + F_{i,j} }{1-F_{ii} } \right) \\ & = F_{i,k}F_{k,j} + F_{i,j} \left( F_{k,j} + F_{k,i} \right) \\ & = \left( \frac{F_{i,k} F_{k,j} }{1-F_{k,k}} + F_{i,j} \right) (1-F_{k,k}) \\ & = Q_{i,j}(1-F_{k,k}) \leq Q_{i,j}\,. \end{align*} \end{proof} \subsection{Hitting probability metric} In this section we establish~\cref{t:Metric}. In particular, we explore the notion that, much like effective resistance, the normalized hitting probabilities matrix provides a natural notion of distance on the digraph (or between states of a Markov Chain). To begin, we recall the definition of $d^\beta$ from~\cref{e:Dist} and note that we have already established the symmetry $d^\beta(i,j) = d^\beta (j,i)$ for all $i,j$. As seen from the statement of~\cref{t:Metric}, we will observe that the triangle inequality holds for all $\beta \ge \sfrac12$ and that positivity holds for all $\beta > \sfrac12$. In the case $\beta = \sfrac12$, $d^{\sfrac12}$ gives a pseudometric structure, as there can indeed exist structures in a directed graph or Markov Chain on which $d^{\sfrac12} (i,j) = 0$ and $i \neq j$. As an example, consider the nodes on a cycle with in degree and out degree $1$. (See~\cref{s:Examples}.) When $d^{\sfrac12}(i,j)=0$ there are specific structures that restrict all random walks leaving $i$ so that they must hit $j$ before returning to $i$. We show in~\cref{t:12bounds} that for any graph, there exists a canonical quotient graph on which $d^{\sfrac12}$ is indeed a metric that is closely related to $d^{\sfrac12}$ on the original graph. Let us now proceed to the proofs. \begin{proof}[Proof of \cref{t:Metric}] First, we show positivity for $i\ne j$. Note that $d^\beta(i,j) > 0$ iff $\Ahtb_{i,j} < 1$. Consider the converse \[ 1 \le \Ahtb_{i,j} = \frac{\phi_i^\beta}{\phi_j^{1-\beta} } Q_{i,j} = \frac{\phi_j^\beta}{\phi_i^{1-\beta} } Q_{j,i}\,, \] that is, \[ \frac{\phi_j^{1-\beta}}{\phi_i^\beta} \le Q_{i,j} \quad \text{ and } \quad \frac{\phi_i^{1-\beta}}{\phi_j^\beta} \le Q_{j,i}\,. \] Then if $\beta > \sfrac12$, we have \[ 1 \ge Q_{i,j} Q_{j,i} \ge \phi_j^{1-2\beta} \phi_i^{1-2\beta} > 1\,, \] a contradiction. For $\beta=\sfrac12$, the last inequality above becomes an equality, so the corresponding argument by contradiction requires only that $\Ahtb[\sfrac12] \le 1$ and thus $d^{\sfrac12}(i,j) \ge 0$. Symmetry follows from~\cref{l:symmetric}, and $d^\beta(i,i)=0$ is immediate, so all that remains is the triangle inequality. To prove the triangle inequality, we observe for $i\neq j\neq k$ that \begin{align} d^\beta(i,j) &= - \log \left( \Ahtb_{i,j} \right) = -\beta\log\phi_i -(\beta-1)\log\phi_j - \log Q_{i,j} \notag \\ &= d^\beta(i,k) + d^\beta(k,j) + (2 \beta - 1) \log\phi_k + [\log Q_{i,k} + \log Q_{k,j} - \log Q_{i,j}]\,, \label{e:triangle-slack} \end{align} which, applying~\cref{Qlemma}, proves that the triangle inequality holds for all $\beta \geq \sfrac12$. \end{proof} We observe that the $2 \beta -1$ coefficient of $\log \phi_k$ in~\cref{e:triangle-slack} vanishes when $\beta=\sfrac12$, hence the triangle inequality is as tight as possible (since $Q_{i,k}Q_{k,j}/Q_{i,j} = 1$, {\it e.g.\/},~for a directed cycle graph). From the above argument, we can see that the only obstruction to $d^{\frac12}$ being a metric is if there is a pair $i,j$ such that \[ \Ahtb[\sfrac12]=\sqrt{ \frac{\phi_j}{\phi_i} } Q_{j,i} = \sqrt{ \frac{\phi_i}{\phi_j} } Q_{i,j} = 1\,. \] In this case, \[ Q_{j,i} = \sqrt{ \frac{\phi_i}{\phi_j} } \quad\text{ and }\quad Q_{i,j} = \sqrt{ \frac{\phi_j}{\phi_i} }\,. \] Thus, $Q_{i,j} Q_{j,i} =1$, which, as they are both probabilities, means in fact $Q_{i,j} = Q_{j,i} =1$. Hence, also $\phi_i = \phi_j$. \begin{obs} The condition $\phi_i = \phi_j$ is not an extra restriction beyond $Q_{i,j} = Q_{j,i} =1$: if $Q_{i,j} =Q_{j,i}=1$ then a random walker must visit $i$ every time it visits $j$ (and vice versa) and hence the invariant probabilities of sites $i$ and $j$ must be equivalent. \end{obs} \subsection{Structure theory of digraphs where \texorpdfstring{$d^{\sfrac12}$}{d} is not a metric}% \label{structure} In this subsection, we investigate the structure of graphs where $d^{\sfrac12}$ is not a metric, which we refer to as ($d^{\sfrac12}$-) degenerate. This is useful for understanding our metric embedding and is foundational for~\cref{quotients}, where we derive the quotienting procedure to repair graph degeneracies. In this section, we first give a general construction to produce degenerate graphs and show that all degenerate graphs can be constructed in this way. Next, we give a general decomposition of degenerate graphs into equivalence classes and their segments. \subsubsection{A general construction for degenerate graphs} A simple example of a graph with $Q_{i,j}=Q_{j,i}=1$ is a closed, directed cycle. However, we also have the following much more general construction: Take any two directed acyclic graphs (DAGs), $G_1$ and $G_2$. Connect all the leaves (sinks/nodes of our-degree zero) of $G_1$ to $i$ and all the leaves of $G_2$ to $j$. Connect $j$ to all the roots (sources/nodes of in-degree 0) of $G_1$ and $i$ to all the roots of $G_2$. Possibly add edges between $i$ and $j$. Then, for each node $k$ except $i$ and $j$, replace it with an arbitrary strongly connected graph $H_k$ (corresponding to an irreducible Markov chain), replacing each edge to (from) $k$ with at least one edge to (from) a node in $H_k$. The resulting graph is strongly connected and has $i$ and $j$ only reachable through each other. In fact, all graphs with $Q_{i,j}=Q_{j,i}=1$ can be constructed this way. To see this, note that $i$ and $j$ must have positive in and out degree by strong connectedness. Let $C_i$ be those nodes reachable from $i$ without passing through $j$, and define $C_j$ similarly. Consider the following claim: \begin{claim} $C_i \cup C_j \cup \{i,j\}$ includes all nodes of the graph, and $C_i\cap C_j$ is empty. \end{claim} \begin{proof} For the first part, consider a fixed node $k$ which is not $i$ or $j$, together with a shortest path $C_{ik}$ from $i$ to $k$. If $C_{ik}$ does not pass through $j$, then $k\in C_i$; otherwise, $k$ is reachable from $j$ without passing through $i$ since $C_{ik}$ is a shortest path. For the second part, assume otherwise; that is, pick $k \in C_i \cap C_j$. Then there exist (1) a path $C_{ik}$ from $i$ to $k$ not passing through $j$, (2) a path $C_{jk}$ from $j$ to $k$ not passing through $i$, and (3) a path $C_{kj}$ from $k$ to $j$ (by strong connectedness). Assume WLOG that $C_{kj}$ passes through $j$ only at the end. $C_{kj}$ cannot pass through $i$ since otherwise $C_{ik}+C_{kj}$ contains a walk from $i$ to $i$ without passing through $j$, violating $Q_{i,j}=1$. But then $C_{jk}+C_{kj}$ would be a walk from $j$ to $j$ that does not pass through $i$, contradicting $Q_{j,i}=1$. \end{proof} Since $C_i$ and $C_j$ can only be connected through $i$ and $j$, removing $i$ and $j$ disconnects these two sets. Now consider the subgraphs induced by $C_i$ and $C_j$, respectively. As can be done with any directed graph, we reduce each of these subgraphs to their quotients under the mutual reachability equivalence relation, yielding a pair of DAGs. The next subsection generalizes this decomposition to account for all nodes for which $d^{\sfrac12}$ vanishes rather than a single pair. \subsubsection{Decomposition into equivalence classes and segments} Consider an equivalence class $\alpha=\{a_1,a_2,\ldots,a_K\}$ of nodes under the equivalence relation $i\sim j \Leftrightarrow d^{\sfrac12}_{i,j}=0$. We refer to a node in a non-singleton equivalence class as \emph{($d^{\sfrac12}$-) degenerate}. A graph is $d^{\sfrac12}$-degenerate if it has a degenerate node. \begin{definition}\leavevmode \begin{itemize} \item A \emph{walk} is a sequence of nodes $\{i_1,i_2,\ldots,i_K\}$ such that $P_{i_k,i_{k+1}}\!\!>\!0$, for $1 \le k<K$. \item A walk is \emph{closed} if $i_K = i_1$. \item A closed walk is a \emph{commute} from $i_1$ if $i_{k} \ne i_1$, for $1<k<K $. \item A walk is a \emph{path} if $i_k\ne i_{k'}$ when $k'\ne k$. A commute is a \emph{cycle} if it is a path when the last element is removed. \end{itemize} \end{definition} \begin{lemma} A commute from $a_k\in \alpha$ must include each of the other members of $\alpha$ exactly once in an order that depends only on the graph. \end{lemma} \begin{proof} The proof is in three assertions. We assume without loss of generality that commutes are from $a_1$. First, each $a_k$ is visited. This is the same as claiming that $Q_{a_1,a_k} = 1$, which was shown in the proof of~\cref{t:Metric}. Second, each $a_k$ is visited at most once, since $Q_{a_k,a_1} = 1$. Lastly, each $a_k$ is visited in a fixed order: Let $J_1$ and $J_2$ be commutes from $a_1$ that visit, respectively, $a_2$ before $a_3$ and vice versa. Then let $J_1'$ and $J_2'$ be the sub-walks from $a_1$ to $a_2$ and from $a_2$ to $a_1$ in $J_1$ and $J_2$, respectively. The concatenation of $J_1'$ and $J_2'$ is thus a commute from $a_1$ that does not visit $a_3$, a contradiction. \end{proof} In the rest of~\cref{structure}, we assume that equivalence classes under $\sim$ are sorted so that they must be visited in the order by commutes from their first element. Similarly, if the $K$ elements of an equivalence class are numbered $a_1,\ldots,a_K$, we naturally identify $a_x = a_{x \!\!\!\mod\! K}$. \begin{lemma}% \label{seg_lem} Given an equivalence class $\alpha$ under $\sim$, for each $j\in G - \alpha$ there is a unique $k$ such that all walks $J$ containing $j$ with $\alpha\cap J \ne \varnothing$ include either $a_{k-1}$ before $j$ or $a_{k}$ after $j$. \end{lemma} \begin{proof} It is enough to consider paths. Suppose $J_1$ and $J_2$ are two paths from $j$ which reach $a_k$ and $a_{k'}$, respectively, before reaching any other elements of $\alpha$, with $k\ne k'$. By strong connectivity, we can select a shortest path from $a_k$ to $j$ to extend $J_1$ to a cycle from $j$, which we call $J_3$. That is, if we select a shortest (in number of distinct steps) path, $\gamma$, from $a_k$ to $j$ then $J_1 \cup \gamma= J_3$ is the required extension of $J_1$. Now, $J_3$ can be cyclically reordered to be a commute from $a_k$. Thus, $J_3$ includes $a_{k'}$, and since $\gamma$ was shortest possible, it includes $a_{k'}$ exactly once. Let $J_4 \subset J_3$ be the sub-walk from $a_{k'}$ to $j$. Then concatenating $J_4$ and $J_2$ gives a commute from $a_{k'}$ that does not include $a_k$, a contradiction. Thus, $j$ has a unique successor $a_k$ in $\alpha$. The conclusion that there is a unique predecessor of $j$ in $\alpha$ follows by reversing the direction of all edges and re-applying the above argument. It must be $a_{k-1}$ since $a_k$ is the first member of the equivalence class encountered in any commute from $j$. \end{proof} \begin{definition} We will here refer to the equivalence classes on $G-\alpha$ induced by \cref{seg_lem} as \emph{($\alpha$-) segments} of $G$.\footnote{Alternatively, we could define segments more generally with respect to any node set $\alpha$. Then the segment corresponding to $i \in \alpha$ is the set of nodes reachable from $i$ without passing through any other elements of $\alpha$. From this perspective, the absolute segments described later are simply the intersection of the segments with respect to all the equivalence classes.} \end{definition} \begin{lemma} Given distinct equivalence classes $\alpha$ and $\beta$ under $\sim$, every element of $\alpha$ must lie within a single segment induced by $\beta$. \end{lemma} \begin{proof} Let $\alpha = \{a_1,a_2,\ldots,a_{K_{\alpha}}\}$ and $\beta = \{b_1,b_2,\ldots,b_{K_{\beta}}\}$. Suppose, by way of contradiction, that $a_1$ lies between $b_{k_1}$ and $b_{k_1+1}$ and $a_2$ lies between $b_{k_2}$ and $b_{k_2+1}$ for $k_1 \ne k_2$. By strong connectedness, there exists a (shortest) path from $b_{k_1}$ to $a_1$ to $b_{k_1+1}$. If $b_{k_1+1}$ and $b_{k_2}$ are distinct nodes, there also exists a shortest path from $b_{k_1+1}$ to $b_{k_2}$. Since $Q_{b_{k_2},a_2} < 1$, there exists a shortest path from $b_{k_2}$ to $b_{k_2+1}$ not passing though $a_2$. Finally, if $b_{k_2+1}$ and $b_{k_1}$ are distinct nodes, there exists a shortest path from $b_{k_2+1}$ to $b_{k_1}$. Concatenating all these paths gives a commute from $a_1$ to itself not passing though $a_2$, a contradiction. \end{proof} The foregoing lemmata show that the nontrivial equivalence classes in a $d^{\sfrac12}$-degenerate digraph induce a structure of equivalence cycles and their segments, with distinct equivalence cycles restricted to lie within the segments of each other. This has potential application in segmentation of directed graphs and will be an important technical tool in the proofs in the next subsection. \subsection{Quotients of \texorpdfstring{$d^{\sfrac12}$}{d}-degenerate Markov chains}% \label{quotients} Next, we develop a way to transform a Markov chain $X$ for which $d^{\sfrac12}$ is not a metric into a quotient Markov chain $X'$, for which $d^{\sfrac12}$ is a metric. \begin{rem} In~\cref{quotients}, we identify singleton classes with their member. Additionally, we append a prime to any symbol when it is meant to refer to $X'$ rather than $X$. \end{rem} The quotient graph is given by the following construction, which has appeared in~\cite{mitavskiy08} as well as in~\cite{madras_2002_decomposition,martin_2000_staircase,caracciolo_1992_tempering}, and possibly other places. \begin{definition}% \label{d:quotient} Given a Markov chain $X$ and an equivalence relation on the states of $X$, the \emph{quotient Markov chain} has one state for each equivalence class, and the transition probabilities are given by \[ P_{U,V}' = \frac{1}{\phi_U} \sum_{i\in U} \phi_i P_{i,V} = \frac{1}{\phi_U} \sum_{i\in U} \sum_{j\in V} \phi_i P_{i,j}\,, \] where $\phi_U = \sum_{i \in U} \phi_i$. \end{definition} The map that sends $X$ to $X'$ is denoted $\iota$. It can be shown~\cite{mitavskiy08} that the invariant measure on $X'$ evaluated at state $U$ is $\phi'_U = \phi_U$. Furthermore, $P$ carries information about the equilibration rate in ergodic chains~\cite{martin_2000_staircase,madras_2002_decomposition,caracciolo_1992_tempering}, although we do not use this fact in this paper. When applying~\cref{d:quotient} to $\sim$, the definition reduces to \[ P_{U,V}' = \frac{1}{|U|} \sum_{i\in U} \sum_{j\in V} P_{i,j}\,, \] since $\phi$ is constant within equivalence classes (see proof of~\cref{t:Metric}). \begin{lemma}[Quotienting one class at a time] Let $\sim$ induce the non-singleton classes $\{\alpha_1,\alpha_2,\ldots,\alpha_L\}$. For a node set $S$, let $\sim_{S}$ be the relation with non-singleton class $S$, keeping all other nodes in individual (singleton) classes. One can then produce a graph with the same nodes as $P'$ by performing a series of quotienting operations $P \underset{\sim_{\alpha_1}}{\to} P_1 \underset{\sim_{\alpha_2}}{\to} \cdots \underset{\sim_{\alpha_L}}{\to} P_L$. Then $P_L = P'$, after identifying nested classes with the nodes in them, {\it e.g.\/}, $\{\{a\},\{b\}\} \to \{a,b\}$.% \label{l:order} \end{lemma} \begin{proof} The proof is by induction on $L$. If $L=1$, the result is vacuously true. So assume the result is true for graphs having $L$ non-singleton equivalence classes, and we proceed to establish the result for $L+1$ non-singleton classes. Let $G$ have classes $\{\alpha_1,\ldots,\alpha_{L+1}\}$ under $\sim$. Then we apply $\sim_{\alpha_1}$ to get $P_1$ and then use the inductive assumption to conclude that $P_{L+1} = P_1'$. So we need to prove that $P_1'=P'$. We have \[ {P_1'}_{\alpha,\beta} = \begin{cases} P_{\alpha,\beta} & \alpha\ne\alpha_1,\beta\ne\alpha_1 \\ \sum_{j\in\beta} {P_1}_{\alpha_1,j} & \alpha=\alpha_1,\beta\ne\alpha_1 \\ \frac1{|\alpha|} \sum_{i\in\alpha} {P_1}_{i,\alpha_1} & \alpha\ne\alpha_1,\beta=\alpha_1\\ {P_1}_{\alpha_1,\alpha_1} & \alpha=\alpha_1=\beta\, , \end{cases} \] where we have implicitly used the fact that $\phi_{P_1}$ has the form given in~\cref{l:order}. Expanding further gives \[ {P_1'}_{\alpha,\beta} = \begin{cases} P_{\alpha,\beta} & \alpha\ne\alpha_1,\beta\ne\alpha_1 \\ \sum_{j\in\beta} \frac1{|\alpha_1|}\sum_{i\in\alpha_1} P_{i,j} & \alpha=\alpha_1,\beta\ne\alpha_1 \\ \frac1{|\alpha|} \sum_{i\in\alpha} \sum_{j\in\alpha_1} P_{i,j} & \alpha\ne\alpha_1,\beta=\alpha_1\\ \frac1{|\alpha_1|} \sum_{i\in\alpha_1,j\in\alpha_1} P_{i,j} & \alpha=\alpha_1=\beta\,. \end{cases} \] Rearranging sums \[ {P_1'}_{\alpha,\beta} = \frac1{|\alpha|} \sum_{i\in\alpha,j\in\beta} P_{i,j} = P_{\alpha,\beta}\,, \] as expected. \end{proof} \begin{lemma}% \label{l:single-collapse} Collapsing a single equivalence class $\alpha$ respects $Q$ in the following sense. Let $i$ and $j$ be two non-equivalent nodes. \begin{itemize} \item If $i$ and $j$ lie in the same $\alpha$-segment, then $Q_{i,j} = Q'_{i,j}$. \item If $i$ and $j$ lie in different $\alpha$-segments, then $\frac{1}{2} Q_{i,j} < Q'_{i,j} < Q_{i,j}$. \item If $i\in\alpha$, then $Q_{i,j} = |\alpha| Q'_{\alpha,j}$. \item If $j\in\alpha$, then $Q_{i,j} = Q'_{i,\alpha}$. \end{itemize} \end{lemma} \begin{proof} Let $\alpha=\{a_1,\ldots,a_K\}$, where $K = |\alpha|$. It is clear that $Q_{i,j}$ is unaffected by taking quotients if $i$ and $j$ lie in the same $\alpha$-segment or if $j\in\alpha$. For $i=a_k\in\alpha$, we know that $Q_{a_k,j} = Q_{a_{\ell},j}$ for all $\ell$, so WLOG assume that $j$ lies in the segment between $a_k$ and $a_{k+1}$. Now, let us denote by $Q_{i_1,i_2,i_3}$ the probability of a random walker starting at $i_1$ and reaching $i_2$ before reaching $i_3$ (in particular, $Q_{i,j} = Q_{i,j,i}$). Then, \begin{align*} Q'_{\alpha,j} &= P'_{\alpha,j} + \sum_{i'\ne j,\alpha} P'_{\alpha,i'} Q'_{i',j,\alpha} = \frac1K P_{i,j} + \frac1K \sum_{\ell=1}^K \sum_{i'\ne j, i'\notin \alpha} P_{a_{\ell},i'} Q_{i',j,\alpha} \\ &= \frac1K P_{i,j} + \frac1K \sum_{i'\ne j, i' \notin \alpha} P_{a_k,i'} Q_{i',j,\alpha} = \frac1K P_{i,j} + \frac1K \sum_{i'\ne j,i} P_{i,i'} Q_{i',j,i} = \frac1K Q_{i,j}\,. \end{align*} Finally let $i$ and $j$ be such that any path from $i$ to $j$ must pass through $a_1,\ldots,a_k$ before encountering $j$. Then the following reasoning applies. Let $a=a_1,b=a_K$, $x = Q_{a,b,j}$ and $y = Q_{b,i,a}$. Then we have \begin{align*} Q_{i,j} &= Q_{i,a,i} Q_{a,j,i}\,, \\ Q_{a,j,i} &= (1-x) + x Q_{b,j,i}\,, \\ Q_{b,j,i} &= (1-y)Q_{a,j,i}\,. \end{align*} Solving for $Q_{i,j}$ yields \[ Q_{i,j} = Q_{i,a}\frac{1-x}{1-x+xy}\,. \] On the quotiented graph, we also have \begin{align*} Q'_{i,j} = Q_{i,a} Q'_{\alpha,j,i}, \ \ Q'_{\alpha,j,i} &= \frac{1}{2} \left[ 1-x + x Q'_{\alpha,j,i}\right] + \frac{1}{2} \left[ (1-y) Q'_{\alpha,j,i} \right]\,. \end{align*} Hence, $ Q'_{i,j} = Q_{i,a}\frac{1-x}{1-x+y} $ and thus $Q'_{i,j} < Q_{i,j}$. Furthermore, we can bound the ratio \begin{equation} \frac{Q_{i,j}}{Q'_{i,j}} = \frac1{1-\frac{(1-x)y}{1-x+y}}\,. \label{rat} \end{equation} Since the function $g(x_1,x_2) = \frac{x_1x_2}{x_1+x_2}$ is bounded above by $\frac{1}{2}$ on ${(0,1)}^2$,~\cref{rat} cannot exceed $2$, which gives the bound. The bound is tight because all values of $x$ and $y$ can be attained when considering arbitrary weighted graphs. (A graph with only the four nodes $a,b,i,j$ mentioned in the proof and edges $a\to j$, $a\to b$, $j\to b$, $b\to i$, $b\to a$, and $i\to a$ suffices to attain all possible values of $x,y$.) \end{proof} \begin{definition} An \emph{absolute segment} is a maximal set of nodes which lie in the same segment with respect to all non-singleton equivalence classes. \end{definition} \begin{lemma} $\iota$ respects $Q$ in the following sense for nodes $i$ and $j$ in distinct equivalence classes $\alpha$ and $\beta$: \begin{itemize} \item If $i$ and $j$ lie in the same absolute segment, then $Q_{i,j} = Q'_{i,j}$. \item Otherwise, $\frac{1}{2^{c}|\alpha|} Q_{i,j} \le Q'_{\alpha,\beta} < Q_{i,j}$, where $c$ is the number of equivalence classes with respect to which $i$ and $j$ lie in different segments. (In particular, $c<L$.) Equality holds only when $c=0$. \end{itemize} \end{lemma} \begin{proof} If $i$ is degenerate, first collapse $\alpha$, scaling $Q_{i,j}$ by $|\alpha|$. Next, collapse all other equivalence classes one at a time, further scaling $Q_{i,j}$ by the appropriate factor in $(\frac{1}{2},1)$ whenever $i$ and $j$ lie in different segments with respect to the collapsing class. \end{proof} From this lemma we immediately get the following theorem. \begin{theorem}\label{t:12bounds} $X'$ is a metric space with metric $(d')^{\sfrac12}$. In particular, for $i\in\alpha$ and $j\in\beta$, with $\alpha\ne\beta$: \begin{itemize} \item If $i$ and $j$ lie in the same absolute segment, then $d^{\sfrac12}_{i,j} = (d')_{i,j}^{\sfrac12}$. \item Otherwise $d_{i,j} < d_{\alpha,\beta}' \le d_{i,j} + \frac{1}{2} \log{|\alpha||\beta|} + c\log 2$, where $c$ is the number of equivalence classes with respect to which $i$ and $j$ lie in different segments. Equality holds only when $c=0$. \end{itemize} \end{theorem} Thus, $\iota$ pushes apart the different absolute segments. All other distances are unaffected. \begin{rem} $\iota$ is analogous to a rigid motion on each absolute segment, in that none of the in-absolute-segment distances are distorted. \end{rem} \section{Computational methods}\label{s:CompMeth} To compute the normalized hitting probabilities matrix and metric structure on a Markov chain (or network) consisting of $n$ nodes/states with probability transition matrix $P$, we require only the computation of the invariant measure and the $Q$ matrix. The invariant measure can be computed using iterative eigenvector methods, which need $O(m)$ operations per iteration for $m$ edges. We briefly recall the work in~\cite[Theorem 5]{Thiede_2015}, that shows the $Q$ matrix can be computed in $O(n^3)$ time. The key idea from~\cite[Lemma 5]{Thiede_2015} is that one can compute \begin{equation} \label{eq:Q} Q_{i,j} (P) = \frac{e_i^T {(I - P_j)}^{-1} P_j e_j}{e_i^T {(I - P_j)}^{-1} e_i} = \frac{ {M(j)}^{-1}_{i,j}}{ {M(j)}^{-1}_{i,i}}\,, \end{equation} where $e_j\in \mathbb R^n$ is the vector with a $1$ in the $j$th entry and zeros elsewhere, $P_j = (I-e_j e_j^T) P \in \mathbb R^{n\times n}$, and the invertible matrix $M(j) = I - P + e_j e_j^T P\in \mathbb R^{n\times n}$. See Theorem $5$ of~\cite{Thiede_2015} for full details, but this identity follows from realizing that as defined $M(j)$ is invertible with inverse \[ M(j)^{-1} = \begin{pmatrix} (I-P_j)^{-1} & (I-P_j)^{-1} P_j e_j \\ 0 & 1 \end{pmatrix} \] given in block form on the $e_j^\perp$, $e_j$ basis. If we then compute $M(1)^{-1}$ on the way to obtaining the first column $Q_{i,1} = {M(1)}^{-1}_{i1}/{M(1)}^{-1}_{ii}$, then $M(j)$ is a rank-$2$ perturbation of $M(1)$ and we can apply the Sherman-Morrison-Woodbury identity to compute $M(j)^{-1}$. Since we only access $2n-2$ elements of $M(j)^{-1}$, the full $O(n^2)$-time Sherman-Morrison-Woodbury update is not needed, and we can get the $j$th column $Q_{i,j}$ in $O(n)$ computations from ${M(1)}^{-1}$. A \textsc{MATLAB} implementation of this procedure, along with code for all of the numerical experiments described in the paper, is available at~\url{https://github.com/zboyd2/hitting_probabilities_metric}. The matrix $Q$ encodes the hitting probabilities of a random walk on the nodes of a graph and the order of the method we present here is very well documented in \cite{Thiede_2015}. However, there are several results that consider the computational complexity of the related problem of commute times, see for instance the works \cite{li2010random,boley2011commute}. The computational cost of computing hitting probabilities through inversion of the Laplacian has been explored further in \cite{golnari2019markov,cohen2016faster,cohen2018solving}, resulting in some cases in which the method may be improved to better than $O(n^3)$. As we are mostly interested in the construction of the metric here, we will not further explore the question of optimal order of the computation. \section{Examples}\label{s:Examples}% We consider examples of Markov Chains and directed graphs to illustrate the proposed metric. We start with simple graphs for which the calculations can be performed exactly. We then numerically explore a variety of synthetic graphs and a real-world example defined from New York City taxi cab data. \subsection{Exact formulations}% \label{sec:glued} Here we consider some simple graphs on which the invariant measure and hitting probabilities can be computed exactly to help us understand $\Ahtb$ and $d^{\beta}$. \begin{enumerate} \item \emph{Directed cycles:} Consider a directed cycle on $n$ nodes. Then $\phi_i = \sfrac{1}{n}$ for all $i$, and $Q_{i,j} = 1$ for all $i\ne j$. Therefore, $\Ahtb$ is a weighted clique, and $d^{\beta}$ has all points equidistant. For $\beta=\sfrac12$, the weights equal to $1$ and all nodes are identified with each other in the metric topology. \item \emph{Complete graphs:} Consider a complete graph on $n>2$ nodes. Then $\phi_i = \sfrac{1}{n}$ for all $i$, and $Q_{i,j} = \mathrm{const} < 1$ for all $i \ne j$. Therefore, $\Ahtb$ is a weighted clique. Unlike the directed cycle case, the weights in the clique are $<1$ for all $\beta\geq\sfrac12$. \item \emph{Glued cycles:} Consider graphs of the type depicted in~\cref{fig:glue}, namely graphs composed of $n_b$ ``backbone nodes'' forming a directed chain, which then branches into $C$ chains of length $n_c$, each of which finally connects back to the beginning of the backbone chain. Intuitively, a random walker on this graph transitions between $C+1$ groups of nodes, namely, each of the $C$ branches and the backbone. As illustrated in~\cref{tab:glue}, our metric captures this intuition by placing each node very close to the others on its chain. This is in contrast to commute-time-based metrics, where the length of the chain must be taken into account. (See~\cref{fig:effRes}.) In~\cref{sec:num_ex}, we consider some numerical results based on this example. \end{enumerate} \begin{table} \centering \begin{tabular}{llcccc} \toprule $i$ & $j$ & $Q$ & $\frac{\Ahtb[\beta]}{(n_b + n_c)^{1-2\beta}}$ & $\Ahtb[\sfrac12]$ & $d^\frac12$ \\ \midrule branch & same branch & $1$ & $\sfrac1{C^{2\beta-1}}$ & $1$ & $0$ \\\addlinespace[2pt] branch & different branch & $\sfrac12$ & $\sfrac1{2 C^{2\beta-1}}$ & $\sfrac12$ & $\log 2$ \\\addlinespace[2pt] backbone & branch & $\sfrac1C$ & $\sfrac1{C^{\beta}}$ & $C^{-\sfrac12}$ & $\sfrac12 \log C$ \\\addlinespace[2pt] branch & backbone & $1$ & $\sfrac1{C^{\beta}}$ & $C^{-\sfrac12}$ & $\sfrac12 \log C$ \\\addlinespace[2pt] backbone & backbone & $1$ & $1$ & $1$ & $0$ \\\addlinespace[2pt] \bottomrule \end{tabular} \caption{Values of $Q$, $\Ahtb$, and $d$ evaluated at distinct nodes $i$ and $j$ for the glued cycles example from~\cref{sec:glued}. We include extra columns for the case $\beta=\sfrac12$, which is particularly interpretable. Observe that neither $\Ahtb$ nor $d^{\beta}$ depends on $n_b$ or $n_t$ (except up to scaling), which is a manifestation of their blindness to walk length. Also, the nodes that are closest together are those which lie on common chains. Note that we scaled $\Ahtb$ for visual clarity. The invariant measure is easily verified to be $\left(n_b + n_c\right)^{-1}$ on the backbone and $\left(C (n_b + n_c)\right)^{-1}$ elsewhere.}% \label{tab:glue} \end{table} \subsection{Synthetic numerical examples}% \label{sec:num_ex} We consider four examples. The first two demonstrate that the spectrum of $\Ahtb$ (for $\beta = \sfrac12$ or $\beta=1$) identifies cyclic and clique-like sets in a useful manner. We compare to two alternative symmetrizations and another metric. The second example additionally shows the scalability of our approach. In the third example, we explore when it is advantageous to use $d$ for visualization and clustering purposes, using a directed planted partition model for ground truth comparisons. In dense, difficult-to-detect regimes, our method is more accurate than clustering using the input adjacency matrix directly. Finally, in the fourth example, we compare $d^{\sfrac12}$, $d^{1}$, and spatial distance for geometric graphs, finding that our distance captures comparable information to the spatial distance, with the similarity being especially tight when $\beta=\sfrac12$. \subsubsection{Glued-cycles networks}% \label{sec:gluedcycles} For the two-glued-cycles networks illustrated in~\cref{fig:glue}, we construct a probability transition matrix $P$ by taking a uniform edge weight for all connected vertices and performing a row normalization. We then compute the Fiedler eigenvector corresponding to the second smallest eigenvalue of the graph Laplacian (sometimes called the Fiedler vector) for different symmetrized adjacency matrices. For the adjacency matrices $A$ constructed below, we calculate the graph Laplacian $L=D-A$, with $D$ the diagonal matrix of node degrees (row or column sums of $A$). The examples here are two directly glued cycles, as well as two glued cycles with a bidirectional edge between the cycles. In the first case, the results are all very similar regardless of the symmetrization, but for the second case the results differ significantly. In each case, we group the nodes based on whether the corresponding vector element is positive, negative, or zero. \begin{figure} \centering \begin{tikzpicture}[scale=0.75] \begin{scope}[every node/.style={circle,thick,draw}] \node (A) at (0,0) {}; \node (B) at (0,1) {}; \node (C) at (0,2) {}; \node (D) at (1,2.5) {}; \node (E) at (1.5,1.5) {}; \node (F) at (1.5,.5) {} ; \node (G) at (1,-.5) {} ; \node (H) at (-1,2.5) {} ; \node (I) at (-1.5,1.5) {} ; \node (J) at (-1.5,.5) {} ; \node (K) at (-1,-.5) {} ; \end{scope} \begin{scope}[>={Stealth[black]}, every node/.style={fill=white,circle}, every edge/.style={draw=red,very thick}] \path [->] (A) edge (B); \path [->] (B) edge (C); \path [->] (C) edge (D); \path [->] (D) edge (E); \path [->] (E) edge (F); \path [->] (F) edge (G); \path [->] (G) edge (A); \path [->] (C) edge (H); \path [->] (H) edge (I); \path [->] (I) edge (J); \path [->] (J) edge (K); \path [->] (K) edge (A); \end{scope} \end{tikzpicture} \ \ \ \begin{tikzpicture}[scale=0.75] \begin{scope}[every node/.style={circle,thick,draw}] \node (A) at (0,0) {}; \node (B) at (0,1) {}; \node (C) at (0,2) {}; \node (D) at (1,2.5) {}; \node (E) at (1.5,1.5) {}; \node (F) at (1.5,.5) {} ; \node (G) at (1,-.5) {} ; \node (H) at (-1,2.5) {} ; \node (I) at (-1.5,1.5) {} ; \node (J) at (-1.5,.5) {} ; \node (K) at (-1,-.5) {} ; \end{scope} \begin{scope}[>={Stealth[black]}, every node/.style={fill=white,circle}, every edge/.style={draw=red,very thick}] \path [<->] (F) edge (J); \path [->] (A) edge (B); \path [->] (B) edge (C); \path [->] (C) edge (D); \path [->] (D) edge (E); \path [->] (E) edge (F); \path [->] (F) edge (G); \path [->] (G) edge (A); \path [->] (C) edge (H); \path [->] (H) edge (I); \path [->] (I) edge (J); \path [->] (J) edge (K); \path [->] (K) edge (A); \end{scope} \end{tikzpicture} \ \ \ \begin{tikzpicture}[scale=0.75] \begin{scope}[every node/.style={circle,thick,draw}] \node[fill=magenta] (A) at (0,0) {}; \node[fill=magenta] (B) at (0,1) {}; \node[fill=magenta] (C) at (0,2) {}; \node[fill=blue] (D) at (1,2.5) {}; \node[fill=blue] (E) at (1.5,1.5) {}; \node[fill=blue] (F) at (1.5,.5) {} ; \node[fill=blue] (G) at (1,-.5) {} ; \node[fill=green] (H) at (-1,2.5) {} ; \node[fill=green] (I) at (-1.5,1.5) {} ; \node[fill=green] (J) at (-1.5,.5) {} ; \node[fill=green] (K) at (-1,-.5) {} ; \end{scope} \begin{scope}[>={Stealth[black]}, every node/.style={fill=white,circle}, every edge/.style={draw=red,very thick}] \path [->] (A) edge (B); \path [->] (B) edge (C); \path [->] (C) edge (D); \path [->] (D) edge (E); \path [->] (E) edge (F); \path [->] (F) edge (G); \path [->] (G) edge (A); \path [->] (C) edge (H); \path [->] (H) edge (I); \path [->] (I) edge (J); \path [->] (J) edge (K); \path [->] (K) edge (A); \end{scope} \end{tikzpicture} \\ \vspace{.2cm} \begin{tabular}{cccc} \begin{tikzpicture}[scale=0.75] \begin{scope}[every node/.style={circle,thick,draw}] \node[fill=blue] (A) at (0,0) {}; \node[fill=blue] (B) at (0,1) {}; \node[fill=green] (C) at (0,2) {}; \node[fill=green] (D) at (1,2.5) {}; \node[fill=green] (E) at (1.5,1.5) {}; \node[fill=blue] (F) at (1.5,.5) {} ; \node[fill=blue] (G) at (1,-.5) {} ; \node[fill=green] (H) at (-1,2.5) {} ; \node[fill=green] (I) at (-1.5,1.5) {} ; \node[fill=blue] (J) at (-1.5,.5) {} ; \node[fill=blue] (K) at (-1,-.5) {} ; \end{scope} \begin{scope}[>={Stealth[black]}, every node/.style={fill=white,circle}, every edge/.style={draw=red,very thick}] \path [<->] (F) edge (J); \path [->] (A) edge (B); \path [->] (B) edge (C); \path [->] (C) edge (D); \path [->] (D) edge (E); \path [->] (E) edge (F); \path [->] (F) edge (G); \path [->] (G) edge (A); \path [->] (C) edge (H); \path [->] (H) edge (I); \path [->] (I) edge (J); \path [->] (J) edge (K); \path [->] (K) edge (A); \end{scope} \end{tikzpicture} & \begin{tikzpicture}[scale=0.75] \begin{scope}[every node/.style={circle,thick,draw}] \node[fill=green] (K) at (-1,-.5) {} ; \node[fill=green] (A) at (0,0) {}; \node[fill=green] (B) at (0,1) {}; \node[fill=green] (C) at (0,2) {}; \node[fill=blue] (D) at (1,2.5) {}; \node[fill=blue] (E) at (1.5,1.5) {}; \node[fill=blue] (F) at (1.5,.5) {} ; \node[fill=green] (G) at (1,-.5) {} ; \node[fill=blue] (H) at (-1,2.5) {} ; \node[fill=blue] (I) at (-1.5,1.5) {} ; \node[fill=blue] (J) at (-1.5,.5) {} ; \end{scope} \begin{scope}[>={Stealth[black]}, every node/.style={fill=white,circle}, every edge/.style={draw=red,very thick}] \path [<->] (F) edge (J); \path [->] (A) edge (B); \path [->] (B) edge (C); \path [->] (C) edge (D); \path [->] (D) edge (E); \path [->] (E) edge (F); \path [->] (F) edge (G); \path [->] (G) edge (A); \path [->] (C) edge (H); \path [->] (H) edge (I); \path [->] (I) edge (J); \path [->] (J) edge (K); \path [->] (K) edge (A); \end{scope} \end{tikzpicture} & \begin{tikzpicture}[scale=0.75] \begin{scope}[every node/.style={circle,thick,draw}] \node[fill=magenta] (A) at (0,0) {}; \node[fill=magenta] (B) at (0,1) {}; \node[fill=magenta] (C) at (0,2) {}; \node[fill=blue] (D) at (1,2.5) {}; \node[fill=blue] (E) at (1.5,1.5) {}; \node[fill=blue] (F) at (1.5,.5) {} ; \node[fill=blue] (G) at (1,-.5) {} ; \node[fill=green] (H) at (-1,2.5) {} ; \node[fill=green] (I) at (-1.5,1.5) {} ; \node[fill=green] (J) at (-1.5,.5) {} ; \node[fill=green] (K) at (-1,-.5) {} ; \end{scope} \begin{scope}[>={Stealth[black]}, every node/.style={fill=white,circle}, every edge/.style={draw=red,very thick}] \path [<->] (F) edge (J); \path [->] (A) edge (B); \path [->] (B) edge (C); \path [->] (C) edge (D); \path [->] (D) edge (E); \path [->] (E) edge (F); \path [->] (F) edge (G); \path [->] (G) edge (A); \path [->] (C) edge (H); \path [->] (H) edge (I); \path [->] (I) edge (J); \path [->] (J) edge (K); \path [->] (K) edge (A); \end{scope} \end{tikzpicture} & \begin{tikzpicture}[scale=0.75] \begin{scope}[every node/.style={circle,thick,draw}] \node[fill=magenta] (A) at (0,0) {}; \node[fill=magenta] (B) at (0,1) {}; \node[fill=magenta] (C) at (0,2) {}; \node[fill=blue] (D) at (1,2.5) {}; \node[fill=blue] (E) at (1.5,1.5) {}; \node[fill=blue] (F) at (1.5,.5) {} ; \node[fill=blue] (G) at (1,-.5) {} ; \node[fill=green] (H) at (-1,2.5) {} ; \node[fill=green] (I) at (-1.5,1.5) {} ; \node[fill=green] (J) at (-1.5,.5) {} ; \node[fill=green] (K) at (-1,-.5) {} ; \end{scope} \begin{scope}[>={Stealth[black]}, every node/.style={fill=white,circle}, every edge/.style={draw=red,very thick}] \path [<->] (F) edge (J); \path [->] (A) edge (B); \path [->] (B) edge (C); \path [->] (C) edge (D); \path [->] (D) edge (E); \path [->] (E) edge (F); \path [->] (F) edge (G); \path [->] (G) edge (A); \path [->] (C) edge (H); \path [->] (H) edge (I); \path [->] (I) edge (J); \path [->] (J) edge (K); \path [->] (K) edge (A); \end{scope} \end{tikzpicture} \\ $A = \max(P,P^T)$ & Chung's $L$~\cite{Chung_2005} & $A =\Ahtb[1]$ & $A=\Ahtb[\sfrac12]$ \end{tabular} \caption{(Top Left) Two-glued-cycles example from~\cref{sec:glued} with $n_b=3$, $n_c=4$, and $C=2$. The ``backbone'' nodes run along the center, and the two partial cycles split off from and then return to it. (Top Middle) Similar two-glued-cycles network with a bidirectional edge. (Top Right) Sign of the Fiedler vector of the Laplacian for several different symmetrizations. (Bottom) The sign of the Fiedler vector of the Laplacian for several different symmetrizations. The sign of the Fiedler vectors is encoded as ($-$, green), ($0$, magenta) and ($+$, blue). \label{fig:glue} \end{figure} For the two glued cycles without the bidirectional edge, the naive symmetrizations of the directed adjacency matrix, either $A = (P+P^T)/2$ or $A=\max\left(P,P^T\right)$, have a Fiedler vector that is $0$ on the spine and splits each cycle into signed components, see the top right plot in~\cref{fig:glue}. However, in the bottom left component~\cref{fig:glue}, for the two-glued-cycles network with the bidirectional edge, the naive symmetrization splits the network horizontally, which is reasonable, since the resulting graph cut is small, although this (by construction) does not reflect the coherent, directed structure of the original graph. One way to account for directed structure in a way that minimizes equilibrium flux across the cut was suggested by Fan Chung~\cite{Chung_2005} (cited in \cref{s:RelWork}), defining the Laplacian by $ L = I - \frac{1}{2} \left[ \Phi^{\sfrac 1 2} P \Phi^{-\sfrac 1 2} + \Phi^{-\sfrac 1 2} P^T \Phi^{\sfrac 1 2} \right], $ where $\Phi = \textrm{diag}(\phi) \in \mathbb R^{n \times n}$. Chung uses $L$ to establish a Cheeger-type inequality for digraphs, which is used to study the rate of convergence for Markov chains. Using Chung's Laplacian again gives a comparable outcome for the two glued cycles example (\cref{fig:glue}), but in the example with the bidirectional edge, this symmetrization places most of the non-backbone nodes in one class and all backbone nodes in the other (second plot in~\cref{fig:glue}). The normalized hitting probabilities matrices $\Ahtb[1]$ and $\Ahtb[\sfrac12]$ each distinguish between the two branches, with the backbone set equal to zero in both the cases of the glued cycles and the glued cycles with a bidirectional edge as seen in~\cref{fig:glue}. Thus, all three approaches uncover different structure in the two-glued-cycle graph with the bidirectional edge, with the naive symmetrization yielding small undirected cuts, Chung's approach yielding (perhaps) two different dynamical states, and $\Ahtb[\beta]$ showing all three chains in a natural way for both $\beta = \sfrac12, 1$. Finally, we compare the total effective resistance metric of~\cite{Young_2016b} to our metric on the example of the two-glued-cycles network (with no bidirectional edge). As one might expect given the relation ship between effective resistance and commute times in the undirected case, the total effective resistance of~\cite{Young_2016b} is sensitive to cycle length. \Cref{fig:effRes} demonstrates that the commute time approach views the distances on each cycle quite differently and that the relative distances from the total effective resistance metric are more difficult to interpret in the second loop. \begin{figure} \centering \begin{tabular}{ccc} \begin{tikzpicture}[xscale=.35,yscale=.4] \Vertices[RGB=true,size=.1]{nodesd12.dat} \Edges[Direct=True,lw=.01,opacity=.2]{edges.dat} \end{tikzpicture} & \begin{tikzpicture}[xscale=.35,yscale=.4] \Vertices[RGB=true,size=.1]{nodesd1.dat} \Edges[Direct=True,lw=.01,opacity=.2]{edges.dat} \end{tikzpicture} & \begin{tikzpicture}[xscale=.35,yscale=.4] \Vertices[RGB=true,size=.1]{nodesR.dat} \Edges[Direct=True,lw=.01,opacity=.2]{edges.dat} \end{tikzpicture} \\ $d^{\sfrac12}$ & $d^1$ & Total effective resistance \end{tabular} \caption{Two glued cycles with $n_c=55$ and $n_b=5$, with nodes colored by distance from a node on the far right. Blue denotes small distances. The metric $d^{\sfrac12}$ has three levels of distance corresponding to nodes on the same branch, backbone, and opposite branch, respectively. The metric $d^1$ is similar, except nodes on the same branch are not distinguished from backbone nodes. Finally, the total effective resistance metric from~\cite{Young_2016b} gives a smoother notion of distance on the right branch and backbone, but on the left branch, proceeding counterclockwise, one finds the distance decreasing and then increasing again, which is somewhat difficult to interpret. This example shows how different resistance/commute time are from hitting-probability distance.}% \label{fig:effRes} \end{figure} \subsubsection{Cycle adjoined to directed Erd\H{o}s--R\'enyi Graph} Consider the following construction, illustrated in~\cref{fig:er}. Let $n=n_{\mathrm{er}} + n_{\mathrm{cycle}}$, and let the \begin{wrapfigure}[15]{r}{.35\textwidth} \centering \includegraphics[width=.35\textwidth]{er_plus_cycle.pdf} \vspace*{-0.2in} \caption{Erd\H{o}s-R\'enyi plus cycle example.} \label{fig:er} \end{wrapfigure} first $n_{\mathrm{er}}$ nodes form an unweighted, directed ER graph with connection probability $p$. The remaining $n_{\mathrm{cycle}}$ nodes form an unweighted, directed cycle. An adjacency matrix for the ER graph and cycle are connected by adding $2\, \mathrm{round}(n p)-1$ edges of weight $w$ to each cycle node from randomly selected nodes in the ER graph.\footnote{These edges are drawn with replacement with multi-edges merged to a single edge of weight $w$. Results were similar when we added the weights instead.} Finally, a single, bidirectional edge of weight $1$ is added from one cycle node to one ER node. Normalizing the rows to form a probability transition matrix, a random walker on this graph would transition between the ER and cycle subgraphs, where the cycle subgraph is difficult to escape quickly because of the single exit. For the particular choice of $n_{\mathrm{er}}=20$, $n_{\mathrm{cycle}}=8$, $p=.5$, and $w=3$, we find that the Fiedler vector of (the Laplacian associated with) $\Ahtb[\sfrac12]$ is positive on the cycle nodes and negative elsewhere. In contrast, the Fiedler vector of the naive symmetrization $A=(P+P^T)/2$ or Chung's $L$~\cite{Chung_2005} does not separate the cycle and ER nodes. Scaling up to $n=n_{\mathrm{er}}+n_{\mathrm{cycle}} = 7,200 + 2,800 = 10,000$ nodes keeping the other parameters the same ($\approx 38.7$ million edges) gives similar eigenvector results. The computation takes 31 seconds on a Lenovo ThinkStation P410 desktop with Xeon E5--1620V4 3.5 GHz CPU and 16 GB RAM using {\sc MATLAB} R2019a Update 4 (9.6.0.1150989) 64-bit (glnxa64): 18 seconds to compute $Q$, 6 seconds to compute $\phi$, 2 seconds to form $\Ahtb[\sfrac{1}{2}]$, and 5 seconds to compute the Fiedler vector. \subsubsection{Cluster detection and visualization for digraphs}\label{sec:kmeans} We next use $d$ for clustering and dimension reduction. We consider directed graphs generated by a planted partition model with nodes grouped into three ground truth communities and form a uniformly weighted adjacency matrix by connecting an edge from $i$ to $j$ with probability $p_{\mathrm{in}}$ if $i$ and $j$ are in the same community and $p_{\mathrm{out}}$ ($<p_{\mathrm{in}}$) otherwise. A probability transition matrix can then be formed using row normalization. We then attempt to recover the ground truth node assignments. \begin{wrapfigure}[15]{r}{.32\textwidth} \centering \includegraphics[width=.3\textwidth]{pca_layout.pdf} \caption{PCA embedding of $d^{\sfrac12}$, colored by ground truth community.} \label{fig:pca} \end{wrapfigure} The difficulty of this problem is generally understood in terms of $\Delta = p_{\mathrm{in}} - p_{\mathrm{out}}$ and $\rho = \frac{p_{\mathrm{in}} + 2 p_{\mathrm{out}}}{3}$. Small values of $\Delta$ correspond to more difficult clustering problems that may be solved less accurately (relative to the ground truth). In this example we attempt to cluster the nodes into $k=3$ clusters using several approaches: (1) principal component analysis\footnotemark (PCA)~\cite{pca} on the adjacency matrix, $A$, followed by $k$-means clustering on the first $k-1$ PCA vectors; (2) PCA on $d^{\sfrac12}$ followed by $k$-means; and (3) $k$-medoids on $d^{\sfrac12}$. (The $k$-medoids algorithm is similar in spirit to the $k$-means unsupervised clustering algorithm but applies in arbitrary metric spaces, see for instance \cite{kaufmann1987clustering,park2009simple}.) Results are shown in~\cref{fig:kmedoids,fig:regions}. \footnotetext{Specifically, we used the PCA routine from MATLAB R2019a Update 4 (9.6.0.1150989) 64-bit (glnxa64). As expected, this gives different results in general when applied to a matrix versus its transpose. In this case, the matrix is stochastically equivalent with its transpose, and in the NY taxi example below, the PCA-based plots are similar regardless of whether the transpose is used.} \begin{figure}[!ht] \centering \footnotesize \def<desired width>{\linewidth} \input{clusteringstudyinput.tex} \caption{Results of (top row) PCA on $A$ followed by $k$-means, (middle) PCA on $d^{\sfrac12}$ followed by $k$-means, and (bottom) $k$-medoids on $d^{\sfrac12}$ on 300-node graphs generated using the directed planted partition model with three clusters, as described in~\cref{sec:kmeans}. We varied the mean edge density, $\rho$, and cluster quality, $\Delta=p_{\mathrm{in}}-p_{\mathrm{out}}$. Since results depend on the random initialization, we report best of 5 runs for each entry. If any generated graph was not strongly connected, we did not try to cluster it. The left column is the accuracy (purity) of the recovered partition, and the right value is the empirical $p$ value of the accuracy relative to 4,000 random partitions obtained by drawing each community label uniformly at random. Notably, method (2) has the best performance for dense, weakly-clustered graphs. [Note that the triangular blue region on the lower left of each plot represents a $(\rho,\Delta)$ parameter combination that cannot exist.]}% \label{fig:kmedoids} \end{figure} \begin{figure} \centering \footnotesize \def<desired width>{\linewidth} \input{clusteringstudy1input.tex} \caption{(Left) Regions where the methods from~\cref{fig:kmedoids} perform best. Here, light blue is method (1), green is method (2), and yellow is method (3). (Middle) Difference in accuracy between the best and second-best methods. (Right) Ratio of $p$ value of the best and second best methods. from~\cref{fig:kmedoids}. Combining these plots, we see that there is a significant parameter regime consisting of dense, difficult to detect structure, where using $d^{\sfrac12}$ instead of $A$ enhances the spectral detection of structure by 5--20\% for graphs where method (1) is recovering essentially no structure in $A$. Note that the y axis is different from~\cref{fig:kmedoids}.}% \label{fig:regions} \end{figure} We find that method (1) works best on sparse or well-separated clusters, method (2) works best with dense, difficult-to-detect clusters, and method (3) has no clear advantage. More specifically, using $d^{\sfrac12}$ in method (2) enhances our ability to get a better-than-chance clustering in dense networks.\footnote{We also tried using the shortest commute and generalized effective resistance metrics~\cite{Young_2016b,Young_2016a} as substitute for the hitting time metric in this example and found similar improvements over using the raw adjacency matrix. In particular, the shortest commute was the most effective metric for this task (although this metric is not robust, so the real-life performance may be different).} (We note that spectral methods in undirected graphs give asymptotically optimal almost-exact recovery but are not optimal for harder cases where only better-than-chance recoverability is possible~\cite{abbe}. This is consistent with~\cref{fig:kmedoids,fig:regions}.) Finally, we can also use PCA on $d^{\sfrac12}$ to visualize the directed network. The first and second principal components, generated using {\sc MATLAB}'s built-in routine, are plotted in~\cref{fig:pca}, clearly showing the separation into three clusters, which are in accordance with the three ground-truth communities. \subsubsection{Distances on geometric graphs} Given known convergence properties of various graph models to continuum problems (e.g. \cite{trillos2016consistency,trillos2018variational,singer2012vector,singer2017spectral,Osting_2017}), we are motivated by the question of how our distance metric compares to a standard notion of distance when the network arises from a natural geometric setting. For instance, as mentioned in the introduction, \cite{singer2012vector} proves that the notion of diffusion distance converges to that of geodesic distance as a point cloud samples a closed manifold at higher and higher densities. In~\cref{fig:geodistance}, we consider distances computed using our metric structure in a family of geometric graphs constructed using Euclidean distances to determine edge weights. The geometric graphs considered are \begin{enumerate}[(a)] \item A random point cloud on a flat torus ${[0,2 \pi]}^2$ with $36^2$ points, \item A random point cloud on a flat torus with a hole ${[0,2 \pi]}^2 \setminus B((\pi,\pi),\pi/2)$ with $36^2$ points (distances relative to a point in the bottom left of the torus), \item An $H$ shaped domain $({[0,2 \pi]}^2 \cap \{|x_1-\pi|\geq \pi/2\}) \cup ({[0,2 \pi]}^2 \cap \{|x_2-\pi|\leq \pi/4\})$ with $36^2$ points (distances relative to a point in the bottom right of the $H$), \item A random point cloud on the circle of length $2 \pi$ with $1000$ points, \item A random point cloud on a sphere of radius $1$ in $\mathbb{R}^3$ with $1000$ points, \item A square $10 \times 10$ lattice on the flat torus ${[0,2 \pi]}^2$. \end{enumerate} For the regular lattice example, the edge weights are only carried on nearest neighbor vertices. In all other cases, we consider the edge weights to be of the form $e^{-\gamma d_{\text{Euc}}{(x_i,x_j)}^2}$, where $d_{\text{Euc}}$ is just the Euclidean distance metric (determined with periodicity if the domain is periodic, {\it i.e.\/}, we take shortest path distance in the flat torus). We have chosen the scale factor $\gamma = 1$ uniformly throughout. Once the geometric graph is constructed, we computed the pairwise Euclidean distances, as well as the pairwise distances $d^{\sfrac12}$ and $d^1$ for comparison. To assist with interpretation and comparison, we have ordered the vertices in~\cref{fig:geodistance} from closest to farthest relative to the $d^{\sfrac12}$ metric and plotted for each distance function the rescaled distances $(d-d_{\min})/(d_{\max} - d_{\min})$ to normalize all of them to the same scale. Throughout, we note that $d^{\sfrac12}$ is a reasonable fit to the measured Euclidean distances, while $d^1$ seems to do well only when the geometry is such that the invariant measure normalization (that is, the choice of $\beta$) does not matter as much. Note that the distance $d^{\sfrac12}$ and $d^1$ are identical on the square lattice, up to scaling. In this case, we are really studying the structure of the $Q$ hitting probability matrix. Our results give some preliminary indication that in the consistency limit the $d^{\sfrac12}$ metric may converge to the Euclidean distance while the $d^1$ metric converges to something else entirely. However, we leave this pursuit for future analytical studies. \begin{figure}[t!] \centering \begin{subfigure}[t]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{FlatTorusPC_alt.pdf} \caption{Flat Torus} \end{subfigure}\quad \begin{subfigure}[t]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{FlatTorusHolePC_alt.pdf} \caption{Flat Torus with a Hole} \end{subfigure}\quad \begin{subfigure}[t]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{HdomainPC_alt.pdf} \caption{$H$ shaped domain} \end{subfigure} \\ \begin{subfigure}[t]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{CirclePC_alt.pdf} \caption{Circle} \end{subfigure}\quad \begin{subfigure}[t]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{SpherePC_alt.pdf} \caption{Sphere} \end{subfigure}\quad \begin{subfigure}[t]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{SquareLattice.pdf} \caption{Square Lattice} \end{subfigure} \caption{Normalized distance plots comparing the scaled distances from one node in a geometric graph computed using Euclidean distances (Black), $d^{\sfrac12}$ (Blue) and $d^1$ (Red). The geometric graphs from top left to bottom right are {\bf (a)} A random point cloud on a flat torus {\bf (b)} A random point cloud on a flat torus with a hole, {\bf (c)} A random point cloud on an $H$ shaped domain, {\bf (d)} A random point cloud on the circle, {\bf (e)} A random point cloud on a sphere, {\bf (f)} A square lattice on the flat torus. Note, in all subplots, we have ordered the vertices from closest to farthest from a reference node given by the first vertex generated relative to the $d^{\sfrac12}$ metric.} \label{fig:geodistance} \end{figure} \subsection{Real-world example: the New York City taxi network}% \label{sec:taxi}% Consider the movement over time of a New York City taxi, which we interpret as a Markov chain where the states are neighborhoods and $P_{i,j}$ is the probability that a trip begun in neighborhood $i$ ends in neighborhood $j$. Using publicly available data from the New York City Taxi and Limousine Commission,\footnote{Accessed at~\url{https://www1.nyc.gov/site/tlc/about/tlc-trip-record-data.page} in April 2020.} we computed an adjacency matrix where $A_{i,j}$ is the number of Yellow Taxi trips in January 2019 that started at $i$ and ended at $j$, where $i$ and $j$ are chosen from 262 neighborhoods\footnote{Two additional neighborhoods are marked ``unknown'' and appear to designate out-of-city or out-of-state endpoints. We excluded these from our analysis.} spread across the city's five boroughs (Manhattan, Staten Island, Queens, Brooklyn, and the Bronx). We also included trips to and from Newark Liberty International Airport (EWR) in New Jersey. We restricted our analysis to the 250-neighborhood strongly connected component The data is dominated by degree, as shown in~\cref{nyc_heatmap}, with the busier Manhattan neighborhoods having tens of thousands of trips, and the Staten Island neighborhoods having median out-degree of 4. Traffic is also organized by borough, although the distinctions between the spatially adjacent Brooklyn, Queens, and Bronx boroughs are perhaps less apparent, and they might be properly considered as peripheral to the Manhattan core. Staten Island is notable for its remoteness, which is reflected in the sparsity of $A$ in that block. In~\cref{nyc_heatmap}, we compare $A$ with $d^{\sfrac12}$ and $d^1$. While $d^{\sfrac12}$ highlights Manhattan in a manner similar to $A$, it does not distinguish much between Queens, Brooklyn, and the Bronx, showing them instead as a single interconnected group. In contrast, Staten Island is very clearly highlighted as its own, close group, which is reasonable given the geographic proximity of these neighborhoods and the fact that a disproportionately large number of trips involving Staten Island both started and ended there. Although the purpose of this example is not to provide an optimal clustering of the data, we note that Staten Island does represent a difficult cluster to detect, and arguably is not even a cluster, since there are only eight interior edges (counting multiplicity but excluding self-edges) and 309 incoming or outgoing edges, all of which is hidden in over 1 million edges (again, counting multiplicity). Note that the fact that Staten Island is highlighted by $d^{\sfrac12}$ is not simply because of degree scaling, as a heat map of $P$ does not highlight Staten Island as a block. The true explanation seems to involve two factors: (1) Staten Island has eight non-diagonal in-edges 13 neighborhoods,\footnote{Staten Island has 20 neighborhoods, but 7 have 0 out-degree and are thus excluded from the strongly connected component.} and the median out-degree is 4. Thus, a taxi that does enter Staten Island has a relatively large likelihood of visiting another Staten Island location next, relative to taxis starting at other neighborhoods. (2) The average frequency of visiting Staten Island at all is so low that the pattern of visiting is almost memoryless, with taxis leaving Staten Island having plenty of time to mix in other areas before visiting Staten Island again, so that the probability of leaving Staten Island and then reaching another Staten Island location before returning to the first one is about $\frac12$, despite the low degree of Staten Island neighborhoods. In contrast, Staten Island is far from other locations, especially Manhattan, since by~\cref{d2sym}, mutually high hitting probabilities are required for closeness, but the probability of starting in a Manhattan neighborhood and reaching Staten Island before returning is very low. The distance $d^1$ places the Manhattan nodes close to most other nodes, especially each other, while the Staten Island nodes are far from everything, especially each other. Since $d^1_{i,j} = -\log(\phi_i) - \log(Q_{i,j})$, this distance is small only when (1) $\phi_i$ is large and (2) $Q_{i,j}$ is far from zero. Thus, the Staten Island nodes, which have small values of $\phi$, cannot be close to anything, and the Manhattan nodes, which have the largest values of $\phi$, can be close to other nodes, depending on $Q_{i,j}$. Empirically, $Q_{i,j}$ is usually not very small, with 77\% of the entries in $Q$ being at least $0.1$, which explains Manhattan's overall closeness to other nodes. The fact that the Manhattan nodes are closer to each other than to other nodes is accounted for by the fact that $Q_{i,j}$ for $i$ in Manhattan is generally larger if $j$ is also in Manhattan, which might be expected. (The medians differ by a factor of $5.4$.) A similar observation explains why the Staten Island nodes are considered farther from each other than they are from nodes in the other boroughs. Finally, we used $d^{\sfrac12}$ to perform PCA, with the first two principal components (PCs) visualized in~\cref{nyc_pca}. These two PCs explained 64\% and 34\% of the variation, respectively, with the first PC being closely related to out-degree (Pearson correlation with $\log k_{\mathrm{out}}$ is $.96$) and the second PC being well-correlated (Pearson correlation $.9978$) with the column means of $d^{\sfrac12}$. So over 98\% of the variance is explained by these two PCs. Interestingly, both the highest- and lowest-degree nodes were on average far from other nodes. Recalling that $d^{\sfrac12}_{i,j}$ is the negative log of the geometric mean of $Q_{i,j}$ and $Q_{j,i}$ (see~\cref{d2sym}), closeness requires that both of these factors be high. If $i$ is a high-degree core node, then $Q_{i,j}$ is small for most $j$. In contrast, a mid-degree peripheral node in Queens, the Bronx, or Brooklyn, enjoys reasonable values of $Q_{i,j}$ for other peripheral nodes $j$, since once a taxi enters the Manhattan core, it is likely to visit a significant portion of the other nodes before returning to $i$. Finally, if a node's degree is too small, the probability of a taxi reaching it at all is too small for the hitting probabilities to be high. For comparison, performing PCA directly on $A$ gives a similar first PC, with a different second PC that explains about half as much variance as the second PC of $d^{\sfrac12}$. The second PC is nearly constant, except on Manhattan, where it correlates with the East-West coordinate. \begin{figure}[t] \centering \footnotesize \begin{tabular}{@{\hskip 0in}c@{\hskip 0.18in}c@{\hskip 0.18in}c@{\hskip 0in}} \def<desired width>{.31\linewidth} \input{nyc_adjacencyinput.tex} & \def<desired width>{.29\linewidth} \input{nyc_distanceinput.tex} & \def<desired width>{.30\linewidth} \input{nyc_d1input.tex}\\ $A$ & $d^{\sfrac12}$ & $d^1$ \end{tabular} \caption{An example based on New York City taxi transit data, where nodes are 250 neighborhoods and $A_{i,j}$ is the number of trips from $i$ to $j$. (Left) Heatmap of $A$ with the nodes sorted by borough in this order: EWR airport (one node), Staten Island, Manhattan, Queens, Brooklyn, and the Bronx. The taxi traffic is dominated by the Manhattan block in the upper left, with Queens, Brooklyn, and the Bronx forming three blocks further down and to the right. (Middle) A similarly arranged heatmap of $d^{\sfrac12}$. The Manhattan neighborhoods are close together, and the smaller upper left block corresponding to Staten Island is distinguished as a coherent submodule, despite having only 8 interior edges. Queens, Brooklyn, and the Bronx form a large block in the lower right. See~\cref{sec:taxi} for an explanation of these differences. (Right) A heatmap of $d^{1}$. We observe that in this normalization Staten Island is quite far from everything, including itself. Manhattan is relatively close to almost everything, especially itself. This is exactly what we should expect because $d^1_{i,j}$ is small only when both $\phi_i$ and the $i\rightarrow j$ hitting probability are large. [In the right two heatmaps, the diagonal is set to a non-zero value to improve contrast.]}% \label{nyc_heatmap} \end{figure} \begin{figure}[t] \centering \begin{tabular}{cc} \begin{tikzpicture}[scale=0.42] \begin{axis}[ cycle list={ {red,mark=*}, {blue,mark=*}, {yellow,mark=*}, {green,mark=*}, {brown,mark=*}, {orange,mark=* }, legend entries={Manhattan, Staten Island, Queens, Brooklyn, Bronx,EWR}, legend cell align=left, only marks, mark size=1pt, width=\textwidth, ticks=none, xlabel={Principal component 1 (log scale)}, ylabel={Principal component 2}, title={Principal component analysis of $d^{\sfrac12}$ for New York City Taxis}, ] \addplot table [col sep=comma] {pca3.dat}; \addplot table [col sep=comma] {pca2.dat}; \addplot table [col sep=comma] {pca4.dat}; \addplot table [col sep=comma] {pca5.dat}; \addplot table [col sep=comma] {pca6.dat}; \addplot table [col sep=comma] {pca1.dat}; \end{axis} \end{tikzpicture} & \begin{tikzpicture}[scale=0.42] \begin{axis}[ cycle list={ {red,mark=*}, {blue,mark=*}, {yellow,mark=*}, {green,mark=*}, {brown,mark=*}, {orange,mark=* }, legend entries={Manhattan, Staten Island, Queens, Brooklyn, Bronx,EWR}, legend cell align=left, only marks, mark size=1pt, width=\textwidth, ticks=none, xlabel={Principal component 1 (log scale)}, ylabel={Principal component 2}, title={Principal component analysis of $A$ for New York City Taxis}, ] \addplot table [col sep=comma] {pcaA3.dat}; \addplot table [col sep=comma] {pcaA2.dat}; \addplot table [col sep=comma] {pcaA4.dat}; \addplot table [col sep=comma] {pcaA5.dat}; \addplot table [col sep=comma] {pcaA6.dat}; \addplot table [col sep=comma] {pcaA1.dat}; \end{axis} \end{tikzpicture} \end{tabular} \caption{(Left) These two PCs explain $98.3\%$ of the variance. The first PC has a $.96$ correlation coefficient with $\log k_{\mathrm{out}}$, and the second PC as a correlation coefficient of $.9978$ with the column means of $d^{\sfrac12}$, which we interpret as the average distance to other nodes. Notably, the highest-degree nodes also have high average distance to other nodes. This is also true of the lowest-degree nodes, while the mid-degree nodes in Queens, Brooklyn, and the Bronx are closer to other nodes on average. We interpret this by noting that, while high-degree nodes are common endpoints for trips, (so $Q_{i,j}$ might be high when $j$ is a high-degree node), they have a lot of self-loops, and the taxis that leave them tend to return relatively quickly (so $Q_{j,i}$ is low for most $i$). Using~\cref{d2sym}, we see that $d^{\sfrac12}_{i,j}$ will then not be very small for high-degree $i$. The mid-degree nodes, in contrast, send a lot of taxis into the Manhattan core, which are likely to mix through the city for a long time before returning (so $Q_{i,j}$ is not very small for almost all destinations $j$). (Right) PCA on $A$ gives a similar first PC. The second PC is nearly constant except on Manhattan, where it is correlated with the East-West coordinate (Pearson .42, p=.0004). The second PC explains about half as much variance for $A$ as for $d^{\sfrac12}$.}% \label{nyc_pca} \end{figure} \section{Conclusion}\label{s:Disc} Given a probability transition matrix for an ergodic, finite-state, time-homogeneous Markov chain, we have constructed a family of (possibly pseudo-)metrics on the state space, which we refer to as hitting probability distances. Alternatively, this construction gives a metric on the nodes of a strongly connected, directed graph. In the cases where we do not obtain a proper metric, the degeneracies give global structural information, and we can quotient them away. Our metrics can be computed in $O(n^3)$ time and $O(n^2)$ space, in one example scaling up $10,000$ nodes and $\approx 38M$ edges on a desktop computer. Our metric captures different information compared to other directed graph metrics and captures multiscale structure in the taxi example. We have considered the utility of this metric for structure detection, dimension reduction, and visualization, finding in each case advantages of our method compared to existing techniques. Some other possible applications include efficient nearest-neighbor search, new notions of graph curvature~\cite{van_Gennip_2014}, Cheeger inequalities, and provable optimality of weak recovery for dense, directed communities. Additionally, in our experiments, we observed that several eigenvalues of the symmetrized adjacency matrix contained useful information about structure such as cycles, and it would be good to understand better which structures get encoded in leading eigenspaces. Empirically, it is important to know how commonly $d^{\sfrac12}$ is degenerate, and what useful structure is revealed in practice. A natural theoretical question is consistency of the distances in the large graph limit as we approach a natural geometric object embedded in a standard Euclidean space~\cite{Osting_2017,singer2012vector,singer2017spectral,trillos2018variational,trillos2016consistency,Yuan2020}. In terms of possible improvements to our method, an effective means of thresholding the symmetrized hitting probability matrix could improve scalability. A natural question to pursue in a variety of settings would be the sparsification of $\Ahtb$ and its implications for spectral analysis and clustering applications. In particular, the potentially sparse $P$ will map into a full (but symmetric) matrix $\Ahtb$. In large systems the $O(n^2)$ storage requirement may become a burden. Hence, it is natural to ask: If we sparsify the $\Ahtb$ matrix to have a comparable number of edges to that of the original $P$, how much information can be stably preserved in the spectrum? This will be a topic of future work on the hitting probability matrices we have constructed. \bibliographystyle{siamplain}
{ "timestamp": "2021-01-19T02:43:57", "yymm": "2006", "arxiv_id": "2006.14482", "language": "en", "url": "https://arxiv.org/abs/2006.14482", "abstract": "The shortest-path, commute time, and diffusion distances on undirected graphs have been widely employed in applications such as dimensionality reduction, link prediction, and trip planning. Increasingly, there is interest in using asymmetric structure of data derived from Markov chains and directed graphs, but few metrics are specifically adapted to this task. We introduce a metric on the state space of any ergodic, finite-state, time-homogeneous Markov chain and, in particular, on any Markov chain derived from a directed graph. Our construction is based on hitting probabilities, with nearness in the metric space related to the transfer of random walkers from one node to another at stationarity. Notably, our metric is insensitive to shortest and average walk distances, thus giving new information compared to existing metrics. We use possible degeneracies in the metric to develop an interesting structural theory of directed graphs and explore a related quotienting procedure. Our metric can be computed in $O(n^3)$ time, where $n$ is the number of states, and in examples we scale up to $n=10,000$ nodes and $\\approx 38M$ edges on a desktop computer. In several examples, we explore the nature of the metric, compare it to alternative methods, and demonstrate its utility for weak recovery of community structure in dense graphs, visualization, structure recovering, dynamics exploration, and multiscale cluster detection.", "subjects": "Social and Information Networks (cs.SI); Machine Learning (cs.LG); Numerical Analysis (math.NA); Probability (math.PR); Machine Learning (stat.ML)", "title": "A metric on directed graphs and Markov chains based on hitting probabilities", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.982013788985129, "lm_q2_score": 0.8152324938410783, "lm_q1q2_score": 0.8005695501806732 }
https://arxiv.org/abs/1602.03246
Reliability Polynomials of Simple Graphs having Arbitrarily many Inflection Points
In this paper we show that for each $n$, there exists a simple graph whose reliability polynomial has at least $n$ inflection points.
\section{Introduction} The reliability of a graph $G$ is the probability that the graph remains connected when each edge is included, or ``functions", with independent probability $p$. Equivalently, we can say that each edge fails with probability $q = 1-p$. This function can be written as a polynomial in either $p$ or $q$, though for our purposes it will be convenient to use $q$; for instance, if $f(q) = R(G)(q)$ for a nontrivial graph $G$, then we have $f(0) = 1$ and $f(1) = 0$. Since the first derivative of $R(G)(q)$ is always negative on $(0,1)$, it is natural to consider whether the second derivative is ever zero, i.e., whether $R(G)(q)$ has any inflection points. It is typical for the reliability of a graph to have at least one inflection point, and families of simple graphs with reliability polynomials having two inflection points have been found \cite{BKK}. In \cite{GM}, Graves and Milan show that there exist non-simple graphs whose reliability polynomials have at least $n$ inflection points for any integer $n$. They point out that no example is known of a simple graph whose reliability polynomial has more than two inflection points. What we show in this paper is that for each $n$, there exists a simple graph whose reliability polynomial has at least $n$ inflection points. \section{Preliminaries} Our proof consists of two major parts: we first demonstrate that there exist reliability polynomials whose second derivative satisfies certain bounds, and then from this collection of polynomials we form products which have arbitrarily many inflection points. The \emph{one-point union} of graphs $G$ and $H$, denoted $G*H$, is the graph union where exactly one vertex is chosen from each graph and the chosen vertices are identified. Regardless of the choices made the reliability polynomial of the one-point union of graphs is the product of the reliability polynomials of each graph. Ultimately, the graphs we use will be one-point unions of complete graphs, and in the following section we demonstrate that the second derivative of such graphs can be made arbitrarily small outside a given interval. In this section we establish a few facts about reliability polynomials in general. The reliability polynomial of a graph $G$ can be written in the form \begin{equation}\label{reldefn} R(G)(q) = \sum_{i=0}^m N_i (1-q)^i q^{m-i}, \end{equation} where $m$ is the number of edges in $G$, and $N_i$ counts the number of connected spanning subgraphs of $G$ with $i$ edges. We first prove a fact about all polynomials having a form similar to \eqref{reldefn}. \begin{prop}\label{derivbound} Let $f(q) = \sum_{i=0}^m N_i(1-q)^iq^{m-i}$, where the $N_i$ are either all non-negative or all non-positive. Then for $q \in [0,1]$, \[q(1-q)|f'(q)| \leq m|f(q)|.\] \end{prop} \begin{proof} We first compute \begin{equation}\label{relderiv} f'(q) = \sum_{i=0}^m N_i ((m-i)(1-q)^i q^{m-i-1} - i(1-q)^{i-1}q^{m-i}). \end{equation} Then we have \begin{gather*} q(1-q)|f'(q)| = \left|\sum_{i=0}^m (m-i)N_i(1-q)^{i+1} q^{m-i} - iN_i(1-q)^{i}q^{m-i+1}\right|\\ \leq m(1-q)\left|\sum_{i=0}^m N_i(1-q)^iq^{m-i}\right| + mq\left|\sum_{i=0}^m N_i(1-q)^iq^{m-i}\right| \leq m|f(q)|. \end{gather*} Note that the first inequality, where we bound the coefficients uniformly by $m$, uses the hypothesis that the $N_i$ have the same sign. \end{proof} In the form \eqref{reldefn} of a reliability polynomial, the coefficient $N_i$ counts a subset of the subgraphs of $G$ with $i$ edges, and thus we have $0 \leq N_i \leq \binom{m}{i}$. Then clearly the above theorem applies when $f$ is a reliability polynomial. However, we can also consider \begin{equation} (1-R(G))(q) = \sum_{i=0}^m \left(\binom{m}{i} - N_i\right)(1-q)^i q^{m-i}. \end{equation} By the above observation, the coefficients of this polynomial are non-negative, and so \hyperref[derivbound]{Lemma \ref*{derivbound}} applies to $1-R(G)$ as well. Next we show that it can also be applied to $R(G)'(q)$. \begin{prop} Let $f(q) = \sum_{i=0}^m N_i(1-q)^iq^{m-i}$ be a reliability polynomial. Then $f'(q)$ can be written in the same form, and all of the coefficients are non-positive. \end{prop} \begin{proof} We already computed the derivative of such a function in \eqref{relderiv}; collecting like terms gives \begin{equation}\label{relderivcollect} f'(q) = \sum_{i=0}^{m-1} ((m-i)N_i - (i+1)N_{i+1})(1-q)^iq^{m-i-1}. \end{equation} Recall that $N_i$ represents the number of connected spanning subgraphs of $G$ with $i$ edges. Thus we can think of $(m-i)N_i$ as counting the pairs consisting of a connected spanning subgraph of size $i$ together with a particular edge not in the subgraph. Similarly, $(i+1)N_{i+1}$ counts the number of pairs consisting of a connected spanning subgraph of size $i+1$ and an edge in the subgraph. However, since adding an edge to a connected spanning graph gives another connected spanning graph, $(m-i)N_i \leq (i+1) N_{i+1}$. Thus each of the coefficients in the expression for $f'(q)$ is non-positive. \end{proof} We note that, by the preceding argument, the coefficient of $(1-q)^i q^{m-i-1}$ in $R(G)'(q)$ counts the number of pairs consisting of a connected spanning subgraph of size $i+1$ and a bridge in the subgraph. This gives the following result. \begin{prop}\label{endpointderiv} Let $f(q)$ be the reliability polynomial of a graph $G$. If $G$ has no bridges, then $f'(0) = 0$. If $G$ has at least 3 vertices, then $f'(1) = 0$. \end{prop} \begin{proof} From \eqref{relderivcollect}, we see that $f'(0) = 0$ when the coefficient of $(1-q)^{m-1}$ is zero. Since there is only a single subgraph of size $m$, namely the graph $G$, this is equivalent to $G$ having no bridges. Similarly, $f'(1) = 0$ whenever the coefficient of $q^{m-1}$ is zero. In a subgraph with one edge, that edge is a bridge; thus, $f'(1) = 0$ if and only if there are no connected spanning subgraphs with one edge, which is clearly true if the graph $G$ has 3 or more vertices. \end{proof} \section{Bounding the Reliability of Complete Graphs} For the first half of the proof, we will work directly with the reliability polynomials of complete graphs. We make use of a recurrence relation given in Colbourn's book \cite{Colbourn}, which we restate here. If $r_n(q)$ is the reliability of the complete graph $K_n$, then \begin{align*} r_1 &= 1;\\ r_n &= 1 - \sum_{k=1}^{n-1} \binom{n-1}{k-1}q^{k(n-k)}r_k. \end{align*} To make these polynomials easier to work with, we bound them by simpler polynomials. \begin{lem} Let $\alpha = \frac{1}{8}$, and let $r_n$ denote $R(K_n)$. For $q \in [0,\alpha]$ and $n \geq 2$, we have \[1 - (n+1)q^{n-1} \leq r_n(q) \leq 1 - (n-1)q^{n-1}.\] \end{lem} \begin{proof} We proceed by induction on $n$. The base case where $n=2$ is clear, since $r_2 = 1-q$. Now, suppose the claim is true for $2 \leq k \leq n-1$. We first note that, since $r_n = 1 - \sum_{k=1}^{n-1} \binom{n-1}{k-1}q^{k(n-k)}r_k$, we can rewrite the inequality we would like to prove as \[(n-1)q^{n-1} \leq \sum_{k=1}^{n-1} \binom{n-1}{k-1}q^{k(n-k)}r_k \leq (n+1)q^{n-1}.\] We will first prove the left hand side of the inequality. From the induction hypothesis, $r_{n-1} \geq 1 - nq^{n-2}$, and the fact that $r_k \geq 0$ for all $k$, we have \[\sum_{k=1}^{n-1} \binom{n-1}{k-1}q^{k(n-k)}r_k \geq q^{n-1} + (n-1)q^{n-1}(1 - nq^{n-2}).\] If we can show that $q^{n-1}(1 - n(n-1)q^{n-2}) \geq 0$, it will follow that the right hand side is greater than $(n-1)q^{n-1}$, which was what we wanted. If we suppose $q \leq \frac{1}{6}$, then it follows that $n(n-1)q^{n-2} \leq 1$ for $n \geq 3$, and the claim holds. We now proceed to the right hand side of the inequality. Since $r_k \leq 1$ for all $k$, and $n(n-k) \geq 2n-4$ when $2 \leq k \leq n-2$, we have \begin{align*} \sum_{k=1}^{n-1} \binom{n-1}{k-1}q^{k(n-k)}r_k &\leq \sum_{k=1}^{n-1} \binom{n-1}{k-1}q^{k(n-k)}\\ &\leq nq^{n-1} + \sum_{k=2}^{n-2} \binom{n-1}{k-1}q^{k(n-k)} \\ &\leq nq^{n-1} + 2^{n-1} q^{2n-4}. \end{align*} To show that this is less than or equal to $(n+1)q^{n-1}$, it suffices to show that $2^{n-1}q^{n-3} \leq 1$. If $n \geq 4$, then it is enough to take $q \leq \frac{1}{8}$. We may verify directly that \[ r_3 = 1 - q^2 - 2q^2(1 - q) = 1 - 3q^2 + 2q^3 \leq 1 - 2q^2 \] provided $q \leq \frac{1}{2}$, which is clearly satisfied since $q \leq \frac{1}{8}$. This completes our induction proof, and so the bounds hold for all $n \geq 2$. \end{proof} We also prove another bound which will be used during the proof of the following theorem. \begin{lem}\label{roverp2} Let $r_n(q) = R(K_n)(q)$, with $n \geq 3$. Then for $q \in [0,1]$, \[\frac{r_n(q)}{(1-q)^2} \leq \frac{1}{2}\binom{n}{2}^2.\] \end{lem} \begin{proof} Let $m = \binom{n}{2}$ denote the number of edges of $K_n$. For $2 \leq i \leq m$, we have \[\binom{m}{i} = \frac{m(m-1)}{i(i-1)}\binom{m-2}{i-2} \leq \frac{m^2}{2}\binom{m-2}{i-2}.\] Recall that we can write $r_n(q) = \sum_{i=0}^{m} N_i (1-q)^iq^{m-i}$. Since $N_i$ is the number of connected spanning subgraphs of $K_n$ of size $i$, we have $0 \leq N_i \leq \binom{m}{i}$ for all $i$, and $N_i = 0$ for $0 \leq i < n-1$. Since $n \geq 3$, we can say \begin{gather*} \frac{r_n(q)}{(1-q)^2} = \sum_{i=0}^{m-2} N_{i+2}(1-q)^iq^{m-i-2} \leq \sum_{i=0}^{m-2} \binom{m}{i+2}(1-q)^iq^{m-i-2}\\ \leq \frac{m^2}{2} \sum_{i=0}^{m-2} \binom{m-2}{i}(1-q)^iq^{m-2-i} = \frac{m^2}{2} = \frac{1}{2}\binom{n}{2}^2.\qedhere \end{gather*} \end{proof} Now we have enough in place to prove our first theorem. \begin{thm} Let $0 < a < b < \frac{1}{8}$, and let $\epsilon > 0$. Then there exists a graph $G$ such that $|R(G)''(q)| \leq \epsilon$ when $q$ in $[0,1] \setminus [a,b]$. \end{thm} \begin{proof} We first recall that $r_n$ denotes the reliability of the complete graph $K_n$ and the inequalities \[1-(n+1)q^{n-1} \leq r_n(q) \leq 1-(n-1)q^{n-1}\] are valid for $q \in [0,1/8]$. Since we are only interested in upper bounds, we reformulate the left hand side as $1 - r_n(q) \leq (n+1)q^{n-1}$. We compute \[(r_n^\ell)''(q) = \ell(\ell-1)r_n^{\ell-2}(q)r_n'^2(q) + \ell r_n^{\ell-1}(q)r_n''(q),\] from which we now obtain various bounds. Note that the coefficients of $1-r_n(q)$ are all non-negative. From our lemmas, we have \begin{align*} &q(1-q)|r_n'(q)| \leq \binom{n}{2}|r_n(q)|;\\ &q(1-q)|r_n'(q)| = q(1-q)|(1-r_n)'(q)| \leq \binom{n}{2}|1-r_n(q)|;\\ &q(1-q)|r_n''(q)| \leq \binom{n}{2}|r_n'(q)|. \end{align*} For $q \in [0,1/8]$, we have \begin{align*} q^2(1-q)^2|(r_n^{\ell})''(q)| &\leq \ell(\ell-1)\frac{n^2(n-1)^2}{4}|r_n^{\ell-2}(q)||1-r_n(q)|^2\\ &\quad + \ell \frac{n^2(n-1)^2}{4}|r_n^{\ell-1}(q)||1-r_n(q)|\\ &\leq \frac{\ell^2n^6}{4}(1-(n-1)q^{n-1})^{\ell-2}q^{2n-2}\\ &\quad + \frac{\ell n^5}{4}(1-(n-1)q^{n-1})^{\ell-1}q^{n-1}, \end{align*} where we have, in addition to using the lemmas, simplified $(n-1)(n+1) \leq n^2$. We note that $(1-q)^2 \geq \frac{1}{4}$ when $q \leq \frac{1}{8}$, and so we may rearrange this inequality to obtain \begin{align*} |(r_n^{\ell})''(q)| &\leq \ell^2n^6(1-(n-1)q^{n-1})^{\ell-2}q^{2n-4}\\ &\quad +\ell n^5(1-(n-1)q^{n-1})^{\ell-1}q^{n-3}. \end{align*} We denote the two terms in the above bound by $f(q)$ and $g(q)$, respectively. Note that $f$ and $g$ are both dependent on our choice of $\ell$ and $n$; from here on it will be useful to assume that $n \geq 4$. We'd like to demonstrate that $f(q)+g(q) < \epsilon$ on $[0,a]$ and $[b,1/8]$. To do this, we first show that $f+g$ is increasing on the first interval and decreasing on the second; then it suffices to show that $f(a)+g(a) < \epsilon$ and $f(b)+g(b) < \epsilon$. Taking the derivative of $f$ and factoring, we see that the sign of $f'(q)$ is the same as that of the expression \[-(\ell-2)(n-1)^2q^{n-1} + (2n-4)(1-(n-1)q^{n-1}).\] This expression is non-negative precisely when \[((\ell-2)(n-1)^2 + (2n-4)(n-1))q^{n-1} \leq 2n-4.\] After some estimation, we see that if $\ell n(n-1)q^{n-1} \leq 2(n-2)$, then $f'(q) \geq 0$. Since $n \geq 4$, it suffices to show $\ell nq^{n-1} \leq 1$. The sign of $g'(q)$ is the same as that of \[-(\ell-1)(n-1)^2q^{n-1} + (n-3)(1-(n-1)q^{n-1});\] similarly, if $\ell n(n-1)q^{n-1} \leq n-3$, then $g'(q) \geq 0$. Since $n \geq 4$, it suffices to show that $\ell nq^{n-1} \leq \frac{1}{4}$; if this is true, then the condition above holds as well, and so $(f+g)'(q) \geq 0$. Moreover, since $\ell nq^{n-1} \leq \ell na^{n-1}$ for $q \in [0,a]$, it is sufficient to show that $\ell na^{n-1} \leq \frac{1}{4}$. By the same sort of approximation, we see that if $\ell q^{n-1} \geq 1$, then $(f+~g)'(q) \leq 0$. Again, if this is true at $b$, then it clearly holds for all $q \in [b,1/8]$ as well. For $q \in [1/8,1]$, we have $\frac{1}{q^2} \leq 64$. Using our usual bounds, along with \hyperref[roverp2]{Lemma \ref*{roverp2}}, we see \begin{align*} |(r_n^{\ell})''(q)| &\leq \frac{1}{q^2} \ell^2 \binom{n}{2}^2 r_n^{\ell-1}(q) \frac{r_n(q)}{(1-q)^2}\\ &\leq 16\ell^2 \binom{n}{2}^4 r_n^{\ell-1}(q) \leq \ell^2n^8 r_n^{\ell-1}(q)\\ &\leq \ell^2n^8 r_n^{\ell-1}(1/8) \leq \ell^2n^8 (1-(n-1)(1/8)^{n-1})^{\ell-1}. \end{align*} Note that we were able to replace $q$ with $1/8$ since $q \in [1/8,1]$ and $r_n$ is decreasing. We restate our conditions here: we want to find $\ell \geq 1$ and $n \geq 4$ such that \begin{itemize} \item $\ell na^{n-1} \leq \frac{1}{4}$, \item $\ell b^{n-1} \geq 1$, \item $f(a)+g(a) \leq \epsilon$, \item $f(b)+g(b) \leq \epsilon$, and \item $\ell^2n^8 (1-(n-1)(1/8)^{n-1})^{\ell-1} \leq \epsilon$. \end{itemize} If we can find $\ell,n$ satisfying these conditions, then the arguments above show that $|R(K_n^{\ell})''(q)| \leq \epsilon$ for $q \in [0,1] \setminus [a,b]$. To show that all of the above inequalities can be satisfied, we define a sequence of $\ell_i$ and $n_i$ such that the quantities on the left become arbitrarily small (or large, in the case of the second one) for sufficiently large $i$. We begin by choosing positive integers $N,k$ such that $a < N^{-1/k} < b$; this is possible because there is an integer between $b^{-k}$ and $a^{-k}$ for sufficiently large $k$. Then, we let $\ell_i = N^i$ and $n_i = ik$. Consider the first expression: we rewrite \[\ell_i n_i a^{n_i-1} = \frac{ik}{a}(Na^k)^i,\] which becomes small as $i$ goes to infinity, by an application of l'Hospital's rule and the fact that $Na^k < 1$. Similarly, we can write \[\ell_i b^{n_i-1} \geq (Nb^k)^i,\] which tends to infinity since $Nb^k > 1$. For the next two inequalities, we analyze $f$ and $g$ separately. First, we have \begin{align*} \log(f(q)) &= 2\log(\ell_i) + 6\log(n_i) \\ &\quad + (\ell_i-2)\log(1-(n_i-1)q^{n_i-1}) + (2n_i-4)\log(q). \end{align*} Note that $\log(1 - (n_i-1)q^{n_i-1}) \leq -(n_i-1)q^{n_i-1}$; since we'd like to show that $\log(f(q)) \to -\infty$ as $i \to \infty$, we may make this replacement. Thus we must show that the expression \begin{align*} 2i\log(N) + 6\log(ik) - (N^i-2)(ik-1)q^{ik-1} + 2i\log(q^k) - 4\log(q) \end{align*} becomes arbitrarily small as $i$ tends to infinity. The term $4\log(q)$ is constant with respect to $i$, and so after some rearrangement, we need only consider \begin{align*} &2i\log(N) + 6\log(ik) - (N^i-2)(ik-1)q^{ik-1} + 2i\log(q^k)\\ &\leq 2i\log(Nq^k) + 6\log(ik) - N^{-1}(Nq^k)^i \end{align*} For $q=a$, the first term tends to negative infinity, and dominates the second, while the third term is also negative; thus $\log(f(a))$ can be made arbitrarily small, and $f(a) < \frac{\epsilon}{2}$ for sufficiently large $i$. For $q=b$, we have $Nq^k > 1$, and so the third term dominates the first two terms, and again we can make $f(b)$ arbitrarily small. Proceeding similarly, we collect the nonconstant terms of $\log(g(q))$, and make the same upward approximation as before : \begin{align*} \log(\ell) &+ 5\log(n) - (\ell-1)(n-1)q^{n-1} + n\log(q)\\ &= i\log(Nq^k) + 5\log(ik) - (N^i-1)(ik-1)q^{ik-1}\\ &\leq i\log(Nq^k) + 5\log(ik) - N^{-1}(Nq^k)^i. \end{align*} The arguments showing that $g(a)$ and $g(b)$ become arbitrarily small for sufficiently large $i$ are identical to those given above for $f$. Finally, we consider the expression in the fifth inequality. We have \begin{align*} \log(\ell^2&n^8 (1-(n-1)(\tfrac{1}{8})^{n-1})^{\ell-1})\\ &\leq 2i\log(N) + 8\log(ik) -(N^i-1)(ik-1)(\tfrac{1}{8})^{ik-1}\\ &\leq 2i\log(N) + 8\log(ik) -N^{-1}(N(\tfrac{1}{8})^k)^i. \end{align*} Since $N(\frac{1}{8})^k > 1$, the last term dominates the others, and evidently this can also be made arbitrarily small. Thus, if we take $i$ sufficiently large, we see that $G = K_{n_i}^{\ell_i}$ satisfies the desired condition. \end{proof} \section{Main Result} Now we proceed to the proof that there are reliability polynomials with arbitrarily many inflection points. We choose a collection of intervals \[I_{k,m} = (a_{k,m},b_{k,m}) \subset [0,1],\] for $k \geq 0$ and $m \geq 1$, such that \begin{itemize} \item $0 < a_{0,1} < b_{0,1} < a_{1,1} < b_{1,1} < a_{2,1} < \cdots$, and $b_{k,1} < \frac{1}{8}$ for all $k \geq 0$; \item $I_{k,m+1} \subset I_{k,m}$; \item $\ell(I_{k,m}) = b_{k,m} - a_{k,m} \leq 2^{-m}$. \end{itemize} For $n \geq 3$, $m \geq 1$, $K_n^m$ has at least three vertices and no bridges. By our previous theorem, along with \hyperref[endpointderiv]{Lemma \ref*{endpointderiv}}, we can find a collection of reliability polynomials $s_{k,m}: [0,1] \to [0,1]$, satisfying the following properties. \begin{itemize} \item $s_{k,m}(0) = 1$ and $s_{k,m}(1) = 0$; \item $s_{k,m}'(0) = s_{k,m}'(1) = 0$, and $s_{k,m}'(q) \leq 0$ for all $q$; \item $|s_{k,m}''(q)| \leq 2^{-m-1}$ for $q \notin I_{k,m}$. \end{itemize} We now collect the consequences as a series of lemmas. \begin{lem} If $q \leq a_{k,m}$, then $|s_{k,m}'(q)| \leq 2^{-m-1}$ and $|1 - s_{k,m}(q)| \leq 2^{-m-1}$. If $q \geq b_{k,m}$, then $|s_{k,m}'(q)| \leq 2^{-m-1}$ and $|s_{k,m}(q)| \leq 2^{-m-1}$. \end{lem} \begin{proof} Consider the first statement; we suppose otherwise and apply the mean value theorem. That is, we suppose there is a $c \in [0,a_{k,m}]$ such that $|s_{k,m}'(c)| \geq 2^{-m-1}$. But this implies that there is a $d \in [0,c]$ with \[|s_{k,m}''(d)| = \frac{|s_{k,m}'(c) - s_{k,m}'(0)|}{|c - 0|} > 2^{-m-1},\] contradicting our assumption about the $s_{k,m}$. Similarly applying the mean value theorem again shows that if there was a $c \in [0,a_{k,m}]$ with $|s_{k,m}(c) - 1| \geq 2^{-m-1}$ there would be a $d$ at which the above bound is not satisfied, another contradiction. The arguments for $q \in [b_{k,m},1]$ are similar. \end{proof} Now we'd like to show how {\em large} the derivatives must be on the intervals $I_{k,m}$. In particular, we'd like to find a single point in each interval at which both the first and second derivatives are ``sufficiently large". \begin{lem} In each $I_{k,m}$, there is a point $q_{k,m}$ satisfying $s_{k,m}'(q_{k,m}) < -2^{m - 2}$ and $s_{k,m}''(q_{k,m}) > 2^{2m - 2}$. \end{lem} \begin{proof} We note that for every $k \geq 0$, $m \geq 1$, we have $s_{k,m}(a_{k,m}) \geq \frac{3}{4}$ and $s_{k,m}(b_{k,m}) \leq \frac{1}{4}$; in particular, the difference is at least $\frac{1}{2}$. Then by the mean value theorem, there is a $c \in I_{k,m}$ where \[|s_{k,m}'(c)| = \frac{|s_{k,m}(b_{k,m})-s_{k,m}(a_{k,m})|}{|b_{k,m}-a_{k,m}|} \geq \frac{2^{-1}}{2^{-m}} = 2^{m - 1}.\] Note that since $s_{k,m}'(q) \leq 0$ for all $q$, we have $s_{k,m}'(c) = -2^{m-1}$. Now let $d$ be the smallest number such that $d > c$ and $s_{k,m}'(d) = -2^{m - 2}$; applying the mean value theorem again shows that there is a $q_{k,m} \in [c,d] \subset I_{k,m}$ satisfying \[s_{k,m}''(q_{k,m}) = \frac{s_{k,m}'(d) - s_{k,m}'(c)}{d - c} \geq \frac{2^{m-2}}{2^{-m}} = 2^{2m - 2}.\] Note that by the definition of $d$, we must have $s_{k,m}'(q_{k,m}) \leq s_{k,m}'(d) = -2^{m - 2}$. \end{proof} Now we show that there exist products of $s_{k,m}$ (that is, reliability functions of one point unions of complete graphs) with arbitrarily many inflection points. This proof is modeled after the proof of the main result given in \cite{GM}. \begin{thm} There exist functions $g_k : [0,1] \to [0,1]$ for $k \geq 0$, such that $g_k = s_{k,m_k} g_{k-1}$ for $k \geq 1$, and $g_k$ has at least $2k$ inflection points. \end{thm} \begin{proof} We prove this by induction on $k$. We are going to show that there is a sequence of reliability polynomials $g_k$, integers $m_k$, and points $q_k \in I_{k,m_k}$, for $k \geq 0$, such that \begin{itemize} \item $g_k''(q_i) > 0$ for $i=0,...,k$; \item $\int_{q_{i-1}}^{q_i} g_k''(q) dq < 0$ for $i=1,...,k$; \item $g_k = s_{k,m_k} g_{k-1}$ if $k \geq 1$. \end{itemize} We begin with the base case $k=0$. We simply let $g_0(q) = s_{0,1}$, $q_0 = q_{0,1}$. By our lemma, $s_{0,1}''(q_{0,1}) > 0$, and so the first condition holds; the other conditions hold vacuously. Now suppose that we have found a $g_{k-1}$ satisfying the above properties, and we would like to find an $m_k$ such that $g_k = s_{k,m_k} g_{k-1}$ also satisfies them. We first consider \begin{equation} (s_{k,m}g_{k-1})' = s_{k,m}'g_{k-1} + s_{k,m}g_{k-1}'. \end{equation} For each $i < k$, $q_i < a_{k,1}$, so $\underset{m \to \infty}{\lim} s_{k,m}'(q_i) = 0$ and $\underset{m \to \infty}{\lim} s_{k,m}(q_i) = 1$. It follows that \[\underset{m \to \infty}{\lim} (s_{k,m}g_{k-1})'(q_i) = g_{k-1}'(q_i).\] Since \[\int_{q_{i-1}}^{q_i} (s_{k,m}g_{k-1})''(q) dq = (s_{k,m}g_{k-1})'(q_i) - (s_{k,m}g_{k-1})'(q_{i-1}),\] we have \[\lim_{m \to \infty} \int_{q_{i-1}}^{q_i} (s_{k,m}g_{k-1})''(q) dq = \int_{q_{i-1}}^{q_i} g_{k-1}''(q) dq < 0.\] Now we look at $(s_{k,m}g_{k-1})'(q_{k,m})$, noting that the precise location of $q_{k,m}$ depends on $m$. By construction, we have $s_{k,m}'(q_{k,m}) \leq -2^{m-2}$, but $|s_{k,m}(q_{k,m})| \leq 1$; thus, by taking $m$ sufficiently large, we can guarantee that $(s_{k,m}g_{k-1})'(q_{k,m}) - (s_{k,m}g_{k-1})'(q_{k-1}) < 0$. Next we need to consider \begin{equation} (s_{k,m}g_{k-1})'' = s_{k,m}''g_{k-1} + 2s_{k,m}'g_{k-1}' + s_{k,m}g_{k-1}''. \end{equation} For $q < a_{k,1}$, we have $\underset{m \to \infty}{\lim} (s_{k,m}g_{k-1})''(q) = g_{k-1}''(q)$, since $\underset{m \to \infty}{\lim} s_{k,m}''(q) = 0$, $\underset{m \to \infty}{\lim} s_{k,m}'(q) = 0$, and $\underset{m \to \infty}{\lim} s_{k,m}(q) = 1$. Thus by the induction hypothesis, for sufficiently large $m$ we have $(s_{k,m}g_{k-1})''(q_i) > 0$ for $i=0,...,k-1$. Now we need only consider $(s_{k,m}g_{k-1})''(q_{k,m})$. Since $g_{k-1}$ is a reliability polynomial, $g_{k-1}' \leq 0$, so the second term in the expansion of $(s_{k,m}g_{k-1})''$ is positive. Since $s_{k,m} < 1$, the third term is bounded by the maximum of $|g_{k-1}''|$ on [0,1]. Finally, since $g_{k-1}$ is bounded away from 0 on $I_{k,1}$ and $\underset{m \to \infty}{\lim} s_{k,m}''(q_{k,m}) = \infty$, it follows that $(s_{k,m}g_{k-1})''(q_{k,m}) > 0$ for sufficiently large $m$. If we let $m_k$ be sufficiently large such that all of the above constructions hold, and define $q_k = q_{k,m_k}$ and $g_k = s_{k,m_k} g_{k-1}$, then $g_k$ satisfies the induction hypothesis; thus we have constructed the desired $g_k$ for all $k \geq 0$. We know that $g_k''(q_{i-1}) > 0$ and $g_k''(q_i) > 0$ for each $i = 1, \dots ,k$; but $\int_{q_{i-1}}^{q_i} g_k'' dq < 0$ tells us that $g_k''$ is negative somewhere on $[q_{i-1},q_i]$, which implies that there are at least 2 inflection points on this interval. Thus $g_k$ has at least $2k$ inflection points, which was what we wanted to show. \end{proof} Note that the polynomial $g_k$ constructed in the above theorem is the product of the reliability polynomials of a number of complete graphs, and thus it is the reliability polynomial of the one-point union of those graphs. In Figure 1, the second derivatives of three such graphs are shown, demonstrating reliability polynomials of simple graphs having 3,4, and 5 inflection points. \begin{center} \begin{figure} \includegraphics[width=6cm]{c3ip.pdf} \includegraphics[width=6cm]{c4ip.pdf} \includegraphics[width=6cm]{c5ip.pdf} \caption{\label{Examples} The graphs of the second derivatives of $R(K_{2}^{5} * K_{3}^{4} * K_{4}^{3} * K_{5}^{92}), R(K_{2}^{2} * K_{4}^{2} * K_{14}^{750}),$ and $R(K_{2}^{5} * K_{3}^{4} * K_{5}^{116} * K_{14}^{100,000,000})$ respectively.} \end{figure} \end{center} \section{Acknowledgements} This research was conducted during the 2015 Research Experience for Undergraduates at University of Texas at Tyler under the direction of Dr. David Milan. We would like to thank the organizers of the UT Tyler REU for their guidance during this project. This REU was supported by the National Science Foundation (DMS-1359101). \bibliographystyle{amsplain}
{ "timestamp": "2016-02-11T02:02:54", "yymm": "1602", "arxiv_id": "1602.03246", "language": "en", "url": "https://arxiv.org/abs/1602.03246", "abstract": "In this paper we show that for each $n$, there exists a simple graph whose reliability polynomial has at least $n$ inflection points.", "subjects": "Combinatorics (math.CO)", "title": "Reliability Polynomials of Simple Graphs having Arbitrarily many Inflection Points", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9820137910906876, "lm_q2_score": 0.8152324871074608, "lm_q1q2_score": 0.8005695452846877 }
https://arxiv.org/abs/1007.4222
Strict inequality in the box-counting dimension product formulas
It is known that the upper box-counting dimension of a Cartesian product satisfies the inequality $\dim_{B}\left(F\times G\right)\leq \dim_{B}\left(F\right) + \dim_{B}\left(G\right)$ whilst the lower box-counting dimension satisfies the inequality $\dim_{LB}\left(F\times G\right)\geq \dim_{LB}\left(F\right) + \dim_{LB}\left(G\right)$. We construct Cantor-like sets to demonstrate that both of these inequalities can be strict.
\part{Use this type of header for very long papers only} \section{Preliminaries} \label{intro} \noindent In a metric space $X$ the Hausdorff dimension of a compact set $F\subset X$ is defined as the supremum of the $d\geq 0$ such that the $d$-dimensional Hausdorff measure $\mathcal{H}^{d}\left(F\right)$ of $F$ is infinite. The Hausdorff dimension takes values in the non-negative reals and extends the elementary integer-valued topological dimension in the sense that for a large class of `reasonable' sets these two values coincide. Sets with non-coinciding Hausdorff and topological dimensions are called `fractal', a term coined by Mandelbrot in his original study of such sets \cite{Mandelbrot75}. Hausdorff introduced this generalised dimension in \cite{Hausdorff18} and its subsequent extensive use in geometric measure theory is developed by Federer \cite{BkFederer69} and Falconer \cite{BkFalconer85}. The fact that the Hausdorff dimension satisfies $\dim_{H}\left(F\times G\right)\geq \dim_{H}\left(F\right) + \dim_{H}\left(G\right)$ for the Cartesian product of sets was proved in full generality in \cite{Marstrand53} (and later summarised in \cite{BkFalconer03} \S 7.1 `Product Formulae') after some partial results: The inequality was proved in \cite{BesicovitchMoran45} with the restriction that $0<\mathcal{H}^{s}\left(F\right),\mathcal{H}^{t}\left(G\right)<\infty$ for some $s,t$ and was extended to a larger class of sets in \cite{Eggleston53}. The paper \cite{BesicovitchMoran45} also provides an example for which there is a strict inequality in the product formula and again this is summarised in \cite{BkFalconer03} \S 7.1. \paragraph*{} In this paper we prove similar product inequalities for the upper and lower box-counting dimensions which are less familiar generalisations of dimension (treated briefly in \cite{BkFalconer03}; see \cite{BkRobinson10} for a more detailed exposition) and have applications to dynamical systems (see, for example, \cite{BkRobinson01}). Our main result is an example analogous to that in \cite{BesicovitchMoran45} which demonstrates that the box-counting product inequalities can be strict. In a metric space $X$ the upper and lower box-counting dimensions of a compact set $F\subset X$ are defined by \begin{align} \dim_{B}\left(F\right) &= \limsup_{\delta\searrow 0}\frac{\log\left(N\left(F,\delta\right)\right)}{-\log\delta}\label{dimB def} \intertext{and} \dim_{LB}\left(F\right) &= \liminf_{\delta\searrow 0}\frac{\log\left(N\left(F,\delta\right)\right)}{-\log\delta}\label{dimLB def} \end{align} respectively, where $N\left(F,\delta\right)$ is the smallest number of sets with diameter at most $\delta$ which form a cover (called a $\delta$-cover) of $F$. Essentially, if $N\left(F,\delta\right)$ scales like $\delta^{-\varepsilon}$ as $\delta\rightarrow 0$ then these quantities capture $\varepsilon$ which gives an indication of how many more sets are required to cover $F$ as the length-scales decrease and so encodes how `spread out' the set $F$ is at small length-scales. These limits are unchanged if we replace $N\left(F,\delta\right)$ with one of many similar quantities (discussed by Falconer in \cite{BkFalconer03} \S 3.1 `Equivalent Definitions'). Of these quantities we will also make use of the largest number of disjoint closed balls of diameter $\delta$ with centres in $F$, which we denote $M\left(F,\delta\right)$. \paragraph*{} We first recall the standard (see, for example, \cite{BkFalconer03} or \cite{BkRobinson10}) proof of the box-counting product inequalities when $F$ and $G$ are compact sets in metric spaces $X$ and $Y$ respectively, although the inequality \eqref{dimLB product inequality} is less familiar (Robinson provides a proof in \cite{BkRobinson10}). We endow the product space $X\times Y$ with the usual Euclidean metric $d_{X\times Y}=\sqrt{d_{X}^{2}+d_{Y}^{2}}$, but the proof can be adapted for a variety of product metrics (see \cite{BkRobinson10}). \begin{theorem}\label{product inequalities theorem} For compact sets $F\subset X$ and $G\subset Y$ the box-counting dimensions of the product set $F\times G$ satisfy the inequalities \begin{align} \dim_{B}\left(F\times G\right)&\leq \dim_{B}\left(F\right) + \dim_{B}\left(G\right)\label{dimB product inequality}\\ \dim_{LB}\left(F\times G\right)&\geq \dim_{LB}\left(F\right) + \dim_{LB}\left(G\right)\label{dimLB product inequality} \end{align} \end{theorem} \begin{proof} Suppose $\set{U_{i}}_{i=1}^{n_{1}}$ and $\set{V_{j}}_{j=1}^{n_{2}}$ are $\delta$-covers of $F$ and of $G$ respectively then the set of products $\set{U_{i}\times V_{j}\vert i=1\ldots n_{1}, j=1\ldots n_{2}}$ is a cover of $F\times G$ with a total of $n_{1}n_{2}$ elements and the diameter of each $U_{i}\times V_{j}$ is no greater than $\sqrt{2}\delta$. As there exist $\delta$-covers of $F$ and $G$ consisting of $N\left(F,\delta\right)$ and $N\left(G,\delta\right)$ elements respectively this construction gives a $\sqrt{2}\delta$-cover of $F\times G$ consisting of $N\left(F,\delta\right)N\left(G,\delta\right)$ elements, hence \begin{equation}\label{minimal cover inequality} N\left(F\times G,\sqrt{2}\delta\right) \leq N\left(F,\delta\right)N\left(G,\delta\right). \end{equation} Next, if both $\set{x_{i}}_{i=1}^{n_{1}}\subset F$ and $\set{y_{j}}_{j=1}^{n_{2}}\subset G$ are sets of centres of disjoint balls with diameter $\delta$ then the balls with radius $\delta$ centred on the $n_{1}n_{2}$ points $\set{\left(x_{i},y_{j}\right)\vert i=1\ldots n_{1}, j=1\ldots n_{2}}\subset F\times G$, are also disjoint. As there exist sets of disjoint balls of diameter $\delta$ with centres in $F$ and $G$ consisting of $M\left(F,\delta\right)$ and $M\left(G,\delta\right)$ elements respectively the above construction gives $M\left(F,\delta\right)M\left(G,\delta\right)$ disjoint balls of diameter $\delta$ with centres in $F\times G$, hence \begin{equation}\label{maximal disjoint inequality} M\left(F\times G,\delta\right) \geq M\left(F,\delta\right)M\left(G,\delta\right). \end{equation} From \eqref{dimB def} and \eqref{minimal cover inequality} we see that \begin{align} \dim_{B}\left(F\times G\right) &= \limsup_{\delta\searrow 0}\frac{\log\left(N\left(F\times G,\sqrt{2}\delta\right)\right)}{-\log\left(\sqrt{2}\delta\right)}\notag\\ &\leq \limsup_{\delta\searrow 0} \left[\frac{\log\left(N\left(F,\delta\right)\right)}{-\log\left(\sqrt{2}\delta\right)} + \frac{\log\left(N\left(G,\delta\right)\right)}{-\log\left(\sqrt{2}\delta\right)}\right]\label{limsup of sum}\\ &\leq \limsup_{\delta\searrow 0} \frac{\log\left(N\left(F,\delta\right)\right)}{-\log\delta -\log\sqrt{2}} + \limsup_{\delta\searrow 0} \frac{\log\left(N\left(G,\delta\right)\right)}{-\log\delta -\log\sqrt{2}}\notag\\ & = \dim_{B}\left(F\right) + \dim_{B}\left(G\right). \nota \intertext{From \eqref{dimLB def} and \eqref{maximal disjoint inequality} we have} \dim_{LB}\left(F\times G\right) &= \liminf_{\delta\searrow 0}\frac{\log\left(M\left(F\times G,\delta\right)\right)}{-\log\delta}\notag\\ &\geq \liminf_{\delta\searrow 0}\left[\frac{\log\left(M\left(F,\delta\right)\right)}{-\log\delta} + \frac{\log\left(M\left(G,\delta\right)\right)}{-\log\delta}\right]\label{liminf of sum}\\ &\geq \liminf_{\delta\searrow 0}\frac{\log\left(M\left(F,\delta\right)\right)}{-\log\delta} + \liminf_{\delta\searrow 0}\frac{\log\left(M\left(G,\delta\right)\right)}{-\log\delta} \notag \\ &= \dim_{LB}\left(F\right) + \dim_{LB}\left(G\right). \nota \end{align} \end{proof} It is known that there are sets with unequal upper and lower box-counting dimension (see exercise 3.8 of \cite{BkFalconer03} or \cite{BkRobinson10} \S 3.1), however if these values coincide for a set $F$ we define their common value as the box-counting dimension of $F$. If sets $F$ and $G$ have well-defined box-counting dimension then the box-counting dimension of their product is also well behaved. \begin{corollary} If $\dim_{B}\left(F\right)=\dim_{LB}\left(F\right)$ and $\dim_{B}\left(G\right)=\dim_{LB}\left(G\right)$ then \[\dim_{B}\left(F\times G\right) = \dim_{LB}\left(F\times G\right) = \dim_{B}\left(F\right)+\dim_{B}\left(G\right).\] \end{corollary} \begin{proof} From the inequalities \eqref{dimB product inequality} and \eqref{dimLB product inequality} we have \[ \dim_{B}\left(F\times G\right) \leq \dim_{B}\left(F\right) + \dim_{B}\left(G\right) = \dim_{LB}\left(F\right) + \dim_{LB}\left(G\right) \leq \dim_{LB}\left(F\times G\right) \] but from the definition of the box-counting dimension $\dim_{LB}\left(F\times G\right)\leq \dim_{B}\left(F\times G\right)$ so we must have equality throughout the above. \end{proof} \paragraph*{} In the following construction both sets $F$ and $G$ have non-coinciding upper and lower box-counting dimensions so that as $\delta\rightarrow 0$ the box-counting functions $\frac{\log\left(N\left(F,\delta\right)\right)}{-\log\delta}$ and $\frac{\log\left(N\left(G,\delta\right)\right)}{-\log\delta}$ oscillate between two values. Further, by ensuring that these functions oscillate with different phases (see figure \ref{figure - product graph}) we can produce strict inequalities after \eqref{limsup of sum} and \eqref{liminf of sum} and so yield strict inequality in both product formulas, that is \[ \dim_{LB}\left(F\times G\right) < \dim_{LB}\left(F\right) + \dim_{LB}\left(G\right) < \dim_{B}\left(F\right) + \dim_{B}\left(G\right) < \dim_{B}\left(F\times G\right). \] \begin{figure} \vspace*{8pt} \includegraphics[scale=0.5]{productgraph.eps} \vspace*{8pt} \caption{The box-counting functions for the sets $F$ and $G$ constructed in the next section (explicitly computed here for small $\delta$) oscillate between $\frac{\log\left(2\right)}{\log\left(3\right)}$ and $\frac{\log\left(2\right)}{\log\left(7\right)}$. The $x$-axis is scaled so that the slow oscillation can be graphed, and this oscillation continues as $\log\left(\log\left(-\log\left(\delta\right)\right)\right)\rightarrow \infty$, that is as $\delta\rightarrow 0$. The differing phases guarantee that the sum of these functions doesn't approach either $\frac{\log\left(2\right)}{\log\left(3\right)}+\frac{\log\left(2\right)}{\log\left(3\right)}$ or $\frac{\log\left(2\right)}{\log\left(7\right)}+\frac{\log\left(2\right)}{\log\left(7\right)}$.} \label{figure - product graph} \end{figure} To this end we construct variations of the Cantor middle-third set from the initial interval $\left[0,1\right]$ except at each stage we use one of the three generators $\gen_{3},\gen_{5}$ or $\gen_{7}$ which remove the middle $\frac{1}{3},\frac{3}{5}$ or $\frac{5}{7}$ of each interval respectively. Note that if we exclusively use the $\gen_{3}$ generator we produce the usual Cantor middle-third set, which has lower and upper box-counting dimension $\frac{\log\left(2\right)}{\log\left(3\right)}$ and if we exclusively use the $\gen_{7}$ generator we produce a similar Cantor set with lower and upper box-counting dimension $\frac{\log\left(2\right)}{\log\left(7\right)}$. By switching generators at certain stages of our construction we can cause $\frac{\log\left(N\left(F,\delta\right)\right)}{-\log\delta}$ to oscillate between these values, providing that we apply the generators a sufficiently large number of times. \paragraph*{} To simplify notation, we choose a sequence of integers $K_{j}=10^{2^{j}}$ which increases sufficiently quickly that \begin{align} \sum_{i=0}^{j} K_{i} &< K_{j+1} \label{Kj upper bound}\\ \sum_{i=0}^{j} K_{i} - K_{i-1} &>K_{j-1} \label{Kj lower bound}\\ \frac{\log\left(7\right)}{\log\left(3\right)}K_{j}&<K_{j+1}\label{Kj log3log7} \intertext{and} \frac{\sum_{i=0}^{j-1} K_{i}}{K_{j}} &\rightarrow 0 \quad \text{as $j\rightarrow \infty$} \label{Kj ratio limit} \end{align} \section{Constructing sets $F$ and $G$} We construct two sets $F$ and $G$ using the following iterative procedure: For $F$ apply the following generator at the $j^{\text{th}}$ stage \[ \begin{cases} \gen_{3} & K_{6n}<j\leq K_{6n+1}\quad \text{for some} \quad n\in\mathbb{N}\\ \gen_{7} &K_{6n+1}<j\leq K_{6n+2}\quad \text{for some} \quad n\in\mathbb{N}\\ \gen_{5} & \text{otherwise}, \end{cases} \] and for $G$ apply the following generator at the $j^{\text{th}}$ stage \[ \begin{cases} \gen_{3} & K_{6m+3}<j\leq K_{6m+4}\quad \text{for some} \quad m\in\mathbb{N}\\ \gen_{7} & K_{6m+4}<j\leq K_{6m+5}\quad \text{for some} \quad m\in\mathbb{N}\\ \gen_{5} & \text{otherwise}. \end{cases} \] Let $f_{3}\left(j\right)$ be the number of times $\gen_{3}$ has been applied and $f_{7}\left(j\right)$ the number of times $\gen_{7}$ has been applied in the construction $F$ by stage $j$. With this notation $\gen_{5}$ has been applied $j-f_{3}\left(j\right)-f_{7}\left(j\right)$ times by stage $j$. Similarly define $g_{3}\left(j\right)$ and $g_{7}\left(j\right)$ for the construction of $G$. Clearly these functions are non-decreasing and we refrain from writing them explicitly except to note that \begin{equation}\label{explicit f(Kj)} \begin{aligned} f_{3}\left(K_{6n+1}\right)&=\sum_{i=0}^{n} K_{6i+1} - K_{6i} & f_{7}\left(K_{6n+2}\right)&=\sum_{i=0}^{n} K_{6i+2} - K_{6i+1}\\ g_{3}\left(K_{6m+4}\right)&=\sum_{i=0}^{m} K_{6i+4} - K_{6i+3} & g_{7}\left(K_{6m+5}\right)&=\sum_{i=0}^{m} K_{6i+5} - K_{6i+4} \end{aligned} \end{equation} and \begin{equation}\label{f constant on intervals} \begin{split} f_{3}\left(j\right)=f_{3}\left(K_{6n+1}\right) \quad\text{for}\quad K_{6n+1}<j\leq K_{6n+6}\\ f_{7}\left(j\right)=f_{7}\left(K_{6n+2}\right) \quad\text{for}\quad K_{6n+2}<j\leq K_{6n+7}\\ g_{3}\left(j\right)=g_{3}\left(K_{6m+3}\right) \quad\text{for}\quad K_{6m+3}<j\leq K_{6m+8}\\ g_{7}\left(j\right)=g_{7}\left(K_{6m+3}\right) \quad\text{for}\quad K_{6m+4}<j\leq K_{6m+9} \end{split} \end{equation} Denote the sets at the $j^{th}$ stage of the construction of $F$ and $G$ by \begin{itemize} \item[] $F_{j}$, which consists of $2^{j}$ intervals of length $3^{-f_{3}\left(j\right)}7^{-f_{7}\left(j\right)}5^{-j+f_{3}\left(j\right)+f_{7}\left(j\right)}$ and \item[] $G_{j}$, which consists of $2^{j}$ intervals of length $3^{-g_{3}\left(j\right)}7^{-g_{7}\left(j\right)}5^{-j+g_{3}\left(j\right)+g_{7}\left(j\right)}$ \end{itemize} so that $F$ and $G$ are defined by $F=\bigcap F_{j}$ and $G=\bigcap G_{j}$. Note that for every $j$ the endpoints of the intervals in $F_{j}$ and $G_{j}$ are in $F$ and $G$ respectively as the generators only remove the middle of each interval. \begin{proposition}\label{N=M=2j prop} For $\delta$ such that \begin{equation}\label{length-scale for stage j} 3^{-f_{3}\left(j\right)}7^{-f_{7}\left(j\right)}5^{-j+f_{3}\left(j\right)+f_{7}\left(j\right)}\leq\delta< 3^{-f_{3}\left(j-1\right)}7^{-f_{7}\left(j-1\right)}5^{-\left(j-1\right)+f_{3}\left(j-1\right)+f_{7}\left(j-1\right)} \end{equation} we have $N\left(F,\delta\right)=M\left(F,\delta\right)=2^{j}$. \end{proposition} We refer to those $\delta$ in the range \eqref{length-scale for stage j} as length-scales corresponding to stage $j$ in the construction of $F$. Clearly every $1>\delta>0$ is a length scale corresponding to exactly one stage $j_{\delta}$ and $j_{\delta}\rightarrow\infty$ as $\delta\rightarrow 0$. We refer to length-scales corresponding to the construction of $G$ in an analogous fashion. \begin{proof} For $\delta$ in the range \eqref{length-scale for stage j} the obvious cover consisting of all intervals in $F_{j}$ gives $N\left(F,\delta\right)\leq 2^{j}$. The opposite inequality comes from the fact that a set with diameter $\delta$ in this range intersects at most one $\left(j-1\right)^{\text{th}}$ stage interval $I$ but cannot cover both $j^{\text{th}}$ stage subintervals of $I$ (see figure \ref{figure - 2jballsneeded}) so at least $2\times 2^{j-1}=2^{j}$ elements are needed to form a cover of $F$ .\\ Next, $\delta$ in the range \eqref{length-scale for stage j} is less than the length of the intervals in $F_{j-1}$ so balls of diameter $\delta$ centred on the end points of the intervals of $F_{j-1}$ are disjoint and have centres in $F$ (see figure \ref{figure - 2jdisjointballscover}). This gives two disjoint balls for each interval, hence $M\left(F,\delta\right)\geq 2\times 2^{j-1}=2^{j}$. For the opposite inequality suppose for a contradiction that $M\left(F,\delta\right)> 2^{j}$, then at least one of the $2^{j-1}$ intervals in $F_{j-1}$ contains the centres of least three disjoint balls with centres in $F$. Let $I$ be such an interval. At the next stage of the construction $I$ is split into two sub-intervals, one of which contains the centres of at least two of these three disjoint balls. However, this $j^{\text{th}}$ stage subinterval has length no greater than $\delta$ so two closed balls of diameter $\delta$ centred in this interval cannot be disjoint (see figure \ref{figure - 2jdisjointballsmax}), which is a contradiction. \end{proof} \begin{figure} \vspace*{8pt \includegraphics[scale=0.7]{2jballsneeded.eps} \vspace*{8pt} \caption{A section of the sets $F_{j-1}$ and $F_{j}$ (shown in black) to illustrate that each set (shown as grey ellipses) with diameter $\delta<3^{-f_{3}\left(j-1\right)}7^{-f_{7}\left(j-1\right)}5^{-\left(j-1\right)+f_{3}\left(j-1\right)+f_{7}\left(j-1\right)}$ (i.e. less than the length of the intervals of $F_{j-1}$) can neither intersect two intervals of $F_{j-1}$ nor cover two intervals of $F_{j}$.} \label{figure - 2jballsneeded} \end{figure} \begin{figure} \vspace*{8pt \includegraphics[scale=0.7]{2jdisjointballscover.eps} \vspace*{8pt} \caption{A section of the set $F_{j-1}$ (shown in black) to illustrate that balls (shown in grey) with diameter $\delta<3^{-f_{3}\left(j-1\right)}7^{-f_{7}\left(j-1\right)}5^{-\left(j-1\right)+f_{3}\left(j-1\right)+f_{7}\left(j-1\right)}$ (i.e. less than the length of the intervals of $F_{j-1}$) with centres the endpoints of the intervals $F_{j-1}$ are disjoint, giving two disjoint balls for each interval of $F_{j-1}$.} \label{figure - 2jdisjointballscover} \end{figure} \begin{figure} \vspace*{8pt \includegraphics[scale=0.7]{2jdisjointballsmax.eps} \vspace*{8pt} \caption{A sub-interval $I\subset F_{j-1}$ (shown in black) which contains the centres of three balls (shown in grey) with diameter $\delta\geq 3^{-f_{3}\left(j\right)}7^{-f_{7}\left(j\right)}5^{-j+f_{3}\left(j\right)+f_{7}\left(j\right)}$ (i.e. greater than the the length of the intervals of $F_{j}$) with centres in $F$. As the centres are also contained in $F_{j}$ (also shown in black) at least one interval of $F_{j}$ contains the centres of two of these balls. Consequently, the distance between the centres is at most the length of an interval of $F_{j}$ which is at most $\delta$ which is the sum of the radii of the closed balls so the two balls are not disjoint.} \label{figure - 2jdisjointballsmax} \end{figure} Adapting the argument we can prove a similar proposition for the set $G$. \begin{proposition}\label{N=M=2j prop G} For $\delta$ such that \begin{equation} 3^{-g_{3}\left(j\right)}7^{-g_{7}\left(j\right)}5^{-j+g_{3}\left(j\right)+g_{7}\left(j\right)}\leq\delta< 3^{-g_{3}\left(j-1\right)}7^{-g_{7}\left(j-1\right)}5^{-\left(j-1\right)+g_{3}\left(j-1\right)+g_{7}\left(j-1\right)} \end{equation} we have $N\left(G,\delta\right)=M\left(G,\delta\right)=2^{j}$. \end{proposition} By taking logarithms we obtain the following more useful form of propositions \ref{N=M=2j prop} and \ref{N=M=2j prop G} \begin{multline}\label{delta implies N(F,delta)} f_{3}\left(j-1\right)\left[\log\left(3\right)-\log\left(5\right)\right]+f_{7}\left(j-1\right)\left[\log\left(7\right)-\log\left(5\right)\right] + \left(j-1\right)\log\left(5\right)\\ < -\log\delta \leq f_{3}\left(j\right)\left[\log\left(3\right)-\log\left(5\right)\right]+f_{7}\left(j\right)\left[\log\left(7\right)-\log\left(5\right)\right] + j\log\left(5\right)\\ \Rightarrow \log\left(N\left(F,\delta\right)\right) = \log\left(M\left(F,\delta\right)\right) = j\log\left(2\right) \end{multline} and \begin{multline}\label{delta implies N(G,delta)} g_{3}\left(j-1\right)\left[\log\left(3\right)-\log\left(5\right)\right]+g_{7}\left(j-1\right)\left[\log\left(7\right)-\log\left(5\right)\right] + \left(j-1\right)\log\left(5\right)\\ < -\log\delta \leq g_{3}\left(j\right)\left[\log\left(3\right)-\log\left(5\right)\right]+g_{7}\left(j\right)\left[\log\left(7\right)-\log\left(5\right)\right] + j\log\left(5\right)\\ \Rightarrow \log\left(N\left(G,\delta\right)\right) = \log\left(M\left(G,\delta\right)\right) = j\log\left(2\right) \end{multline} \paragraph*{} The essential feature of our construction is that the sets $F$ and $G$ at some length-scales look like the Cantor middle-third set, while at other length-scales look like the Cantor middle-$\frac{5}{7}^{\text{th}}$ set. Further, this `local' behaviour is maintained over sufficient length-scales that the box-counting limits of $F$ and $G$ at these length-scales approach the box-counting dimensions of the relevant Cantor set, which we will establish in the following section. We conclude this section by proving that at any length-scale the sets $F$ and $G$ do not both look like the middle-third set nor do they both look like the middle-$\frac{5}{7}^{\text{th}}$ set. \begin{lemma}\label{delta not in both length-scales upper} No $\delta$ is both a length-scale corresponding to some stage in $\left(K_{6n},K_{6n+2}\right]$ in the construction of $F$ for some $n\in\mathbb{N}$ and a length-scale corresponding to some stage in $\left(K_{6m+3},K_{6m+5}\right]$ in the construction of $G$ for some $m\in\mathbb{N}$. \end{lemma} \begin{proof} Assume for a contradiction that $\delta$ is such a length-scale, that is \begin{multline}\label{fg inequality 1} 3^{-f_{3}\left(K_{6n+2}\right)}7^{-f_{7}\left(K_{6n+2}\right)}5^{-K_{6n+2}+f_{3}\left(K_{6n+2}\right)+f_{7}\left(K_{6n+2}\right)}\\ \leq \delta < 3^{-f_{3}\left(K_{6n}\right)}7^{-f_{7}\left(K_{6n}\right)}5^{-K_{6n}+f_{3}\left(K_{6n}\right)+f_{7}\left(K_{6n}\right)} \end{multline} and \begin{multline}\label{fg inequality 2} 3^{-g_{3}\left(K_{6m+5}\right)}7^{-g_{7}\left(K_{6m+5}\right)}5^{-K_{6m+5}+g_{3}\left(K_{6m+5}\right)+g_{7}\left(K_{6m+5}\right)}\\ \leq \delta < 3^{-g_{3}\left(K_{6m+3}\right)}7^{-g_{7}\left(K_{6m+3}\right)}5^{-K_{6m+3}+g_{3}\left(K_{6m+3}\right)+g_{7}\left(K_{6m+3}\right)} \end{multline} We first demonstrate that this could only hold if $n=m$. From \eqref{fg inequality 1} and \eqref{fg inequality 2} we have $7^{-K_{6n+2}} \leq \delta < 3^{-K_{6n}}$ and $7^{-K_{6m+5}} \leq \delta < 3^{-K_{6m}}$ respectively, which in turn yield $7^{-K_{6n+2}} < 3^{-K_{6m}}$ and $7^{-K_{6m+5}} < 3^{-K_{6n}}$. Taking logarithms we get \[ K_{6m} < K_{6n+2}\frac{\log\left(7\right)}{\log\left(3\right)}< K_{6n+3}\quad\text{and}\quad K_{6n} < K_{6m+5}\frac{\log\left(7\right)}{\log\left(3\right)}< K_{6m+6} \] where the final inequalities follow from the growth property \eqref{Kj log3log7}. We conclude that $6m< 6n+3$ and $6n< 6m+6$, which implies that $m=n$. With this restriction, if $\delta$ is such a length-scale then the lower bound from \eqref{fg inequality 1} and the upper bound from \eqref{fg inequality 2} imply \begin{multline*} 3^{-f_{3}\left(K_{6n+2}\right)}7^{-f_{7}\left(K_{6n+2}\right)}5^{-K_{6n+2}+f_{3}\left(K_{6n+2}\right)+f_{7}\left(K_{6n+2}\right)} \\ < 3^{-g_{3}\left(K_{6n+3}\right)}7^{-g_{7}\left(K_{6n+3}\right)}5^{-K_{6n+3}+g_{3}\left(K_{6n+3}\right)+g_{7}\left(K_{6n+3}\right)} \end{multline*} For clarity we suppress the argument of the functions and take logarithms which yields \begin{equation}\label{supressed argument condition} \left[\log\left(5\right)-\log\left(3\right)\right]\left(f_{3}-g_{3}\right) + \left[\log\left(5\right)-\log\left(7\right)\right]\left(f_{7}-g_{7}\right) + \log\left(5\right)\left(K_{6n+3}-K_{6n+2}\right) < 0. \end{equation} The bounds in \eqref{Kj upper bound} and \eqref{Kj lower bound} give \begin{align} g_{3}\left(K_{6n+3}\right)-f_{3}\left(K_{6n+2}\right) &= \sum_{i=0}^{n-1} K_{6i+4} - K_{6i+3} - \sum_{i=0}^{n} K_{6i+1}-K_{6i}\notag\\ &< K_{6n-1} - K_{6n} < 0 \label{g3-f3} \intertext{and} f_{7}\left(K_{6n+2}\right)-g_{7}\left(K_{6n+3}\right) &= \sum_{i=0}^{n} K_{6i+2} - K_{6i+1} - \sum_{i=0}^{n-1} K_{6i+5} - K_{6i+4}\notag\\ &< K_{6n+3} - K_{6n-2}. \label{f7-g7} \end{align} Consequently we drop the positive first term of \eqref{supressed argument condition} which implies \begin{align*} \left[\log\left(5\right)-\log\left(7\right)\right]\left(K_{6n+3} - K_{6n-2}\right) + \log\left(5\right)\left(K_{6n+3}-K_{6n+2}\right) &< 0 \intertext{which is that} \left[2\log\left(5\right)-\log\left(7\right)\right]K_{6n+3} + \left[\log\left(7\right)-\log\left(5\right)\right]K_{6n-2} - \log\left(5\right)K_{6n+2} &< 0 \end{align*} After dropping the positive middle term and rearranging we get \begin{align*} K_{6n+3}&< \frac{\log\left(5\right)}{\left[2\log\left(5\right)-\log\left(7\right)\right]}K_{6n+2} \intertext{so that} K_{6n+3}&< \frac{\log\left(7\right)}{\log\left(3\right)}K_{6n+2} \end{align*} however, by condition \eqref{Kj log3log7} the right hand side is less than $K_{6n+3}$ giving the required contradiction. \end{proof} Consequently, at any length-scale the sets $F$ and $G$ do not both look like the Cantor middle-third set. This lemma gives the following useful corollary: \begin{corollary}\label{subsequence corollary upper} Every sequence $\set{\delta_{i}}$ with $\delta_{i}\rightarrow 0$ either contains a subsequence $\set{\delta_{i_{n}}}$ with each $\delta_{i_{n}}$ corresponding to some stage $j_{\delta_{i_{n}}}\in \left(K_{6n+2},K_{6n+6}\right]$ in the construction of $F$ or contains a subsequence $\set{\delta_{i_{m}}}$ corresponding to some stage $j_{\delta_{i_{m}}}\in \left(K_{6m+5},K_{6m+9}\right]$ in the construction of $G$. \end{corollary} \begin{proof} If the sequence $\set{\delta_{i}}$ did not contain such a subsequence, then there is a $\Delta>0$ such that each $\delta_{i}<\Delta$ is neither a length-scale corresponding to some stage $j\in\left(K_{6n+2},K_{6n+6}\right]$ in the construction of $F$ nor a length-scale corresponding to some stage $j\in \left(K_{6m+5},K_{6m+9}\right]$ in the construction of $G$. Consequently, $\delta_{i}$ is a length scale corresponding to a stage $j\in\left(K_{6n},K_{6n+2}\right]$ in the construction of $F$ and also a length-scale corresponding to a stage $j\in\left(K_{6m+3},K_{6m+5}\right]$ which, from lemma \ref{delta not in both length-scales upper}, is contradictory. \end{proof} The corresponding lemma that $F$ and $G$ do not both look like the Cantor middle-$\frac{5}{7}^{\text{th}}$ set at any length-scale and the corresponding corollary for subsequences are proved in a similar way. \begin{lemma}\label{delta not in both length-scales lower} No $\delta$ is both a length-scale corresponding to some stage in $\left(K_{6n+1},K_{6n+3}\right]$ in the construction of $F$ for some $n\in\mathbb{N}$ and a length-scale corresponding to some stage in $\left(K_{6m+4},K_{6m+6}\right]$ in the construction of $G$ for some $m\in\mathbb{N}$, \end{lemma} \begin{corollary}\label{subsequence corollary lower} Every sequence $\set{\delta_{i}}$ with $\delta_{i}\rightarrow 0$ either contains a subsequence $\set{\delta_{i_{n}}}$ with each $\delta_{i_{n}}$ corresponding to some stage $j_{\delta_{i_{n}}}\in \left(K_{6n+3},K_{6n+7}\right]$ in the construction of $F$ or contains a subsequence $\set{\delta_{i_{m}}}$ corresponding to some stage $j_{\delta_{i_{m}}}\in \left(K_{6m},K_{6m+4}\right]$ in the construction of $G$. \end{corollary} \section{Calculating box-counting dimensions} In order to establish the box-counting dimensions of $F$ and $G$ we need the following proposition on the behaviour of the generator-counting functions at the limit: \begin{proposition}\label{f Kj ratio limit} \[ \frac{f_{3}\left(K_{6n+l}\right)}{K_{6n+l}} \rightarrow \begin{cases} 1 & l=1\\ 0 & 0\leq l < 6,\; l\neq 1 \end{cases}\quad \frac{f_{7}\left(K_{6n+l}\right)}{K_{6n+l}} \rightarrow \begin{cases} 1 & l=2\\ 0 & 0\leq l < 6,\; l\neq 2 \end{cases} \] \[ \frac{g_{3}\left(K_{6m+l}\right)}{K_{6m+l}} \rightarrow \begin{cases} 1 & l=4\\ 0 & 0\leq l < 6,\; l\neq 4 \end{cases}\quad \frac{g_{7}\left(K_{6m+l}\right)}{K_{6m+l}} \rightarrow \begin{cases} 1 & l=5\\ 0 & 0\leq l < 6,\; l\neq 5 \end{cases} \] as $n,m\rightarrow \infty$. \end{proposition} \begin{proof} From \eqref{explicit f(Kj)} we have \[ \frac{f_{3}\left(K_{6n+1}\right)}{K_{6n+1}} = \frac{\sum_{i=0}^{n} K_{6i+1} - K_{6i}}{K_{6n+1}} = 1 + \frac{\sum_{i=0}^{n-1} K_{6i+1}}{K_{6n+1}} - \frac{\sum_{i=0}^{n} K_{6i}}{K_{6n+1}} \] which converges to 1 as $n\rightarrow \infty$ by \eqref{Kj ratio limit}. By \eqref{f constant on intervals} we have $f_{3}\left(K_{6n+l}\right)=f_{3}\left(K_{6n+1}\right)$ for $2\leq l<6$ so \[ \frac{f_{3}\left(K_{6n+l}\right)}{K_{6n+l}} = \frac{f_{3}\left(K_{6n+1}\right)}{K_{6n+l}} = \frac{\sum_{i=0}^{n} K_{6i+1} - K_{6i}}{K_{6n+l}} \] which converges to 0 as $n\rightarrow \infty$ by \eqref{Kj ratio limit}. Similarly, $f_{3}\left(K_{6n+0}\right) = f_{3}\left(K_{6\left(n-1\right)+1}\right)$ so \[ \frac{f_{3}\left(K_{6n+0}\right)}{K_{6n+0}} = \frac{f_{3}\left(K_{6\left(n-1\right)+1}\right)}{K_{6n}} = \frac{\sum_{i=0}^{n-1} K_{6i+1} - K_{6i}}{K_{6n}} \rightarrow 0 \] The remaining results follow analogously. \end{proof} \begin{lemma}\label{upper box lemma} $\dim_{B}\left(F\right)=\dim_{B}\left(G\right)=\frac{\log\left(2\right)}{\log\left(3\right)}.$ \end{lemma} \begin{proof} First, we show that $\dim_{B}\left(F\right)\leq \frac{\log\left(2\right)}{\log\left(3\right)}$. Writing $j_{\delta}$ for the stage associated with the length-scale $\delta$ we have from \eqref{delta implies N(F,delta)} \begin{align*} \frac{\log\left(N\left(F,\delta\right)\right)}{-\log \delta}&< \frac{j_{\delta}\log\left(2\right)}{f_{3}\left(j_{\delta}-1\right)\left[\log\left(3\right)-\log\left(5\right)\right] + f_{7}\left(j_{\delta}-1\right)\left[\log\left(7\right)-\log\left(5\right)\right] + \left(j_{\delta}-1\right)\log\left(5\right)}\\ &\leq \frac{j_{\delta}\log\left(2\right)}{f_{3}\left(j_{\delta}-1\right)\left[\log\left(3\right)-\log\left(5\right)\right] + \left(j_{\delta}-1\right)\log\left(5\right)} \intertext{and as $f_{3}\left(j_{\delta}-1\right)< j_{\delta}-1$} &< \frac{j_{\delta}\log\left(2\right)}{\left(j_{\delta}-1\right)\left[\log\left(3\right)-\log\left(5\right)\right] + \left(j_{\delta}-1\right)\log\left(5\right)} = \frac{j_{\delta}\log\left(2\right)}{\left(j_{\delta}-1\right)\log\left(3\right)}\\ \end{align*} which converges to $\frac{\log\left(2\right)}{\log\left(3\right)}$ as $\delta\rightarrow 0$. Similarly we can show $\dim_{B}\left(G\right)\leq\frac{\log\left(2\right)}{\log\left(3\right)}$ \paragraph*{} Next, if we take the sequence $\set{\delta_{n}}$ with \[ -\log\delta_{n} = f_{3}\left(K_{6n+1}\right)\left[\log\left(3\right)-\log\left(5\right)\right]+ f_{7}\left(K_{6n+1}\right)\left[\log\left(7\right)-\log\left(5\right)\right]+ K_{6n+1}\log\left(5\right) \] we have from \eqref{delta implies N(F,delta)} that $\log\left(N\left(F,\delta_{n}\right)\right)= K_{6n+1}\log\left(2\right)$. Consequently, \begin{align*} \frac{\log\left(N\left(F,\delta_{n}\right)\right)}{-\log \delta_{n}} &= \frac{K_{6n+1}\log\left(2\right)}{f_{3}\left(K_{6n+1}\right)\left[\log\left(3\right)-\log\left(5\right)\right]+ f_{7}\left(K_{6n+1}\right)\left[\log\left(7\right)-\log\left(5\right)\right]+ K_{6n+1}\log\left(5\right)}\\ &= \frac{\log\left(2\right)}{\frac{f_{3}\left(K_{6n+1}\right)}{K_{6n+1}}\left[\log\left(3\right)-\log\left(5\right)\right]+\frac{f_{7}\left(K_{6n+1}\right)}{K_{6n+1}}\left[\log\left(7\right)-\log\left(5\right)\right]+ \log\left(5\right)}\\ &\rightarrow \frac{\log\left(2\right)}{\log\left(3\right)-\log\left(5\right) + 0 + \log\left(5\right)} = \frac{\log\left(2\right)}{\log\left(3\right)} \end{align*} as $n\rightarrow\infty$ by the convergence results \eqref{f Kj ratio limit} so that $\dim_{B}\left(F\right)\geq \frac{\log\left(2\right)}{\log\left(3\right)}$\\ A similar sequence gives the corresponding inequality for $G$. \end{proof} \begin{lemma}\label{lower box lemma} $\dim_{LB}\left(F\right)=\dim_{LB}\left(G\right)=\frac{\log\left(2\right)}{\log\left(7\right)}.$ \end{lemma} \begin{proof} For all $\delta>0$ the implication \eqref{delta implies N(F,delta)} gives \begin{align*} \frac{\log\left(N\left(F,\delta\right)\right)}{-\log \delta} &\geq \frac{j_{\delta}\log\left(2\right)}{ f_{3}\left(j_{\delta}\right)\left[\log\left(3\right)-\log\left(5\right)\right]+ f_{7}\left(j_{\delta}\right)\left[\log\left(7\right)-\log\left(5\right)\right]+j_{\delta}\log\left(5\right)}\\ &\geq \frac{j_{\delta}\log\left(2\right)}{ f_{7}\left(j_{\delta}\right)\left[\log\left(7\right)-\log\left(5\right)\right]+j_{\delta}\log\left(5\right)}\\ &> \frac{j_{\delta}\log\left(2\right)}{ j_{\delta}\left[\log\left(7\right)-\log\left(5\right)\right]+j_{\delta} \log\left(5\right)}=\frac{\log\left(2\right)}{\log\left(7\right)} \end{align*} Next, if we take the sequence $\set{\delta_{n}}$ with \[ -\log\delta_{n} = f_{3}\left(K_{6n+2}\right)\left[\log\left(3\right)-\log\left(5\right)\right]+ f_{7}\left(K_{6n+2}\right)\left[\log\left(7\right)-\log\left(5\right)\right]+ K_{6n+2}\log\left(5\right) \] we have \begin{align*} \frac{\log\left(N\left(F,\delta_{n}\right)\right)}{-\log \delta_{n}} &= \frac{K_{6n+2}\log\left(2\right)}{ f_{3}\left(K_{6n+2}\right)\left[\log\left(3\right)-\log\left(5\right)\right]+ f_{7}\left(K_{6n+2}\right)\left[\log\left(7\right)-\log\left(5\right)\right]+ K_{6n+2}\log\left(5\right)}\\ &= \frac{\log\left(2\right)}{\frac{f_{3}\left(K_{6n+2}\right)}{K_{6n+2}}\left[\log\left(3\right)-\log\left(5\right)\right]+\frac{f_{7}\left(K_{6n+2}\right)}{K_{6n+2}}\left[\log\left(7\right)-\log\left(5\right)\right]+ \log\left(5\right)}\\ &\rightarrow \frac{\log\left(2\right)}{0 + \log\left(7\right)-\log\left(5\right) + \log\left(5\right)} = \frac{\log\left(2\right)}{\log\left(7\right)} \end{align*} as $n\rightarrow\infty$ by the convergence results \eqref{f Kj ratio limit}. Hence $\dim_{LB}\left(F\right)=\frac{\log\left(2\right)}{\log\left(7\right)}$, and similarly $\dim_{LB}\left(G\right)=\frac{\log\left(2\right)}{\log\left(7\right)}$. \end{proof} Consequently, both $F$ and $G$ have unequal upper and lower box-counting dimensions: \begin{corollary} $\dim_{LB}\left(F\right)=\dim_{LB}\left(G\right) < \dim_{B}\left(F\right)=\dim_{B}\left(G\right)$. \end{corollary} \paragraph*{} Whilst the above lemmas demonstrate that for $F$ and $G$ there are sequences of length-scales $\set{\delta_{n}}$ with $\lim_{n\rightarrow\infty}\frac{\log\left(N\left(F,\delta_{n}\right)\right)}{-\log\delta_{n}}$ equal to $\frac{\log\left(2\right)}{\log\left(3\right)}$ or equal to $\frac{\log\left(2\right)}{\log\left(7\right)}$ we now show that for a large class of sequences (in fact the very sequences that corollaries \ref{subsequence corollary upper} and \ref{subsequence corollary lower} produce) this limit, if it exists, is bounded by $\frac{\log\left(2\right)}{\log\left(5\right)}$. \begin{lemma}\label{deltan N(F,deltan) converge to log2log5} Suppose $\set{\delta_{n}}$ is a sequence such that each length-scale $\delta_{n}$ corresponds to the construction of $F$ at some stage $j_{n}\in \left(K_{6n+2},K_{6n+6}\right]$, then if the limit exists \[\lim_{n\rightarrow\infty}\frac{\log\left(N\left(F,\delta_{n}\right)\right)}{-\log\delta_{n}}\leq\frac{\log\left(2\right)}{\log\left(5\right)}. \] \end{lemma} Essentially, these stages are sufficiently far from the range $\left(K_{6n},K_{6n+1}\right]$ where $\gen_{3}$ is applied so that the set $F$ does not look like the Cantor middle-third set at these stages. The proof relies on the fact that by stage $j_{n}$ the generator $\gen_{3}$ has not been applied for at least the last $K_{6n+2}-K_{6n+1}$ stages. \begin{proof} For each $n\in\mathbb{N}$ from \eqref{delta implies N(F,delta)} \begin{align*} \frac{\log\left(N\left(F,\delta_{n}\right)\right)}{-\log\delta_{n}} &\leq \frac{j_{n}\log\left(2\right)}{f_{3}\left(j_{n}-1\right)\left[\log\left(3\right)-\log\left(5\right)\right]+f_{7}\left(j_{n}-1\right)\left[\log\left(7\right)-\log\left(5\right)\right] + \left(j_{n}-1\right)\log\left(5\right)}\\ &\leq \frac{j_{n}\log\left(2\right)}{f_{3}\left(j_{n}-1\right)\left[\log\left(3\right)-\log\left(5\right)\right]+ \left(j_{n}-1\right)\log\left(5\right)} \intertext{From \eqref{f constant on intervals} we have $f_{3}\left(j_{n}-1\right)=f_{3}\left(K_{6n+1}\right)$ so that} &=\frac{j_{n}\log\left(2\right)}{f_{3}\left(K_{6n+1}\right)\left[\log\left(3\right)-\log\left(5\right)\right]+ \left(j_{n}-1\right)\log\left(5\right)}\\ &=\frac{\log\left(2\right)}{\frac{f\left(K_{6n+1}\right)}{j_{n}}\left[\log\left(3\right)-\log\left(5\right)\right]+\log\left(5\right)-\frac{1}{j_{n}}\log\left(5\right)}. \intertext{Next, as $j_{n}>K_{6n+2}$} &\leq \frac{\log\left(2\right)}{\frac{f_{3}\left(K_{6n+1}\right)}{K_{6n+2}}\left[\log\left(3\right)-\log\left(5\right)\right]+\log\left(5\right)-\frac{1}{j_{n}}\log\left(5\right)}\\ &\rightarrow \frac{\log\left(2\right)}{\log\left(5\right)} \end{align*} as $n\rightarrow\infty$ by the convergence result \eqref{f Kj ratio limit}. \end{proof} The corresponding result for $G$, proved in a similar way, is as follows. \begin{lemma}\label{deltan N(G,deltan) converge to log2log5} Suppose $\set{\delta_{m}}$ is a sequence such that each length-scale $\delta_{m}$ corresponds to the construction of $G$ at some stage $j_{m}\in \left(K_{6m+5},K_{6m+9}\right]$, then if the limit exists \[\lim_{m\rightarrow\infty}\frac{\log\left(N\left(G,\delta_{m}\right)\right)}{-\log\delta_{m}}\leq\frac{\log\left(2\right)}{\log\left(5\right)}. \] \end{lemma} The following results for lower bounds are also proved similarly. \begin{lemma}\label{deltan M(F,deltan) converge to log2log5} Suppose $\set{\delta_{n}}$ is a sequence such that each length-scale $\delta_{n}$ corresponds to the construction of $F$ at some stage $j_{n}\in \left(K_{6n+3},K_{6n+7}\right]$, then if the limit exists \[\lim_{n\rightarrow\infty}\frac{\log\left(M\left(F,\delta_{n}\right)\right)}{-\log\delta_{n}}\geq\frac{\log\left(2\right)}{\log\left(5\right)} \] \end{lemma} and \begin{lemma}\label{deltan M(G,deltan) converge to log2log5} Suppose $\set{\delta_{m}}$ is a sequence such that each length-scale $\delta_{m}$ corresponds to the construction of $G$ at some stage $j_{m}\in \left(K_{6m},K_{6m+4}\right]$, then if the limit exists \[\lim_{m\rightarrow\infty}\frac{\log\left(M\left(G,\delta_{m}\right)\right)}{-\log\delta_{m}}\geq\frac{\log\left(2\right)}{\log\left(5\right)}. \] \end{lemma} Finally, we find a bound on the box-counting dimensions of the product $F\times G$. \begin{theorem}\label{strict upper box product inequality} $\dim_{B}\left(F\times G\right)\leq \frac{\log\left(2\right)}{\log\left(3\right)} + \frac{\log\left(2\right)}{\log\left(5\right)}.$ \end{theorem} \begin{proof} We have from \eqref{limsup of sum} that \[ \limsup_{\delta\rightarrow 0} \frac{\log\left(N\left(F\times G,\delta\right)\right)}{-\log \delta} \leq \limsup_{\delta\rightarrow 0}\left[\frac{\log\left(N\left(F,\delta\right)\right)}{-\log\delta} + \frac{\log\left(N\left(G,\delta\right)\right)}{-\log\delta}\right] \] so it is sufficient to show that the right hand side is no greater than $\frac{\log\left(2\right)}{\log\left(3\right)} + \frac{\log\left(2\right)}{\log\left(5\right)}$. Suppose that $\set{\delta_{i}}$ is a sequence with $\delta_{i}\rightarrow 0$ such that the limits $\lim_{i\rightarrow \infty}\frac{\log\left(N\left(F,\delta_{i}\right)\right)}{-\log \delta_{i}}$ and $\lim_{i\rightarrow \infty}\frac{\log\left(N\left(G,\delta_{i}\right)\right)}{-\log \delta_{i}}$ exist. Corollary \ref{subsequence corollary upper} guarantees that this sequence either contains a subsequence $\set{\delta_{i_{n}}}$ satisfying the hypothesis of lemma \ref{deltan N(F,deltan) converge to log2log5} or contains a subsequence $\set{\delta_{i_{m}}}$ satisfying the hypothesis of lemma \ref{deltan N(G,deltan) converge to log2log5} so at least one of $\frac{\log\left(N\left(F,\delta_{i_{n}}\right)\right)}{-\log \delta_{i_{n}}}$ and $\frac{\log\left(N\left(G,\delta_{i_{n}}\right)\right)}{-\log \delta_{i_{n}}}$ converges to $\frac{\log\left(2\right)}{\log\left(5\right)}$. Using the upper box-counting dimension from lemma \ref{upper box lemma} to bind the other term yields \[ \lim_{n\rightarrow \infty} \frac{\log\left(N\left(F,\delta_{i_{n}}\right)\right)}{-\log \delta_{i_{n}}} + \frac{\log\left(N\left(G,\delta_{i_{n}}\right)\right)}{-\log \delta_{i_{n}}} \leq \frac{\log\left(2\right)}{\log\left(3\right)} + \frac{\log\left(2\right)}{\log\left(5\right)} \] a bound which also hold for the original sequence $\set{\delta_{i}}$. As $\set{\delta_{i}}$ was an arbitrary convergent sequence, \[ \limsup_{\delta\rightarrow 0}\left[\frac{\log\left(N\left(F,\delta\right)\right)}{-\log\delta} + \frac{\log\left(N\left(G,\delta\right)\right)}{-\log\delta}\right] \leq \frac{\log\left(2\right)}{\log\left(3\right)} + \frac{\log\left(2\right)}{\log\left(5\right)} \] \end{proof} \begin{corollary} $\dim_{B}\left(F\times G\right)< \dim_{B}\left(F\right) + \dim_{B}\left(G\right)$ \end{corollary} \begin{theorem} $\dim_{LB}\left(F\times G\right)\geq \frac{\log\left(2\right)}{\log\left(7\right)} + \frac{\log\left(2\right)}{\log\left(5\right)}.$ \end{theorem} \begin{proof} From \eqref{liminf of sum} we have \[ \liminf_{\delta\rightarrow 0} \frac{\log\left(M\left(F\times G,\delta\right)\right)}{-\log\delta} \geq \liminf_{\delta\rightarrow 0} \left[\frac{\log\left(M\left(F,\delta\right)\right)}{-\log\delta} + \frac{\log\left(M\left(G,\delta\right)\right)}{-\log\delta} \right] \] so it is sufficient to prove that the right hand side is no less than $\frac{\log\left(2\right)}{\log\left(7\right)} + \frac{\log\left(2\right)}{\log\left(5\right)}$. Suppose that $\set{\delta_{i}}$ is a sequence with $\delta_{i}\rightarrow 0$ such that the limits $\lim_{i\rightarrow \infty}\frac{\log\left(N\left(F,\delta_{i}\right)\right)}{-\log \delta_{i}}$ and $\lim_{i\rightarrow \infty}\frac{\log\left(N\left(G,\delta_{i}\right)\right)}{-\log \delta_{i}}$ exist. In a similar fashion to theorem \ref{strict upper box product inequality}, corollary \ref{subsequence corollary lower} and lemmas \ref{deltan M(F,deltan) converge to log2log5} and \ref{deltan M(G,deltan) converge to log2log5} guarantee that at least one of $\lim_{i\rightarrow\infty}\frac{\log\left(M\left(F,\delta_{i_{n}}\right)\right)}{-\log\delta_{i_{n}}}$ and $\lim_{i\rightarrow\infty}\frac{\log\left(M\left(G,\delta_{i_{n}}\right)\right)}{-\log\delta_{i_{n}}}$ is no less than $\frac{\log\left(2\right)}{\log\left(5\right)}$ so \[ \liminf_{\delta\rightarrow 0} \left[\frac{\log\left(M\left(F,\delta\right)\right)}{-\log\delta} + \frac{\log\left(M\left(G,\delta\right)\right)}{-\log\delta}\right] \geq \frac{\log\left(2\right)}{\log\left(5\right)} + \frac{\log\left(2\right)}{\log\left(7\right)} \] \end{proof} \begin{corollary} $\dim_{LB}\left(F\times G\right) > \dim_{LB}\left(F\right) + \dim_{LB}\left(G\right)$ \end{corollary} \begin{acknowledgements}\label{ackref} I am greatly indebted to my PhD supervisor, James Robinson, for highlighting that there was no example of a strict product inequality in the box-counting literature and to the generosity of Isabelle Harding and Matt Gibson in providing accommodation whilst much of this work was completed. \end{acknowledgements}
{ "timestamp": "2010-07-27T02:00:26", "yymm": "1007", "arxiv_id": "1007.4222", "language": "en", "url": "https://arxiv.org/abs/1007.4222", "abstract": "It is known that the upper box-counting dimension of a Cartesian product satisfies the inequality $\\dim_{B}\\left(F\\times G\\right)\\leq \\dim_{B}\\left(F\\right) + \\dim_{B}\\left(G\\right)$ whilst the lower box-counting dimension satisfies the inequality $\\dim_{LB}\\left(F\\times G\\right)\\geq \\dim_{LB}\\left(F\\right) + \\dim_{LB}\\left(G\\right)$. We construct Cantor-like sets to demonstrate that both of these inequalities can be strict.", "subjects": "Metric Geometry (math.MG)", "title": "Strict inequality in the box-counting dimension product formulas", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9820137937226358, "lm_q2_score": 0.8152324848629215, "lm_q1q2_score": 0.8005695452261689 }
https://arxiv.org/abs/2107.09450
Monochromatic Edges in Complete Multipartite Hypergraphs
Consider the following problem. In a school with three classes containing $n$ students each, given that their genders are unknown, find the minimum possible number of triples of same-gender students not all of which are from the same class. Muaengwaeng asked this question and conjectured that the minimum scenario occurs when the classes are all boy, all girl and half-and-half. In this paper, we solve many generalizations of the problem including when the school has more than three classes, when triples are replaced by groups of larger sizes, when the classes are of different sizes, and when gender is replaced by other non-binary attributes.
\section{Introduction} A \textit{hypergraph} is a pair $(V, E)$ where $V$ is a finite set of vertices and $E$ is a collection of subsets of $V$. Each subset in $E$ is called an \textit{edge}. An \textit{$r$-uniform} hypergraph contains only edges of size $r$ and if it contains all possible edges of size $r$, this $r$-uniform hypergraph is said to be \textit{complete}. For decades, many researchers have been studying hypergraphs as a generalization of graphs and many theorems in graph theory have been extended to hypergraphs (see \cite{Frank}, \cite{Gyarfas}, \cite{Katona}, \cite{Keevash} and \cite{Mubayi}). Graph coloring which is a popular topic is also a part of those interesting extensions (see \cite{Berge} and \cite{Erdos}). Given a hypergraph, an \emph{$m$-coloring} is an assignment of a color to each vertex of the hypergraph from $m$ available colors. An edge is said to be \textit{monochromatic} if all vertices in it have the same color. Some later researches (see \cite{Bujt} and \cite{Cowen}) focused on \textit{proper coloring} and \textit{defective coloring} which involve colorings where monochromatic edges do not exist or exist only in some limited amount. The complexity of general hypergraphs has led to study focusing on hypergraphs which have an orderly and symmetric structure, for example, $k$-partite hypergraphs. A \textit{balanced $k$-partite $r$-uniform hypergraph} has $k$ vertex classes $V_1, V_2, … , V_k$ of the same size. There are two natural generalizations of $k$-partite graphs to hypergraphs. First, each edge is an $r$-subset of $\displaystyle\bigcup_{i=1}^{k}V_i$, all of whose vertices are from different classes. The second definition of edges is that each edge is an $r$-subset, not all of whose vertices are from the same class. In the paper, we will use the latter definition. The problem in the abstract can be restated in the language of hypergraphs as follows. \begin{problem} Which $2$-coloring minimizes the number of monochromatic edges of a balanced complete tripartite $3$-uniform hypergraph? \end{problem}This question was asked by Muaengwaeng \cite{Muaeng}. Given that those two colors are red and blue, she conjectured that the minimum coloring occurs when those three classes are all blue, all red and half-and-half. In this paper, we solve many generalizations of Problem $1$. First, we study a balanced complete $k$-partite $r$-uniform hypergraph. \begin{theorem}\label{thm:1} Let $n \geqslant r \geqslant 3$ and $k \geqslant 2$. The $2$-coloring minimizing the number of monochromatic edges of a balanced complete $k$-partite $r$-uniform hypergraph with $n$ vertices in each class is as follows: \begin{enumerate} \item Color all vertices of the first $ \lfloor \frac{k}{2} \rfloor $ classes with red. \item Color all vertices of the last $ \lfloor \frac{k}{2} \rfloor $ classes with blue. \item If there is another class, color the vertices of that class such that the number of red and blue vertices are as equal as possible. \end{enumerate}Moreover, this coloring is unique up to a permutation of colors and classes. \end{theorem} The case where $r=k=3$ was presented in a conference \cite{Boon} by the second author. The key idea is to calculate the change in the number of monochromatic edges when a vertex is recolored. We use this to find the minimum coloring among those with a fixed number of red vertices. Then we compare these minimum colorings. The proof of Theorem $2$ gives a clue on how to prove a more general case when each class does not contain the same number of vertices. \begin{theorem}\label{thm:2} For any unbalanced complete tripartite $3$-uniform hypergraph $H$ with the numbers of vertices of the first, second and third classes, $ n_1 \leqslant n_2 \leqslant n_3$ where $n_3 \geqslant 3$ and $n_1+n_2+n_3 = N$, a $2$-coloring minimizing the number of monochromatic edges of $H$ is as follows: \begin{enumerate} \item If $n_1 + n_2 > n_3$, color all vertices in the second and third classes with blue and red, respectively, then and color $\left\lceil \frac{N^2-3N-n^2_1-2n_1n_3+3n_1+4n_3}{2(N-n_1)}\right\rceil-n_3$ vertices in the first class with red. \item If $n_1+n_2 \leqslant n_3$, then color all vertices in the first and second classes with red and color the third class with blue. \end{enumerate}Moreover, each coloring is unique up to a permutation of colors unless $(n_1,n_2,n_3)=(2,n,n+1)$ for $n \geqslant 2$ in which case there is another extremal coloring namely a coloring such that vertices in the third class are all red, all vertices in the second class are all blue and the first class has one red and one blue vertices. \end{theorem} The proof consists of two parts. First, by swapping a red vertex and a blue vertex in different classes, we conclude that a minimum coloring must be in one of the $12$ canonical forms. Then we compares the numbers of monochromatic edges between the forms. Finally, we study the number of monochromatic edges of balanced complete $k$-partite $r$-uniform hypergraphs except, this time, up to three colors will be available. The problem of minimizing the number of monochromatic edges gets more complicated as it is not a simple two-way comparison between red and blue. The results are divided upon the remainder of the number of vertex classes, $k$, divided by $3$. We are able to solve the cases $k\equiv 0,1 \ (\textrm{mod}\ 3)$. \begin{theorem}\label{thm:3} Let $n \geqslant r \geqslant 3$ and $k \geqslant 3$. For any balanced complete $k$-partite $r$-uniform hypergraph, $H$, with $n$ vertices in each class, if $k\equiv 0\ (\textrm{mod}\ 3)$, the $3$-coloring minimizing the number of monochromatic edges of $H$ is as follows: \begin{enumerate} \item Color all vertices of the first $\frac{k}{3}$ classes with red. \item Color all vertices of the next $\frac{k}{3}$ classes with blue. \item Color all vertices of the last $\frac{k}{3}$ classes with green. \end{enumerate} If $k\equiv 1 \ (\textrm{mod}\ 3)$, the $3$-coloring minimizing the number of monochromatic edges of $H$ is as follows: \begin{enumerate} \item Color all vertices of the first $\left\lfloor\frac{k}{3}\right\rfloor$ classes with red. \item Color all vertices of the next $\left\lfloor\frac{k}{3}\right\rfloor$ classes with blue. \item Color all vertices of the next $\left\lfloor\frac{k}{3}\right\rfloor$ classes with green. \item Color all vertices of the last class such that the number of red, blue and green vertices are as equal as possible. \end{enumerate} Moreover, each coloring is unique up to a permutation of colors. \end{theorem} We use a similar idea to the proof of Theorem $3$, but we need to develop some new lemmas to construct the canonical forms of the colorings. The rest of this paper is organized as follows. In Section \ref{sec2}, we introduce some notations and useful properties that will be used throughout the paper. Later, we consider some straightforward cases. Sections \ref{sec:thm:1}, \ref{sec:thm:2} and \ref{sec:thm:3} are devoted to proving Theorems \ref{thm:1}, \ref{thm:2} and \ref{thm:3}, respectively. Finally, we conclude in Section \ref{sec:conclude} with a discussion of some open problems. \section{Preliminaries} First, we will introduce some notations that will be used throughout the paper. Later, we will mainly discuss some useful properties of binomial coefficients and some trivial cases of the problem. \label{sec2} \subsection{The number of monochromatic edges of a $2$-coloring of balanced complete $k$-partite $r$-uniform hypergraphs} Let $H$ be a balanced complete $k$-partite $(r+1)$-uniform hypergraph with $n$ vertices in each class. We consider $(r+1)$-uniform instead of $r$-uniform hypergraph for simple calculation. Let $c$ be a coloring of $H$ with $x_i$ red vertices in the $i^{th}$ class and let $X = x_1+x_2+\cdots +x_k$. Let $M(H,c)$ be the number of monochromatic edges of $H$ with coloring $c$. Then, \begin{align*} M(H,c) = \left[ { X \choose r+1}-\sum^{k}_{i=1}{x_i \choose r+1} \right] + \left[ { kn-X \choose r+1}-\sum^{k}_{i=1}{n-x_i \choose r+1} \right]. \end{align*} This function is the main method to count the number of monochromatic edges. However, this function alone is not enough for comparing the numbers of monochromatic edges of all colorings. Let $\bigtriangleup_iM(H,c)$ be the change in the number of monochromatic edges when a blue vertex in the $i^{th}$ class is recolored (if possible). The change is equal to the difference between the number of monochromatic edges containing the vertex that will be recolored, before and after the recoloring. Then, \begin{align*} \bigtriangleup_iM(H,c)= \left[ {X \choose r} - {x_i \choose r}\right] - \left[ {kn-X-1 \choose r} - {n-x_i-1 \choose r}\right]. \end{align*} We sometimes simply write $\bigtriangleup M(H,c)$ instead of $\bigtriangleup_iM(H,c)$ if the class of the color-changing vertex is clear. \subsection{The number of monochromatic edges of a $2$-coloring of unbalanced complete tripartite $3$-uniform hypergraphs} Let $H$ be an unbalanced complete tripartite $3$-uniform hypergraph with the numbers of vertices of the first, second and third classes equal to $n_1 \leqslant n_2 \leqslant n_3$, respectively, and let $N= n_1+n_2+n_3$. Let $c$ be the coloring of $H$ with the number of red vertices of the first, second and third classes equal to $x_1,x_2$ and $x_3$, respectively, and let $X=x_1+x_2+x_3$. Then, \begin{align*} M(H,c)= \left[ { X \choose 3}-\sum^{3}_{i=1}{x_i \choose 3} \right] + \left[ { N-X \choose 3}-\sum^{3}_{i=1}{n_i-x_i \choose 3} \right], \end{align*} and \begin{align*} \bigtriangleup_iM(H,c)= \left[ {X \choose 2} - {x_i \choose 2}\right] - \left[ {N-X-1 \choose 2} - {n_i-x_i-1 \choose 2}\right]. \end{align*} \subsection{The number of monochromatic edges of a $3$-coloring of balanced complete $k$-partite $r$-uniform hypergraphs} Let $H$ be a balanced complete $k$-partite $(r+1)$-uniform hypergraph with $n$ vertices in each class. Let $c$ be the $3$-coloring of $H$ with the numbers of red, blue and green vertices of the $i^{th}$ class equal to $r_i, b_i$ and $g_i$, respectively, and let $R$, $B$ and $G$ be the total numbers of red, blue and green vertices, respectively. Note that $ R= \sum^{k}_{i=1}{r_i}, B= \sum^{k}_{i=1}{b_i}$ and $ G= \sum^{k}_{i=1}{g_i}$. Moreover, $n= r_i+b_i+g_i$ for each $i = 1,2,\ldots,k$ and $kn = R+B+G$. Then, \begin{align*} M(H,c)&= \left[ { R \choose r+1}-\sum^{k}_{i=1}{r_i \choose r+1} \right] + \left[ { B \choose r+1}-\sum^{k}_{i=1}{b_i \choose r+1} \right]+ \left[ { G \choose r+1}-\sum^{k}_{i=1}{g_i \choose r+1} \right]. \end{align*} We write $\bigtriangleup_iM(H,c)$ for the change in the number of monochromatic edges when a blue vertex in the $i^{th}$ class is recolored to red (if possible). Then, \begin{align*} \bigtriangleup_iM(H,c) = \left[ {R \choose r} - {r_i \choose r}\right] - \left[ {B-1 \choose r} - {b_i-1 \choose r}\right]. \end{align*} The change can be calculated similarly for any recoloring with other color combinations. \subsection{Properties of binomial coefficients} We will additionally introduce some standard tools that will be applied throughout this paper. \begin{proposition} For any non-negative integers $a,b,c$ and $d$ with $c \leqslant a \leqslant b \leqslant d$, if $a+b\leqslant c+d$, then ${a \choose r} + {b \choose r} \leqslant {c \choose r} + {d \choose r}$ for any positive integer $r$. Moreover, the equality holds if and only if $(a=c$ and $b=d)$ or $d<r$. \end{proposition} Note that the inequality trivially holds when all upper indices of binomial coefficient terms are less than the lower index. This trivial condition will be found occasionally throughout our proofs of the main theorems. \begin{proposition} For any non-negative integers $x_1, x_2 ,\ldots,x_n$ whose sum is constant and for any non-negative integer $r$, $ \sum_{i=1}^n{x_i\choose r}$ is smallest if and only if $x_1, x_2 ,\ldots,x_n$ are as equal as possible or $ max\{x_1,x_2,\ldots,x_n\}<r$ . \end{proposition} \begin{proposition} For any non-negative integers $x_1, x_2 ,\ldots,x_n$ whose sum is constant and for any non-negative integer $r$, $ \sum_{i=1}^n{x_i\choose r}$ is largest if and only if all but one $x_i$ are zeros or $\sum_{i=1}^n x_i < r$. \end{proposition} Proposition $4$ is the main tool to compare binomial coefficients while Propositions $5$ and $6$ are generalizations of Proposition $4$ which we will apply to prove some trivial cases of the problems in the next subsections. \subsection{Colorings of hypergraphs with the size of each class fewer than the size of an edge} In our theorems, we assume that the size of each class must be at least the size of an edge, otherwise, an edge cannot be contained in a class and our hypergraphs are just complete hypergraphs. In this subsection, we will note that the problem is trivial when $n<r$ and determine a coloring that have the minimum number of monochromatic edges. Let $H$ be a complete $r$-uniform hypergraphs $H$ and let $c$ be a coloring of $H$. Then, \begin{align*} M(H,c)={R \choose r} + {B \choose r} \end{align*} if $c$ is a $2$-coloring with $R$ red and $B$ blue vertices, and \begin{align*} M(H,c)={R \choose r} + {B \choose r}+{G \choose r} \end{align*} if $c$ is a $3$-coloring with $R$ red, $B$ blue and $G$ green vertices. By Proposition $5$, $M(H,c)$ is smallest if $R$, $B$ (and $G$) are as equal as possible. Hence, a coloring such that the numbers of vertices of each color are as equal as possible has the minimum number of monochromatic edges. \subsection{Colorings of hypergraphs with the number of classes fewer than or divisible by the number of colors} In this subsection, we will consider colorings that the number of classes is fewer than or divisible by the number of colors and determine the colorings with the minimum number of monochromatic edges. We consider not only the colorings with $2$ or $3$ colors, but also $m$-colorings with $m$ greater than $3$. First, we consider the hypergraphs with the number of classes fewer than the number of colors. There are no monochromatic edges of a fixed color only when all vertices of that color are contained in at most one class, or the number of vertices of that color is fewer than $r$. Hence, the colorings such that each color appears in at most one class, or appears on fewer than $r$ vertices, are the only colorings with no monochromatic edges. Such colorings exist when the number of classes is fewer than the number of colors. The determination of the minimum coloring of the remaining case is straightforward from the Propositions $5$ and $6$. \begin{proposition} Let $n \geqslant r$ and let $k$ be divisible by $m$. The $m$-coloring minimizing the number of monochromatic edges of a balanced complete $k$-partite $r$-uniform hypergraph with $n$ vertices in each class is the coloring with equal numbers of vertices of each color and no polychromatic class. \end{proposition} \begin{proof} Suppose that $ n \geqslant r$. Let $H$ be a balanced complete $k$-partite $r$-uniform hypergraph with $n$ vertices in each class and let $c$ be an $m$-coloring of $H$ such that $k$ is divisible by $m$. Let $x_{li}$ be the number of vertices with the $l^{th}$ color in the $i^{th}$ class and let $X_l$ be the total number of vertices with the $l^{th}$ color. Then, \begin{align*} M(H,c)&= \sum^{m}_{l=1} \left[ {X_l \choose r} - \sum^{k}_{i=1} {x_{li} \choose r} \right]. \end{align*} We will show that the coloring $c^*$ with equal numbers of vertices of each color and no polychromatic class is the minimum coloring by directly comparing the numbers of monochromatic edges of $c$ and $c^*$. By Propositions $5$ and $6$, \begin{align*} M(H,c)&=\sum^{m}_{l=1} \left[ {X_l \choose r} - \sum^{k}_{i=1} {x_{li} \choose r} \right]\\ &\geqslant m{\frac{1}{m} \sum_{l=1}^m X_l \choose r} - \sum^{k}_{i=1} {\sum^{m}_{l=1} x_{li} \choose r }\\ &=m {\frac{kn}{m} \choose r} - \sum^{k}_{i=1} {n \choose r}=M(H,c^*). \end{align*} The equality holds only when ($X_i= \frac{kn}{m}$ for each $i$ and each class is monochromatic) or $n<r$, but the latter is impossible. Hence, $c^*$ is the unique coloring with the minimum number of monochromatic edges. \end{proof} The proof in this subsection is a straightforward comparison due to the simplicity of color distribution. However, in other general cases, they are much more complicated. \section{Proof of Theorem~\ref{thm:1}} \label{sec:thm:1} \begin{proof}[Proof of Theorem~\ref{thm:1}] Let $H$ be a balanced complete $k$-partite $(r+1)$-uniform hypergraph with $n \geqslant r+1$ vertices in each class where $k \geqslant 2$ and $r \geqslant 2$. Let $c$ be a coloring of $H$ with $x_i$ red vertices in the $i^{th}$ class and let $X = x_1+x_2+\cdots+x_k.$ We may assume that $X\leqslant \lfloor \frac{kn}{2} \rfloor$ , otherwise, we relabel the names of the colors. Note that if $k$ is even, the proof is completed by Proposition $7$. However, we will not need to assume that $k$ is odd in the following proof. We will calculate $M(H,c)$ in a new manner by summing each change when a blue vertex is recolored into red one by one starting from the all blue hypergraph until we reach $c$. Let $c_0$ be the all blue coloring and let $c_j$ be the coloring after the $j^{th}$ change of $H$. Thus, \begin{align*} M(H,c) = M(H,c_0) + \sum^{X-1}_{j=0} \bigtriangleup M(H,c_j)={kn \choose r+1}-k{n \choose r+1} +\sum^{X-1}_{j=0} \bigtriangleup M(H,c_j). \end{align*} We suppose that the vertices in the first class of the all blue hypergraph will be recolored first to match the first class of $c$ and then continue to the next class. Note that $c_j$ has $j$ red vertices. Let the $i^{th}$ class be the class containing the blue vertex that will be recolored and $x$ be the number of red vertices in that class. Then, from Section $2.1$, \begin{align*} \bigtriangleup M(H,c_j) &= \left[ {j \choose r} - {x \choose r}\right] - \left[ {kn-j-1 \choose r} - {n-x-1 \choose r}\right]\\ &= \left[ {j \choose r} - {kn-j-1 \choose r}\right] - \left[ {x \choose r} - {n-x-1 \choose r}\right]. \end{align*} Note that while each vertex in the changing class is being recolored, the term $x$ ascends from $0$ to $x_{i}-1$. Thus, \begin{align*} M(H,c) &={kn \choose r+1}-k{n \choose r+1} + \sum^{X-1}_{j=0} \bigtriangleup M(H,c_j)\\ &= {kn \choose r+1}-k{n \choose r+1} + \sum^{X-1}_{j=0}\left[ {j \choose r} - {kn-j-1 \choose r}\right]- \sum^{k}_{i=1}\sum^{x_i-1}_{x=0}\left[ {x \choose r} - {n-x-1 \choose r}\right]. \end{align*} In this way, if we consider only the colorings with $X$ red vertices, then the terms \[ {kn \choose r+1}-k{n \choose r+1} + \sum^{X-1}_{j=0}\left[ {j \choose r} - {kn-j-1 \choose r}\right] \] in the function $M(H,c)$ are constant. Only the term \[\sum^{k}_{i=1}\sum^{x_i-1}_{x=0}\left[ {x \choose r} - {n-x-1 \choose r}\right]\] is distinct and we denote this term by $S(x_1,x_2,\ldots,x_k)$. Hence, the coloring with maximum value of $S(x_1,x_2,\ldots,x_k)$ will have the minimum number of monochromatic edges. \begin{claim*} Among the colorings with a constant total number $X$ of red vertices, the coloring $c^*_X$ with the minimum number of polychromatic classes has the minimum number of monochromatic edges. Moreover, the minimum coloring is unique up to a permutation of classes. \end{claim*} \begin{proof} Consider a coloring with the number of red vertices in the $i^{th}$ class equal to $x_i$ where $x_1 +x_2+\cdots+x_k = X$ and $x_1 \geqslant x_2 \geqslant \cdots\geqslant x_k$. Suppose that the coloring is not $c^*_X$. Therefore, there exist classes $l < m$ such that $x_l \neq n$ and $x_m \neq 0$. Next, we will compare the terms \[S(x_1, \ldots,x_l,\ldots,x_m,\ldots,x_k)\] and \[S(x_1, \ldots,x_l+1,\ldots,x_m-1,\ldots,x_k).\] which corresponds to swapping a red vertex from the $m^{th}$ class with a blue vertex from the $l^{th}$ class. Thus, since $x_l \geqslant x_m$, \begin{align*}S(x_1&, \ldots,x_l+1,\ldots,x_m-1,\ldots,x_k) - S(x_1, \ldots,x_l,\ldots,x_m,\ldots,x_k)\\ &=\sum_{x=0}^{x_l + 1 -1}\left[ {x \choose r} - {n-x-1 \choose r}\right] + \sum_{x=0}^{x_m -1 -1}\left[ {x \choose r} - {n-x-1 \choose r}\right]\\ &\;\;\;\;-\sum_{x=0}^{x_l-1}\left[ {x \choose r} - {n-x-1 \choose r}\right]-\sum_{x=0}^{x_m-1}\left[ {x \choose r} - {n-x-1 \choose r}\right]\\ &= \left[{x_l \choose r} -{x_m-1 \choose r} \right] + \left[{n-x_m \choose r}-{n-x_l-1\choose r}\right] \geqslant 0. \end{align*} The equality holds only when all upper indices of the binomial coefficient terms are less than $r$. The swapping resulted in fewer or equal monochromatic edges. We will continue swapping as long as possible to reduce the number of polychromatic classes. The inequality is strict at some point since either $x_l$ will eventually equal to $n-1$ in which case $x_l=n-1 \geqslant r$, or $x_m$ will eventually equal to $1$ in which case $n-x_m=n-1 \geqslant r$. This implies that the original coloring has strictly more monochromatic edges than some coloring. Hence, $c^*_X$ is the unique coloring with the minimum number of monochromatic edges among those with $X$ red vertices. \end{proof} We have determined the minimum coloring for each value of $X$. Next, we will make comparisons between colorings with different values of $X$. We will show that $M(H,c^*_X) > M(H, c^*_{X+1})$ for all $X \leqslant \lfloor\frac{kn}{2}\rfloor -1$. Let $c^*_X$ be the minimum coloring with $X \leqslant \lfloor\frac{kn}{2}\rfloor-1$ red vertices. Suppose that the polychromatic class of $c^*_X$ is the $i^{th}$ class, but if $c^*_X$ has no polychromatic class, suppose the $i^{th}$ class is an all blue class. Observe that \begin{align*} M(H,c^*_{X+1})-M(H,c^*_X) &=\bigtriangleup_i M(H,c^*_X)\\ &= \left[ {X \choose r} - {x_i \choose r}\right] - \left[ {kn-X-1 \choose r} - {n-x_i-1 \choose r}\right]\\ &= \left[ {X \choose r} + {n-x_i-1 \choose r}\right] - \left[ {kn-X-1 \choose r} + {x_i \choose r}\right]. \end{align*} The following proof will be divided into two cases according to the value of $x_i $. \textit{Case 1}: $\frac{n}{2} \leqslant x_i < n $. Thus, \begin{align*} X- (kn-X-1)=2\left(X+\frac{1}{2} -\frac{kn}{2}\right) \leqslant 2\left(\left\lfloor\frac{kn}{2}\right\rfloor-\frac{kn}{2}-\frac{1}{2}\right)< 0, \end{align*}and \begin{align*}(n-x_i-1)-x_i = 2\left(\frac{n}{2}-x_i-\frac{1}{2}\right)< 0. \end{align*} Hence, $\bigtriangleup_i M(H,c^*_X) \leqslant 0$ but $\bigtriangleup_i M(H,c^*_X) \neq 0$ since one of the upper indices is at least $r$. Indeed, $kn-X-1 \geqslant \left\lceil\frac{kn}{2}\right\rceil \geqslant n > r$ because $k \geqslant 2$. \textit{Case 2}: $0 \leqslant x_i < \frac{n}{2}$. We will show that $\bigtriangleup_i M(H,c^*_X) < 0$ by Proposition 4. As in Case 1, \[X- (kn-X-1)<0.\] Moreover, \begin{align*}(n-x_i-1)-x_i = 2\left(\frac{n}{2}-x_i-\frac{1}{2}\right)\geqslant 0. \end{align*} Suppose that there are $k^*$ red classes in $c^*_X$, i.e., $X = k^*n + x_i$. Since $X\leqslant \lfloor\frac{kn}{2}\rfloor-1$, we have $k^* \leqslant \frac{k-1}{2}$. Thus, \begin{align*} (X +n-x_i-1)-(x_i+kn-X-1)=2\left(k^*n - \frac{k-1}{2}n\right)\leqslant 0. \end{align*}\noindent Similarly, $kn-X-1>r$. Hence, by Proposition 4, $\bigtriangleup_i M(H,c^*_X) < 0.$ By the two cases, $c^*_X$ contains strictly more monochromatic edges than $c^*_{X+1}$ does, given that $X \leqslant \lfloor\frac{kn}{2}\rfloor -1$. Consequently, we can conclude that the unique coloring with the minimum number of monochromatic edges among all minimum colorings $c^*_X$ is $c^*_{\lfloor\frac{kn}{2}\rfloor}$. Together with the claim, $c^*_{\lfloor\frac{kn}{2}\rfloor}$ is the unique coloring with the minimum number of monochromatic edges.\end{proof} Note that, in the claim, we determine $c^*_X$ by means of determining the coloring with maximum $S(x_1,x_2,\ldots,x_k)$. On the contrary, we could determine the coloring with a constant number $X$ of red vertices that has the maximum number of monochromatic edges by showing conversely that the coloring with minimum $S(x_1,x_2,\ldots,x_k)$ is the coloring such that $x_1$, $x_2, \ldots, x_k$ are as equal as possible. However, this is out of our topic. \section{Proof of Theorem~\ref{thm:2}} \label{sec:thm:2} \begin{proof}[Proof of Theorem~\ref{thm:2}] Let $H$ be an unbalanced complete tripartite $3$-uniform hypergraph with the numbers of vertices of the first, second and third classes $ n_1 \leqslant n_2 \leqslant n_3$ and let $N=n_1+n_2+n_3$. Let $c$ be the coloring of $H$ with the number of red vertices of the first, second and third classes equal to $x_1,x_2$ and $x_3$, respectively, and let $X=x_1+x_2+x_3$. We divide the proof into two subsections according to the size of the smallest class. In the first subsection, a similar idea as in the proof of Theorem~\ref{thm:1} is extended to determine the minimum coloring when the number of vertices of each class is at least $3$. The second subsection is mainly about hypergraphs with some small classes. \subsection{Hypergraphs with $n_1 \geqslant 3$} Assume that $n_1 \geqslant 3$. Let $\bigtriangleup_{i{i'}} M(H,c)$ be the change in the number of monochromatic edges if a blue vertex in the $i^{th}$ class is recolored into red and a red vertex in the ${i'}^{th}$ class is recolored into blue. The process will be called \textit{swapping} which results in a new coloring, say $c'$. We compute $\bigtriangleup_{i{i'}} M(H,c)$ by comparing the number of monochromatic edges containing those vertices undergone swapping before and after swapping process. Thus, \begin{align*} \bigtriangleup_{i{i'}}M(H,c)&= \left[ {x_1+x_2+x_3-1 \choose 2}-{x_i \choose 2} + {N-x_1-x_2-x_3-1 \choose 2} - {n_{i'}-x_{i'}\choose 2} \right]\\ &\;\;\;\;- \left[ {x_1+x_2+x_3-1 \choose 2}-{x_{i'}-1 \choose 2} + {N-x_1-x_2-x_3-1 \choose 2} - {n_i-x_i-1\choose 2} \right]\\ &= \left[ {x_{i'}-1 \choose 2} + {n_i-x_i-1\choose 2} \right] - \left[ {x_i \choose 2} + {n_{i'}-x_{i'}\choose 2} \right]. \end{align*} A \textit{successful swapping} is a swapping in such a way that the number of monochromatic edges is reduced, i.e., $\bigtriangleup_{i{i'}} M(H,c) < 0$. \begin{lemma} If $\bigtriangleup_{i{i'}} M(H,c) \leqslant 0$, then $\bigtriangleup_{i{i'}} M(H,c') < 0$. \end{lemma}\begin{proof} Observe that \begin{align*} \bigtriangleup_{i{i'}}M(H,c')&= \left[ {(x_{i'}-1)-1 \choose 2} + {n_i-(x_i+1)-1\choose 2} \right] - \left[ {x_i+1 \choose 2} + {n_{i'}-(x_{i'}-1)\choose 2} \right]\\ &\leqslant \left[ {x_{i'}-1 \choose 2} + {n_i-x_i-1\choose 2} \right] - \left[ {x_i \choose 2} + {n_{i'}-x_{i'}\choose 2} \right]=\bigtriangleup_{i{i'}} M(H,c) \leqslant 0. \end{align*} The equality holds only when ($x_{i'}-1<2$ and $n_i-x_i-1<2$) and ($x_i+1<2$ and $n_{i'}-x_{i'}+1<2$) and $\bigtriangleup_{i{i'}} M(H,c)=0$. Suppose that $\bigtriangleup_{i{i'}}M(H,c')=0$. We have that $x_i=0$ and $x_{i'}<3$. Thus, $n_i=n_i-x_i<3$ and $n_{i'}-2 \leqslant n_{i'} -x_{i'}<1 $, i.e., $n_{i'}<3$ which contradicts with $3 \leqslant n_1\leqslant n_2\leqslant n_3$. \end{proof} Lemma 8 means that if a swapping can be done without increasing the number of monochromatic edges, another swapping in the same direction will be successful (if there are red and blue vertices to be swapped). The process of successful swappings will terminate when one of the two classes (or both) is monochromatic. \begin{lemma} If $\bigtriangleup_{i{i'}} M(H,c) \geqslant 0$, then $\bigtriangleup_{{i'}i} M(H,c) < 0$. \end{lemma} \begin{proof}Observe that \begin{align*} \bigtriangleup_{{i'}i} M(H,c) &= \left[ {x_i-1 \choose 2} + {n_{i'}-x_{i'}-1\choose 2} \right] - \left[ {x_{i'} \choose 2} + {n_i-x_i\choose 2} \right]\\ &\leqslant \left[ {x_i \choose 2} + {n_{i'}-x_{i'}\choose 2} \right] - \left[ {x_{i'}-1 \choose 2} + {n_i-x_i-1\choose 2} \right]\\ &= -\bigtriangleup_{i{i'}} M(H,c) \leqslant 0. \end{align*} The equality holds only when ($x_i<2$ and $n_{i'}-x_{i'}<2$) and ($x_{i'}<2$ and $n_i-x_i<2$) and $\bigtriangleup_{i{i'}} M(H,c)=0$. Suppose that $\bigtriangleup_{{i'}i}M(H,c)=0$. We have that $x_i \leqslant 1$ and $x_{i'}\leqslant 1$. Thus, $n_i-1\leqslant =n_i-x_i<2$ and $n_{i'}-1\leqslant n_{i'} -x_{i'}<2 $, i.e., $n_i<3$ and $n_{i'}<3$ which contradicts with $3 \leqslant n_1\leqslant n_2\leqslant n_3$. \end{proof} Note that, for any coloring $c$, if $c$ contains two classes, the $i^{th}$ and ${i'}^{th}$, which are polychromatic, then a swapping can be done in two directions as follows. \begin{enumerate}\item Swapping a red vertex of the $i^{th}$ class with a blue vertex of the ${i'}^{th}$ class. \item Swapping a blue vertex of the $i^{th}$ class with a red vertex of the ${i'}^{th}$ class. \end{enumerate} By Lemma 9, one of the two directions is successful. Moreover, by Lemma 8, we can continue swapping in the same direction until one of the two classes is monochromatic and get fewer monochromatic edges. Hence, the coloring with minimum number of monochromatic edges among colorings with constant number of red vertices, must have at most one polychromatic class. We will list all these forms in the following table which will be the candidates for the coloring with minimum number of monochromatic edges. \begin{center} \begin{tabular}{ |P{3cm}|P{3cm}|P{3cm}|P{3cm}|} \hline Canonical forms & $1^{st}$ Class & $2^{nd}$ Class & $3^{rd}$ Class\\ \hline $F_1$& polychromatic& blue&blue\\ $F_2$& blue& polychromatic&blue\\ $F_3$& blue& blue&polychromatic\\ $F_4$& red&polychromatic&blue\\ $F_5$& red&blue&polychromatic\\ $F_6$& polychromatic&red&blue\\ $F_7$& blue&red&polychromatic\\ $F_8$& polychromatic&blue &red\\ $F_9$& blue& polychromatic&red\\ $F_{10}$& red& red&polychromatic\\ $F_{11}$& red& polychromatic&red\\ $F_{12}$& polychromatic& red&red\\ \hline \end{tabular} \end{center} The first column illustrates the list of $12$ canonical forms and the remaining columns describe the colors of vertices in those classes. The terms \textit{red} and \textit{blue} mean all vertices in those classes are monochromatic of red and blue, respectively. On the other hand, \textit{polychromatic} means that this class is allowed to be polychromatic but it may be monochromatic. Note that a coloring can be considered to be in several canonical forms, for example, the all blue coloring is of the form $F_1$, $F_2$ or $F_3$. We may assume that $X \leqslant\lfloor\frac{N}{2}\rfloor$. Consequently, both $F_{11}$ and $F_{12}$ are out of our interest since the total numbers of red vertices, which are $n_1+x_2+n_3$ and $x_1+n_2+n_3$, respectively, exceed $\lfloor\frac{N}{2}\rfloor$ as shown: \begin{align*} n_1+x_2+n_3 = \frac{n_1+n_3+2x_2}{2}+\frac{n_1+n_3}{2} > \frac{n_2}{2} +\frac{n_1+n_3}{2} \geqslant \left\lfloor\frac{n_1+n_2+n_3}{2}\right\rfloor \end{align*} and \begin{align*} x_1+n_2+n_3 = \frac{n_2+n_3+2x_1}{2}+\frac{n_2+n_3}{2} > \frac{n_1}{2} +\frac{n_2+n_3}{2} \geqslant \left\lfloor\frac{n_1+n_2+n_3}{2}\right\rfloor. \end{align*} Next, we will focus on the possibility of $F_{10}$. If $c$ has total red vertices to be $X=n_1+n_2+x_3 \leqslant \lfloor\frac{n_1+n_2+n_3}{2}\rfloor $. Then, \begin{align*} n_1 + n_2 \leqslant 2\left(\left\lfloor\frac{n_1+n_2+n_3}{2}\right\rfloor-\frac{n_1+n_2}{2}\right)\leqslant 2\left(\frac{n_1+n_2+n_3}{2}-\frac{n_1+n_2}{2}\right)=n_3. \end{align*} The necessary condition of a coloring $c$ of a hypergraph $H$ to be in the form $F_{10}$ is that $n_1+n_2 \leqslant n_3$. Note that the condition $n_1+n_2 \leqslant n_3 $ is equivalent to $n_1+n_2 \leqslant \lfloor\frac{n_1+n_2+n_3}{2}\rfloor \leqslant n_3$ and we call a hypergraph with this condition \textit{type A}. On the other hand, the condition $n_1+n_2 > n_3 $ is equivalent to $n_3\leqslant \lfloor\frac{n_1+n_2+n_3}{2}\rfloor < n_1+n_2$ and we call a hypergraph with this condition \textit{type B}. Next, we will determine the minimum coloring among those colorings with constant number $X$ of red vertices or $c^*_X$ from the candidates $F_1$ to $F_9$ and $F_{10}$ will be considered only when $H$ is a type A hypergraph. As in the proof of Theorem~\ref{thm:1}, we calculate $M(H,c)$ by summing each change when a blue vertex is recolored into red one by one starting from the all blue hypergraph until we reach $c$. Let $c_0$ be the all blue coloring and let $c_j$ be the coloring after the $j^{th}$ change of $H$. Thus, \begin{align*} M(H,c) = M(H,c_0) + \sum^{X-1}_{j=0} \bigtriangleup M(H,c_j)= {N \choose 3}-\sum_{i=1}^3{n_i \choose 3}+\sum^{X-1}_{j=0} \bigtriangleup M(H,c_j). \end{align*} We suppose that the vertices in the first class of the all blue hypergraph will be recolored first to match the first class of $c$ and then continue to the next class. Note that $c_j$ has $j$ red vertices. Let the $i^{th}$ class be the class containing the blue vertex that will be recolored and $x$ be the number of red vertices in that class. Then, from Section $2.2$, \begin{align*} \bigtriangleup M(H,c_j) &= \left[ {j \choose 2} - {x \choose 2}\right] - \left[ {N-j-1 \choose r} - {n_i-x-1 \choose 2}\right]\\ &= \left[ {j \choose 2} - {N-j-1 \choose 2}\right] - \left[ {x \choose 2} - {n_i-x-1 \choose 2}\right]. \end{align*} Note that while each vertex in the changing class is being recolored, the term $x$ ascends from $0$ to $x_i-1$. Thus, \begin{align*} M(H,c) &= {N \choose 3}-\sum_{i=1}^3{n_i \choose 3}+\sum^{X-1}_{j=0} \bigtriangleup M(H,c_j)\\ &={N \choose 3}-\sum_{i=1}^3{n_i \choose 3}+\sum^{X-1}_{j=0}\left[ {j \choose 2} - {N-j-1 \choose 2}\right]- \sum^{3}_{i=1}\sum^{x_i-1}_{x=0}\left[ {x \choose 2} - {n_i-x-1 \choose 2}\right]. \end{align*} Similarly, if we consider only the coloring with $X$ red vertices, then the terms \[{N \choose 3}-\sum_{i=1}^3{n_i \choose 3}+\sum^{X-1}_{j=0}\left[ {j \choose 2} - {N-j-1 \choose 2}\right]\] in the function $M(H,c)$ are constant. Only the term \[\sum^{3}_{i=1}\sum^{x_i-1}_{x=0}\left[ {x \choose 2} - {n_i-x-1 \choose 2}\right]\] is distinct and we denote this term by $S(x_1,x_2,x_3)$. Hence, the coloring with maximum value of $S(x_1,x_2,x_3)$ will have the minimum number of monochromatic edges. Remark that if $x_i = n_i$, then the term $\sum^{x_i-1}_{x=0}\left[ {x \choose 2} - {n_i-x-1 \choose 2} \right]$ cancels itself out. Hence, $S(n_1,x_2,x_3)=S(0,x_2,x_3)$ and similarly when $x_2=n_2$ or $x_3=n_3$. Next, we will determine $c^*_X$ by considering and comparing only among possible canonical forms according to the value $X$ and type of $H$. We will divide into several cases. \textit{Case 1}: $0 \leqslant X < n_1$. \begin{center} \begin{tabular}{ |P{3cm}|P{1.5cm}|P{1.5cm}|P{1.5cm}|P{3cm}| } \hline Possible forms & $x_1$ & $x_2$ & $x_3$ & $S(x_1,x_2,x_3)$\\ \hline $F_1$& $X$& $0$&$0$&$S(X,0,0)$\\ $F_2$& $0$& $X$&$0$&$S(0,X,0)$\\ $F_3$& $0$& $0$&$X$&$S(0,0,X)$\\ \hline \end{tabular} \end{center} \noindent We will compare among colorings in the forms $F_1$, $F_2$ and $F_3$. Note that if $n_1=n_2$, $F_1$ and $F_2$ are the same. Similarly, if $n_1=n_2=n_3$, $F_1,F_2$ and $F_3$ are the same. Then, \begin{align*} S(0,X,0) = \sum_{j=0}^{X-1}\left[ {j \choose 2}-{n_2-j-1\choose 2} \right]\leqslant \sum_{j=0}^{X-1}\left[ {j \choose 2}-{n_1-j-1\choose 2} \right] = S(X,0,0). \end{align*} \noindent The equality holds only when $n_1=n_2$ or $n_2-1<2$. Since $3 \leqslant n_2 \leqslant n_3$, we can conclude that a coloring in the form $F_1$ has fewer monochromatic edges than $F_2$ and similarly for $F_3$. Hence, $c^*_X$ is in the form $F_1$. \textit{Case 2}: $n_1 \leqslant X < \frac{n_1 + n_2}{2}$. \begin{center} \begin{tabular}{ |P{3cm}|P{1.5cm}|P{1.5cm}|P{1.5cm}|P{3cm}| } \hline Possible forms & $x_1$ & $x_2$ & $x_3$&$S(x_1,x_2,x_3)$\\ \hline $F_2$& $0$& $X$&$0$&$S(0,X,0)$\\ $F_3$& $0$& $0$&$X$&$S(0,0,X)$\\ $F_4$& $n_1$& $X-n_1$&$0$&$S(n_1,X-n_1,0)$\\ $F_5$& $n_1$& $0$&$X-n_1$&$S(n_1,0,X-n_1)$\\ \hline \end{tabular} \end{center} In this case, we must have that $n_1 < n_2$. Similarly to \textit{Case 1}, we have that $F_2$ has fewer number monochromatic edges than $F_3$. Next, we will compare between colorings in the forms $F_4$ and $F_5$. If $n_2=n_3$, then both forms are the same. Thus, \begin{align*} S(n_1,0,X-n_1) =S(0,0,X-n_1)\leqslant S(0,X-n_1,0)=S(n_1,X-n_1,0). \end{align*} The equality holds only when $n_2=n_3$ or $n_3-1<2$. Since $3\leqslant n_3$, we have that $F_4$ has fewer number monochromatic edges than $F_5$. Finally, we will compare between colorings in the forms $F_2$ and $F_4$. Note that $X-n_1 < \frac{n_1 + n_2}{2} - n_1 = n_2 - \frac{n_1 + n_2}{2} < n_2 -X $. Then, \begin{align*} S(0,X,0)&=\sum_{j=0}^{X-1}\left[ {j \choose 2}-{n_2-j-1\choose 2} \right]\\ &=\sum_{j=0}^{(X-n_1)-1}\left[ {j \choose 2}-{n_2-j-1\choose 2} \right]+\left[{X-n_1 \choose 2} +{X-n_1 +1 \choose 2} +\cdots+ {X-1 \choose 2} \right]\\ &\;\;\;\;- \left[{n_2-X\choose2}+{n_2-X+1\choose2}+\cdots+{n_1+n_2-X-1\choose2}\right]\\ &\leqslant \sum_{j=0}^{(X-n_1)-1}\left[ {j \choose 2}-{n_2-j-1\choose 2} \right] = S(0,X-n_1,0) = S(n_1, X-n_1,0). \end{align*} The equality holds only when $n_1+n_2-X-1 <2$. Since $X< \frac{n_1 + n_2}{2}$, this occurs only when $n_1+n_2 \leqslant 3$, i.e., $n_1 \leqslant n_2 <3$. Since $3 \leqslant n_2$, we have that $F_4$ gives fewer number monochromatic edges than $F_2$ and $c^*_X$ is in the form $F_4$. \textit{Case 3}: $\frac{n_1 + n_2}{2} \leqslant X < n_2$. \begin{center} \begin{tabular}{ |P{3cm}|P{1.5cm}|P{1.5cm}|P{1.5cm}|P{3cm}| } \hline Possible forms & $x_1$ & $x_2$ & $x_3$&$S(x_1,x_2,x_3)$\\ \hline $F_2$& $0$& $X$&$0$&$S(0,X,0)$\\ $F_3$& $0$& $0$&$X$&$S(0,0,X)$\\ $F_4$& $n_1$& $X-n_1$&$0$&$S(n_1,X-n_1,0)$\\ $F_5$& $n_1$& $0$&$X-n_1$&$S(n_1,0,X-n_1)$\\ \hline \end{tabular} \end{center} Similarly, we must have $n_1 < n_2$ in this case. The comparisons between $F_2$ and $F_3$ and between $F_4$ and $F_5$ are similar to the previous case. Note that $X-n_1 \geqslant \frac{n_1 + n_2}{2} - n_1 = n_2 - \frac{n_1 + n_2}{2} \geqslant n_2 -X $. Next, we will show the comparison between $F_2$ and $F_4$ which gives a contrary result as shown:\begin{align*} S(0,X,0)&=\sum_{j=0}^{X-1}\left[ {j \choose 2}-{n_2-j-1\choose 2} \right]\\ &=\sum_{j=0}^{(X-n_1)-1}\left[ {j \choose 2}-{n_2-j-1\choose 2} \right]+\left[{X-n_1 \choose 2} +{X-n_1 +1 \choose 2} +\cdots+ {X-1 \choose 2} \right]\\ &\;\;\;\;- \left[{n_2-X\choose2}+{n_2-X+1\choose2}+\cdots+{n_1+n_2-X-1\choose2}\right]\\ &\geqslant \sum_{j=0}^{(X-n_1)-1}\left[ {j \choose 2}-{n_2-j-1\choose 2} \right] = S(0,X-n_1,0) = S(n_1, X-n_1,0). \end{align*} The equality holds only when $X=\frac{n_1+n_2}{2}$ or $X-1<2$. Since $\frac{n_1 + n_2}{2} \leqslant X < n_2$ and $n_1<n_2$, the condition that $X<3$ occurs only when $n_1=1$ and $n_2=2$ or $3$. Since $3 \leqslant n_1$, we have that $F_2$ gives fewer number monochromatic edges than $F_4$ and $c^*_X$ is in the form $F_2$ when $X>\frac{n_1+n_2}{2}$. If $X=\frac{n_1+n_2}{2}$, both forms give the same number of monochromatic edges. From now on, the cases will be divided by whether the hypergraph is of type A or B. \textit{Case 4A}: $n_2 \leqslant X < n_1+n_2$ and $H$ is a type A hypergraph. \begin{center} \begin{tabular}{ |P{3cm}|P{1.5cm}|P{1.5cm}|P{1.5cm}|P{3cm}| } \hline Possible forms & $x_1$ & $x_2$ & $x_3$&$S(x_1,x_2,x_3)$\\ \hline $F_3$& $0$& $0$&$X$&$S(0,0,X)$\\ $F_4$& $n_1$& $X-n_1$&$0$&$S(n_1,X-n_1,0)$\\ $F_5$& $n_1$& $0$&$X-n_1$&$S(n_1,0,X-n_1)$\\ $F_6$& $X-n_2$& $n_2$&$0$&$S(X-n_2,n_2,0)$\\ $F_7$& $0$& $n_2$&$X-n_2$&$S(0,n_2,X-n_2)$\\ \hline \end{tabular} \end{center} Similarly to \textit{Case 2}, we have that $F_4$ has fewer number monochromatic edges than $F_5$. Next, we will compare between colorings in the forms $F_6$ and $F_7$. If $n_1=n_2=n_3$, then the forms $F_4, F_5, F_6$ and $F_7$ are the same. Thus, \begin{align*} S(0,n_2,X-n_2) =S(0,0,X-n_2) \leqslant S(X-n_2,0,0)=S(X-n_2,n_2,0). \end{align*} The equality holds only when $n_1=n_3$ or $n_3-1<2$. Since $n_3 \geqslant 3$, we have that $F_6$ has fewer number monochromatic edges than $F_7$. Next, we will compare between colorings in the forms $F_4$ and $F_6$. If $n_1=n_2$, then both forms are the same. Suppose that $n_1<n_2$. Thus, since $X< n_1+n_2$, \begin{align*} S(n_1, X-n_1,0) &= S(0,X-n_1,0 \\&= \left[{0 \choose 2} + {1 \choose 2} +\cdots+{X-n_2-1 \choose 2}\right]\\ &\;\;\;\;- \left[{n_1+n_2-X \choose 2} + {n_1+n_2-X+1 \choose 2} + \cdots + {n_1 -1 \choose 2}\right]\\ &\;\;\;\;+ \left[{X-n_2 \choose 2} + {X-n_2+1 \choose 2} +\cdots+{X-n_1-1 \choose 2}\right]\\ &\;\;\;\;- \left[{n_1\choose 2} + {n_1+1 \choose 2} + \cdots + {n_2-1 \choose 2}\right]\\ &\leqslant \left[{0 \choose 2} + {1 \choose 2} +\cdots+{X-n_2-1 \choose 2}\right]\\ &\;\;\;\;- \left[{n_1+n_2-X \choose 2} + {n_1+n_2-X+1 \choose 2} + \cdots + {n_1 -1 \choose 2}\right]\\ &= S(X-n_2,0,0) = S(X-n_2,n_2,0). \end{align*} The equality holds only when $n_2-1<2$. Since $ n_2\geqslant 3$, we have that $F_6$ has fewer number monochromatic edges than $F_4$. Finally, we will compare between colorings in the forms $F_3$ and $F_6$. Then, since $n_2 \leqslant X < n_1+n_2 \leqslant n_3$, \allowdisplaybreaks \begin{align*} S(0,0,X) &=\left[{0 \choose 2} + {1 \choose 2} +\cdots+{X-1 \choose 2}\right]- \left[{n_3-X \choose 2} + {n_3-X+1 \choose 2} + \cdots + {n_3 -1 \choose 2}\right]\\ &\leqslant\left[{0 \choose 2} + {1 \choose 2} +\cdots+{X-1 \choose 2}\right]- \left[{n_3-X \choose 2} + {n_3-X+1 \choose 2} + \cdots + {n_3 -1 \choose 2}\right]\\ &\;\;\;\;- \left[{X-n_2 \choose 2} + {X-n_2+1 \choose 2} +\cdots+{X-1 \choose 2}\right]\\ &\;\;\;\;+ \left[{n_3-n_2\choose 2} + {n_3+n_2+1 \choose 2} + \cdots + {n_3-1 \choose 2}\right]\\ &\;\;\;\;- \left[{n_1+n_2-X \choose 2} + {n_1+n_2-X+1 \choose 2} +\cdots+{n_3-X-1 \choose 2}\right]\\ &\;\;\;\;+ \left[{n_1\choose 2} + {n_1+1 \choose 2} + \cdots + {n_3-n_2-1 \choose 2}\right]\\ &= \left[{0 \choose 2} + {1 \choose 2} +\cdots+{X-n_2-1 \choose 2}\right]\\ &\;\;\;\;- \left[{n_1+n_2-X \choose 2} + {n_1+n_2-X+1 \choose 2} + \cdots + {n_1 -1 \choose 2}\right]\\ &= \sum_{j=0}^{(X-n_2)-1}\left[ {j \choose 2}-{n_1-j-1\choose 2} \right]= S(X-n_2,0,0) = S(X-n_2,n_2,0). \end{align*} The equality holds only when $n_3-1<2$. Since $n_3 \geqslant 3$, $F_6$ gives fewer number monochromatic edges than $F_3$ and $c^*_X$ is in the form $F_6$. \textit{Case 5A}: $n_1+n_2 \leqslant X \leqslant \lfloor{\frac{N}{2}}\rfloor$ and $H$ is a type A hypergraph. \begin{center} \begin{tabular}{ |P{3cm}|P{1.5cm}|P{1.5cm}|P{2cm}|P{4cm}| } \hline Possible forms & $x_1$ & $x_2$ & $x_3$&$S(x_1,x_2,x_3)$\\ \hline $F_3$& $0$& $0$&$X$&$S(0,0,X)$\\ $F_{10}$& $n_1$& $n_2$&$X-n_1-n_2$&$S(n_1,n_2,X-n_1-n_2)$\\ \hline \end{tabular} \end{center} Note that this is only the case that we will consider $F_{10}$. We only compare between colorings in the forms $F_3$ and $F_{10}$. Then, since $X \leqslant \lfloor{\frac{n_1+n_2+n_3}{2}}\rfloor$, \begin{align*} S(0,0,X)&= \left[{0 \choose 2} + {1 \choose 2} +\cdots+{X-1 \choose 2}\right]- \left[{n_3-X \choose 2} + {n_3-X+1 \choose 2} +\cdots+{n_3-1 \choose 2}\right]\\ &= \left[{0 \choose 2} + {1 \choose 2} +\cdots+{X-n_1-n_2-1 \choose 2}\right]\\ &\;\;\;\;- \left[{n_1+n_2+n_3-X \choose 2} + {n_1+n_2+n_3-X+1 \choose 2} + \cdots + {n_3 -1 \choose 2}\right]\\ &\;\;\;\;+ \left[{X-n_1-n_2\choose2}+{X-n_1-n_2+1\choose2}+\cdots+{X-1\choose2}\right]\\ &\;\;\;\;-\left[{n_3-X\choose2}+{n_3-X+1\choose2}+\cdots+{n_1+n_2+n_3-X-1\choose2}\right]\\ &\leqslant \left[{0 \choose 2} + {1 \choose 2} +\cdots+{X-n_1-n_2-1 \choose 2}\right]\\ &\;\;\;\;- \left[{n_1+n_2+n_3-X \choose 2} + {n_1+n_2+n_3-X+1 \choose 2} + \cdots + {n_3 -1 \choose 2}\right]\\ &= S(0,0,X-n_1-n_2) = S(n_1,n_2,X-n_1-n_2). \end{align*} The equality holds only when $X=\frac{N}{2}$ or $N-X-1<2$. Since $X \leqslant \lfloor{\frac{N}{2}}\rfloor$, the condition that $N-X<3$ occurs only when $N \leqslant 4$. This is impossible because $n_3 \geqslant 3$. If $X= \frac{N}{2}$, then both forms have the same number of monochromatic edges. However, if we relabel the names of the colors, then both forms are the same. Hence, $F_{10}$ gives fewer number monochromatic edges than $F_3$ and $c^*_X$ is in the form $F_{10}$. Next, we will focus on the other cases of type B hypergraphs. \textit{Case 4B}: $n_2 \leqslant X < n_3$ and $H$ is a type B hypergraph. \begin{center} \begin{tabular}{ |P{3cm}|P{1.5cm}|P{1.5cm}|P{1.5cm}|P{3cm}| } \hline Possible canonical forms & $x_1$ & $x_2$ & $x_3$&$S(x_1,x_2,x_3)$\\ \hline $F_3$& $0$& $0$&$X$&$S(0,0,X)$\\ $F_4$& $n_1$& $X-n_1$&$0$&$S(n_1,X-n_1,0)$\\ $F_5$& $n_1$& $0$&$X-n_1$&$S(n_1,0,X-n_1)$\\ $F_6$& $X-n_2$& $n_2$&$0$&$S(X-n_2,n_2,0)$\\ $F_7$& $0$& $n_2$&$X-n_2$&$S(0,n_2,X-n_2)$\\ \hline \end{tabular} \end{center} Similarly to \textit{Case 4A}, $F_4$ and $F_6$ have fewer monochromatic edges than $F_5$ and $F_7$, respectively, and $F_6$ has fewer monochromatic edges than $F_4$. Next, we have to compare between $F_3$ and $F_6$ to determine that which form $c^*_X$ is. Due to complexity of the comparison, we will conclude that $c^*_X$ is in the form $F_3$ or $F_6$. \textit{Case 5B}: $n_3 \leqslant X \leqslant \lfloor{\frac{n_1+n_2+n_3}{2}}\rfloor $ and $H$ is a type B hypergraph. \begin{center} \begin{tabular}{ |P{3cm}|P{1.5cm}|P{1.5cm}|P{1.5cm}|P{3cm}| } \hline Possible canonical forms & $x_1$ & $x_2$ & $x_3$&$S(x_1,x_2,x_3)$\\ \hline $F_4$& $n_1$& $X-n_1$&$0$&$S(n_1,X-n_1,0)$\\ $F_5$& $n_1$& $0$&$X-n_1$&$S(n_1,0,X-n_1)$\\ $F_6$& $X-n_2$& $n_2$&$0$&$S(X-n_2,n_2,0)$\\ $F_7$& $0$& $n_2$&$X-n_2$&$S(0,n_2,X-n_2)$\\ $F_8$& $X-n_3$& $0$&$n_3$&$S(X-n_3,0,n_3)$\\ $F_9$& $0$& $X-n_3$&$n_3$&$S(0,X-n_3,n_3)$\\ \hline \end{tabular} \end{center} Similarly to \textit{Case 4A}, $F_4$ and $F_6$ have fewer monochromatic edges than $F_5$ and $F_7$, respectively, and $F_6$ has fewer monochromatic edges than $F_4$. Next, we will compare between colorings in the forms $F_8$ and $F_9$. If $n_1=n_2$, then both forms are the same. Thus, \begin{align*} S(0,X-n_3,n_3) =S(0,X-n_3,0)=S(X-n_3,0,0)==S(X-n_3, 0,n_3). \end{align*} The equality holds only when $n_1=n_2$ or $n_2-1 <2$. Since $3 \leqslant n_2$, we have that $F_8$ gives fewer number monochromatic edges than $F_9$. Finally, we will compare between colorings in the forms $F_6$ and $F_8$. If $n_2=n_3$, then both forms are the same. We may suppose that $n_2<n_3$. Thus, since $X \leqslant \lfloor{\frac{N}{2}}\rfloor $, \begin{align*} S(X-n_2, n_2,0)&= S(X-n_2,0,0)\\ &= \left[{0 \choose 2} + {1 \choose 2} +\cdots+{X-n_3-1 \choose 2}\right]\\ &\;\;\;\;+ \left[{X-n_3 \choose 2} + {X-n_3+1 \choose 2} +\cdots+{X-n_2-1 \choose 2}\right]\\ &\;\;\;\;- \left[{n_1+n_2 -X\choose 2} + {n_1+n_2-X+1 \choose 2} + \cdots + {n_1+n_3-X-1 \choose 2}\right]\\ &\;\;\;\;- \left[{n_1+n_3-X \choose 2} + {n_1+n_3-X+1 \choose 2} + \cdots + {n_1 -1 \choose 2}\right]\\ &\leqslant \left[{0 \choose 2} + {1 \choose 2} +\cdots+{X-n_3-1 \choose 2}\right]\\ &- \left[{n_1+n_3-X \choose 2} + {n_1+n_3-X+1 \choose 2} + \cdots + {n_1 -1 \choose 2}\right]\\ &= S(X-n_3,0,0) = S(X-n_3,0,n_3). \end{align*} The equality holds only when $X=\frac{N}{2}$ or $n_1+n_3-X-1<2$. First, if $X=\frac{N}{2}$, then both forms have the same number of monochromatic edges. However, if we relabel the names of the colors, then both forms are the same. We will focus on the condition of $n_1+n_3-X<3$ where $3 \leqslant n_1 \leqslant n_2<n_3\leqslant X < \frac{N}{2}$ and $n_1+n_2>n_3$. Suppose that $n_1+n_3-X<3$. If $N$ is even, then $3 > n_1+n_2-X \geqslant n_1+n_3-\left(\frac{N}{2}-1\right)=\frac{n_1+n_3-n_2}{2}+1$. Thus, $n_1+n_3-n_2 <4$. Since $1 \leqslant n_3-n_2$, we have that $n_1<3$ which contradicts with $3\leqslant n_1$. If $N$ is odd, then $3 > n_1+n_2-X \geqslant n_1+n_3-\left(\frac{N-1}{2}\right)=\frac{n_1+n_3-n_2+1}{2}$. Thus, $n_1+n_3-n_2 <5$. Since $1 \leqslant n_3-n_2$, we have that $n_1<4$. Consequently, it is only possible when $n_1=3$. Note that $n_2 < n_3$ and $n_1 + n_2=3+n_2 > n_3$. For $N = 3+n_2+n_3$ to be odd, $n_2$ and $n_3$ must have the same parity. Thus, $n_3=n_2+2$ and $5>n_1+n_3-n_2=3+n_2+2-n_2=5$, which is a contradiction. Now, we have that $n_1+n_3-X \geqslant 3$. Hence, $F_8$ gives fewer number monochromatic edges than $F_6$ and $c^*_X$ is in the form $F_8$. To sum up, we have already determined (as shown in the table below) the canonical forms $c^*_X$ that has minimum number of monochromatic edges for each condition of hypergraphs and range of the number of red vertices $X$. \begin{center} \begin{tabular}{ |P{1cm}|P{3.5cm}|P{1.7cm}|P{1cm}|P{3.5cm}|P{1.7cm}| } \hline \multicolumn{6}{|c|}{List of best canonical forms} \\ \hline \multicolumn{3}{|c|}{Type A Hypergraphs $n_1+n_2 \leqslant n_3$} & \multicolumn{3}{|c|}{Type B Hypergraphs $n_1+n_2>n_3$} \\ \hline Cases&Number of red vertices & Canonical form&Cases &Number of red vertices&Canonical form\\ \hline $1$&$0 \leqslant X < n_1$& $F_1$&$1$&$0 \leqslant X < n_1$ &$F_1$\\ $2$&$n_1 \leqslant X < \frac{n_1+n_2}{2}$ &$F_4$&$2$&$n_1 \leqslant X < \frac{n_1+n_2}{2}$& $F_4$\\ $3$&$X = \frac{n_1+n_2}{2}$& $F_2$ or $ F_4$&$3$&$X = \frac{n_1+n_2}{2}$ &$F_2$ or $ F_4$\\ $ $&$\frac{n_1+n_2}{2} < X < n_2$& $F_2$&$ $&$\frac{n_1+n_2}{2} < X < n_2$ &$F_2$\\ $4A$&$n_2 \leqslant X < n_1 +n_2$& $F_6$&$4B$&$n_2 \leqslant X <n_3$&$F_3$ or $F_6$\\ $5A$&$n_1 + n_2 \leqslant X \leqslant \lfloor{\frac{N}{2}}\rfloor$&$F_{10}$&$5B$&$n_3 \leqslant X \leqslant \lfloor{\frac{N}{2}}\rfloor$ & $F_8$\\ \hline \end{tabular} \end{center} Note that, in \textit{Case 3}, $c^*_X$ is only in the form $F_2$ when $\frac{n_1+n_2}{2} < X < n_2$. However, if $X=\frac{n_1+n_2}{2}$, then $F_2$ and $F_4$ give the same number of monochromatic edges. Moreover, in \textit{Case 4B}, we have not compared the colorings in the form $F_3$ and $F_6$. The uniqueness of $c^*_X$ of the other cases will be also considered. We can see that the inequalities in some cases are equal when the sizes of some parts are equal which means that those canonical forms are equivalent. For example, in \textit{Case 1}, if $n_2=n_3$, $F_2$ is equivalent to $F_3$. The remaining inequalities are equal when $X$ is equal to some certain value, for example, in \textit{Case 5A} and \textit{Case 5B} when $X = \frac{N}{2}$; the colorings in the forms $F_3$ and $F_{10}$ are isomorphic up to a permutation of the name of colors and so are colorings in the forms $F_6$ and $F_8$ in \textit{Case 5B}. Hence, apart from \textit{Case 3} where $X=\frac{n_1+n_2}{2}$ and \textit{Case 4B}, a coloring with red vertices in the form, according to the previous table, has fewer monochromatic edges than other colorings with the same amount of red vertices and the uniqueness follows. Next, we will make comparisons between colorings with different values of $X$. We will show that any $c^*_X$ with $X\leqslant \lfloor{\frac{N}{2}}\rfloor -1$ has strictly more monochromatic edges than some colorings. We, hence, would like to show that $M(H,c^*_{X+1})-M(H,c^*_{X}) < 0$ for each $0 \leqslant X < \lfloor{\frac{N}{2}}\rfloor-1$. Note that, if $c^*_X$ and $c^*_{X+1}$ are in the same canonical form, we will consider $\displaystyle \bigtriangleup M(H,c^*_X)$ instead. Again, we will divide into several cases conforming to the value of $X$ and the type of $H$. \textit{Case 1}: $0 \leqslant X < n_1$. We have that $c^*_X$ is in the form $F_1$ with $x_1 = X$, $x_2 =0$ and $x_3=0$. Then,\begin{align*} \bigtriangleup_1M(H,c^*_X)= \left[ {X \choose 2} - {X \choose 2}\right] - \left[ {N-X-1 \choose 2} - {n_1-X-1 \choose 2}\right]\leqslant 0. \end{align*} The equality holds only when $N-X-1<2$, which is impossible. \textit{Case 2}: $n_1 \leqslant X < \frac{n_1 + n_2}{2}$. We have that $c^*_X$ is in the form $F_4$ with $x_1 = n_1$, $x_2 =X-n_1$ and $x_3=0$. Then, \begin{align*} \bigtriangleup_2M(H,c^*_X)= \left[ {X \choose 2}+{n_1+n_2-X-1 \choose 2}\right]-\left[ {X-n_1 \choose 2} + {N-X-1 \choose 2} \right]. \end{align*} We will show that $\bigtriangleup_2M(H,c^*_X)<0$ by Proposition 4. We have $X + (n_1+n_2-X-1) = n_1+n_2 -1 \leqslant n_3 +n_2 -1 =(X - n_1) +(N-X-1)$, $X-n_1 < X$ and $n_1+n_2-X-1 < N -X -1$. Since $2 \leqslant N-X-1$, we have that $\displaystyle \bigtriangleup_2M(H,c^*_X)<0$. \textit{Case 3}: $\frac{n_1 + n_2}{2} \leqslant X < n_2$. We have that $c^*_X$ is in the form $F_2$ or $F_4$. We will show that $\bigtriangleup_2M(H,c^*_X) < 0$ for both forms. If $X =\frac{n_1 + n_2}{2}$ and $ c^*_X$ is in the form $F_4$ with $x_1 = n_1$, $x_2=X-n_1$ and $x_3 = 0$, then $\bigtriangleup_2M(H,c^*_X) < 0$ similarly as in \textit{Case 2}. If $\frac{n_1 + n_2}{2} \leqslant X < n_2$ and $c^*_X$ is in the form $F_2$ with $x_1 = 0$, $x_2 =X$ and $x_3=0$, then \begin{align*} \bigtriangleup_2M(H,c^*_X) = \left[ {X \choose 2} - {X \choose 2}\right] - \left[ {N-X-1 \choose 2} - {n_2-X-1 \choose 2}\right]\leqslant 0. \end{align*} The equality holds only when $N-X-1<2$, which is impossible. Again, from this point, the cases will be divided by whether the hypergraph is of type A or B. \textit{Case 4A}: $n_2 \leqslant X < n_1+n_2$ and $H$ is a type A hypergraph. We have that $c^*_X$ is in the form $F_6$ with $x_1 = X-n_2$, $x_2 =n_2$ and $x_3=0$. Then, \begin{align*} \bigtriangleup_1M(H,c^*_X) =\left[{X\choose 2}+{n_1+n_2-X-1 \choose2} \right] -\left[ {X-n_2 \choose2}+{N-X-1\choose 2} \right]. \end{align*} We will show that $\displaystyle \bigtriangleup_1M(H,c^*_X)<0$ by Proposition 4. We have $X+(n_1+n_2-X-1)=n_1+n_2-1 \leqslant n_1+n_3-1=(X-n_2)+(N-X-1)$, $X-n_2 < X$ and $n_1+n_2-X-1<N-X-1$. Since $2 \leqslant N-X-1$, we have that $\displaystyle \bigtriangleup_1M(H,c^*_X)<0$. \textit{Case 5A}: $n_1+n_2 \leqslant X \leqslant \lfloor{\frac{N}{2}}\rfloor$ and $H$ is a type A hypergraph. We have that $c^*_X$ is in the form $F_{10}$ with $x_1 = n_1$, $x_2=n_2$ and $x_3 = X-n_1-n_2$. Then, \begin{align*} \bigtriangleup_3M(H,c^*_X)=\left[{X\choose 2} - {X-n_1-n_2 \choose2}\right] - \left[{N-X-1\choose 2} - {N-X-1 \choose2}\right] \geqslant 0. \end{align*} The equality holds only when $X<2$, which is impossible. This yields a contrary result: The number of monochromatic edges increases when the number of red vertices increases. Hence, $F_{10}$ gives the minimum number of monochromatic edges when $X=n_1+n_2$ instead. Next, we will consider the last two cases of type B hypergraphs. \textit{Case 4B}: $n_2 \leqslant X < n_3$ and $H$ is a type B hypergraph. We have that $c^*_X$ is in the form $F_3$ or $F_6$. We will show that $\bigtriangleup_3M(H,c^*_X) < 0$ for $F_3$ and $\bigtriangleup_1M(H,c^*_X) < 0$ for $F_6$. If $c^*_X$ is in the form $F_3$ with $x_1 = 0$, $x_2=0$ and $x_3 = X$, then \begin{align*} \bigtriangleup_3M(H,c^*_X) = \left[ {X \choose 2} - {X \choose 2}\right] - \left[ {N-X-1 \choose 2} - {n_3-X-1 \choose 2}\right]\leqslant 0. \end{align*} Note that the inequality is equal when $N-X-1<2$ which is impossible. Next, if $c^*_X$ is in the form $F_6$ with $x_1 = X-n_2$, $x_2=n_2$ and $x_3 = 0$, then \begin{align*} \bigtriangleup_1M(H,c^*_X) = \left[{X \choose 2} +{n_1 +n_2-X-1\choose 2} \right]-\left[{N-X-1 \choose 2} + {X-n_2 \choose 2}\right]. \end{align*} We will show that $\displaystyle \bigtriangleup_1M(H,c^*_X) < 0$ by Proposition 4. We have $X+(n_1+n_2-X-1) = n_1+n_2 -1 \leqslant n_1+n_3 -1 = (N-X-1) +(X-n_2) $, $X > X-n_2$ and $n_1 +n_2-X-1 < N-X-1$. Since $2\leqslant N-X-1$, we have that $\displaystyle \bigtriangleup_1M(H,c^*_X) < 0$. \textit{Case 5B}: $n_3 \leqslant X \leqslant \lfloor{\frac{N}{2}}\rfloor $ and $H$ is a type B hypergraph. We have that $c^*_X$ is in the form $F_{8}$ with $x_1 = X-n_3$, $x_2=0$ and $x_3 = n_3$. Then, \begin{align*} \bigtriangleup_1M(H,c^*_X)&=\left[{X\choose 2} - {X-n_3 \choose2}\right] - \left[{N-X-1\choose 2} - {n_1-(X-n_3)-1 \choose2}\right]\\ &= \frac{n_1^2+2n_1n_3+2X(N-n_1)-3n_1+3N-N^2-4n_3}{2}. \end{align*} Since we cannot apply any lemmas to $\bigtriangleup_1M(H,c^*_X)$, we expand the binomial coefficient terms and see when $\displaystyle \bigtriangleup_1M(H,c^*_X)$ is fewer than $0$. Consequently, $\displaystyle \bigtriangleup_1M(H,c^*_X) < 0$ if and only if \[X< \left \lceil\frac{N^2-3N-n_1^2-2n_1n_3+3n_1+4n_3}{2(N-n_1)}\right\rceil.\] Write $X'$ for $\left\lceil\frac{N^2-3N-n_1^2-2n_1n_3+3n_1+4n_3}{2(N-n_1)}\right\rceil$. We have that the $c^*_{X'}$ has the minimum number of monochromatic edges among all colorings in the form $F_8$ and we will show that $X'\leqslant\frac{N}{2}$. \begin{align*} X'=\left\lceil\frac{N^2-3N-n_1^2-2n_1n_3+3n_1+4n_3}{2(N-n_1)}\right\rceil \leqslant \frac{N^2 -Nn_1- 2n_3+2n_3}{2(N-n_1)}= \frac{N}{2}. \end{align*} Hence, we have shown all the comparisons between colorings and we can conclude that: \begin{enumerate} \item If $H$ is a type A hypergraph, then the coloring with $n_1 + n_2$ red vertices in the form $F_{10}$ has the minimum number of monochromatic edges. \item If $H$ is a type B hypergraph, then the coloring with $X'$ red vertices in the form $F_8$ has the minimum number of monochromatic edges. \end{enumerate} We have already proved that those minimum colorings are the unique colorings that has the minimum number of monochromatic edges among colorings with the same red vertices. Furthermore, we have shown that $\bigtriangleup M(H,c^*_X)$ is fewer than zero when $X$ is fewer than $n_1+n_2$ in a type A hypergraph and when $X$ is less than $X'$ in a type B hypergraph. Hence, those minimum colorings are the unique colorings that has minimum number of monochromatic edges among all colorings. \subsection{Hypergraphs with $n_1 < 3$ or $n_2 < 3$} In this subsection, we will prove the remaining cases which are unbalance complete tripartite hypergraphs with some classes smaller than $3$. These cases are easy and straightforward but contain fuzzy details. First, we will consider an unbalanced complete tripartite $3$-uniform hypergraphs with $n_1 \leqslant n_2 < n_3 \geqslant 3$. There are three possibilities for these hypergraphs: \textit{Case i}: $n_1 = n_2 =1$ and $n_3 \geqslant 3$. Since we have that the first two classes are smaller than $3$, no edge can be contained in the first two classes. Thus, \[M(H,c)={X\choose 3}+{N-X\choose 3}-{x_3 \choose 3}-{n_3-x_3\choose3}.\] We have that $M(H,c) \geqslant 0$. Suppose that $M(H,c)=0$. Since $X \geqslant x_3$ and $N-X \geqslant n_3-x_3$, we have that ${X\choose 3}={x_3 \choose 3}$ and ${N-X\choose 3}={n_3-x_3\choose3}$. Since $N=n_1+n_2+n_3=1+1+n_3\geqslant 5$, at least $3$ vertices are colored the same say red, i.e., $X\geqslant 3$. Then, $X=x_3$ which implies that $N-X=n_3+2-X>n_3-x_3$. Consequently, $3 > N-X = n_3+2-x_3$, i.e., $n_3=x_3$. This means that $c$ is a coloring such that the third class contains all red vertices and no blue vertex and the first two classes are all blue. Hence, we have already determined the minimum coloring and showed that it is unique up to a permutation of colors and classes. \textit{Case ii}: $n_1=1, n_2 =2$ and $n_3 \geqslant 3$. Again, no edge can be contained in the first two classes. Thus, \[M(H,c)={X\choose 3}+{N-X\choose 3}-{x_3 \choose 3}-{n_3-x_3\choose3}.\] We will show that $M(H,c) \geqslant 1$. We have that $X \geqslant x_3$ and $N-X \geqslant n_3-x_3$. Since $N=n_1+n_2+n_3=1+1+n_3\geqslant 6$, at least $3$ vertices are colored the same say red, i.e., $X\geqslant 3$. If $X>x_3$, then $M(H,c)>0$. Suppose that $X=x_3$. Then, $N-X=n_3+3-x_3>n_3-x_3 \geqslant 3$. Hence, $M(H,c)>0$. Next, suppose that $M(H,c)=1$. If $X>x_3$, then \begin{align*} 1\leqslant {X-1 \choose 2} \leqslant{X\choose 3}-{x_3\choose 3}\leqslant{X\choose 3}+{N-X\choose 3}-{x_3 \choose 3}-{n_3-x_3\choose3}= M(H,c). \end{align*} This implies that $X=3$ and so $N-X \geqslant 3$, moreover, ${N-X\choose 3}={n_3-x_3\choose3}$. Hence, $N-X=n_3-x_3$. Now, we have $N-3=N-X=n_3-x_3=N-3-x_3$, i.e., $x_3=0$. This means $c$ is a coloring such that the third class contains no red vertices and the first two classes are all red. Suppose that $X=x_3$. Then, \begin{align*} 1=M(H,c)={X\choose 3}+{N-X\choose 3}-{x_3 \choose 3}-{n_3-x_3\choose3}={n_3-x_3+3\choose 3}-{n_3-x_3\choose3}. \end{align*} It is only possible when $n_3-x_3=0$ or $c$ is a coloring such that the third class contains all red vertices and no blue vertex and the first two classes are all blue. Note that if we relabel the names of colors, then both colorings are the same. Hence, we have already determined the minimum coloring and showed that it is unique up to a permutation of colors and classes. \textit{Case iii}: $n_1=2, n_2 =2$ and $n_3 \geqslant 3$. Similarly, no edges can be contained in the first two classes. Thus, \[M(H,c)={X\choose 3}+{N-X\choose 3}-{x_3 \choose 3}-{n_3-x_3\choose3}.\] We will show that $M(H,c) \geqslant 4$. We have that $X \geqslant x_3$ and $N-X \geqslant n_3-x_3$. If there is a color, say red, such that all vertices of that color are only in the third class, then, we have $X=x_3$ and $N-X=n_3+4-x_3\geqslant 4$. Thus, \begin{align*} M(H,c)&= {N-X\choose 3}-{N-X-4\choose3}\\ &={N-X-1 \choose 2}+{N-X-2 \choose 2}+{N-X-3 \choose 2}+{N-X-4 \choose 2}\\ &\geqslant {3 \choose 2}+{2 \choose 2} = 4. \end{align*} The equality holds only when $N-X=4$, i.e., $n_3=x_3$. Hence, if $M(H,c)=4$, then $c$ is a colorings such that the third class contains all red vertices but no blue vertex and the first two classes are all blue. Suppose that there is no color such that all vertices of that color are only in the third class, i.e., $X>x_3$ and $N-X>n_3-x_3$. Since $N=n_1+n_2+n_3=2+2+n_3\geqslant 7$, at least $4$ vertices are colored the same say red, i.e., $X\geqslant 4$. If $X \geqslant 5$, then \begin{align*} M(H,c)\geqslant {X\choose 3}-{x_3\choose3}\geqslant {X\choose 3}-{X-1\choose3}={X-1 \choose 2} \geqslant {4 \choose 2} = 6. \end{align*} If $X=4$, then $N-X \geqslant 3$ and \begin{align*} M(H,c)\geqslant {X\choose 3}-{x_3\choose3} + 1\geqslant {X\choose 3}-{X-1\choose3}+1={X-1 \choose 2}+1= 4. \end{align*} The equality holds only when $N-X =3$ and $x_3=X-1$, i.e., $N=7$ and $x_3=3$. This means that $n_3=3$. Hence, if $M(H,c)=4$, then $c$ is a coloring of $H$ which has $7$ vertices such that the third class is all red, the second class is all blue and the first class has one red and one blue vertices. This implies that when $n_3=3$, minimum colorings are not unique. Finally, the last class is a hypergraph such that only the first class is smaller than $3$. \textit{Case iv}: $n_1<3$ and $3 \leqslant n_2 \leqslant n_3$. Fortunately, this case conforms to almost all cases in the previous subsection since we assume that $n_2 \geqslant 3$. However, there are two points that we use the fact that $n_1\geqslant 3$. The first one is in the last part of \textit{Case 3} where we compare and determine which form of $F_2$ and $F_4$ has fewer monochromatic edges. Since we do not have that $n_1\geqslant 3$, it is possible that both forms have the same number of monochromatic edges. This is not problematic because both of them have strictly more monochromatic edges than some coloring according to the next comparisons. The next case is in the last part of \textit{Case 5B} where we compare and determine which form of $F_6$ and $F_8$ has fewer monochromatic edges. Again, since we do not have that $n_1\geqslant 3$, we may have that $n_1+n_3-X<3$ where $n_2<n_3\leqslant X < \frac{n_1+n_2+n_3}{2}$ and $n_1+n_2>n_3$. If this condition occurs, we will have that $F_6$ and $F_8$ have the same number of monochromatic edges. Since $n_1<3$, the condition is possible only when ($N$ is even and $n_1<3$) or ($N$ is odd and $n_1 \leqslant 3$). Suppose that $N$ is even. If $n_1=1$, then we have that $n_3 < n_1+n_2 = n_2+1$. Since $n_2<n_3$, there is no choice for $n_3$. If $n_1=2$, then we have that $n_3 < n_1+n_2= n_2+2$. Since $n_2<n_3$, $n_3=n_2+1$. However, for $N$ to be even, $n_2$ and $n_3$ must have different parity, which is impossible. Suppose that $N$ is odd. Again, it is impossible for $n_1=1$. Since $n_1<3$, we have $n_1=2$. Similarly, $n_3=n_2+1$ and $N=n_1+n_2+n_3=2+n_2+n_2+1=2(n_2+1)+1$ which is odd. Consequently, the condition is possible only here and $F_6$ and $F_8$ have the same number of monochromatic edges. Hence, if $H$ is a hypergraph with $n_1=2$ and $ 3\leqslant n_2=n_3-1$, then there are only two colorings (each unique up to a permutation of colors and classes) that have minimum number of monochromatic which are colorings with $X=\lfloor{\frac{N}{2}}\rfloor$ in the form $F_6$ and $F_8$. Now, we have determined the minimum colorings of all unbalanced complete tripartite $3$-uniform hypergraphs. \end{proof} \section{Proof of Theorem~\ref{thm:3}} \label{sec:thm:3} \begin{proof}[Proof of Theorem~\ref{thm:3}] Assume that $ k\geqslant 2$. Let $H$ be a balanced complete $k$-partite $(r+1)$-uniform hypergraph with $n \geqslant r+1$ vertices in each class and let $N=kn$. Let $c$ be a red/blue/green coloring of $H$ with the numbers of red, blue and green vertices of the $i^{th}$ class equal to $r_i, b_i$ and $g_i$, respectively, and let $R$, $B$ and $G$ be the total numbers of red, blue and green vertices, respectively. Let $\bigtriangleup_{i{i'}} M(H,c,r,b)$ be the change in the number of monochromatic edges if a red vertex in the $i^{th}$ class is recolored into blue and a blue vertex in the ${i'}^{th}$ class is recolored into red. The definitions are similar for other colors combinations. The process will be called a \textit{swapping} which results in a new coloring, say $c'$. As a result of the process, the number of red vertices in the $i^{th}$ class decreases by $1$ and the number of red vertices in the ${i'}^{th}$ class increases by $1$ while the total number of red vertices remains the same. In other words, the coloring $c'$ has $r_i - 1$ and $r_{i'}+1$ red vertices in the $i^{th}$ and ${i'}^{th}$ classes, respectively. Likewise, the coloring $c'$ has $b_i+1$ and $b_{i'}-1$ blue vertices in the $i^{th}$ and ${i'}^{th}$ classes, respectively. We can compute $\bigtriangleup_{i{i'}} M(H,c,r,b)$ by comparing the numbers of monochromatic edges containing those vertices undergone swapping before and after the swapping process. Thus, \begin{align*} \bigtriangleup_{i{i'}}M(H,c,r,b)&= \left[ {B-1 \choose r}-{b_i \choose r} + {R-1 \choose r}- {r_{i'}\choose r} \right]\\ &\;\;\;\;- \left[ {R-1 \choose r}-{r_i-1 \choose r} + {B-1 \choose r} - {b_{i'}-1\choose r} \right]\\ &= \left[ {r_i-1 \choose r} + {b_{i'}-1\choose r} \right] - \left[ {b_i \choose r} + {r_{i'}\choose r} \right]. \end{align*} A \textit{successful swapping} is a swapping in such a way that the number of monochromatic edges is reduced, i.e., $\bigtriangleup_{i{i'}} M(H,c,r,b) < 0$. Note that if $0<r_i<n$, $<b_i<n$, $r_{i'}=n-1$ and $b_{i'}=1$, then \begin{align*} \bigtriangleup_{i{i'}}M(H,c,r,b)&= \left[ {r_i-1 \choose r} + {b_{i'}-1\choose r} \right] - \left[ {b_i \choose r} + {r_{i'}\choose r} \right]\\ &=\left[ {r_i-1 \choose r} + {1-1\choose r} \right] - \left[ {b_i \choose r} + {n-1\choose r} \right]. \end{align*} Since $r_i<n$, $0<b_i$ and $n-1 \geqslant r$, we have that $\bigtriangleup_{i{i'}}M(H,c,r,b)<0$. This implies that a swapping resulting in fewer polychromatic classes is always successful. \begin{lemma} If $\bigtriangleup_{i{i'}} M(H,c,\text{r},\text{b}) \leqslant 0$, then $\bigtriangleup_{i{i'}} M(H,c',\text{r},\text{b}) \leqslant 0$. \end{lemma} The proof of this lemma is similar to that of Lemma 8. Lemma 10 means that if a swapping can be done without increasing the number of monochromatic edges, another swapping in the same direction will be successful (if there is a red and blue vertices to be swapped). The process of successful swappings will terminate when the $i^{th}$ class has no red vertex or the ${i'}^{th}$ class has no blue vertex. \begin{lemma} If $\bigtriangleup_{i{i'}} M(H,c,\text{r},\text{b}) \geqslant 0$, then $\bigtriangleup_{i{i'}} M(H,c,\text{b},\text{r}) \leqslant 0$. \end{lemma} The proof of this lemma is similar to that of Lemma 9. However, in contrast to Lemma $9$, it is possible that the equality holds. Note that if $c$ contains two classes, the $i^{th}$ and ${i'}^{th}$, such that both of them contain at least one red vertex and one blue vertex, then there are two directions of swapping as follows: \begin{enumerate}\item Swapping a red vertex of the $i^{th}$ class with a blue vertex of the ${i'}^{th}$ class. \item Swapping a blue vertex of the $i^{th}$ class with a red vertex of the ${i'}^{th}$ class. \end{enumerate} By Lemma 11, one of the two directions can be achieved without increasing the number of monochromatic edges. Moreover, by Lemma 10, we can continue swapping in the same direction until the $i^{th}$ class has no red vertex or the ${i'}^{th}$ class has no blue vertex and the number of monochromatic edges does not increase. Note that we get the same result when considering a swapping relating to other color combinations. Hence, the coloring with minimum number of monochromatic edges among colorings with constant number of red, blue and green vertices, is the coloring such that, for any two classes, they must have at most one color of vertices to be in common. We will list all these forms in the following table. \begin{center} \begin{tabular}{ |P{3cm}|P{10cm}| } \hline Canonical forms & Descriptions \\ \hline $F_1$& The first class contains three colors while the other classes are monochromatic.\\ \hline $F_2$& The first class contains a pair of colors while the other classes are monochromatic.\\ \hline $F_3$& The first and second classes contain different pairs colors while the other classes are monochromatic.\\ \hline $F_4$& The first, second and third classes contain different pairs colors while the other classes are monochromatic.\\ \hline $F_5$& All classes are monochromatic.\\ \hline \end{tabular} \end{center} The first column illustrates the list of $5$ canonical forms and the second column describes the colors of vertices in each class. Suppose that $r_i \leqslant b_{i'}$. Let $\bigtriangleup_{i{i'}} M_T(H,c,r,b)$ be the change in the number of monochromatic edges if all red vertices in the $i^{th}$ class are recolored into blue and $r_i$ blue vertices in the ${i'}^{th}$ class are recolored into red. The process will be called \textit{total swapping} which results in a new coloring, say $c'$. We can compute $\bigtriangleup_{i{i'}} M_T(H,c,r,b)$ by summing the change in the number of monochromatic edges of the swappings. Thus, \begin{align*} \bigtriangleup_{i{i'}} &M_T(H,c,r,b)= \sum_{k=0}^{r_i-1}\left[ {r_i-1-k \choose r} + {b_{i'}-1-k\choose r} \right] - \left[ {b_i+k\choose r} + {r_{i'}+k\choose r} \right]\\ &=\left[{0 \choose r}+{1 \choose r}+\cdots+{r_i-1 \choose r}\right]+\left[{b_{i'}-r_i \choose r}+{b_{i'}-r_i+1 \choose r}+\cdots+{b_{i'}-1 \choose r}\right] \\ &\;\;\;\;-\left[{b_i \choose r}+{b_i+1 \choose r}+\cdots+{b_i+r_i-1 \choose r}\right]-\left[{r_{i'} \choose r}+{r_{i'}+1 \choose r}+\cdots+{r_{i'}+r_i-1 \choose r}\right]. \end{align*} A \textit{successful total swapping} is a total swapping in such a way that the number of monochromatic edges is reduced. \begin{lemma} Suppose that $1 \leqslant r_i \leqslant b_{i'}$. Then, $\bigtriangleup_{i{i'}} M_T(H,c,r,b)\leqslant0$ if $c$ satisfies at least one of the following conditions: \begin{enumerate} \item $b_{i'} < r_i+b_i$. \item $b_{i'} < r_i+r_{i'}$. \end{enumerate} \end{lemma} \begin{proof} Suppose that $ 1 \leqslant r_i \leqslant b_{i'}$. If $b_{i'} < r_i+b_i$, then \begin{align*} \bigtriangleup_{i{i'}} &M_T(H,c,r,b)\\ &\leqslant\left[{0 \choose r}+{1 \choose r}+\cdots+{r_i-1 \choose r}\right]-\left[{r_{i'} \choose r}+{r_{i'}+1 \choose r}+\cdots+{r_{i'}+r_i-1 \choose r}\right]\leqslant 0. \end{align*}The equality holds only when $b_i+r_i-1<r$ and $r_{i'}+r_i-1<r$, i.e., $b_i+r_i<r+1$ and $r_i+r_{i'}<r+1$. If $b_{i'} < r_i+r_{i'}$, then \begin{align*} \bigtriangleup_{i{i'}} M_T(H,c,r,b)& \leqslant\left[{0 \choose r}+{1 \choose r}+\cdots+{r_i-1 \choose r}\right]-\left[{b_i \choose r}+{b_i+1 \choose r}+\cdots+{b_i+r_i-1 \choose r}\right]\\ &\leqslant 0. \end{align*}The equality holds only when $b_i+r_i-1<r$ and $r_{i'}+r_i-1<r$, i.e., $b_i+r_i<r+1$ and $r_i+r_{i'}<r+1$. \end{proof} Note that if we consider a class with only two colors say $r_i+b_i=n \geqslant r+1$, then $\bigtriangleup_{i{i'}} M_T(H,c,r,b)$ is strictly less than zero in both cases. We will use Lemma $12$ to get more information from the canonical forms. From this point, the term \textit{quadruple} refers to a collection of $4$ values which are the numbers of vertices of any pair of colors in any pair of classes, e.g. $(r_i,b_i,r_{i'},b_{i'})$. Consequently, in each coloring in the forms any canonical forms, all quadruples contain at least one zero. Suppose that $c$ contains the $i^{th}$ and ${i'}^{th}$ classes such that they have a color in common but the $i^{th}$ class contains a color that the ${i'}^{th}$ class does not, say we focus on $(r_i\neq 0,b_i\neq 0,r_{i'} =0,b_{i'}\neq 0)$. If $c$ is a minimum coloring with $r_i+b_i=n$, then we can conclude by Lemma $12$ that \begin{enumerate} \item if $r_i \leqslant b_{i'}$, then $b_{i'} \geqslant r_i+b_i$ and $b_{i'} \geqslant r_i +r_{i'}=r_i$, and \item if $r_i \geqslant b_{i'}$, then $r_i \geqslant b_i + b_{i'}$ and $r_i \geqslant r_{i'}+b_{i'}=b_{i'} $. \end{enumerate} We will call these \textit{quadruple conditions}. Next, we will focus on the possibility of $F_4$. Suppose that $c$ is the coloring that has minimum number of monochromatic edges which is in the form $F_4$ such that, WLOG, the first class has $g_1 \neq 0$ green and $r_1 \neq 0$ red vertices, the second class has $g_2 \neq 0$ green and $b_2 \neq 0$ blue vertices and the third class has $r_3 \neq 0$ red and $b_3 \neq 0$ blue vertices, where $g_1$ is the maximum among those values. We will apply quadruple conditions to the quadruple $(g_1,b_1=0,g_2,b_2)$. Since $g_1 \geqslant b_2$ and $g_2+b_2=n$, $n = g_2 + b_2 \leqslant g_1 = n - r_1 < n$, which is a contradiction. Hence, an minimum coloring cannot be in the form $F_4$ and this form will be out of our interest. Consequently, we will search for the minimum colorings only from $F_1, F_2, F_3$ and $F_5$. We will divide this into $2$ cases upon the remainder of the number $k$ of vertex classes divided by $3$. For simplicity, we define a \textit{type i} hypergraph to be a hypergraph with $k\equiv i \ (\textrm{mod}\ 3)$ classes for $i=0,1,2$. We will consider a type 0 hypergraph first. \textit{Case 0}: $H$ is a type 0 hypergraph. Since the number of classes is divisible by the number of colors, it is done by Proposition 7. \textit{Case 1}: $H$ is a type 1 hypergraph. \textit{Case 1.1}: The coloring $c$ is in the form $F_3$. In this case, our aim is to show that all colorings in the form $F_3$ have more monochromatic edges than some coloring in the form $F_2$. We have that $c$ has two classes that are polychromatic, say the first class contains green and red vertices and the second class contains green and blue vertices, while the rest classes are monochromatic. The number of vertices of each color and the number of monochromatic classes of each color are shown in the table below. \begin{center} \begin{tabular}{ |P{1.3cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}| } \hline \multicolumn{4}{|c|}{Polychromatic Classes} & \multicolumn{3}{|c|}{Monochromatic Classes} \\ \hline Classes&Red vertices & Blue vertices&Green vertices &Red classes &Blue classes&Green classes\\ \hline $1$&$r_1 \neq 0$& $0$&$g_1 \neq 0$&$k_r$ &$k_b$&$k_g$\\ $2$&$0$ &$b_2 \neq 0$&$g_2 \neq 0$&$ $& $ $& $ $\\ \hline \end{tabular} \end{center} We will show that the coloring such that $k_r$, $k_b$ and $k_g$ are as equal as possible has smaller number of monochromatic edges than the coloring those of which are not. WLOG, suppose that $c$ has $k_g$ green classes and $k_r$ red classes such that $k_g - k_r \geqslant 2$. We will recolor all vertices in a green class into red and get a new coloring $c'$. Thus, \begin{align*} M(H,c)&= { nk_r+ r_1 \choose r+1}+ { nk_b+ b_2 \choose r+1}+ { nk_g+ g_1+g_2 \choose r+1}\\ &\;\;\;\;-{r_1 \choose r+1}-{b_2 \choose r+1}-{g_1 \choose r+1}-{g_2 \choose r+1} -(k-2){n \choose r+1} \end{align*}and \begin{align*} M(H,c')&= { n(k_r+1)+ r_1 \choose r+1}+ { nk_b+ b_2 \choose r+1}+ { n(k_g-1)+ g_1+g_2 \choose r+1}\\ &\;\;\;\;-{r_1 \choose r+1}-{b_2 \choose r+1}-{g_1 \choose r+1}-{g_2 \choose r+1} -(k-2){n \choose r+1}. \end{align*}Then, \begin{align*} M(H,c')-M(H,c)&= \left[ { n(k_r+1)+ r_1 \choose r+1}+ { n(k_g-1)+ g_1+g_2 \choose r+1}\right]\\&\;\;\;\;-\left[{ nk_r+ r_1 \choose r+1}+ { nk_g+ g_1+g_2 \choose r+1}\right]. \end{align*} We will show that $M(H,c')-M(H,c) < 0$ by Proposition 4. We have $\left[n(k_r+1)+ r_1\right]+\left[n(k_g-1)+ g_1+g_2\right] = \left[nk_r+ r_1\right]+ \left[nk_g+ g_1+g_2\right]$, $nk_r+ r_1 < n(k_r+1)+ r_1 \leqslant n(k_g-1)+r_1 < nk_g+ g_1+g_2$ and $nk_r+ r_1 < n(k_r+1)+ g_1+g_2 \leqslant n(k_g-1)+g_1+g_2 < nk_g+ g_1+g_2$. Since $nk_g+g_1+g_2>r+1$, we have that $M(H,c')-M(H,c) < 0$. Note that we only use the fact that $r_1 < n$ and $g_1+g_2 >0$. Since $(k-2)\equiv 2 \ (\textrm{mod}\ 3)$, colorings such that $k_r$, $k_b$ and $k_g$ are as equal as possible must have one of these conditions: \begin{enumerate} \item $k_r+1=k_b=k_g$, \item $k_r=k_b+1=k_g$, \item $k_r=k_b=k_g+1$. \end{enumerate} We will show that a coloring with the last condition has more number of monochromatic edges than a coloring with either of the first two conditions (with the same polychromatic classes). Suppose that $c$ is a coloring such that $k_r=k_b=k_g+1$. We consider $(g_1,r_1,g_2,0)$ of $c$. Since $g_1+r_1=n$, if $g_2 \geqslant r_1$, then $g_2 \geqslant r_1+g_1$. This leads to a contradiction since $n = r_1+g_1 \leqslant g_2 = n-b_2 <n$. Hence, $g_2 < r_1$ and consequently $g_1+g_2 \leqslant r_1$ by quadruple conditions. Next, we will increase the number of green class. WLOG, we will recolor all vertices in a red class into green and get a new coloring $c'$ with condition (1). Then, \begin{align*} M(H,c)&= { nk_r+ r_1 \choose r+1}+ { nk_b+ b_2 \choose r+1}+ { nk_g+ g_1+g_2 \choose r+1}\\ &\;\;\;\;-{r_1 \choose r+1}-{b_2 \choose r+1}-{g_1 \choose r+1}-{g_2 \choose r+1} -(k-2){n \choose r+1} \end{align*}and \begin{align*} M(H,c')&= { n(k_r-1)+ r_1 \choose r+1}+ { nk_b+ b_2 \choose r+1}+ { n(k_g+1)+ g_1+g_2 \choose r+1}\\ &\;\;\;\;-{r_1 \choose r+1}-{b_2 \choose r+1}-{g_1 \choose r+1}-{g_2 \choose r+1} -(k-2){n \choose r+1}. \end{align*}Then, \begin{align*} M(H,c')-M(H,c)&= \left[ { n(k_r-1)+ r_1 \choose r+1}+ { n(k_g+1)+ g_1+g_2 \choose r+1}\right]\\&\;\;\;\;-\left[{ nk_r+ r_1 \choose r+1}+ { nk_g+ g_1+g_2 \choose r+1}\right]. \end{align*} We will show that $M(H,c')-M(H,c) < 0$ by Proposition $4$. We have $\left[n(k_r-1)+ r_1\right]+\left[n(k_g+1)+ g_1+g_2\right] =\left[ nk_r+ r_1\right]+ \left[nk_g+ g_1+g_2\right]$. Since $g_1+g_2 \leqslant r_1$, then $nk_g+ g_1+g_2 \leqslant nk_g +r_1 = n(k_r-1) +r_1 < nk_r+r_1$ and $nk_g+ g_1+g_2 < n(k_g+1)+ g_1+g_2 \leqslant n(k_g+1)+r_1 = nk_r+ r_1$. Since $nk_g+g_1+g_2>r+1$, we have that $M(H,c')-M(H,c) < 0$. Finally, we will consider a coloring with condition (1) or (2). By symmetry, assume that $c$ is a coloring with $k_r+1=k_b=k_g$. We will recolor a green vertex in the first class of $c$ into red. Then, \begin{align*} \bigtriangleup_1M(H,c) = \left[ {nk_r+r_1 \choose r}- {nk_g+g_1+g_2-1 \choose r} \right] + \left[{g_1-1 \choose r} -{r_1 \choose r} \right]. \end{align*} We have $nk_r+r_1 = n(k_g-1)+r_1 < nk_g < nk_g+g_1+g_2-1$ and $g_1-1<g_1+g_2 \leqslant r_1$. Since $nk_g+g_1+g_2-1>r$, we have that $\bigtriangleup_1M(H,c) < 0$. This is true for any arbitrary value of $r_1 > 0$. Hence, we will recolor a green vertex into red until the first class is a red class which is a coloring in the form $F_2$ and get fewer number of monochromatic edges. Consequently, we can conclude that $c$ has more monochromatic edges than a coloring in the form $F_2$. \textit{Case 1.2}: The coloring $c$ is in the form $F_5$. In this case, our aim is to show that all colorings in the form $F_5$ have more monochromatic edges than some coloring in the form $F_2$. We have that all classes of $c$ are monochromatic with $k_r$, $k_b$ and $k_g$ red, blue and green classes, respectively. We can show that the coloring that $k_r$, $k_b$ and $k_g$ are as equal as possible has smaller number of monochromatic edges similarly as in \textit{Case 1.1}. Since $k\equiv 1 \ (\textrm{mod}\ 3)$, colorings that $k_r$, $k_b$ and $k_g$ are as equal as possible must have one of these conditions: \begin{enumerate} \item $k_r-1=k_b=k_g$, \item $k_r=k_b-1=k_g$, \item $k_r=k_b=k_g-1$. \end{enumerate} By symmetry, suppose that $c$ is a coloring such that $k_r-1=k_b=k_g$. We will show that if a red vertex in a red class, say the first class, is recolored into green, the number of monochromatic edges will decrease. The new coloring is not in the form $F_5$ but in the form $F_2$ instead. Then, \begin{align*} \bigtriangleup_1M(H,c) = \left[ {nk_g \choose r}+{n-1 \choose r} \right] - \left[ {nk_r-1 \choose r} +{0 \choose r} \right]. \end{align*} we will show that $\bigtriangleup_1M(H,c) < 0$ by Proposition $4$. We have $nk_g+ (n-1)= n(k_g+1) -1 = (nk_r-1)+0$ and $0 < n-1 < nk_g <nk_r-1$. Since $nk_r-1>r$, we have that $\bigtriangleup_1M(H,c) < 0$. Consequently, we can conclude that $c$ has more monochromatic edges than a coloring in the form $F_2$. \textit{Case 1.3}: The coloring $c$ is in the form $F_2$. In this case, our aim is to show that all colorings in the form $F_2$ have more monochromatic edges than some coloring in the form $F_1$. We have that $c$ has a class that is polychromatic, WLOG, say the first class contains red and blue vertices while the rest classes are monochromatic. Note that the number of vertices of each color and the number of monochromatic classes of each color are shown in the table below. \begin{center} \begin{tabular}{ |P{1.3cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}| } \hline \multicolumn{4}{|c|}{Polychromatic Classes} & \multicolumn{3}{|c|}{Monochromatic Classes} \\ \hline Classes&Red vertices & Blue vertices&Green vertices &Red classes &Blue classes&Green classes\\ \hline $1$&$r_1 \neq 0$& $b_1 \neq 0$&$0$&$k_r$ &$k_b$&$k_g$\\ \hline \end{tabular} \end{center} We can show that the coloring such that $k_r$, $k_b$ and $k_g$ are as equal as possible has smaller number of monochromatic edges similarly as in \textit{Case 1.1}. Hence, we will focus on a coloring that $k_r$, $k_b$ and $k_g$ are as equal as possible. Since $(k-1)\equiv 0 \ (\textrm{mod}\ 3)$, colorings such that $k_r$, $k_b$ and $k_g$ are as equal as possible must have $k_r=k_b=k_g$. Suppose that $c$ is a coloring such that $k_r=k_b=k_g$. By symmetry, suppose that $r_1 > 1$. We will show that if a red vertex in the first class is recolored into green, the number of monochromatic edges will decrease. The new coloring is not in the form $F_2$ but in the form $F_1$ instead. Then, \begin{align*} \bigtriangleup_1M(H,c) &= \left[ {nk_g \choose r} - {0 \choose r}\right] - \left[ {nk_r+r_1-1 \choose r} - {r_1-1 \choose r}\right]\\ &= \left[ {nk_g \choose r}+{r_1-1 \choose r} \right] - \left[ {nk_r +r_1-1 \choose r} +{0 \choose r} \right]. \end{align*} We will show that $\bigtriangleup_1M(H,c) < 0$ by Proposition $4$. We have $nk_g+ (r_1-1)= (nk_r +r_1 -1)+0$ and $0 < r_1-1 < nk_g <nk_r+r_1-1$. Since $nk_r+r_1-1>r$, we have that $\bigtriangleup_1M(H,c) < 0$. Consequently, we can conclude that $c$ has more monochromatic edges than a coloring in the form $F_1$. \textit{Case 1.4}: The coloring $c$ is in the form $F_1$. In this case, our aim is to show that a coloring in the form $F_1$ such that the numbers of red, blue and green vertices are as equal as possible has the minimum number of monochromatic edges. We have that $c$ has a class that is polychromatic, WLOG, say the first class contains red, blue and green vertices while the rest classes are monochromatic. Note that the number of vertices of each color and the number of monochromatic classes of each color are shown in the table below. \begin{center} \begin{tabular}{ |P{1.3cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}|P{1.4cm}| } \hline \multicolumn{4}{|c|}{Polychromatic Classes} & \multicolumn{3}{|c|}{Monochromatic Classes} \\ \hline Classes&Red vertices & Blue vertices&Green vertices &Red classes &Blue classes&Green classes\\ \hline $1$&$r_1 \neq 0$& $b_1 \neq 0$&$g_1 \neq 0$&$k_r$ &$k_b$&$k_g$\\ \hline \end{tabular} \end{center} We can show that the coloring that $k_r$, $k_b$ and $k_g$ are as equal as possible has smaller number of monochromatic edges similarly as in \textit{Case 1.1}. Since $(k-1)\equiv 0 \ (\textrm{mod}\ 3)$, colorings such that $k_r$, $k_b$ and $k_g$ are as equal as possible must have $k_r=k_b=k_g$. Suppose that $c$ is a coloring such that $k_r=k_b=k_g$. We will show that if we recolor the vertices in such a way that $r_1, b_1$ and $g_1$ are as equal as possible then the number of monochromatic edges will decrease. WLOG, assume that $r_1-g_1 \geqslant 2$. We will show that if a red vertex in the first class is recolored into green, the number of monochromatic edges will decrease. Then, \begin{align*} \bigtriangleup_1M(H,c) &= \left[ {nk_g +g_1 \choose r} - {g_1 \choose r}\right] - \left[ {nk_r+r_1-1 \choose r} - {r_1-1 \choose r}\right]\\ &= \left[ {nk_g +g_1 \choose r}+{r_1-1 \choose r} \right] - \left[ {nk_r +r_1-1 \choose r} +{g_1 \choose r} \right]. \end{align*} We will show that $\bigtriangleup_1M(H,c) < 0$ by Proposition $4$. We have $(nk_g+g_1)+ (r_1-1)= (nk_r +r_1 -1)+g_1$ and $g_1 < r_1-1 < nk_g+g_1 <nk_r+r_1-1$. Since $nk_r+r_1-1>r$, we have that $\bigtriangleup_1M(H,c) < 0$. Consequently, we can conclude that $c$ is not a minimum coloring. To sum up, we have proved that all colorings in the forms $F_2, F_3$ and $F_5$ have strictly more number of monochromatic edges than some coloring in the form $F_1$. Moreover, a coloring in the form $F_1$ where the numbers of red, blue and green vertices in the first class are not as equal as possible has strictly more monochromatic edges than the coloring $c^*$ where the numbers of red, blue and green classes are as equal as possible and the number of red, blue and green vertices in the first class are as equal as possible. Hence, this is the minimum coloring and it is unique up to a permutation of colors and classes. \end{proof} \section{Concluding remarks} \label{sec:conclude} In this paper, we considered $2$-colorings of balanced complete $k$-partite $r$-uniform hypergraphs and determined which one has the minimum number of monochromatic edges. The proof may give a clue for further generalization to an unbalanced hypergraph with arbitrary sizes of classes. We observed that the minimum coloring can only be in certain forms which are called canonical forms of the colorings. We studied the canonical forms of $2$-colorings of unbalanced complete tripartite $3$-uniform hypergraphs. Finally, we continued to determine the extermal $3$-coloring of balanced complete $k$-partite $r$-uniform hypergraphs when $k \equiv 0,1 \mod{3}$. In Theorem \ref{thm:2}, we determined the minimum coloring for unbalanced $3$-uniform hypergraphs. In the proof, almost all comparisons have been done without expanding the binomial coefficient terms and they also hold if we try to generalize the proof to $r$-uniform hypergraphs with arbitrary $r \leqslant n_1$. However, in the \textit{Case 5B}, it cannot be done without expanding those binomial coefficient terms where, in this case, we expand with $r+1=3$ as the lower index. Hence, the proof works only for $3$-uniform hypergraphs. We believe that the minimum coloring for $r$-uniform hypergraphs with arbitrary $r \leqslant n_1$ differs from those in Theorem \ref{thm:2}. Another generalized case is unbalanced complete hypergraphs with several vertex classes ($k>3$). The problem seems to be much more complicated because Theorem \ref{thm:2} demonstrates that the extremal coloring varies depending on the relationship among the sizes of the vertex classes. \begin{problem} What is the minimum $2$-coloring of an unbalanced complete $k$-partite $r$-uniform hypergraph? \end{problem} We have determined the extermal $3$-coloring of balanced complete $k$-partite $r$-uniform hypergraphs only for $k\equiv 0,1 \mod{3}$. \begin{problem} What is the minimum $3$-coloring of a balanced complete $k$-partite $r$-uniform hypergraph where $k\equiv 2 \mod{3}$? \end{problem} For $k \equiv 2 \mod{3}$, if we apply the same ideas as in the proof of Theorem \ref{thm:3}, we can conclude that the minimum coloring is in the form $F_3$ instead of $F_1$. However, the comparisons between colorings in the form $F_3$ are rather challenging and we believe that they require some further comparison tools. One might expect the minimum coloring to have approximately $\frac{\abs{V(H)}}{3}$ vertices in each color as in Theorem $4$. However, this is not the case, for example, when the number of vertices in each class is much greater than $k$. For $m>3$, we have studied only some trivial cases for $m$-colorings of the hypergraphs in Section $2.6$. \begin{problem} What is the minimum $m$-coloring of a balanced complete $k$-partite $r$-uniform hypergraph ? \end{problem} This is a generalization of the hypergraphs in Problem $15$ which is extremely complex as we believe that the minimum coloring varies depending on the relationship between number of colors and the number of classes. However, the proof of Theorem \ref{thm:3} might be useful to determine the minimum $m$-coloring of balanced complete $k$-partite $r$-uniform hypergraphs with $k\equiv 1 \mod{m}$, and we speculate that it would be in the form similar to the coloring in the form $F_1$. Finally, there is another natural definition of $k$-partite hypergraphs where each edge is an $r$-subset containing vertices from different classes. \begin{problem} With the above definition of $k$-partite hypergraphs, what is the minimum coloring of a balanced complete $k$-partite $r$-uniform hypergraph? \end{problem}
{ "timestamp": "2021-07-21T02:18:35", "yymm": "2107", "arxiv_id": "2107.09450", "language": "en", "url": "https://arxiv.org/abs/2107.09450", "abstract": "Consider the following problem. In a school with three classes containing $n$ students each, given that their genders are unknown, find the minimum possible number of triples of same-gender students not all of which are from the same class. Muaengwaeng asked this question and conjectured that the minimum scenario occurs when the classes are all boy, all girl and half-and-half. In this paper, we solve many generalizations of the problem including when the school has more than three classes, when triples are replaced by groups of larger sizes, when the classes are of different sizes, and when gender is replaced by other non-binary attributes.", "subjects": "Combinatorics (math.CO)", "title": "Monochromatic Edges in Complete Multipartite Hypergraphs", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9877587257892506, "lm_q2_score": 0.8104789086703225, "lm_q1q2_score": 0.8005576141072601 }
https://arxiv.org/abs/1807.01492
The minimal $k$-dispersion of point sets in high-dimensions
In this manuscript we introduce and study an extended version of the minimal dispersion of point sets, which has recently attracted considerable attention. Given a set $\mathscr P_n=\{x_1,\dots,x_n\}\subset [0,1]^d$ and $k\in\{0,1,\dots,n\}$, we define the $k$-dispersion to be the volume of the largest box amidst a point set containing at most $k$ points. The minimal $k$-dispersion is then given by the infimum over all possible point sets of cardinality $n$. We provide both upper and lower bounds for the minimal $k$-dispersion that coincide with the known bounds for the classical minimal dispersion for a surprisingly large range of $k$'s.
\section{Introduction and main results} A classical problem in computational geometry and complexity asks for the size of the largest empty and axis-parallel box given a point configuration in the cube $[0,1]^2$. From a complexity point of view this \emph{maximum empty rectangle problem} was already studied by Naamad, Lee, and Hsu \cite{NLH1984} who provided an $\mathcal O(n^2)$-time algorithm as well as an $\mathcal O(n\log^2(n))$-expected-time algorithm (when the points are drawn independently and uniformly at random) to actually find such a rectangle. Further results in this direction have been obtained by Chazelle, Drysdale, and Lee in \cite{CDL1986}. In the last decade the study of high-dimensional (geometric) structures has become increasingly important and it has been realized by now that the presence of high dimensions forces a certain regularity in the geometry of the space while, on the other hand, it unfolds various rather unexpected phenomena. The maximum empty rectangle problem studied in \cite{NLH1984} has its natural multivariate counterpart. Let $d,n\in\ensuremath{{\mathbb N}}$ and denote by $\mathscr B^d_{\text{ax}}$ the collection of axis-parallel boxes in $[0,1]^d$. Then the ($n$-th) minimal dispersion is defined to be the quantity \begin{equation}\label{eq:min:1} \disp^*(n,d) := \inf_{\mathscr P\subset[0,1]^d\atop{\#\mathscr P=n}}\,\sup_{\mathbb B\in\mathscr B_{\text{ax}}\atop{\mathbb B\cap \mathscr P=\emptyset}} \mathrm{vol}_d(\mathbb B), \end{equation} where $\#$ denotes the cardinality of a set and $\mathrm{vol}_d(\cdot)$ the $d$-dimensional Lebesgue measure. In other words, the minimal dispersion depicts the size of the largest empty, axis-parallel box amidst \emph{any} point set of cardinality $n$ in the $d$-dimensional cube. The interest in this notion is in parts motivated by applications in approximation theory, more precisely, in problems concerning the approximation of high-dimensional rank one tensors \cite{BDDG2014,NR2016} and in Marcinkiewicz-type discretizations of the uniform norm of multivariate trigonometric polynomials \cite{T2018}. Even though the minimal dispersion is conceptually quite simple and the problem of determining (or estimating) its order attracted considerable attention in the past $3$ years (see, e.g., \cite{AHR17, DJ2013, K2018, R2018,S2018,U2018, UV2018}), its behavior, simultaneously in the number of points $n$ and in the dimension $d$, is still not completely understood. Let us briefly describe the current state of the art. Aistleitner, Hinrichs, and Rudolf proved in \cite[Theorem 1]{AHR17} that, for all $d,n\in\ensuremath{{\mathbb N}}$, \[ \disp^*(n,d) \geq \frac{\log_2(d)}{4(n+\log_2(d))}\,. \] In particular, this shows that the volume of the largest empty box increases with the dimension $d$. In the same paper, the authors communicate an upper bound, which is attributed to Larcher (see \cite[Section 4]{AHR17}) and shows that \[ \disp^*(n,d) \leq \frac{2^{7d+1}}{n}\,. \] This improves upon a bound of Rote and Tichy \cite[Proposition 3.1]{RT1996} when $d\geq 54$. Under the assumption that the number of points satisfies $n> 2d$, it was recently proved by Rudolf \cite[Corollary 1]{R2018} that \[ \disp^*(n,d) \leq \frac{4d}{n} \log_2\Big(\frac{9n}{d}\Big)\,. \] In contrast to the probabilistic methods, several authors then provided explicit constructions of sets with small dispersion and small number of points. For example, Krieg \cite[Theorem]{K2018} shows that for $\varepsilon\in(0,1)$ and $d\ge 2$ there is a sparse grid with the number of points bounded by $(2d)^{\log_2(1/\varepsilon)}$ and dispersion at most $\varepsilon$. Temlyakov proved in \cite[Theorem 2.1 and Theorem 4.1]{T2018'} that Fibonacci and Frolov point sets achieve the dispersion of optimal order $n^{-1}$, but without paying extra attention to the dependence on $d$, see also~\cite{U2018b}. An essential breakthrough was achieved by Sosnovec in \cite[Theorem 2]{S2018}. He provided a randomized construction of a set with at most $c_\varepsilon \log_2(d)$ points in $[0,1]^d$ with dispersion at most $\varepsilon\in(0,1/4]$. Here, $c_\varepsilon\in(0,\infty)$ is a quantity depending only on $\varepsilon.$ The $\varepsilon$-dependence was then refined by the third and fourth author in \cite[Theorem 1]{UV2018}. More precisely, for every $\varepsilon\in (0,1/2)$ and $d\ge 2$, they provided a randomized construction of a point set $\mathscr P$ with $$ \#{\mathscr P}\le 2^7\frac{(1+\log_2(\varepsilon^{-1}))^2}{\varepsilon^2}\log_2 (d) $$ and dispersion at most $\varepsilon$. This result can be also reformulated as \begin{equation}\label{eq:UV} \disp^*(n,d)\le c\log_2(n)\sqrt{\frac{\log_2(d)}{n}} \end{equation} for $n,d\ge 2$ and some absolute constant $c\in(0,\infty)$. In this manuscript, we generalize the notion of dispersion by introducing a quantity, which measures the size of the largest box amidst a point set containing \emph{at most} $k$ points. The minimal dispersion \eqref{eq:min:1} then corresponds to the case $k=0$. For $d,n\in\ensuremath{{\mathbb N}}$ and $k\in\ensuremath{{\mathbb N}}\cup \{0\}$, we define the $k$-dispersion of a point set $\mathscr P_n=\{x_1,\dots,x_n\}\subset [0,1]^d$ to be the quantity \[ k\!\operatorname{-disp}(\mathscr P_n,d) := \sup\Big\{\mathrm{vol}_d(\mathbb B)\,\colon\, \mathbb B\in \mathscr B_{\text ax}^d \text{ with }\# (\mathscr P_n\cap \mathbb B)\leq k\Big\}, \] where $\mathscr B_{\text ax}^d$ is the collection of all axis-parallel boxes inside the $d$-dimensional cube $[0,1]^d$. The minimal $k$-dispersion is defined to be the infimum over all possible point sets of cardinality $n$, i.e., \[ k\!\operatorname{-disp}^*(n,d) := \inf_{\substack{\mathscr P_n\subset [0,1]^d\\ \#\mathscr P_n=n}} k\!\operatorname{-disp}(\mathscr P_n,d). \] Our first main result provides an upper bound on the minimal $k$-dispersion. \begin{thmalpha}\label{thm:main 1} There exists a constant $C\in(0,\infty)$ such that for any $d\geq 2$ and all $k,n\in\ensuremath{{\mathbb N}}$ with $k<n/2$, we have \begin{equation}\label{eq:dispk} k\!\operatorname{-disp}^*(n,d) \,\le\, C\,\max\left\{\log_2 (n)\sqrt{\frac{\log_2 (d)}{n}},\, k\,\frac{\log_2(n/k)}{n}\right\}. \end{equation} \end{thmalpha} \begin{remark} The minimal $k$-dispersion is easily seen to be non-decreasing in $k$. On the other hand, a comparison of \eqref{eq:UV} and \eqref{eq:dispk} reveals that the upper bound of minimal dispersion and minimal $k$-dispersion are of the same order for a large range of $k$'s. Indeed, if $k=k(n,d)$ increases with the number of points $n$ and the dimension $d$ while satisfying \[ k(n,d) \leq c \sqrt{n\cdot\log_2(d)}\,, \] for some absolute constant $c\in(0,\infty)$, then \eqref{eq:UV} and \eqref{eq:dispk} provide the same order in $n$ and $d$. Motivated by this result, we conjecture that \eqref{eq:UV} actually offers a lot of space for improvement. \end{remark} The second result establishes the following lower bound on the minimal $k$-dispersion. \begin{thmalpha}\label{thm:main 2} Let $k,n,d\in\ensuremath{{\mathbb N}}$. Then \[ k\!\operatorname{-disp}^*(n,d) \,\ge\, \frac18\,\min\left\{1, \frac{k+\log_2(d)}{n}\right\}. \] \end{thmalpha} \section{Proof of Theorem \ref{thm:main 1} -- the upper bound}\label{sec:thm1-multi} We shall present here the proof for the upper bound of the minimal $k$-dispersion. In the proof, we modify the ideas developed in \cite{S2018} and \cite{UV2018} to our setting. \subsection{The idea of proof} Before we start let us briefly discuss the strategy of the proof. For every $\varepsilon\in(0,1/4)$, we construct a set ${\mathbb X}=\{x^1,\dots,x^n\}\subset[0,1]^d$ with a small number of elements and small $k$-dispersion. This set is constructed (similarly to \cite{S2018} and \cite{UV2018}) by random sampling from a discrete mash in $[0,1]^d$. To allow for independence across the steps of this sampling, it might happen that some points of the mash are actually sampled more than once. Such points were naturally discarded in \cite{UV2018}, because they could not influence if a box ${\mathbb B}\subset[0,1]^d$ intersects ${\mathbb X}$, or not. Here, we need to keep a more detailed track of the intersection of boxes with ${\mathbb X}$. Therefore, we allow for repeated sampling, and ${\mathbb X}$ will actually be a multiset. In Section \ref{subsec:multiset}, we argue in detail that this modification does not alter the minimal $k$-dispersion. It is then not difficult to see that each box of large enough volume contains indeed at least $k$ points from the multiset ${\mathbb X}$ with high probability. Unfortunately, this is still not enough to apply the union bound, as there are infinitely many boxes with large volume. We will therefore divide them into finitely many groups (called $\Omega_{m}(s,p)$ later on), and show that even the intersection of all boxes included in $\Omega_m(s,p)$ still contains at least $k$ points from ${\mathbb X}$ with high probability. At long last, we can apply the union bound over all admissible parameter pairs $(s,p)$. \subsection{A random multiset}\label{subsec:multiset} Let $\varepsilon\in (0,1/4)$ be fixed and put $m=\lceil\log_2(1/\varepsilon)\rceil$. We define a one-dimensional set via \[ \mathbb M_m:=\left\{\frac{1}{2^{m}},\dots,\frac{2^{m}-1}{2^{m}}\right\}\subset [0,1]. \] For $n\in\ensuremath{{\mathbb N}}$, we construct a random multiset $\mathbb X=\{x^1,\dots,x^n\}$ by sampling independently and uniformly at random from the points in $\mathbb M_m^d$, where $n$ will later be the number of points in Theorem \ref{thm:main 1}. Note that a multiset $\mathbb X$ in the cube $[0,1]^d$ is naturally identified with the \emph{multiplicity function} $\mathbb X\colon [0,1]^d\to \ensuremath{{\mathbb N}}_0$, where $\mathbb X(z)$ gives the multiplicity of $z\in [0,1]^d$ in $\mathbb X$. For ${\mathbb B}\in {\mathscr B}_{ax}$, we define \[ \#\bigr(\mathbb X \cap {\mathbb B}\bigl) \,:=\, \sum_{z\in {\mathbb B}} \mathbb X(z)\quad\text{and}\quad \#{\mathbb X} \,:=\, \sum_{z\in[0,1]^d} \mathbb X(z)\,, \] and the $k$-dispersion of the multiset ${\mathbb X}$ as $$ k\!\operatorname{-disp}_m({\mathbb X},d):= \sup\Big\{\mathrm{vol}_d(\mathbb B)\,\colon\, \mathbb B\in \mathscr B_{\text ax}^d \text{ with }\# ({\mathbb X}\cap \mathbb B)\leq k\Big\}. $$ Finally, we take the infimum over all possible multisets of cardinality $n$ and obtain \[ k\!\operatorname{-disp}_m^*(n,d) := \inf_{\substack{{\mathbb X}\subset [0,1]^d\\ \#{\mathbb X}=n}} k\!\operatorname{-disp}_m({\mathbb X},d). \] As each classical set ${\mathscr P}\in[0,1]^d$ is also a multiset (with the multiplicity function bounded by one), we immediately obtain that \[ k\!\operatorname{-disp}_m^*(n,d)\le k\!\operatorname{-disp}^*(n,d). \] On the other hand, if ${\mathbb X}=\{x^1,\dots,x^n\}\subset [0,1]^d$ is a multiset, then we consider the sets $\{x^1+\xi^1,\dots,x^n+\xi^n\}$ with $\|\xi^j\|_\infty\le \delta$ (for $\delta\in(0,\infty)$), where $\xi^1,\dots,\xi^n$ are independent random vectors which are uniformly distributed over $[-\delta,\delta]^d$. If we then let $\delta\to 0$, it follows that $k\!\operatorname{-disp}^*(n,d)\le k\!\operatorname{-disp}_m^*(n,d)$. \subsection{The partitioning scheme}\label{subsec:partition} We now introduce a set $\Omega_m$ containing all those boxes $\mathbb B$ with `large' volume. For $m\in\ensuremath{{\mathbb N}}$, we define \[ \Omega_m:=\Big\{\mathbb B=I_1\times\dots\times I_d\subset[0,1]^d\,\colon\,\mathrm{vol}_d(\mathbb B)>\frac{1}{2^m}\Big\}\,. \] As already described before, our approach will later be based on a union bound over all the boxes $\mathbb B\in\Omega_m$. As there are infinitely many of those boxes, we first divide $\Omega_m$ into finitely many `suitable' subsets. This is done as follows: for $s=(s_1,\dots,s_d)\in\{0,1,\dots,2^{m}-1\}^d$ and $p=(p_1,\dots,p_d)\in\{1/2^{m},\dots,1-1/2^{m}\}^d$, we define the collection $\Omega_m(s,p)$ of subsets of $\Omega_m$ to be \begin{align*} \Omega_m(s,p)&:=\bigg\{\mathbb B=I_1\times\dots\times I_d\in\Omega_m \,\colon\, \forall \ell\in\{1,\dots,d\}: \frac{s_\ell}{2^{m}}<\mathrm{vol}(I_\ell) \le \frac{s_\ell+1}{2^{m}}\\ &\qquad\qquad \text{and}\quad\inf I_\ell\in \Big[p_\ell-\frac{1}{2^{m}},p_\ell\Big)\bigg\}. \end{align*} We first observe that $\Omega_m(s,p)=\emptyset$ if the choice of $s$ does not allow $\Omega_m(s,p)$ to contain any box $\mathbb{B}$ with $\mathrm{vol}_d(\mathbb B)>2^{-m}$. This holds, e.g., if $s_{\ell_0}=0$ for some $\ell_0\in\{1,\dots,d\}$. We define the index set \[ \mathbb I_m \,:=\, \Bigl\{(s,p)\,\colon\, \Omega_m(s,p)\neq\emptyset\Bigr\}, \] which contains those indices $(s,p)$ that are needed for the following considerations, and we bound its cardinality. Let \[ A_m(s) := \# \Big\{\ell\in\{1,\dots,d\}\,\colon\, s_\ell < 2^m-1 \Big\} \] and observe that, by definition, any $\mathbb B\in\Omega_m(s,p)$ must satisfy $$ \frac{1}{2^m} \,<\, \mathrm{vol}_d(\mathbb B) \,\le\, \prod_{\ell=1}^d\frac{s_\ell+1}{2^{m}} \,\le\, \Bigl(1-\frac{1}{2^{m}}\Bigr)^{A_m(s)}. $$ This is a contradiction if $A_m(s)>\log(2)\, m 2^m$. Therefore, $\Omega_m(s,p)\neq\emptyset$ implies that \begin{equation*} A_m(s) \,\le\, \min\Bigl\{\lfloor\log(2)\, m 2^m\rfloor,\, d\Bigr\} \,=:\, A_m, \end{equation*} i.e., there are at most $A_m$ choices of $\ell$ with $s_\ell<2^m-1$. Clearly, there are at most $\binom{d}{A_m} 2^{m A_m}$ choices for $s\in\{0,1,\dots,2^{m}-1\}^d$ with $A_m(s)\le A_m$. Moreover, for given $s$, there are at most $2^{m A_m(s)}$ choices for $p$ with $\Omega_m(s,p)\neq\emptyset$. This follows from the fact that for each $\ell\in\{1,\dots,d\}$, we have at most $2^m-1$ choices for $p_\ell$ (by definition) and, if $s_{\ell_0}=2^m-1$ for some $\ell_0\in\{1,\dots,d\}$, then we have $\Omega_m(s,p)=\emptyset$ unless $p_{\ell_0}=2^{-m}$. For other $p_{\ell_0}$ the boxes cannot be contained in the unit cube. For $m$ such that $A_m<d$, we obtain \begin{equation* \begin{split} \#\mathbb I_m \,&<\, \binom{d}{A_m} \, 2^{2m A_m} \,<\, \biggl(\frac{ed}{A_m}\cdot 2^{2m}\biggr)^{A_m} \\ &<\, \biggl(\frac{4d2^{m+1}}{m}\biggr)^{A_m} \,\le\, \exp\Bigl(m2^{m}\log(2^{m+3}d)\Bigr), \end{split} \end{equation*} where we used that $\log(2)m 2^{m-1}< A_m\le \log(2) m 2^{m}$ for $m\in\ensuremath{{\mathbb N}}$ and $e/\log(2)< 4$. On the other hand, if $A_m=d$, i.e., if $d<\log(2)m2^m$, then we obtain $$ \#\mathbb I_m\le 2^{md}\cdot 2^{md}\le \exp\Big(\log^2(2)\cdot2m2^m\Big). $$ Therefore, for arbitrary $m\in\ensuremath{{\mathbb N}}$, we obtain \begin{equation}\label{eq:card} \#\mathbb I_m \,\le\, \exp\Bigl(m2^{m}\log(2^{m+3}d)\Bigr). \end{equation} \medskip \subsection{The proof} We shall now present the proof of the upper bound on the minimal $k$-dispersion. We do this by proving that our random multiset has small $k$-dispersion with positive probability, which proves the existence of a `good' multiset. The following result is from \cite[Lemma 3]{UV2018}. Note that it is stated there in a different way, but (the end of) its proof clearly shows this variant. For $(s,p)\in\mathbb I_m$, let \[ \mathbb B_m(s,p) \,:=\, \bigcap_{\mathbb B\in \Omega_m(s,p)} \mathbb B \,=\, \prod_{\ell=1}^d \Bigl[p_\ell,p_\ell+\frac{s_\ell-1}{2^{m}}\Bigr]. \] \begin{lem}\label{lem:discrete} Let $m\in\ensuremath{{\mathbb N}}$, $(s,p)\in\mathbb I_m$ and $z$ be uniformly distributed in $\mathbb M_m^d$. Then \begin{equation*} \ensuremath{{\mathbb P}}\big(z\in \mathbb B_m(s,p)\big)\ge \frac{1}{2 ^{m+4}}\,. \end{equation*} \end{lem} \medskip For the random multiset as constructed in Section~\ref{subsec:multiset}, we now estimate the probability that the number of points in $\mathbb X\cap \mathbb B_m(s,p)$ does not exceed $k\in\ensuremath{{\mathbb N}}$, where $k<\frac{n}{2}$ (for the case $k=0$ see~\cite{UV2018}). Let us consider two cases. \vskip 1mm \emph{Case 1:} Assume that $\ensuremath{{\mathbb P}}\big(x^1\in \mathbb B_m(s,p)\big)\le 1/2$. Then we use Lemma~\ref{lem:discrete} and obtain the estimate \[\begin{split} \ensuremath{{\mathbb P}}\big(\#(\mathbb X & \cap \mathbb B_m(s,p))\le k\big) \,=\, \sum_{\ell=0}^k {n\choose \ell} \ensuremath{{\mathbb P}}\big(x^1\in \mathbb B_m(s,p)\big)^{\ell}\ \ensuremath{{\mathbb P}}\big(x^1\not\in \mathbb B_m(s,p)\big)^{n-\ell}\\ &\le\, (k+1){n\choose k}\ensuremath{{\mathbb P}}\big(x^1\not\in\mathbb B_m(s,p)\big)^{n} \,\le\, 2k \,\frac{n^k}{k!}\,\Bigl(1-\ensuremath{{\mathbb P}}\big(x^1\in\mathbb B_m(s,p)\big)\Bigr)^n\\ &\le\, \frac{2n^k}{(k-1)!}\,\Bigl(1-\frac{1}{2^{m+4}}\Bigr)^n \,\le\, \frac{2n^k}{(k-1)!}\, \exp\Big(-\frac{n}{2^{m+4}}\Big). \end{split}\] \vskip 1mm \emph{Case 2:} Assume $\ensuremath{{\mathbb P}}(x^1\in\mathbb B_m(s,p))>1/2$. Then \begin{align*} \ensuremath{{\mathbb P}}(\#(\mathbb X & \cap \mathbb B_m(s,p))\le k) \,=\, \sum_{\ell=0}^k {n\choose \ell} \ensuremath{{\mathbb P}}\big(x^1\in \mathbb B_m(s,p)\big)^{\ell}\ \ensuremath{{\mathbb P}}\big(x^1\not\in\mathbb B_m(s,p)\big)^{n-\ell}\\ &\le\, (k+1){n\choose k} \frac{1}{2^{n-k}} \,\le\, \frac{2n^k}{(k-1)!\, 2^{n-k}} \,\le\, \frac{2n^k}{(k-1)!}\exp\Big(-\frac{n}{2^{m+4}}\Big), \end{align*} where the last inequality follows from the fact that \[ (n-k)\log(2) > \frac{n}{2}\log(2)\ge \frac{n}{2^{4}}\ge \frac{n}{2^{m+4}}. \] \vskip 2mm Putting both cases together, we see that \begin{align}\label{ineq:prob} \ensuremath{{\mathbb P}}\big(\#(\mathbb X\cap \mathbb B_m(s,p))\le k\big) \,\le\, \frac{2n^k}{(k-1)!}\exp\Big(-\frac{n}{2^{m+4}}\Big). \end{align} \smallskip Recall from Section~\ref{subsec:partition} that \[ \Omega_m \,=\, \bigcup_{(s,p)\in\mathbb I_m} \Omega_m(s,p). \] Combining the upper bound \eqref{eq:card} on the cardinality of $\mathbb I_m$ with the estimate in Lemma~\ref{lem:discrete}, we obtain by a union bound that \begin{eqnarray*} \ensuremath{{\mathbb P}}\big(\exists \mathbb B\in\Omega_m\,:\,\#(\mathbb X\cap \mathbb B)\le k\big) & \le &\sum_{(s,p)\in\mathbb I_m} \ensuremath{{\mathbb P}}\big(\exists \mathbb B\in\Omega_m(s,p)\,:\,\#(\mathbb X\cap \mathbb B)\le k\big)\\ &\le&\sum_{(s,p)\in\mathbb I_m} \ensuremath{{\mathbb P}}\big(\#(\mathbb X\cap \mathbb B_m(s,p))\le k\big)\\ &<& \frac{2n^k}{(k-1)!}\,\exp\Bigl(m2^{m}\log(2^{m+3}d)-n2^{-m-4}\Bigr)\,. \end{eqnarray*} The last expression will be smaller than or equal to $1$ if and only if \begin{equation}\label{eq:condition} \begin{split} n2^{-m-4} \,&\geq m2^{m}\log(2^{m+3}d) +\log\Big(\frac{2n^k}{(k-1)!}\Big) \\ \,&= \, m2^{m}\log(2^{m+3}d) + k\, \log\Big(c_k \frac{n}{k}\Big) \end{split} \end{equation} with $c_k:=k\bigl(\frac{2}{(k-1)!}\bigr)^{1/k}$. Note that by Stirling's formula, $c_k\uparrow e$ as $k\to\infty$. To guarantee \eqref{eq:condition}, it is enough to assume that \[ n\geq m 2^{2m+5}\log(2^{m+3}d) \qquad\text{and}\qquad n\geq 2^{m+5}\,k\, \log\Big(e\frac{n}{k}\Big). \] It is easy to prove that the second inequality is implied by $n\ge k m 2^{m+9}> e k (m+5) 2^{m+5}$. Hence, we find an $n\in\ensuremath{{\mathbb N}}$ with \eqref{eq:condition} such that \[ n \,\le\, C\,m\, 2^m\, \max\Bigl\{2^{m}\,\log(2^m d),\, k \Bigr\} \] for some constant $C\le2^9$. This ensures that there exists a realization of the multiset $\mathbb X$ with cardinality $n$ such that, for all boxes $\mathbb B$ with $\mathrm{vol}_d(\mathbb B)>2^{-m}$, we have $\#(\mathbb B\cap \mathbb X)>k$. Using the argument of Section \ref{subsec:multiset}, we obtain \begin{align*} N(2^{-m},d) & := \min \Big\{N\in\ensuremath{{\mathbb N}}\,:\, k\!\operatorname{-disp}(N,d)\leq 2^{-m} \Big\} \\ & \leq C\,m\, 2^m\, \max\Bigl\{2^{m}\,\log(2^m d),\, k \Bigr\}. \end{align*} To finish the proof, let $\varepsilon\in(0,\frac{1}{4})$ and denote by $m:=m_\varepsilon\in\ensuremath{{\mathbb N}}$ the unique integer satisfying \[ \frac{1}{2^m}\le\varepsilon< \frac{1}{2^{m-1}}\,, \] i.e., $m=\lceil\log_2(1/\varepsilon)\rceil$. By this choice of $m$, \begin{align*} m 2^{2m}\log(2^{m}d) \,<\, c_1\,\log_2(d)\bigg(\frac{\log_2(1/\varepsilon)}{\varepsilon}\bigg)^2, \end{align*} and \[ m 2^{m}\,k \,<\, c_2\, \frac{k\cdot\log_2(1/\varepsilon)}{\varepsilon} \] for some constants $c_1,c_2\in(0,\infty)$. This means that, since $N(\cdot,d)$ is decreasing in the first argument, \[ N(\varepsilon,d) \,\le\, C\cdot \max\left\{\log_2(d)\bigg(\frac{\log_2(1/\varepsilon)}{\varepsilon}\bigg)^2,\, \frac{k\cdot\log_2(1/\varepsilon)}{\varepsilon} \right\} \] for some constant $C\in(0,\infty)$. We therefore conclude that \[ k\!\operatorname{-disp}(n,d) \,\le\, C'\,\max\left\{\log_2(n)\sqrt{\frac{\log_2(d)}{n}},\, \frac{k\cdot\log_2(n/k)}{n}\right\} \] with an absolute constant $C'\in(0,\infty)$. \section{Proof of Theorem \ref{thm:main 2} -- the lower bound}\label{sec:thm2} The proof is very much inspired by the proof of the lower bound on the dispersion of Aistleitner, Hinrichs and Rudolf \cite{AHR17}. We recall their argument in a slightly modified form. Given a point set $\mathscr P_n=\{x_1,\dots,x_n\}$ with $x_i=(x_{i,1},\dots,x_{i,d})\in[0,1]^d$, we define the matrix $A=A(\mathscr P_n)\in\ensuremath{{\mathbb R}}^{n\times d}$ by \[ A_{i,j} \,=\, \begin{cases} 1 & :\, x_{i,j} \ge 1/2\,; \\ 0 & \,\text{ otherwise}, \end{cases} \] with $i=1,\dots,n$ and $j=1,\dots,d$. Note that if $A$ contains two equal columns, then the projection of the point set on the two coordinates corresponding to these columns is contained in the union of the lower-left and the upper-right quarter of the unit square. Therefore, the dispersion is at least $1/4$. Likewise, if two columns $c_1,c_2\in\{0,1\}^{n}$ of $A$ satisfy $c_1=1-c_2$, then the projection is contained in the upper-left and the lower-right quarter and the dispersion is at least $1/4$. Recall that $A$ has $d$ columns. It is clear from the pigeon hole principle that there must be two columns that satisfy one of the above conditions whenever $d>2^{n-1}$. This implies \[ 0\!\operatorname{-disp}^*\!\big(\lceil\log_2 d\rceil, d\big) \,\ge\, 1/4. \] We now consider the $k$-dispersion for $k\ge1$. Following the above arguments, if there are two columns $c_1,c_2$ of $A$ that agree (or disagree) in all but $k$ entries, then there exists a box of volume $1/4$ that contains at most $k$ points. Again, from the pigeon hole principle (and just ignoring $k$ rows), we obtain that such columns exist whenever $d>2^{n-k-1}$. This implies \begin{equation}\label{eq:init} k\!\operatorname{-disp}^*\!\big(k+\lceil\log_2 d\rceil, d\big) \,\ge\, 1/4. \end{equation} Finally, note that Lemma~1 from \cite{AHR17} holds also for the $k$-dispersion, i.e., for all $n,k,d,\ell\in\ensuremath{{\mathbb N}}$ we have \begin{equation}\label{eq:rec} k\!\operatorname{-disp}^*(n,d) \,\ge\, \frac{(\ell+1)\, k\!\operatorname{-disp}^*(\ell,d)}{n+\ell+1}. \end{equation} From this we conclude Theorem \ref{thm:main 2}. \begin{proof}[Proof of Theorem \ref{thm:main 2}] For $n\le k+\log_2(d)$, we use \eqref{eq:init} and obtain $k\!\operatorname{-disp}^*(n,d)\ge1/4\ge1/8$. For $n> k+\log_2(d)$, we use \eqref{eq:rec} with $\ell=k+\lceil\log_2(d)\rceil\le n$ and obtain \[ k\!\operatorname{-disp}^*(n,d) \,\ge\, \frac{(\ell+1)\, k\!\operatorname{-disp}^*(\ell,d)}{n+\ell+1} \,\ge\, \frac14\,\frac{(\ell+1)}{n+\ell+1} \,\ge\, \frac{k+\log_2(d)}{8n}. \] \end{proof} \begin{remark} As the method to obtain \eqref{eq:init} is clearly related to packing numbers on the discrete cube $\{0,1\}^n$ with respect to the Hamming metric, one could try to apply more involved methods to obtain better bounds. For example, the well-known \emph{sphere-packing bound} (also known as Hamming bound), see \cite[Theorem 5.2.7]{Lint1992}, states that the maximal size of a $k$-packing of the cube, say $M(n,k)$, satisfies \[ M(n,k) \,\le\, \frac{2^n}{\sum_{t=0}^{\lfloor(k-1)/2\rfloor}\binom{n}{t}}. \] However, using this bound does not lead to any significant improvement. \end{remark} \bibliographystyle{abbrv}
{ "timestamp": "2018-07-05T02:06:07", "yymm": "1807", "arxiv_id": "1807.01492", "language": "en", "url": "https://arxiv.org/abs/1807.01492", "abstract": "In this manuscript we introduce and study an extended version of the minimal dispersion of point sets, which has recently attracted considerable attention. Given a set $\\mathscr P_n=\\{x_1,\\dots,x_n\\}\\subset [0,1]^d$ and $k\\in\\{0,1,\\dots,n\\}$, we define the $k$-dispersion to be the volume of the largest box amidst a point set containing at most $k$ points. The minimal $k$-dispersion is then given by the infimum over all possible point sets of cardinality $n$. We provide both upper and lower bounds for the minimal $k$-dispersion that coincide with the known bounds for the classical minimal dispersion for a surprisingly large range of $k$'s.", "subjects": "Numerical Analysis (math.NA); Classical Analysis and ODEs (math.CA)", "title": "The minimal $k$-dispersion of point sets in high-dimensions", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.987758723627135, "lm_q2_score": 0.8104789040926008, "lm_q1q2_score": 0.8005576078332266 }
https://arxiv.org/abs/1804.09697
Electrostatic Interpretation of Zeros of Orthogonal Polynomials
We study the differential equation $ - (p(x) y')' + q(x) y' = \lambda y,$ where $p(x)$ is a polynomial of degree at most 2 and $q(x)$ is a polynomial of degree at most 1. This includes the classical Jacobi polynomials, Hermite polynomials, Legendre polynomials, Chebychev polynomials and Laguerre polynomials. We provide a general electrostatic interpretation of zeros of such polynomials: the set of real numbers $\left\{x_1, \dots, x_n\right\}$ satisfies $$ p(x_i) \sum_{k = 1 \atop k \neq i}^{n}{\frac{2}{x_k - x_i}} = q(x_i) - p'(x_i) \qquad \mbox{for all}~ 1\leq i \leq n$$ if and only if they are zeros of a polynomial solving the differential equation. We also derive a system of ODEs depending on $p(x),q(x)$ whose solutions converge to the zeros of the orthogonal polynomial at an exponential rate.
\section{Introduction} \subsection{Introduction.} We start by describing an 1885 result of Stieltjes for Jacobi polynomials \cite{stieltjes}. Jacobi polynomials $P_n^{\alpha, \beta}(x)$, for real $\alpha, \beta > -1$, are the unique (up to a constant factor) solutions of the equation $$ (1 - x^2 ) y''(x) - \left( \beta - \alpha - (\alpha + \beta + 2)x\right)y'(x) = n(n + \alpha + \beta + 1) y(x)$$ where $n \in \mathbb{N}$. The solution is a polynomial of degree $n$ having all its zeros in $(-1,1)$. Stieltjes defined a notion of energy for any set $\left\{x_1, \dots, x_n \right\} \subset (-1,1)$ as $$ E = - \sum_{i,j = 1 \atop i \neq j}^{n}{\log{|x_i - x_j|}} - \sum_{i=1}^{n}{\left( \frac{\alpha+1}{2}\log{|x-1|} + \frac{\beta+1}{2}\log{|x+1|}\right)}$$ and showed that the minimal energy is exactly given by the zeros of the Jacobi polynomial. Differentiating this expression in all the $n$ variables leads to a system of equations describing an electrostatic equilibrium $$ \sum_{k=1 \atop k \neq i}^{n}{ \frac{1}{x_k - x_i} } = \frac{1}{2}\frac{\alpha+1}{x_i - 1} + \frac{1}{2} \frac{\beta + 1}{x_i + 1} \qquad \mbox{for all}~1 \leq i \leq n.$$ A traditional application of the formula is to establish monotonicity of zeros with respect to the parameters $\alpha, \beta$. This argument can be found in Szeg\H{o}'s book \cite{szeg} who also discusses Laguerre and Hermite polynomials. These types of questions have been studied by a large number of people, we refer to \cite{ahmed, dimitrov, hendriksen, ismail}, the survey \cite{survey}, the 1978 survey and 2001 book by F. Calogero \cite{calogero, calogero2}, the more recent papers of F. A. Gr\"unbaum \cite{grunbaum, grunbaum2} and M. E. H. Ismail \cite{ismail, ismail2} and references therein. \subsection{Equilibrium.} The purpose of this paper is to prove a simple general result that characterizes zeros of orthogonal polynomials. The statement itself is a bit more general but most interesting when applied to the classical orthogonal polynomials. We first discuss second order equations, then describe system of ODEs converging to zeros and conclude with a short remark on higher order equations. \begin{theorem} Let $p(x), q(x)$ be polynomials of degree at most 2 and 1, respectively. Then the set $\left\{x_1, \dots, x_n\right\}$, assumed to be in the domain of definition, satisfies $$ p(x_i) \sum_{k = 1 \atop k \neq i}^{n}{\frac{2}{x_k - x_i}} = q(x_i) - p'(x_i) \qquad \mbox{for all} \quad 1\leq i \leq n$$ if and only if $$y(x) = \prod_{k=1}^{n}{(x-x_k)} \quad \mbox{solves} \quad - (p(x) y')' + q(x) y' = \lambda y \quad \mbox{for some} ~~ \lambda \in \mathbb{R}.$$ \end{theorem} In classical applications, there is a unique polynomial solution of fixed degree corresponding to a fixed value of $\lambda$. This removes $\lambda$ as a variable and leads to a complete characterization. One direction of the statement (zeros of a polynomial solutions satisfy the system of equations) is a fairly straightforward computation and the outline of the argument can be found, for example, in \cite{grunbaum} or, as a remark, in \cite[\S 3]{ismail}. Moreover, results in this direction can be obtained for much more general differential equations of higher order (see, for example, \cite{calogero, calogero2,avila}). Arguments in the other direction seem to exist only in special cases, as in the work of Stieltjes, and are based on interpreting the system of equations as the critical point of an associated energy functional (see e.g. \cite{ismail, survey}). Our argument proceeds by a different route and bypasses considerations of an underlying energy.\\ We remark that there is no assumption of orthogonality nor is there any restriction on the domain which may be either bounded or unbounded but we do assume that the $n$ points are contained in the domain of definition. Let us quickly consider two special cases: we start with the Hermite differential equation $$ -y'' + x y' = \lambda y.$$ This corresponds to $p(x) \equiv 1$ and $q(x) \equiv x$. We deduce from Theorem 1 that the zeros of the $n-$th solution satisfy a relationship also going back to Stieltjes, see \cite{ahmed1}, $$ \sum_{k = 1 \atop k \neq i}^{n}{\frac{2}{x_k - x_i}} = x_i \qquad \mbox{for all}~1 \leq i \leq n.$$ Returning to a special case of Jacobi polynomials $\alpha = 0 = \beta$ (for simplicity of exposition), we obtain $p(x) = x^2 -1$, $q(x) = 0$ and thus $$ (x_i^2-1) \sum_{k = 1 \atop k \neq i}^{n}{\frac{2}{x_k - x_i}} = 2x_i \qquad \mbox{for all} \quad 1\leq i \leq n$$ which is easily seen to be equivalent to Stieltjes' electrostatic equilibrium. \subsection{A System of ODEs.} One interesting byproduct of our approach is a procedure that allows for the computation of zeros without ever computing the polynomial. More precisely, for any given set $x_1(0) < x_2(0) < \dots < x_n(0)$ (assumed to be in the domain of definition), we consider the system of ordinary differential equations $$ \frac{d}{dt} x_i(t) = -p(x_i) \sum_{k = 1 \atop k \neq i}^{n}{\frac{2}{x_k(t) - x_i(t)}} + p'(x_i(t)) - q(x_i(t)) \qquad \qquad (\diamond)$$ The result above implies that the unique stationary point of this system of ODEs is given by the $n$ zeros of a polynomial solution of degree $n$. We will now show that under standard assumptions on the equation $ - (p(x) y')' + q(x) y' = \lambda y,$ the system converges to this fixed point at an exponential rate. We require somewhat stronger assumptions than we do for Theorem 1 and ask that \begin{enumerate} \item the equation $ - (p(x) y'(x))' + q(x) y'(x) = \lambda y(x)$ is a Sturm-Liouville problem with a discrete set of solutions indexed by $0 = \lambda_0 < \lambda_1 < \lambda_2 < \dots$ and \item the solution corresponding to $y_n$ is a polynomial of degree $n$ for all $n \in \mathbb{N}$ \end{enumerate} These assumptions cover the classical polynomials but are fairly strong, we refer to Bochner's theorem and its nice exposition in \cite[\S 20]{ismail3}. \begin{theorem} The system $(\diamond)$ converges for all initial values $x_1(0) < \dots < x_n(0)$ to the zeros $x_1 < \dots < x_n$ of the degree $n$ polynomial solving the equation. Moreover, $$ \max_{1 \leq i \leq n}{ |x_i(t) - x_i|} \leq c e^{- \sigma_n t},$$ where $c > 0$ depends on everything and $\sigma_n \geq \lambda_n - \lambda_{n-1}$. \end{theorem} We are not aware of any result of this type being known; the connection between partial differential equations and zeros of polynomials (or poles of rational functions) is, of course classical, we refer to the extensive survey of Calogero \cite{calogero} and his more recent book \cite{calogero2}. Our underlying dynamical system is the \textit{backward} heat equation (well-posed for algebraic reasons) and existence of a solution for all time is a consequence of Sturm-Liouville theory. We assume that the initial values are all contained in the domain where the solution is defined. For classical Sturm-Liouville problems, the Weyl law suggests $\lambda_{n} - \lambda_{n-1} \sim n$. The rapid convergence can be easily observed in examples. The assumptions are satisfied for all classical polynomials. We consider, as a specific example, the Laguerre polynomials satisfying $$ - x y'' + (x-1)y' = ny $$ or, in our notation, $p(x) = x = q(x)$. Theorem 2 then implies that for any choice of initial values $x(0) < y(0) < z(0)$, the system of ordinary differential equations \begin{align*} \dot x(t) &= \frac{2x(t)}{x(t) - y(t)} + \frac{2x(t)}{x(t) - z(t)} +1 - x(t) \\ \dot y(t) &= \frac{2y(t)}{y(t) - x(t)} + \frac{2y(t)}{y(t) - z(t)} +1 - y(t) \\ \dot z(t) &= \frac{2z(t)}{z(t) - x(t)} + \frac{2z(t)}{z(t) - y(t)} +1 - z(t) \end{align*} has a solution for all $t > 0$. As $t \rightarrow \infty$, the solutions converge to the zeros of the third Laguerre polynomial $$ L_3(x) = x^3 - 9x^2 + 18x - 6$$ and thus $$ x(t) \rightarrow 0.4157\dots,~ y(t) \rightarrow 2.2942\dots \quad \mbox{and} \quad z(t) \rightarrow 6.2899\dots.$$ Moreover, this convergence happens at an exponential rate (roughly with speed $e^{-t}$). We consider a large-scale example (see Fig. 1) next. The equation $$- \frac{d}{dx}\left( (1-x^2) \frac{d}{dx}y(x)\right) = n(n+1) y(x)$$ is solved by the Legendre polynomials $P_n$ defined on $(-1,1)$. In our notation, we have $p(x) = 1- x^2$ and $q(x) = 0$. We simulate the system of ODEs $$ \frac{d}{dt} x_i(t) = -(1-x_i(t))^2 \sum_{k = 1 \atop k \neq i}^{n}{\frac{2}{x_k(t) - x_i(t)}} - 2x_i(t)$$ for $n=100$ equations. The initial values $\left\{x_1(0), \dots, x_{100}(0)\right\}$ are chosen as uniform random variables in the interval $(-0.1, 0.1)$. Figure 1 shows the evolution of the system and convergence to the zeros of the Legendre polynomial of degree 100. \vspace{-10pt} \begin{center} \begin{figure}[h!] \begin{tikzpicture}[scale=1] \draw [thick, ->] (-4.3,-3) -- (-4.3,3); \draw [thick, ->] (4.3,-0.2) -- (4.7,-0.2); \node at (4.6, -0.5) {$t$}; \filldraw (-4.3, 2.6) circle (0.05cm); \draw [dashed] (-4.3,2.65) -- (4.7,2.65); \draw [dashed] (-4.3,-2.65) -- (4.7,-2.65); \node at (-4.5, 2.6) {1}; \filldraw (-4.3, -2.6) circle (0.05cm); \node at (-4.5, -2.6) {-1}; \node at (-.010,0) {\includegraphics[width=0.7\textwidth]{bigsample.pdf}}; \end{tikzpicture} \caption{Evolution of the system of ODEs for $0 \leq t \leq 0.01$ approaches the zeros of the Legendre polynomial $P_{100}$ in $(-1,1)$.} \end{figure} \end{center} \vspace{-20pt} As indicated in the Theorem, the constant determining the speed of exponential convergence grows linearly in $n$: this is reflected in the rather short time-scale $(0 \leq t \leq 0.01)$ in the picture. \subsection{Higher order equations} Theorem 1 as well as its proof immediately generalizes to polynomial solutions of the equation $$ a_n(x) y^{(n)}(x) + a_{n-1}(x) y^{(n-1)}(x) + a_{1}(x)y'(x) = \lambda y(x),$$ where $a_n$ is a polynomial of degree at most $n$ and the solution $y$ is assumed to only have single zeros: one half of the statement can be found in \cite[Proposition 1]{avila}, the other direction follows from our approach. We are not aware of an extension of Theorem 2 since our proof relies on Sturm-Liouville theory. \section{Proofs} \subsection{Proof of Theorem 1.} \begin{proof} One direction of the argument, showing that the zeros of any polynomial solution of the equation is in the desired electrostatic equilibrium, is a simple computation and known in greater generality \cite{ahmed1,ahmed}. However, it also follows immediately from our argument and is, for completeness sake, sketched at the end of the proof. We assume that the system of equations is satisfied and will prove that the associated polynomial solves the differential equation. Fix $x_1, \dots, x_n$ and consider the candidate polynomial $$ f(x) = \prod_{k=1}^{n}{(x - x_k)}.$$ We want to show that this polynomial satisfies the differential equation $$- (p(x) f')' + q(x) f' = \lambda f \qquad \mbox{for some}~\lambda \in \mathbb{R}.$$ We introduce the function \begin{align*} \frac{\partial}{\partial t} u(t,x) &= \frac{\partial}{\partial x}\left( p(x) \frac{\partial}{\partial x} u(t,x) \right) - q(x) \frac{\partial}{\partial x} u(t,x) \\ u(0,x) &= f(x) \end{align*} This is a parabolic partial differential equation. Generally, unless we specify the sign of $p$, it might be an inverse heat equation and the solution of such an equation need not even exist for a short amount of time. Here, however, since $p$ is at most a polynomial of degree 2 and $q$ is a polynomial of degree at most 1, we see that the right-hand side is always a polynomial of degree at most $n$. In particular, we can rewrite the partial differential equation as a linear system of $n+1$ ordinary differential equations and this guarantees existence for all $t>0$. Suppose $f(x)$ is not a solution of the differential equation. Then $$u_t(0,x) = \frac{\partial}{\partial x}\left( p(x) \frac{\partial}{\partial x} f(x) \right) - q(x) \frac{\partial}{\partial x} f(x)$$ is a polynomial of degree at most $n$ and not identically 0 (since otherwise $f$ would solve the differential equation for $\lambda = 0$). We observe that if this polynomial vanishes in exactly $\left\{x_1, \dots, x_n\right\}$, then it has to be a multiple of $f$ and we have obtained the desired result for some $\lambda \neq 0$. If this is not the case, then $u_t(x_i,0) \neq 0$ for at least one $1 \leq i \leq n$. We fix this value of $i$. Moreover, we note $$f'(x_i) = \prod_{k=1 \atop k \neq i}^{n}{(x_i - x_k)} \neq 0.$$ The implicit function theorem implies that there is a neighborhood of 0 for which there is a function $x_i(t)$ such that $$ u( t, x_i(t)) = 0 \qquad \mbox{and} \qquad \frac{\partial}{\partial t} x_i(t) \big|_{t = 0} \neq 0.$$ We now compute the expression. Differentiation implies $$ 0 = \frac{\partial}{\partial t} u(t,x_i(t))\big|_{t=0} = u_x(0,x_i)\left(\frac{\partial}{\partial t} x_i(t) \big|_{t=0}\right) + u_t(0,x_i).$$ We are interested in the first term, already computed the second term $u_x(x_i,0) = f'(x_i)$ and thus only need to compute the third term. A simple computation shows \begin{align*} u_t(t,x_i)\big|_{t=0} &= \frac{\partial}{\partial x}\left( p(x) \frac{\partial}{\partial x} f(x) \right) - q(x) \frac{\partial}{\partial x} f(x) \big|_{x=x_i} \\ &= p(x) \frac{\partial^2}{\partial x^2} f(x) + (p'(x) - q(x))\frac{\partial}{\partial x} f(x) \big|_{x=x_i}. \end{align*} The first term simplifies to $$ \frac{\partial^2}{\partial x^2} f(x) \big|_{x=x_i} = 2 \sum_{k =1 \atop k \neq i}^{n}{ \prod_{j = 1 \atop j \notin \left\{i, k\right\}}^{n}{(x_i - x_j)}}$$ and altogether we obtain $$ 0 \neq \frac{\partial}{\partial t} x_i(t) \big|_{t = 0} = p(x_i) \sum_{k = 1 \atop k \neq i}^{n}{\frac{2}{x_k - x_i}} + p'(x_i) - q(x_i)$$ which is a contradiction. Conversely, if $f$ is indeed a solution of the equation, then $$ u(t,x) = e^{\lambda t} f(x) \qquad \mbox{and thus} \qquad \frac{\partial}{\partial t} x_i(t) \big|_{t = 0} = 0$$ for all $1 \leq i \leq n$ which implies that the equations are satisfied. \end{proof} \begin{center} \begin{figure}[h!] \begin{tikzpicture}[scale=1] \draw [thick, ->] (-4.3,-3) -- (-4.3,3); \node at (4.7, -2.8) {$t$}; \filldraw (-4.3, 2.6) circle (0.05cm); \draw [thick, ->] (-4.3,-2.6) -- (4.7,-2.6); \node at (-4.7, 2.6) {371}; \filldraw (-4.3, -2.6) circle (0.05cm); \filldraw (4.1, -2.6) circle (0.05cm); \node at (4.1, -2.8) {4}; \node at (-4.1, 2.9) {$x$}; \node at (-4.6, -2.6) {0}; \node at (-0.02,0.04) {\includegraphics[width=0.7\textwidth]{laguerre.pdf}}; \end{tikzpicture} \caption{Another example: starting with initial values $x_i(0) = i$ for $1 \leq i \leq 100$, the associated system of ODEs approaches the zeros of the Laguerre polynomial $L_{100}$. } \end{figure} \end{center} \vspace{-10pt} \subsection{Proof of Theorem 2.} \begin{proof} We assume that the solutions $y_0, y_1, \dots, y_{n-1}, y_n$ satisfy $\mbox{deg} (y_i) = i$ for all $0 \leq i \leq n$. We assume that $p, q$ and the domain of definition are such that classical Sturm-Liouville theory applies. Then the polynomial $y_j$ has exactly $j$ zeros all of which are simple. Suppose $$ x_1(0) < x_2(0) < \dots < x_n(0)$$ is given. We define the function $$ f(x) = \prod_{k=1}^{n}{(x - x_k(0))} \quad \mbox{and write it as} \quad f(x) = \sum_{k=0}^{n}{a_k y_k(x)}$$ for some coefficients $a_0, \dots, a_n$. This is possible because of the assumption on the degrees (an upper triangular matrix is invertible). The strong form of the Sturm Oscillation Theorem, that is not very widely known, states that, as long as not all coefficients $b_i$ vanish, any function of the form $$ \sum_{k=0}^{n-1}{b_k y_k(x)} \qquad \mbox{has at most}~n-1~\mbox{zeros.}$$ We refer to B\'erard \& Helffer \cite{berard} and L\"utzen \cite{lutzen} for the history of this remarkable Theorem that seems to have been forgotten (\cite{berard} gives rigorous proofs in modern language, \cite{stein} gives a quantitative form). Since $f(x)$ has $n$ zeros, the Sturm Oscillation Theorem implies that $a_n \neq 0$. We now define $u(t,x)$ as the solution of the \textit{backward} heat equation \begin{align*} \frac{\partial}{\partial t} u(t,x) &= - \frac{\partial}{\partial x}\left( p(x) \frac{\partial}{\partial x} u(t,x) \right) + q(x) \frac{\partial}{\partial x} u(t,x) \\ u(0,x) &= f(x) \end{align*} and observe that, as explained in the proof of Theorem 1, this equation is well-posed for all $t>0$ since it can be rewritten as a linear system of $n+1$ ordinary differential equations. Linearity implies that the solution is given by $$ u(t,x) = \sum_{k=0}^{n}{a_k e^{\lambda_k t} y_k(x)}.$$ At the same time, as long as the zeros do not collide, we can write $$ u(t,x) = h(t)\prod_{i=1}^{n}{(x - x_i(t))},$$ where $h(t) \neq 0$ for all $t>0$ and the functions $x_i(t)$ satisfy, following the computation done in the proof of Theorem 1 and reversing the sign, the system of ordinary differential equations $$ \frac{d}{dt} x_i(t) = -p(x_i) \sum_{k = 1 \atop k \neq i}^{n}{\frac{2}{x_k(t) - x_i(t)}} + q(x_i(t)) - p'(x_i(t)) \quad \mbox{for all}~1 \leq i \leq n.$$ It is a property of Sturm-Liouville problems that the number of distinct zeros is nonincreasing in time under the forward heat equation; since we are dealing with the backward heat equation, we see that the number of distinct zeros is nondecreasing. Moreover, since we start with a polynomial of degree $n$ and the solution is always a polynomial of degree at most $n$, this implies that zeros can never collide (and that the solution is always exactly of degree $n$). This shows that $$ e^{-\lambda_n t} h(t)\prod_{i=1}^{n}{(x - x_i(t))} = a_n y_n(t) + \sum_{k=0}^{n-1}{a_k e^{(\lambda_k - \lambda_n) t} y_k(x)}.$$ All zeros of $y_n$ are simple, the Inverse Function Theorem now implies that the zeros of $u(t,x)$ converge to the zeros of $y_n$ exponentially quickly in $t$. The speed of convergence depends on the size of $\lambda_{n-1} - \lambda_n$ and the size of $y_n'$ at its zeros; the constant in front will depend on the precise values of the coefficients $\left\{a_0, a_1, \dots, a_n\right\}$. If $a_{n-1} = 0$ or more of the leading terms vanish, then convergence would be even faster. \end{proof} \vspace{-15pt}
{ "timestamp": "2018-05-17T02:08:36", "yymm": "1804", "arxiv_id": "1804.09697", "language": "en", "url": "https://arxiv.org/abs/1804.09697", "abstract": "We study the differential equation $ - (p(x) y')' + q(x) y' = \\lambda y,$ where $p(x)$ is a polynomial of degree at most 2 and $q(x)$ is a polynomial of degree at most 1. This includes the classical Jacobi polynomials, Hermite polynomials, Legendre polynomials, Chebychev polynomials and Laguerre polynomials. We provide a general electrostatic interpretation of zeros of such polynomials: the set of real numbers $\\left\\{x_1, \\dots, x_n\\right\\}$ satisfies $$ p(x_i) \\sum_{k = 1 \\atop k \\neq i}^{n}{\\frac{2}{x_k - x_i}} = q(x_i) - p'(x_i) \\qquad \\mbox{for all}~ 1\\leq i \\leq n$$ if and only if they are zeros of a polynomial solving the differential equation. We also derive a system of ODEs depending on $p(x),q(x)$ whose solutions converge to the zeros of the orthogonal polynomial at an exponential rate.", "subjects": "Classical Analysis and ODEs (math.CA); Spectral Theory (math.SP)", "title": "Electrostatic Interpretation of Zeros of Orthogonal Polynomials", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9848109489630493, "lm_q2_score": 0.8128673201042492, "lm_q1q2_score": 0.8005206368929164 }
https://arxiv.org/abs/1007.3467
Factoring Permutation Matrices Into a Product of Tridiagonal Matrices
Gilbert Strang posited that a permutation matrix of bandwidth $w$ can be written as a product of $N < 2w$ permutation matrices of bandwidth 1. A proof employing a greedy ``parallel bubblesort'' algorithm on the rows of the permutation matrix is detailed and further points of interest are elaborated.
\section{Conjecture and Outline}\label{conjover} This section states the problem, starting with some necessary definitions. As a convention, $M'$ will denote the transpose of a matrix $M$ and $I$ is the identity matrix. \begin{definition}\label{permmatrix} An $n \times n$ \emph{permutation matrix} $P$ contains only 0s and 1s, with only one 1 per row and column. The permutation of a column vector $\vec{x}$, where $\vec{x}' = [1\ 2\ \cdots\ n]$, is the column vector $P\vec{x}$. A matrix $M = [m_{ij}]$ is said to be \emph{of bandwidth $w$}, denoted by $\bandwidth{P} := w$, if $m_{ij} = 0$ whenever $|i - j| > w$. \end{definition} The value of $\bandwidth{M}$ is 0 if $M$ is diagonal, 1 if $M$ is tridiagonal and 2 if $M$ is pentadiagonal. Gilbert Strang posed the following conjecture \cite{Strang : \begin{conjecture}[Strang]\label{strangconjmat} A finite permutation matrix of bandwidth $w > 0$ can be written as the product of at most $2w - 1$ bandwidth--1 permutation matrices. \end{conjecture} \begin{example}\label{examplep} \begin{displaymath} P = \begin{bmatrix} 0 & 0 & 1 & 0\\ 1 & 0 & 0 & 0\\ 0 & 0 & 0 & 1\\ 0 & 1 & 0 & 0 \end{bmatrix} = \begin{bmatrix} 0 & 1 & 0 & 0\\ 1 & 0 & 0 & 0\\ 0 & 0 & 0 & 1\\ 0 & 0 & 1 & 0 \end{bmatrix} \begin{bmatrix} 1 & 0 & 0 & 0\\ 0 & 0 & 1 & 0\\ 0 & 1 & 0 & 0\\ 0 & 0 & 0 & 1 \end{bmatrix} = R_1R_2. \end{displaymath} \end{example} This paper aims to prove Conjecture~\ref{strangconjmat} and explore topics opened during the development of this proof. The next three sections (Sections~\ref{canonperms},~\ref{overtake}, and \ref{proveconj}) will cover the proof of the conjecture, outlining a greedy ``parallel bubblesort''\cite{Strang} strategy to determine a factor per iteration, starting from a specific key class of permutation matrices through progressively larger classes of permutation matrices. Section~\ref{furtherpoints} concludes the paper with some points of further interest. \section{Strang Canonical Matrices}\label{canonperms This section performs three tasks. First, it outlines the general ``parallel bubblesort'' algorithm. Second, it introduces two classes of permutation matrices: Strang canonical matrices and settled matrices. Third, Theorem~\ref{solidrepthm}, the main result of this section, establishes that Conjecture~\ref{strangconjmat} holds for tracefree settled Strang canonical matrices with so-called reducing matrix factors. \begin{definition} If $P$ is an $n \times n$ permutation matrix, then the $m$th row of $P$, $1 \leq m \leq n$ is $\rowindexed{m}{P}$, the $m$th column of $P$ is $\colindexed{m}{P}$ and $[\columnone{1}{P}\ \cdots\ \columnone{n}{P}] := (P\vec{x})'$ where $\vec{x}' = [1\ \cdots\ n]$. If $1 \leq i < j \leq n$, $\rowindexed{i}{P}$ and $\rowindexed{j}{P}$ are an \emph{inverted pair} if $\columnone{i}{P} > \columnone{j}{P}$ and are a \emph{contented pair} otherwise. If $P = \prod\limits_{k = 1}^m T_k$, where $T_k$ are bandwidth--1 permutation matrices, then $\numfactors{P} := m$. \end{definition} If an $n \times n$ permutation matrix $P$ represents a permutation $\sigma$, then $\columnone{i}{P} = \sigma(i)$ for $1 \leq i \leq n$, and each inverted pair of $P$ represents an inversion of $\sigma$. Also, $\bandwidth{P} = \max_{1 \leq i \leq n}|\columnone{i}{P} - i|$. \begin{remark}\label{existproduct} The bubblesort algorithm \cite[p.~40]{Algo} shows that each permutation $\sigma$ is the product of transposition of adjacent elements, $\sigma = \prod \tau_i$. Let $P_\rho$ be the permutation matrix representing $\rho$. Then $P_{\tau_i}$ are all bandwidth--1 permutation matrices, and $P_\sigma = \prod P_{\tau_i}$. Thus, a permutation matrix can always be written as a product of bandwidth--1 permutation matrices. Let $P = \prod\limits_{k = 1}^m T_k$ be a permutation matrix where $T_k$ are bandwidth--1 permutation matrices. Since permutation matrices are unitary and bandwidth--1 permutation matrices are inversions, $${T_k}^2 = I \text{ implies } T_k' = {T_k}^{-1} = T_k \text{ and } P^{-1} = \left(\prod_{k = 1}^m T_k\right)^{-1} = \prod_{k = 1}^m {T_{m + 1 - k}}^{-1} = \prod_{k = 1}^m T_{m + 1 - k} = P'.$$ Thus, $\numfactors{P} = \numfactors{P^{-1}}$. Whenever such a product is defined, denote the indexed matrices $P_0 = P$ and $P_k = T_kP_{k - 1} = \left(\prod\limits_{i = 1}^k T_{k + 1 - i}\right)P$. \end{remark} \begin{lemma}\label{algoterm} If $P = P_0$ is a finite permutation matrix, and $P_k = B_kP_{k - 1}$ where $B_k$ is a nonidentity permutation matrix that performs swaps only on inverted pairs of $P_{k - 1}$, then there is a number $m$ such that $P_m = I$. \end{lemma} \begin{proof} Since $B_k$ makes some inverted pairs of $P_{k - 1}$ contented, the number of inverted pairs of $P_k$ is less than that of $P_{k - 1}$. Since the number of inverted pairs of $P$ is finite, there must be a number $m$ such that $P_m = I$. \end{proof} When $B_k$ only swaps adjacent rows, Lemma~\ref{algoterm} describes ``parallel bubblesorting'', as each bubblesort iteration reduces the number of inversions of a permutation, and $P = \prod\limits_{k = 1}^m B_k$ is the required decomposition. The rest of the paper investigates the greedy selection operation $\{B_k\}$ to ensure that $m = \numfactors{P} < 2w$ where $w = \bandwidth{P}$. Given $P = \prod\limits_{k = 1}^m B_k$ as in Lemma~\ref{algoterm}, for $0 \leq i, j \leq m$, $\rowindexed{m_i}{P_i} = \rowindexed{m_j}{P_j}$ if $\columnone{m_i}{P_i} = \columnone{m_j}{P_j}$. \begin{definition} Treating $P$ as a block diagonal matrix with the finest partition, each diagonal matrix is called a \emph{section} of $P$. A $1 \times 1$ section is \emph{trivial}. A row of a permutation matrix is said to be \emph{positive} (\emph{negative}, \emph{neutral}) if the 1 is to the right of (to the left of, on, respectively) the diagonal. A column of a permutation matrix is said to be \emph{positive} (\emph{negative}, \emph{neutral}) if the 1 is above (below, on, respectively) the diagonal. \end{definition} Each section $S_1, \dots, S_k$ of a permutation matrix $P$ is a permutation matrix, and $\numfactors{P} = \max\{\numfactors{S_1}, \dots, \numfactors{S_k}\}$. \begin{figure} \begin{center} \begin{picture}(55,55)(-30,-30) \put(-20,-20){\makebox(0,0)[c]{$0$}} \put(-20,-10){\makebox(0,0)[c]{$0$}} \put(-20,0){\makebox(0,0)[c]{$1$}} \put(-20,10){\makebox(0,0)[c]{$0$}} \put(-10,-20){\makebox(0,0)[c]{$1$}} \put(-10,-10){\makebox(0,0)[c]{$0$}} \put(-10,0){\makebox(0,0)[c]{$0$}} \put(-10,10){\makebox(0,0)[c]{$0$}} \put(0,-20){\makebox(0,0)[c]{$0$}} \put(0,-10){\makebox(0,0)[c]{$0$}} \put(0,0){\makebox(0,0)[c]{$0$}} \put(0,10){\makebox(0,0)[c]{$1$}} \put(10,-20){\makebox(0,0)[c]{$0$}} \put(10,-10){\makebox(0,0)[c]{$1$}} \put(10,0){\makebox(0,0)[c]{$0$}} \put(10,10){\makebox(0,0)[c]{$0$}} \put(20,-20){\makebox(0,0)[c]{$-$}} \put(20,-10){\makebox(0,0)[c]{$+$}} \put(20,0){\makebox(0,0)[c]{$-$}} \put(20,10){\makebox(0,0)[c]{$+$}} \put(-20,20){\makebox(0,0)[c]{$-$}} \put(-10,20){\makebox(0,0)[c]{$-$}} \put(0,20){\makebox(0,0)[c]{$+$}} \put(10,20){\makebox(0,0)[c]{$+$}} \multiput(-25,-25)(40,0){2}{\line(0,1){40}} \multiput(-25,-25)(37,0){2}{\multiput(0,0)(0,40){2}{\line(1,0){3}}} \end{picture} \caption{Signs of rows and columns of $P$} \label{chargetargetex} \end{center} \end{figure} Each nontrivial section has a positive top row, a negative bottom row, a negative leftmost column and a positive rightmost column. \begin{remark}\label{signcount} Each 1 on the diagonal of a permutation matrix $P$ determines a neutral row and column. Observe that the number of 1s in the upper triangle of $P$ is the sum of the number of its positive and neutral rows, and the sum of the number of its positive and neutral columns. Thus, $P$ has the same number of positive rows and columns. Observing the number of 1s in the lower triangle of $P$ similarly shows that $P$ has the same number of negative rows and columns. \end{remark} \begin{definition} Two permutation matrices are \emph{row-sign-equivalent} (\emph{column-sign-equivalent}) if the signs of their rows (columns) are the same. A section is \emph{row-settled} if all of its positive rows are above its negative rows, \emph{column-settled} if all of its negative columns are to the left of its positive columns and \emph{settled} if it is either row-settled or column-settled. A row-settled (column-settled) matrix has only row-settled (column-settled) sections, and a settled matrix is either row-settled or column-settled. A section is \emph{upper-canonical} (\emph{lower-canonical}), or in \emph{upper-canonical form} (\emph{lower-canonical form}), if its positive (negative) rows are pairwise contented. A section is \emph{Strang canonical} (\emph{half-canonical}), or in \emph{Strang canonical form} (\emph{half-canonical form}), if it is in both (either) upper-canonical and (or) lower-canonical form. A permutation matrix is upper-canonical (lower-canonical), or in upper-canonical form (in lower-canonical form) if all of its sections are in upper-canonical (lower-canonical) form. A permutation matrix is Strang canonical (half-canonical), or in Strang canonical (half-canonical) form, if it is in both (either) upper-canonical and (or) lower-canonical form. \end{definition} The inverse of a row-settled matrix is a column-settled matrix, and vice-versa. The inverse of an upper-canonical matrix is a lower-canonical matrix, and vice versa. If $U$ is an upper-canonical matrix and the 1s in its upper triangle excluding its diagonal are in rows $p_1, \dots, p_k$ such that $p_i < p_{i + 1}$, then $\columnone{p_i}{U} < \columnone{p_{i + 1}}{U}$ for $1 \leq i < k$. If $L$ is a lower-canonical matrix and the 1s in its lower triangle excluding its diagonal are in rows $n_1, \dots, n_\ell$ such that $n_i < n_{i + 1}$, then $\columnone{n_i}{L} < \columnone{n_{i + 1}}{L}$ for $1 \leq i < \ell$. \begin{example} The only $n \times n$, $n \leq 4$, sections that are not in Strang canonical form are \begin{displaymath} \begin{bmatrix} 0 & 0 & 1 & 0\\ 0 & 0 & 0 & 1\\ 0 & 1 & 0 & 0\\ 1 & 0 & 0 & 0 \end{bmatrix}, \begin{bmatrix} 0 & 0 & 0 & 1\\ 0 & 0 & 1 & 0\\ 1 & 0 & 0 & 0\\ 0 & 1 & 0 & 0 \end{bmatrix} \text{ and } \begin{bmatrix} 0 & 0 & 0 & 1\\ 0 & 0 & 1 & 0\\ 0 & 1 & 0 & 0\\ 1 & 0 & 0 & 0 \end{bmatrix}. \end{displaymath} The eight $4 \times 4$ tracefree settled sections are the above three matrices and \begin{displaymath} \begin{bmatrix} 0 & 1 & 0 & 0\\ 0 & 0 & 1 & 0\\ 0 & 0 & 0 & 1\\ 1 & 0 & 0 & 0 \end{bmatrix}, \begin{bmatrix} 0 & 0 & 1 & 0\\ 0 & 0 & 0 & 1\\ 1 & 0 & 0 & 0\\ 0 & 1 & 0 & 0 \end{bmatrix}, \begin{bmatrix} 0 & 1 & 0 & 0\\ 0 & 0 & 0 & 1\\ 1 & 0 & 0 & 0\\ 0 & 0 & 1 & 0 \end{bmatrix}, \begin{bmatrix} 0 & 0 & 1 & 0\\ 1 & 0 & 0 & 0\\ 0 & 0 & 0 & 1\\ 0 & 1 & 0 & 0 \end{bmatrix} \text{ and } \begin{bmatrix} 0 & 0 & 0 & 1\\ 1 & 0 & 0 & 0\\ 0 & 1 & 0 & 0\\ 0 & 0 & 1 & 0 \end{bmatrix}. \end{displaymath} \end{example} \begin{remark}\label{uniqcanonsign} A Strang canonical matrix is uniquely determined by the signs of its rows and columns. The $m$th 1 on its diagonal is at the intersection of its $m$th neutral row and $m$th neutral column. The $m$th 1 in its upper triangle, excluding the diagonal, is at the intersection of its $m$th positive row and $m$th positive column. The $m$th 1 in its lower triangle, excluding the diagonal, is at the intersection of its $m$th negative row and $m$th negative column. \end{remark} \begin{definition} A \emph{reducing swap} $\hat{R}$ of a permutation matrix $P$ is an elementary matrix that swaps adjacent rows of $P$ where the upper row is positive and the lower row is negative. The pair of rows are \emph{reduced} by the swap. The \emph{reducing matrix} $R$ of permutation matrix $P$ is $R = \prod \hat{R}$ over all possible reducing swaps $\hat{R}$ of $P$, and the \emph{reduction} of $P$ is $\reducemat{P} := RP$. \end{definition} Each reducing swap makes an inverted pair contented, and if $R$ is the reducing matrix of a nonidentity permutation matrix, $\bandwidth{R} = 1$. \begin{example} In Example~\ref{examplep}, $P = R_1R_2$, where $R_1$ is the reducing matrix of $P$ and $\reducemat{P} = P_1 = R_1P = R_2$. \end{example} \begin{lemma}\label{canonderangeredlemma} If $P$ is a Strang canonical matrix whose nontrivial sections are tracefree, then $P = \prod\limits_{k = 1}^m R_k$, where $R_k$ is the reducing matrix of $P_{k - 1}$. \end{lemma} \begin{proof}\label{canonderangeredproof} The permutation matrix $P$ and $\reducemat{P}$ have the same Strang canonicity, since reducing swaps only exchange the positions of a positive row and a negative row, so each $P_k$ is Strang canonical. Since the nontrivial sections of $P = P_0$ are tracefree, the only neutral rows of $P$ are in trivial sections. Let $$P_k = \begin{bmatrix}A_k & 0 & B_k\\0 & 1 & 0\\C_k & 0 & D_k\end{bmatrix}, k > 0 \text{ and } P_{k - 1} = \begin{bmatrix}A_{k - 1} & B_{k - 1}\\C_{k - 1} & D_{k - 1}\end{bmatrix}.$$ The indicated neutral row of $P_k$ is in a trivial section if $B_k$ and $C_k$ are zero matrices. Let $A_k$ be $(m - 1) \times (m - 1)$ and $\rowindexed{m}{P_{k - 1}}$ be signed. If $\rowindexed{m}{P_{k - 1}}$ is negative, let $A_{k - 1}$ be $m \times m$. Since $P_{k - 1}$ is Strang canonical, every negative $\rowindexed{m'}{P_{k - 1}}$ with $m' < m$ has its 1 in $A_{k - 1}$. Then $\rowindexed{m - 1}{P_{k - 1}}$ is positive with $\columnone{m - 1}{P_{k - 1}} = m$ and every positive $\rowindexed{m'}{P_{k - 1}}$ with $m' < m$ has its 1 in $A_{k - 1}$. Since each neutral $\rowindexed{m'}{P_{k - 1}}$ with $m' < m$ has its 1 in $A_{k - 1}$, both $B_{k - 1}$ and $C_{k - 1}$ are zero matrices, so $B_k$ and $C_k$ are zero matrices. If $\rowindexed{m}{P_{k - 1}}$ is positive, let $A_{k - 1}$ be $(m - 1) \times (m - 1)$, and it similarly follows that $B_{k - 1}$, $C_{k - 1}$, $B_k$ and $C_k$ are zero matrices. Thus, $P$ and $\reducemat{P}$ are Strang canonical matrices with tracefree nontrivial sections, and the conclusion follows from Lemma~\ref{algoterm}. \end{proof} \begin{theorem}\label{solidrepthm} A tracefree settled Strang canonical matrix of bandwidth $w$ can be written as the product of less than $2w$ bandwidth--1 matrices. \end{theorem} \begin{proof} Let $P$ be a row-settled Strang canonical matrix with $\tracemat{P} = 0$ and $\bandwidth{P} = w$. Thus, it has an $n \times n$ row-settled Strang canonical section $S$ with $\tracemat{S} = 0$ and $\bandwidth{S} = w$. Let $\columnone{m}{S} = n$. Then the upper $m$ rows of $S$ are positive, and the rest are negative, with $\columnone{m + 1}{S} = 1$. From Lemma~\ref{canonderangeredlemma}, a row-settled Strang canonical section is the product of reducing matrices: once a row is reduced, it is reduced by the next reducing matrix, until it is in a trivial section. Hence, \begin{equation} \numfactors{S} = \max_{1 \leq i \leq m} \{\columnone{i}{S} + m - 2 * i\} = \max_{m < i \leq n} \{2 * i - m - 1 - \columnone{i}{S}\}\label{settledcount}\tag{*} \end{equation} $|\columnone{i}{S} - i| \leq w$ indicates the \emph{swap count}, which is the number of reducing swaps for $\rowindexed{i}{S}$ to be placed in the $\columnone{i}{S}$th row, and $m - i$, if $i \leq m$, ($i - (m + 1)$, if $i > m$) is the \emph{delay count}, which is the number of positive (negative) rows that must be reduced before the positive (negative) $\rowindexed{i}{S}$ is first reduced. Since $\columnone{m}{S} = n$, $S$ has $n - m = \columnone{m}{S} - m \leq w$ negative rows; since $\columnone{m + 1}{S} = 1$, $S$ has $m = m + 1 - \columnone{m + 1}{S} \leq w$ positive rows. Thus, if $\bandwidth{S} = w$, then $\numfactors{S} \leq w - 1 + w = 2w - 1$. Since a column-settled Strang canonical matrix $P$ is the inverse of a row-settled Strang canonical matrix $P^{-1}$ and, by Remark~\ref{existproduct}, $\numfactors{P} = \numfactors{P^{-1}}$, the conclusion follows. \end{proof} \begin{remark}\label{eventight} If $C$ is an $n \times n$ circulant nonidentity permutation matrix, then it is Strang canonical, row-settled and column-settled, with $\tracemat{C} = 0$ and $\numfactors{C} = n - 1$, all reducing matrices. Moreover, for some $k$, $1 < k \leq n$, it follows that $\columnone{m}{C} = ((k + m - 2) \bmod n) + 1$, $$\bandwidth{C} = \max\{\columnone{1}{C} - 1, n - \columnone{n}{C}\} = \left\{\begin{array}{cl}k - 1, & \text{ if } k - 1 \geq n / 2\\n - k + 1, & \text{ if } k - 1 \leq n / 2\end{array}\right.,$$ $\colindexed{k}{C}$ is its first positive column and $\rowindexed{n - k + 2}{C}$ is its first negative row. $\numfactors{C}$ is tight with Strang's bound if $n = 2w$ and $w = \bandwidth{C}$. \end{remark} \begin{corollary}\label{circbound} Let $C$ be a circulant matrix, $R$ be a row-settled Strang canonical matrix which is row-sign-equivalent to $C$ and $P$ be a column-settled Strang canonical matrix which is column-sign-equivalent to $C$. Then $\numfactors{C} \geq \numfactors{R}, \numfactors{P}$. \end{corollary} \begin{proof} From Theorem~\ref{solidrepthm}, for each row of $R$ and $C$ with the same index, the delay count is the same, but the swap count is maximum for $C$. So, from Equation~\eqref{settledcount}, $\numfactors{C} \geq \numfactors{R}$. Since $P^{-1}$ is row-settled and Strang canonical and $C^{-1}$ is circulant, $\numfactors{C} = \numfactors{C^{-1}} \geq \numfactors{P^{-1}} = \numfactors{P}$ by Remark~\ref{existproduct}. \end{proof} \section{Overtaking Swaps}\label{overtake} Theorem~\ref{solidrepessthm}, the main result in this section, establishes that Conjecture~\ref{strangconjmat} holds for settled Strang canonical matrices by a comparison with tracefree matrice . \begin{remark}\label{notfixedonly} Since the neutral rows of a permutation matrix $P = P_0$ are pairwise contented, if $P_k = B_kP_{k - 1}$ as in Lemma~\ref{algoterm} and all the signed rows of $P$ are neutral in $P_m$, then $P_m = I$. \end{remark} \begin{definition} An \emph{overtaking swap} of a permutation matrix $P$ is an elementary matrix that swaps an adjacent inverted pair of $P$ where either the upper row is not positive or the lower row is not negative. The upper positive row or the lower negative row \emph{overtakes} the other row by the swap. \begin{align*} \text{If } P = \begin{bmatrix}A & B\\C & D\end{bmatrix} & \text{ where A\text{ is } (m - 1) \times (m - 1), \text{ define } \insertfix{m}{P} := \begin{bmatrix}A & 0 & B\\0 & 1 & 0\\C & 0 & D\end{bmatrix}\\ & \text{ and } \insertfix{m_1, \dots, m_k}{P} := \insertfix{m_k}{\cdots\left(\insertfix{m_1}{P}\right)}.\\ \text{If } P = \begin{bmatrix}A & 0 & B\\0 & 1 & 0\\C & 0 & D\end{bmatrix} & \text{ where A \text{ is } (m - 1) \times (m - 1), \text{ define } \deletefix{m}{P} := \begin{bmatrix}A & B\\C & D\end{bmatrix}\\ & \text{ and } \deletefix{m_1, \dots, m_k}{P} := \deletefix{m_k}{\cdots\left(\deletefix{m_1}{P}\right)}. \end{align*} If $P$ is a nonidentity permutation matrix whose neutral rows have indices $r_1, \dots, r_k$, $r_i > r_{i + 1}$, $\essence{P} := \deletefix{r_1, \dots, r_k}{P}$ is the \emph{essential form} or \emph{essence} of $P$. \end{definition} \begin{remark}\label{fixedinserteffect} If $P$ is an $n \times n$ permutation matrix and $1 \leq m \leq n$, then $\bandwidth{\insertfix{m}{P}} \leq \bandwidth{P} + 1$. \end{remark} \begin{example}\label{transpex} To demonstrate the effect of inserting neutral rows on the number of factors, consider $T_1 = \begin{bmatrix}0 & 1\\1 & 0\end{bmatrix}$, the unique bandwidth--1 section. The factorization of $T_k = \insertfix{2}{T_{k - 1}}$, with matrices containing overtaking swaps underline \ is as follows: \begin{align*} T_2 & = \insertfix{2}{T_1} = \begin{bmatrix} 0 & 0 & 1\\ 0 & 1 & 0\\ 1 & 0 & 0 \end{bmatrix} = \underline{ \begin{bmatrix} 0 & 1 & 0\\ 1 & 0 & 0\\ 0 & 0 & 1 \end{bmatrix}} \begin{bmatrix} 1 & 0 & 0\\ 0 & 0 & 1\\ 0 & 1 & 0 \end{bmatrix} \begin{bmatrix} 0 & 1 & 0\\ 1 & 0 & 0\\ 0 & 0 & 1 \end{bmatrix};\\ T_3 & = \insertfix{2}{T_2} = \begin{bmatrix} 0 & 0 & 0 & 1\\ 0 & 1 & 0 & 0\\ 0 & 0 & 1 & 0\\ 1 & 0 & 0 & 0 \end{bmatrix} = \underline{ \begin{bmatrix} 0 & 1 & 0 & 0\\ 1 & 0 & 0 & 0\\ 0 & 0 & 0 & 1\\ 0 & 0 & 1 & 0 \end{bmatrix}} \begin{bmatrix} 1 & 0 & 0 & 0\\ 0 & 0 & 1 & 0\\ 0 & 1 & 0 & 0\\ 0 & 0 & 0 & 1 \end{bmatrix} \begin{bmatrix} 0 & 1 & 0 & 0\\ 1 & 0 & 0 & 0\\ 0 & 0 & 0 & 1\\ 0 & 0 & 1 & 0 \end{bmatrix};\text{ and }\\ T_4 & = \insertfix{2}{T_3} = \begin{bmatrix} 0 & 0 & 0 & 0 & 1\\ 0 & 1 & 0 & 0 & 0\\ 0 & 0 & 1 & 0 & 0\\ 0 & 0 & 0 & 1 & 0\\ 1 & 0 & 0 & 0 & 0 \end{bmatrix} = \underline{O_1O_2}R_1R_2R_3, \end{align*} where $O_1$ swaps the first and last pairs of rows of $T_4$ and $O_2$ swaps the middle row of $O_1T_4$. \end{example} When a neutral row is overtaken, it assumes the sign of the overtaking row. \begin{theorem}\label{solidrepessthm} A settled Strang canonical matrix of bandwidth $w$ can be written as the product of less than $2w$ bandwidth--1 matrices. \end{theorem} \begin{proof} To show the result for a settled Strang canonical matrix $P$, a circulant matrix $C$ will be used to determine an upper bound for $\numfactors{P}$. Let $C$ be $n \times n$ with $\colindexed{c}{C}$ as its first positive column and $\rowindexed{r}{C}$ as its first negative row. From Remark~\ref{eventight}, $C = \prod\limits_{k = 1}^{n - 1} R_k$ where $R_k$ is the reducing matrix of $C_{k - 1}$. For $P = \insertfix{m}{C}$, let $w = \bandwidth{P} = \bandwidth{C} + 1$. If $m = r$, the neutral row is inserted between the positive and negative rows. The initial reduction $C_1$ is delayed by an overtake of the neutral row by $\rowindexed{m'}{P}$. If $2m < n$, the neutral row is closer to $\rowindexed{1}{P}$. If $2m > n$, it is closer to $\rowindexed{n + 1}{P}$. So the neutral row is overtaken toward whichever row between $\rowindexed{1}{P}$ and $\rowindexed{n + 1}{P}$ it is closer to, or either if $2m = n$, and $\rowindexed{m'}{P}$ has the sign of $n - 2m$. Let the overtaking swap be $O$ and $OP$ replace $P$. \begin{enumerate}[\text{Step} 1] \item If $\rowindexed{r'}{P}$ is first negative row of $P$, then $r' \in \{r, r + 1\}$. Let $k = \min\{r' - 1, n + 1 - r'\}$ and $e = |r' - k| \in \{1, n + 1\}$. For the matrices $O_q$ with $1 \leq q \leq k$, if $\columnone{m_q}{P_{q - 1}} = m$, then $O_q = \insertfix{m_q}{R_q}\hat{O}_q$ with $\hat{O}_q = I$ if $P_{q - 1}$ is Strang canonical or $\hat{O}_q$ is the overtaking swap of $\rowindexed{m_q}{P_{q - 1}}$ otherwise. \item $P_k$ is Strang canonical and $\rowindexed{e}{P_k}$ is neutral. For $q > k$, $O_q$ can be determined by showing that $P_k = \insertfix{e}{\hat{C}_{k - 1}}$ where: \begin{enumerate}[\text{Case} 1] \item If $\rowindexed{m}{P_k}$ is neutral, then $P_k = \insertfix{e, e}{\bar{C}_{k - 1}}$, so $\hat{C} = \insertfix{e}{\bar{C}}$, where $\bar{C}$ is an $n - 1 \times n - 1$ circulant matri . \item If $m$ is between $c$ and $r$ or $m = c$, then $P_k = \insertfix{e}{C_{k - 1}}$, so $\hat{C} = C$. \item Otherwise, $\hat{C}$ is a row-settled Strang canonical matrix which is row-sign-equivalent to $C . \end{enumerate} Then, for $k \leq q \leq \numfactors{\hat{C}}$, $O_{q + 1} = \insertfix{e}{\hat{R}_q}$, where $\hat{R}_q$ is the reducing matrix of $\hat{C}_{q - 1}$ and, from Corollary~\ref{circbound}, $\numfactors{P} = \numfactors{\hat{C}} + 1 \leq n + 1$. \end{enumerate} Upon the completion of the above steps, it can be determined that $P = \prod\limits_{k = 1}^q \bar{O}_k$ where $\bandwidth{\bar{O}_k} = 1$. If $m = r = c$, then $n = 2(m - 1)$ and $\bandwidth{C} = m - 1$. Thus, $\bandwidth{P} = m = w$, $q = n + 1 = 2w - 1$, $\bar{O}_1 = O$ and $\bar{O}_{k + 1} = O_k$ for $1 \leq q \leq n$. Otherwise, $q \in \{n - 1, n\}$ and, by Remark~\ref{eventight}, $2(w - 1) - 1 \geq n - 1$ making $2w - 1 \geq n + 1 > q$. Therefore, for $P = \insertfix{m}{C}$, $\numfactors{P} < 2w$, and this bound is tight only when $m = \frac{n}{2} + 1$. As seen in Example~\ref{transpex}, the parity of the number of neutral rows inserted as a block, say $P = \insertfix{m, \dots, m}{C}$, whether it is an odd or an even number, may affect $\numfactors{P}$ differently. In particular, when $m = \frac{n}{2} + 1$, $\numfactors{\insertfix{m, m}{C}} = \numfactors{\insertfix{m}{C}}$ and $\bandwidth{\insertfix{m, m}{C}} = \bandwidth{\insertfix{m}{C}} + 1$. For $m_1, \dots, m_\ell$ such that $1 < m_{i + 1} < m_i \leq n$, $$\numfactors{\insertfix{m_1, \dots, m_\ell}{C}} - \numfactors{C} = \sum\limits_{i = 1}^\ell \left(\numfactors{\insertfix{m_i}{C}} - \numfactors{C}\right).$$ Multiple blocks of neutral rows inserted to produce $P$ occasionally add a single bandwidth--1 factor, whenever $P_q$ contains a neutral row between rows of the opposite sign, such as, if $r' > r$, for $\insertfix{r, r + 2, r + 2, r + 2}{C}$ and for $\insertfix{r' - r, \dots, r' - r}{C}$ where there are $r + 1$ neutral rows inserted. Therefore, for any circulant nonidentity permutation matrix $C$, given the class $\mathcal{C}_C = \{P : \essence{P} = C\}$, if $d_P = 2\bandwidth{P} - 1 - \numfactors{P}$, then $d_C \geq 0$, $d_C = d_P$ when $C$ is a $2m \times 2m$ matrix and $P = \insertfix{m + 1}{C}$, otherwise $d_P > d_C$ whenever $P \neq C$,. Finally, if $R$ is a row-settled Strang canonical matrix which is row-sign-equivalent to $C$, from Corollary~\ref{circbound} and by following the previous arguments, for every set $\{m_1, \dots, m_t\}$, $\bandwidth{\insertfix{m_1, \dots, m_t}{R}} = \bandwidth{\insertfix{m_1, \dots, m_t}{C}}$ and $\numfactors{\insertfix{m_1, \dots, m_t}{R}} \leq \numfactors{\insertfix{m_1, \dots, m_t}{C}}$. The argument holds for column-settled Strang canonical $R^{-1}$, and the conclusion follows. \end{proof} \begin{remark}\label{fixedlowbound} If $P$ is a settled Strang canonical matrix with $\bandwidth{P} = w$ and $f$ neutral rows, then $\numfactors{P} < 2w - f$, by the proof of Theorem~\ref{solidrepessthm}, noting the tight-bound exception. \end{remark} \section{Opportunistic Overtaking}\label{proveconj} The main result of this section is the completion of the proof of Conjecture~\ref{strangconjmat} with opportunistic-overtaking matrices---a greedy generalization of reducing matrices---along the construction used in Theorem~\ref{solidrepessthm . \begin{definition} If $P$ is a permutation matrix, then $\invertrows{m}{P}$ is the minimal submatrix containing only consecutive rows of $P$ such that $\rowindexed{m}{P}$ and all of the rows of $P$ that are pairwise inverted with $\rowindexed{m}{P}$ are in $\invertrows{m}{P}$. An \emph{inverted block} of a permutation matrix is a maximal submatrix containing only consecutive rows such that all the rows are pairwise inverted. An \emph{opportunistic-overtaking matrix} $O$ of a permutation matrix $P$ is the product of the reducing matrix of $P$ and the overtaking swaps of $P$ such that, for every inverted block of $P$, the only rows that $O$ can leave unswapped are the first and the last rows of the block. The collection of all products of $P$ with any of its opportunistic-overtaking matrices $O$ is denoted by $\oomat{P} \ni OP$. Given a signed $\rowindexed{m}{P}$, \emph{removing its sign} produces a matrix $P' := \unsignrow{m}{P}$. If $m'$ is such that $\columnone{m'}{P} = m$, then: $\rowindexed{k}{P'} = \rowindexed{k}{P}$ when $k \neq m, m'$; $\rowindexed{m'}{P'} = \rowindexed{m}{P}$; and $\rowindexed{m}{P'}$ is neutral. \end{definition} For every $\rowindexed{m}{P}$ not in a trivial section, the top row of $\invertrows{m}{P}$ is positive and the bottom row of $\invertrows{m}{P}$ is negative. If $P$ is upper-canonical and $\rowindexed{m}{P}$ is positive, it is the top row of $\invertrows{m}{P}$. If $P$ is lower-canonical and $\rowindexed{m}{P}$ is negative, it is the bottom row of $\invertrows{m}{P}$. If the rows of an $n \times n$ permutation matrix $P$, from $\rowindexed{i}{P}$ to $\rowindexed{j}{P}$, form an inverted block, then \begin{itemize} \item either $i = 1$ or $\rowindexed{i - 1}{P}$ and $\rowindexed{i}{P}$ are a contented pair, and \item either $j = n$ or $\rowindexed{j}{P}$ and $\rowindexed{j + 1}{P}$ are a contented pair. \end{itemize} If $P$ is a nonidentity permutation matrix and $O$ is any of its opportunistic-overtaking matrices, then $\bandwidth{O} = 1$ and $O$ satisfies the ``parallel bubblesort'' condition of $B_k$ from Lemma~\ref{algoterm}, while providing a locally-optimal, i.e. greedy, condition to determine the next bandwidth--1 factor, in that $P$ and $OP \in \oomat{P}$ share no inverted pairs. The only inverted blocks that a Strang canonical matrix has are reducible pairs and neutral rows with a positive and/or a negative row to overtake it. The product of any permutation matrix and any of its opportunistic-overtaking matrices is a permutation matrix whose inverted blocks have no more than three rows. A permutation matrix $P$ always has a unique reducing matrix. $P$ has a unique opportunistic-overtaking matrix only if each inverted block of $P$ that has more than two rows has a reducible pair, otherwise that block can have two choices of overtaking swaps. An algorithm for determining an opportunistc-overtaking matrix of $P$ is given in the Appendix. \begin{example} If, as in Theorem~\ref{solidrepthm} and the sections in Lemma~\ref{canonderangeredlemma}, $P$ is Strang canonical and $\tracemat{P} = 0$, then its reducible pairs are inverted blocks and $\oomat{P} = \{\reducemat{P}\}$. In the proof of Theorem~\ref{solidrepessthm}, $P_q \in \oomat{P_{q - 1}} . \end{example} \begin{theorem}\label{strangthm} A permutation matrix of bandwidth $w$ can be written as the product of less than $2w$ bandwidth--1 permutation matrices. \end{theorem} \begin{proof}[Proof of Conjecture~\ref{strangconjmat}] The proof will relax the conditions on the permutation matrix and prove that the conjecture holds for each relaxation. Let $P$ be a lower-canonical matrix with $\bandwidth{P} = w$ and $P = \prod\limits_{k = 1}^q O_k$ where $O_kP_{k - 1} = P_k \in \oomat{P_{k - 1}}$. A negative row of $P$ will be in a trivial section in some $P_k$ only by being swapped by $O_k$ with a positive row of $P$. So, let $\rowindexed{m}{P}$ be positive. \begin{enumerate}[\text{Case} 1] \item If $\rowindexed{m}{P} = \rowindexed{m_k}{P_k}$ is never overtaken in $\{O_k\}$, the plan is to localize the determination of the swaps that move $\rowindexed{m}{P}$ in $\{O_k\}$. First, determine $\tilde{P}^m$ such that, through removing the signs of the positive rows that are overtaken by $\rowindexed{m}{P}$. $\tilde{P}^m$ has no such positive row. Next, determine the matrix $P^m$ localizing to the swaps of $\rowindexed{m}{P}$ and the rows that are inverted with it. Then, $P^m$ is column-settled and Strang canonical whose only nontrivial section is from $\rowindexed{m}{P^m}$ to $\rowindexed{\ell_m}{P^m}$, which is row-sign-equivalent to $\invertrows{m}{\tilde{P}^m}$. If, by Theorem~\ref{solidrepessthm}, $P^m = \prod\limits_{k = 1}^{\bar{q}_m} T^m_k$, with $\hat{T}^m_k$ the possibly identity elementary matrix performing the swap in $T^m_k$ that moves $\rowindexed{m_k}{P^m_k} = \rowindexed{m}{P^m}$, can be performed on $\rowindexed{m}{\tilde{P}^m}$, and thus can also be performed on $\rowindexed{m}{P}$. \label{overtakecase There are two scenarios to consider: \begin{enumerate}[\text{Sub-Case} 1] \item Assume that there is a reducible pair in $\invertrows{m}{P}$ that is not in $\invertrows{m}{\tilde{P}^m}$. Since both rows have their signs removed, their reducing swap will be in $O_1$, and $\rowindexed{m}{P}$ will be in a trivial section in $P_{\bar{q}_m + 1}$. By Remark~\ref{fixedlowbound}, $q_m = \bar{q}_m + 1 < 2w - f_m + 1$ with $f_m > 1$. \item Otherwise, $\rowindexed{m}{P}$ will be in a trivial section in $P_{\bar{q}_m}$. By Remark~\ref{fixedlowbound}, $q_m = \bar{q}_m < 2w - f_m$.\label{canonicalalwayssubcase} \end{enumerate} \item Let the $m$th row be overtaken in $\{O_k\}$. From the previous cas , for each $\rowindexed{r'}{P}$ overtaking $\rowindexed{m}{P}$ in $\{O_k\}$, $\rowindexed{r'}{P}$ is in a trivial section in $P_{q_{r'}}$ with $q_{r'} < 2w - f_{r'}$ where $f_{r'}$ is the number of nonnegative rows in $\invertrows{r'}{P}$. Again, there are two scenarios to consider: \begin{enumerate}[\text{Sub-Case} 1] \item If the final swap of $\rowindexed{m}{P}$ in $\{O_k\}$ is with one of the rows overtaking it, say $\rowindexed{r'}{P}$, then $\rowindexed{m}{P}$ is in a trivial section in $P_{q_m}$, where $q_m \leq q_{r'} < 2w - f_{r'}$. \item Otherwise, $\rowindexed{m}{P}$ is positive just before it is in a trivial section, and all rows that can overtake it have overtaken it before it is swapped into a trivial section. Thus, after the last row overtakes it in $P_k$, $\rowindexed{m}{P} = \rowindexed{m_k}{P_k}$ is above its overtaking row $\rowindexed{r_k}{P_k}$, and once $\rowindexed{r_k}{P_k}$ swaps with a row below it, $\rowindexed{m}{P}$ can swap with the row it was overtaken by unless it was first overtaken by $\rowindexed{r_k}{P_k}$ and is contented with $\rowindexed{m}{P}$. Then $\rowindexed{m}{P}$ has a delay count trailing $\rowindexed{r_k}{P_k}$ of at most $f_{r_k}$ and $\rowindexed{m}{P}$ is in a trivial section in $P_{q_m}$, where $q_m \leq q_{r_k} + f_{r_k} < 2w$. \end{enumerate} \end{enumerate} Since $\numfactors{P} = q = \max_{\columnone{m}{P} \neq m} q_m$, then $q < 2w$ and the conjecture holds for lower-canonical matrices. Since an upper-canonical matrix is the inverse of a lower-canonical matrix, the same conclusion follows from Remark~\ref{existproduct}. If $P$ is not half-canonical, $P^m$ can be replaced in Case~\ref{overtakecase} by a column-settled upper-canonical matrix, where the same negative rows of $P^m$ and $P$ constitute an inverted pair, and the results will similarly follow. \end{proof} If, in the above proof, $P$ is Strang canonical, then only Sub-Case~ \ of Case~\ref{overtakecase} holds for each signed row $\rowindexed{m}{P}$. \section{Further Points for Analysis}\label{furtherpoints} Panova \cite{Panova} proved Conjecture~\ref{strangconjmat} through the use of wiring diagrams. This approach is similar to determining \emph{multi-braids . From braid theory, by the Artin relations \cite[Eq.~18, 19]{Braid}, braids that do not share a thread commute: here, any number of commuting braids can be combined, without ambiguity, into a single multi-braid. It is of interest to compare the factors derived from the approach in \cite{Panova}, as with the method of Albert, Li and Yu \cite[Sec.~4]{Strang}, which is not yet readily available, with the opportunistic-overtaking matrices approach. \begin{definition} The \emph{distance table} of an $n \times n$ permutation matrix $P$ is $\disttable{P} := (P - I)\vec{x}$ where $\vec{x}' = [1\ \dots\ n]$. \end{definition} \begin{remark}\label{totswapcount} The total number of reducing and overtaking swaps in the factors of a permutation matrix is the number of inverted pairs of that matrix. This number is also half the sum of absolute values of the entries of its distance table, plus the number of rows that can overtake each signed row and half the number of rows that can overtake each neutral row. Given $P = \prod\limits_{k = 1}^m B_k$, as in Lemma~\ref{algoterm}, if the distance tables are taken as sequences, then, for the following standard norms, \begin{align*} \norm[\ell_1]{\disttable{P_k}} & \leq \norm[\ell_1]{\disttable{P_{k - 1}}},\\ \bandwidth{P_k} = \norm[\ell_\infty]{\disttable{P_k}} & \leq \norm[\ell_\infty]{\disttable{P_{k - 1}}} = \bandwidth{P_{k - 1}},\\ \norm[\ell_2]{\disttable{P_k}} & < \norm[\ell_2]{\disttable{P_{k - 1}}}, \end{align*} indicating that the Manhattan and Chebychev distances cannot increase and that the Euclidean distance always decrease . \end{remark} The previous remark indicates that the subproducts $P_k$ of a given permutation matrix $P$ are ``diffusions'' of the initial state $\disttable{P}$, where a ``parallel bubblesort'' iteration is performed in each ``time-step''. This may be better analyzed if a relevant basis can be found. \begin{definition} A \emph{greedy bubble matrix} $G$ of a permutation matrix $P$ is a product of reducing and overtaking swaps of $P$ such that $P$ and $GP$ have no common inverted pairs. The collection of all products of $P$ with any of its greedy bubble matrices $G$ is denoted by $\greedymat{P} \ni GP$. An \emph{optimal factorization} of a permutation matrix $P$ is $\prod\limits_{k = 1}^m T_k = P$ where $\bandwidth{T_k} = 1$ and each other factorization of $P$ into bandwidth--1 matrices cannot have less factors than $\opfactors{P} := m$. \end{definition} $\oomat{P} \subseteq \greedymat{P}$, but a greedy bubble matrix of $P$ need not include reducing swaps of $P$. A breadth-first spanning-tree algorithm \cite[Sec.~22.2]{Algo} rooted in the identity matrix applied to the Cayley graph of the symmetric group of length $n$, corresponding to set of $n \times n$ permutation matrices, whose connection set is the set of all permutations represented by bandwidth--1 matrices \cite{Cayley} can be used to determine $\opfactors{P}$ for any $n \times n$ permutation matrix $P$. \begin{remark} The number of $n \times n$ permutation matrices of bandwidth $w$, $w \leq 1$, is the $n$th Fibonacci number, $F_n$, where $F_0 = F_1 = 1$. \end{remark} In testing $n \times n$, $n \leq 9$, permutation matrices, the following were observed for every permutation matrix $P$: there is an optimal factorization $P = \prod\limits_{k = 1}^m G_k$, such that $P_k \in \greedymat{P_{k - 1}}$ and $\opfactors{P} \leq n$. The former observation suggests that a greedy algorithm \cite[Ch.~16]{Algo} can determine $\opfactors{P}$; Remark~\ref{totswapcount} indicates that the use of greedy bubble matrices is advantageou . Further observation leads to the following conjecture: \begin{conjecture}\label{greedyconj} A finite permutation matrix of bandwidth $w > 0$ is the product of less than $2w$ greedy bubble matrices. \end{conjecture} Conjecture~\ref{greedyconj} asserts that, if $P_0 = P$ and $P_k \in \greedymat{P_{k - 1}}$, then, for some $m < 2w$, $P_m = I$. Of the tested greedy algorithms on $n \times n$ permutation matrices, $n \leq 9$, $\numfactors{P} \leq \opfactors{P} + \lfloor n / 3\rfloor$. The latter observation seems provable from Theorem~\ref{strangthm} where, if $P$ is a permutation matrix, $\bandwidth{P} = w$, then for every signed $\rowindexed{m}{P}$, $\invertrows{m}{P}$ has at most $2w$ rows. A.~M.~Bruckstein suggests using the sequence of adjacent transpositions to exhaustively generate all permutations of a given length, such the (Steinhaus-)Johnson-Trotter algorithm \cite{SJT}, as suggested in \cite{Even} or in \cite[Table~5]{Thesis}, and the Artin relations \cite{Braid}. D.~Pasechnik suggests that the conjecture does not hold for infinite matrices. \section*{Acknowledgements} The authors would like to thank Alfred Bruckstein and Fr\'{e}d\'{e}rique Oggier for suggesting the use of matrices instead of permutations and their supervision in the editing and reading of the drafts, Dmitrii Pasechnik and Radu Stancu for suggestions that were used in the permutation-oriented drafts, and Li-Lian Wang and Gilbert Strang for their advice and support. \section*{Appendix: Opportunistic-Overtaking Matrix Algorithm} Given: a permutation matrix $P$\\ Output: an opportunistic-overtaking matrix $O$ of $P$ \begin{itemize} \item Initialize $O = I$ and determine the inverted blocks of $P$, $B_1$, $B_2$, \dots, $B_k$ \item For each inverted block of $P$, $B_i$, $1 \leq i \leq k$ \begin{itemize} \item If $B_i$ has a reducible pair, $\rowindexed{m}{P}$ and $\rowindexed{m + 1}{P}$: \item[true:] While $\rowindexed{m - 2}{P}$ is in $B_i$, set $m$ to $m - 2$ \item[false:] Let $\rowindexed{m}{P}$ be the top row of $B_i$ \item While $\rowindexed{m + 1}{P}$ is in $B_i$, swap $\rowindexed{m}{O}$ and $\rowindexed{m + 1}{O}$, then set $m$ as $m + 2$ \end{itemize} \end{itemize}
{ "timestamp": "2010-07-21T02:02:17", "yymm": "1007", "arxiv_id": "1007.3467", "language": "en", "url": "https://arxiv.org/abs/1007.3467", "abstract": "Gilbert Strang posited that a permutation matrix of bandwidth $w$ can be written as a product of $N < 2w$ permutation matrices of bandwidth 1. A proof employing a greedy ``parallel bubblesort'' algorithm on the rows of the permutation matrix is detailed and further points of interest are elaborated.", "subjects": "Combinatorics (math.CO)", "title": "Factoring Permutation Matrices Into a Product of Tridiagonal Matrices", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.984810949854636, "lm_q2_score": 0.8128673178375734, "lm_q1q2_score": 0.800520635385411 }
https://arxiv.org/abs/1011.3486
Computing the $\sin_{p}$ function via the inverse power method
In this paper, we discuss a new iterative method for computing $\sin_{p}$. This function was introduced by Lindqvist in connection with the unidimensional nonlinear Dirichlet eigenvalue problem for the $p$-Laplacian. The iterative technique was inspired by the inverse power method in finite dimensional linear algebra and is competitive with other methods available in the literature.
\section{Introduction} In this paper we present a new method to compute the function $\sin_{p}$, inspired by recent work done by the authors in \cite{BEM}, where an iterative algorithm based on the inverse power method of linear algebra was introduced for the computation of the first eigenvalue and first eigenfunction of the Dirichlet problem for the $p$-Laplacian in arbitrary domains in \mathbb{R} ^{N}$. The functions $\sin_{p}$, $1<p<\infty$, can be thought of as generalizations of the familiar trigonometric functions. They arise in the unidimensional Dirichlet eigenvalue problem for the $p$-Laplacian and were introduced in this capacity in \cite{Lindqvist}, where a power series formula for computing them was also formally given. In \cite{BR1}\textbf{ }$\sin_{p}$\textbf{ }functions were utilized to introduce a generalization of the Pr\"{u}fer transformation and thus represent, in two phase-plane coordinates,\textbf{ }Sturm-Liouville-type problems involving the $N$-dimensional radially symmetric $p$-Laplacian $L_{p}u:=x^{1-N}\left( x^{N-1}\left\vert u^{\prime}\right\vert ^{p-2 u^{\prime}\right) ^{\prime}$, $0\leqslant a<x<b<\infty$.\textbf{ }This approach was numerically implemented in \cite{BR2} for an eigenvalue problem involving $L_{p}$ with separated homogeneous boundary conditions. In that paper an interpolation table for $\sin_{p}$ was obtained by numerically solving an ODE. Also in that paper the authors raised the question of finding a fast and accurate algorithm for computing $\sin_{p}$. Our method depends on the convergence of a sequence of functions whose definition, as in \cite{BEM}, is motivated by an extension of the inverse power method of linear algebra for obtaining the first eigenvalue and first eigenfunction of finite dimensional linear operators. These functions are recursively defined and can be given in integral form, so that they can be obtained by numerical integration. More specifically, recall that it suffices to obtain $\sin_{p}$ in the interval $I_{p}=\left[ 0,\pi_{p}/2\right] $, since it is extended to the interval $\left[ \pi_{p}/2,\pi_{p}\right] $ symmetrically with respect to $\pi_{p}/2$ and afterward to the whole real line \mathbb{R} $ as an odd, $2\pi_{p}$-periodic function (the definition of $\sin_{p}$ as well as the precise value of $\pi_{p}$ are recalled in Section 2). We define the following sequence of (positive) functions $\left\{ \phi_{n}\right\} \subset C^{1}\left( I_{p}\right) $. Set $\phi_{0}\equiv1$ and \[ \left\{ \begin{array} [c]{ll \left( \phi_{n+1}^{\prime}\left\vert \phi_{n+1}^{\prime}\right\vert ^{p-2}\right) ^{\prime}=-\phi_{n}\left\vert \phi_{n}\right\vert ^{p-2} & \text{ \ \ if }x\in I_{p},\\ \phi_{n+1}\left( 0\right) =\phi_{n+1}^{\prime}\left( \pi_{p}/2\right) =0. & \end{array} \right. \] We prove that the scaled sequence $\left\{ \sqrt[p]{p-1}\phi_{n}/\left\Vert \phi_{n}\right\Vert _{\infty}\right\} $ converges uniformly to $\sin_{p}$ in $I_{p}$. The functions $\phi_{n}$ can be written in integral form a \[ \phi_{n+1}\left( x\right) =\int_{0}^{x}\left( \int_{\theta}^{\pi_{p}/2 \phi_{n}\left( s\right) ^{p-1}ds\right) ^{\frac{1}{p-1}}d\theta\text{, \ \ }x\in I_{p}, \] and, therefore, are readily computed using standard efficient numerical methods for definite integrals. This paper is organized as follows. In Section 2, we recall the definition and some basic properties of $\sin_{p}$ which will be used in the sequel. In Section 3, we show how to recursively construct a sequence of functions which converge uniformly to $\sin_{p}$. Finally, in Section 4 we compare the performance of our method with those of \cite{Lindqvist} and \cite{BR2}. \section{The function $\sin_{p}$} For the sake of completeness we recall in this section the definition and some properties of the function $\sin_{p}$. The unidimensional Dirichlet eigenvalue problem for the $p$-Laplacian, $p>1$, i \begin{equation} \left\{ \begin{array} [c]{ll \psi_{p}\left( u^{\prime}\right) ^{\prime}=-\lambda\psi_{p}\left( u\right) & \quad\text{if }a<x<b,\\ u\left( a\right) =u\left( b\right) =0, & \end{array} \right. \label{01 \end{equation} where $\psi_{p}\left( t\right) =t\left\vert t\right\vert ^{p-2}$. It is easy to verify that if $\lambda_{1}$ is the first eigenvalue of \begin{equation} \left\{ \begin{array} [c]{ll \psi_{p}\left( v^{\prime}\right) ^{\prime}=-\lambda\psi_{p}\left( v\right) & \quad\text{if }a<x<m:=\dfrac{a+b}{2},\\ v\left( a\right) =v^{\prime}\left( m\right) =0, & \end{array} \right. \label{02 \end{equation} and $v_{1}$ is the corresponding positive eigenfunction, then $\lambda_{1}$ is also the first eigenvalue for (\ref{01}) wit \[ u_{1}\left( x\right) =\left\{ \begin{array} [c]{ll v_{1}\left( x\right) & \quad\text{if }a\leqslant x\leqslant m,\\ v_{1}\left( a+b-x\right) & \quad\text{if }m\leqslant x\leqslant b, \end{array} \right. \] being the corresponding positive eigenfunction. Moreover, this function is stricly increasing on $[a,m)$, strictly decreasing on $(m,b]$ and has only one maximum point which is reached at $x=m$. Thus, $\left\Vert u_{1}\right\Vert _{\infty}=u_{1}\left( m\right) $. An expression for $\lambda_{1}$ is well known and can be obtained by integration (see \cite{Otani}) as follows. First multiply (\ref{01}) by $u_{1}^{\prime}$ and integrate the resulting equation by parts on $\left[ a,x\right] $ to obtai \begin{equation} \left. \psi_{p}\left( u_{1}^{\prime}\right) u_{1}^{\prime}\right\vert _{a}^{x}-\int_{a}^{x}\psi_{p}\left( u_{1}^{\prime}\right) u_{1 ^{\prime\prime}dx=-\lambda_{1}\int_{a}^{x}\psi_{p}\left( u_{1}\right) u_{1}^{\prime}dx. \label{03 \end{equation} We hav \begin{equation} \left. \psi_{p}\left( u_{1}^{\prime}\right) u_{1}^{\prime}\right\vert _{a}^{x}=\left\vert u_{1}^{\prime}\left( x\right) \right\vert ^{p -\left\vert u_{1}^{\prime}\left( a\right) \right\vert ^{p} \label{04 \end{equation \begin{equation} \int_{a}^{x}\psi_{p}\left( u_{1}\right) u_{1}^{\prime}dx=\int_{u_{1 (a)}^{u_{1}(x)}\psi_{p}\left( s\right) \,ds=\frac{\left\vert u\left( x\right) \right\vert ^{p}}{p}-\frac{\left\vert u\left( a\right) \right\vert ^{p}}{p}, \label{05 \end{equation \begin{equation} \int_{a}^{x}\psi_{p}\left( u_{1}^{\prime}\right) u_{1}^{\prime\prime dx=\int_{u_{1}^{\prime}\left( a\right) }^{u_{1}^{\prime}\left( x\right) }\psi_{p}\left( s\right) \,ds=\frac{\left\vert u^{\prime}\left( x\right) \right\vert ^{p}}{p}-\frac{\left\vert u^{\prime}\left( a\right) \right\vert ^{p}}{p}. \label{06 \end{equation} Substituting (\ref{04}), (\ref{05}) and (\ref{06}) in (\ref{03}) we obtain \[ \left( 1-\frac{1}{p}\right) \left[ \left\vert u_{1}^{\prime}\left( x\right) \right\vert ^{p}-\left\vert u_{1}^{\prime}\left( a\right) \right\vert ^{p}\right] =-\lambda_{1}\left[ \frac{\left\vert u_{1}\left( x\right) \right\vert ^{p}}{p}-\frac{\left\vert u_{1}\left( a\right) \right\vert ^{p}}{p}\right] , \] whence \[ \left[ \left( 1-\frac{1}{p}\right) \left\vert u_{1}^{\prime}\right\vert ^{p}+\lambda_{1}\frac{\left\vert u_{1}\right\vert ^{p}}{p}\right] _{a ^{x}=0. \] This means that \[ \frac{p-1}{p}\left\vert u_{1}^{\prime}\right\vert ^{p}+\frac{\lambda_{1} {p}\left\vert u_{1}\right\vert ^{p}\equiv C, \] where $C$ is a constant and $p^{\prime}=p/\left( p-1\right) $ is the conjugate of $p$. The value of $C$ can be found computing the value of this expression at the maximum point $m$; choosing $u_{1}$ such that $u_{1}\left( m\right) =1$ we fin \[ C=\frac{p-1}{p}\left\vert u_{1}^{\prime}\left( m\right) \right\vert ^{p}+\frac{\lambda_{1}}{p}\left\vert u_{1}\left( m\right) \right\vert ^{p}=\frac{\lambda_{1}}{p}. \] Therefore \begin{equation} \left( p-1\right) \left\vert u_{1}^{\prime}\left( x\right) \right\vert ^{p}+\lambda_{1}\left\vert u_{1}\left( x\right) \right\vert ^{p}=\lambda_{1} \label{07 \end{equation} for all $x\in\left[ a,b\right] $. On the interval $\left[ a,m\right] $ we have $u^{\prime}\geqslant0$, hence we can writ \begin{equation} \frac{u_{1}^{\prime}\left( x\right) }{\sqrt[p]{\left( 1-\left\vert u_{1}\left( x\right) \right\vert ^{p}\right) }}=\sqrt[p]{\frac{\lambda_{1 }{p-1}} \label{08 \end{equation} for all $x\in\left[ a,m\right] $. Integrating this equation on $\left( a,m\right) $ leads t \[ \frac{b-a}{2}\sqrt[p]{\frac{\lambda_{1}}{p-1}}=\int_{u_{1}(a)}^{u_{1}(m) \frac{ds}{\sqrt[p]{1-s^{p}}}=\int_{0}^{1}\frac{ds}{\sqrt[p]{1-s^{p}}}, \] which gives the expressio \begin{equation} \lambda_{1}=(p-1)\left( \frac{2}{b-a}\int_{0}^{1}\frac{ds}{\sqrt[p]{1-s^{p} }\right) ^{p}=\left( \frac{\pi_{p}}{b-a}\right) ^{p}, \label{09 \end{equation} where we se \begin{equation} \pi_{p}:=2\sqrt[p]{p-1}\int_{0}^{1}\frac{ds}{\sqrt[p]{1-s^{p}}}. \label{ppip \end{equation} Making the change of variable $s=\sqrt[p]{t}$ in the last integral and using the classical Beta function $B$ we obtain \[ \int_{0}^{1}\frac{ds}{\sqrt[p]{1-s^{p}}}=\frac{1}{p}\int_{0}^{1}t^{\frac{1 {p}-1}(1-t)^{-\frac{1}{p}}dt=\frac{1}{p}B\left( 1-\frac{1}{p},\frac{1 {p}\right) =\frac{\pi/p}{\sin(\pi/p) \] (Here one use the properties $B(x,y)B(x+y,1-y)=x/x\sin\left( \pi y\right) $ and $B(1,z)=1/z$ with $x=1-1/p$ and $y=z=1/p$). Therefore \begin{equation} \pi_{p}=\frac{2\sqrt[p]{p-1}\left( \pi/p\right) }{\sin(\pi/p)} \label{10 \end{equation} and \[ \lambda_{1}=\left( \frac{2\sqrt[p]{p-1}\left( \pi/p\right) }{(b-a)\sin \left( \pi/p\right) }\right) ^{p}. \] When $a=0$ and $b=\pi_{p}$ we denote the function $\sqrt[p]{p-1}u_{1}$ by $\sin_{p}.$ Thus, $\sin_{p}\left( 0\right) =0=\sin_{p}^{\prime}\left( \pi_{p}/2\right) $, $\lambda_{1}=1$ and from (\ref{07}) \[ \left\vert \sin_{p}^{\prime}\right\vert ^{p}+\frac{\left\vert \sin _{p}\right\vert ^{p}}{p-1}=1. \] It is clear from this equation that $\sin_{p}^{\prime}\left( 0\right) =1.$ We remark that $u=\sin_{p}$ is also the unique solution of the initial value problem \[ \left\vert u^{\prime}\right\vert ^{p}+\frac{\left\vert u\right\vert ^{p} {p-1}=1,\quad u\left( 0\right) =0, \] which can be used to define this function. Alternatively, we can define $\sin_{p}$ on the interval $\left[ 0,\pi _{p}/2\right] $ as an inverse function. In fact, multiplying (\ref{08}) by $\sqrt[p]{p-1}$ and using (\ref{09}) with $a=0$ and $b=\pi_{p}$ we obtain \[ \int_{0}^{\sin_{p}\left( x\right) }\frac{ds}{\sqrt[p]{\left( 1-\frac{s^{p }{p-1}\right) }}=x,\quad\text{for }x\in\left[ 0,\pi_{p}/2\right] , \] that is, $\sin_{p}=\zeta^{-1}$ wher \[ \zeta\left( z\right) : {\displaystyle\int_{0}^{z}} \frac{ds}{\sqrt[p]{\left( 1-\frac{s^{p}}{p-1}\right) }},\quad\text{for z\in\left[ 0,\sqrt[p]{p-1}\right] . \] With this definition, we extend $\sin_{p}$ to the interval $\left[ \pi _{p}/2,\pi_{p}\right] $ symmetrically with respect to $\pi_{p}/2$ and afterward to the whole real line \mathbb{R} $ as an odd, $2\pi_{p}$-periodic function. We list the basic properties of $\sin_{p}$: \begin{enumerate} \item $\sin_{p}\left( 0\right) =0=\sin_{p}\left( \pi_{p}\right) $, $\sin_{p}\left( \pi_{p}/2\right) =\left\Vert \sin_{p}\right\Vert _{\infty }=\sqrt[p]{p-1}$. \item $\sin_{p}\left( x\right) $ is strictly increasing in $\left[ 0,\pi_{p}/2\right] $ and strictly decreasing in $\left[ \pi_{p}/2,\pi _{p}\right] .$ \item $\left\vert \sin_{p}^{\prime}\left( x\right) \right\vert =\sqrt[p]{1-\dfrac{\left\vert \sin_{p}\right\vert ^{p}}{p-1}}.$ \end{enumerate} \section{A sequence uniformly convergent to $\sin_{p}$} Let $I_{p}=\left[ 0,\pi_{p}/2\right] $ and define the following sequence of functions $\left\{ \phi_{n}\right\} \subset C^{1}\left( I_{p}\right) $. Set $\phi_{0}\equiv1$ and \[ \left\{ \begin{array} [c]{ll \left( \psi_{p}\left( \phi_{n+1}^{\prime}\right) \right) ^{\prime =-\psi_{p}\left( \phi_{n}\right) & \text{ \ \ if }x\in I_{p},\\ \phi_{n+1}\left( 0\right) =\phi_{n+1}^{\prime}\left( \pi_{p}/2\right) =0. & \end{array} \right. \] In this section, we prove that the scaled sequence $\left\{ \sqrt[p]{p-1 \phi_{n}/\left\Vert \phi_{n}\right\Vert _{\infty}\right\} $ converges uniformly to $\sin_{p}$ in $I_{p}$. Before proceeding, we recall some basic properties of the $\psi_{p}$ functions: \begin{description} \item[Proposition 3.1.] (Basic properties of $\psi_{p}$) \textit{The following holds:} \end{description} \begin{enumerate} \item $\psi_{p}$\textit{ is continuous, strictly increasing and odd, for each }$p>1.$ \item $\psi_{p}\left( ab\right) =\psi_{p}\left( a\right) \psi_{p}\left( b\right) .$ \item $\psi_{p}\left( \dfrac{a}{b}\right) =\dfrac{\psi_{p}\left( a\right) }{\psi_{p}\left( b\right) }$ \item $\left( \psi_{p}\right) ^{-1}=\psi_{p^{\prime}}.$ \item $\int_{0}^{t}\psi_{p}\left( s\right) ds=\dfrac{\left\vert t\right\vert ^{p}}{p}.$ \end{enumerate} By a straightforward calculation we can find the following recursive integral expression for the $\phi_{n}$-functions: \begin{equation} \phi_{n+1}\left( x\right) =\int_{0}^{x}\psi_{p^{\prime}}\left( \int _{\theta}^{\pi_{p}/2}\psi_{p}\left( \phi_{n}\left( s\right) \right) ds\right) d\theta. \label{11 \end{equation} It is clear from (\ref{11}) that each $\phi_{n}\ $is positive, increasing on $I_{p}$ and reaches its maximum value at $x=\pi_{p}/2$. One can obtain an explicit expression for $\phi_{1}$, the second function in the sequence: \begin{align*} \phi_{1}\left( x\right) & =\int_{0}^{x}\psi_{p^{\prime}}\left( \int_{\theta}^{\pi_{p}/2}\psi_{p}\left( 1\right) ds\right) d\theta\\ & =\int_{0}^{x}\psi_{p^{\prime}}\left( \frac{\pi_{p}}{2}-\theta\right) d\theta\\ & =\int_{\pi_{p}/2-x}^{\pi_{p}/2}\psi_{p^{\prime}}\left( y\right) dy\\ & =\frac{1}{p}\left[ \left( \frac{\pi_{p}}{2}\right) ^{p}-\left( \frac{\pi_{p}}{2}-x\right) ^{p}\right] . \end{align*} Note tha \[ \left\Vert \phi_{1}\right\Vert _{\infty}=\phi_{1}\left( \frac{\pi_{p} {2}\right) =\frac{1}{p}\left( \frac{\pi_{p}}{2}\right) ^{p}=\frac{p-1 {p}\left( \frac{\pi/p}{\sin\left( \pi/p\right) }\right) ^{p}. \] The next $\phi_{n}$-functions however, are very difficult to obtain explicitly by solving the integrals analytically. On the other hand, the integrals can easily be solved numerically. \begin{description} \item[Proposition 3.2.] $\phi_{n+1}\leqslant\left\Vert \phi_{1}\right\Vert _{\infty}\phi_{n}$\textit{ on }$I_{p}.$ \end{description} \noindent\textbf{Proof.} For $n=1$ the result is trivially true since $\phi_{0}\equiv1.$Assuming by induction that $\phi_{n}\leqslant\left\Vert \phi_{1}\right\Vert _{\infty}\phi_{n-1}$, we have \begin{align*} \phi_{n+1}\left( x\right) & =\int_{0}^{x}\psi_{p^{\prime}}\left( \int_{\theta}^{\pi_{p}/2}\psi_{p}\left( \phi_{n}\left( s\right) \right) \,ds\right) \,d\theta\\ & \leqslant\int_{0}^{x}\psi_{p^{\prime}}\left( \int_{\theta}^{\pi_{p}/2 \psi_{p}\left( \left\Vert \phi_{1}\right\Vert _{\infty}\phi_{n-1}\left( s\right) \right) \,ds\right) \,d\theta\\ & =\int_{0}^{x}\psi_{p^{\prime}}\left( \psi_{p}\left( \left\Vert \phi _{1}\right\Vert _{\infty}\right) \int_{\theta}^{\pi_{p}/2}\psi_{p}\left( \phi_{n-1}\left( s\right) \right) \,ds\right) \,d\theta\\ & =\left\Vert \phi_{1}\right\Vert _{\infty}\int_{0}^{x}\psi_{p^{\prime }\left( \int_{\theta}^{\pi_{p}/2}\psi_{p}\left( \phi_{n-1}\left( s\right) \right) \,ds\right) \,d\theta\\ & =\left\Vert \phi_{1}\right\Vert _{\infty}\phi_{n}\left( x\right) . \end{align*} \noindent$\blacksquare$ The following technical lemma, which will be used in the sequel, can be proved via the Cauchy mean value theorem (see \cite{AVV}) and works as a L'H\^{o}pital's rule in order to get monotonicity for a certain quotient function. \begin{description} \item[Lemma 3.3.] \textit{Let }$f,g:\left[ a,b\right] \longrightarrow R$\textit{ be continuous on }$\left[ a,b\right] $\textit{ and differentiable in }$\left( a,b\right) $. \textit{Suppose }$g^{\prime}(x)\neq0$\textit{ for all }$x\in\left( a,b\right) $. \textit{If }$\dfrac{f^{\prime}}{g^{\prime} $\textit{ is (strictly) increasing }[\textit{decreasing}],\textit{ then both }$\dfrac{f(x)-f(a)}{g(x)-g(a)}$\textit{ and }$\dfrac{f(x)-f(b)}{g(x)-g(b) $\textit{are (strictly) increasing }[\textit{decreasing}]\textit{.} \item[Theorem 3.4.] \textit{For each }$n\geqslant1$\textit{ the function }$\dfrac{\phi_{n}}{\phi_{n+1}}$\textit{ is strictly decreasing on }$I_{p $\textit{ and} \end{description} \begin{enumerate} \item[(i)] $\dfrac{1}{\left\Vert \phi_{1}\right\Vert _{\infty}}\leqslant \inf\limits_{I_{p}}\dfrac{\phi_{n}}{\phi_{n+1}}=\dfrac{\phi_{n}\left( \pi _{p}/2\right) }{\phi_{n+1}\left( \pi_{p}/2\right) }=\dfrac{\left\Vert \phi_{n}\right\Vert _{\infty}}{\left\Vert \phi_{n+1}\right\Vert _{\infty}}.$ \item[(ii)] $\left\Vert \dfrac{\phi_{n}}{\phi_{n+1}}\right\Vert _{\infty =\psi_{p^{\prime}}\left( \dfrac{\int_{0}^{\pi_{p}/2}\psi_{p}\left( \phi_{n-1}\left( s\right) \right) ds}{\int_{0}^{\pi_{p}/2}\psi_{p}\left( \phi_{n}\left( s\right) \right) ds}\right) $ \ \ for $n\geqslant1.$ \item[(iii)] $\left\Vert \dfrac{\phi_{n}}{\phi_{n+1}}\right\Vert _{\infty }\leqslant\left\Vert \dfrac{\phi_{n-1}}{\phi_{n}}\right\Vert _{\infty }\leqslant\cdots\leqslant\left\Vert \dfrac{\phi_{1}}{\phi_{2}}\right\Vert _{\infty}<\infty.$ \end{enumerate} \noindent\textbf{Proof.} Since $\phi_{1}$ is strictly increasing, it follows that $1/\phi_{1}$ is strictly decreasing. Assume by induction that $\phi _{n-1}/\phi_{n}$ is strictly decreasing. Since \[ \frac{\phi_{n}\left( x\right) -\phi_{n}\left( 0\right) }{\phi_{n+1 -\phi_{n+1}\left( 0\right) }=\frac{\phi_{n}\left( x\right) }{\phi _{n+1}\left( x\right) }, \] in order to show that $\phi_{n}/\phi_{n+1}$ is strictly decreasing, it suffices in light of the lemma to verify that $\phi_{n}^{\prime}/\phi _{n+1}^{\prime}$ is strictly decreasing on $I_{p}$. But, \[ \frac{\phi_{n}^{\prime}\left( x\right) }{\phi_{n+1}^{\prime}\left( x\right) }=\dfrac{\psi_{p^{\prime}}\left( {\displaystyle\int_{x}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n-1}\left( s\right) \right) ds\right) {\psi_{p^{\prime}}\left( {\displaystyle\int_{x}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n}\left( s\right) \right) ds\right) }=\psi _{p^{\prime}}\left( \frac {\displaystyle\int_{x}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n-1}\left( s\right) \right) ds} {\displaystyle\int_{x}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n}\left( s\right) \right) ds}\right) . \] Since $\psi_{p^{\prime}}$ is strictly increasing and the functions $\int _{x}^{\pi_{p}/2}\psi_{p}\left( \phi_{n-1}\left( s\right) \right) ds$ and $\int_{x}^{\pi_{p}/2}\psi_{p}\left( \phi_{n}\left( s\right) \right) ds$ are null at $x=\pi_{p}/2$, we can apply the lemma again to verify that the quotient of these integral functions is a strictly decreasing function. We hav \[ \frac{\left( {\displaystyle\int_{x}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n-1}\left( s\right) \right) ds\right) ^{\prime }{\left( {\displaystyle\int_{x}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n}\left( s\right) \right) ds\right) ^{\prime} =\frac{\psi_{p}\left( \phi_{n-1}\left( s\right) \right) }{\psi_{p}\left( \phi_{n}\left( s\right) \right) }=\psi_{p}\left( \frac{\phi_{n-1} {\phi_{n}}\right) , \] which is strictly decreasing by the induction hypothesis. The inequality in (i) follows from Proposition 3.2. Before verifying (ii) we remark that $\left\Vert 1/\phi_{1}\right\Vert _{\infty}=\infty$ since $\phi_{1}\left( 0\right) =0.$ In order to prove (ii) we first observe that the monotonicity of $\phi_{n}/\phi_{n+1}$ implies that \[ \left\Vert \dfrac{\phi_{n}}{\phi_{n+1}}\right\Vert _{\infty}=\lim _{x\rightarrow0^{+}}\frac{\phi_{n}\left( x\right) }{\phi_{n+1}\left( x\right) }. \] L'H\^{o}pital's rule then yield \[ \lim_{x\rightarrow0^{+}}\frac{\phi_{n}\left( x\right) }{\phi_{n+1}\left( x\right) }=\lim_{x\rightarrow0^{+}}\frac{\phi_{n}^{\prime}\left( x\right) }{\phi_{n+1}^{\prime}\left( x\right) }=\psi_{p^{\prime}}\left( \frac {\int_{0}^{\pi_{p}/2}\psi_{p}\left( \phi_{n-1}\left( s\right) \right) ds}{\int_{0}^{\pi_{p}/2}\psi_{p}\left( \phi_{n}\left( s\right) \right) ds}\right) <\infty. \] The proof of (iii) is a consequence of the following estimates, valid for $n\geqslant2$ \begin{align*} \left\Vert \frac{\phi_{n}}{\phi_{n+1}}\right\Vert _{\infty} & =\psi _{p^{\prime}}\left( \frac {\displaystyle\int_{0}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n-1}\left( s\right) \right) ds} {\displaystyle\int_{0}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n}\left( s\right) \right) ds}\right) \\ & \leqslant\psi_{p^{\prime}}\left( \frac {\displaystyle\int_{0}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n}\left( s\right) \right) \psi_{p}\left( \dfrac {\phi_{n-1}}{\phi_{n}}\left( s\right) \right) ds} {\displaystyle\int_{0}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n}\left( s\right) \right) ds}\right) \\ & \leqslant\psi_{p^{\prime}}\left( \frac {\displaystyle\int_{0}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n}\left( s\right) \right) \psi_{p}\left( \left\Vert \dfrac{\phi_{n-1}}{\phi_{n}}\right\Vert _{\infty}\right) ds} {\displaystyle\int_{0}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n}\left( s\right) \right) ds}\right) \\ & =\left\Vert \frac{\phi_{n-1}}{\phi_{n}}\right\Vert _{\infty}\psi _{p^{\prime}}\left( \frac {\displaystyle\int_{0}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n}\left( s\right) \right) ds} {\displaystyle\int_{0}^{\pi_{p}/2}} \psi_{p}\left( \phi_{n}\left( s\right) \right) ds}\right) \\ & =\left\Vert \frac{\phi_{n-1}}{\phi_{n}}\right\Vert _{\infty}. \end{align*} \noindent$\blacksquare$ \begin{description} \item[Theorem 2.4.] \textit{Let }$u_{n}:=\dfrac{\phi_{n}}{\left\Vert \phi _{n}\right\Vert _{\infty}}\in C^{1}\left( I_{p}\right) ,$\textit{ for }$n\geqslant1.$\textit{ Then the sequence }$\left\{ u_{n}\left( x\right) \right\} _{n\geqslant1}$\textit{ is decreasing for each }$x\in I_{p}$\textit{ and \[ \sqrt[p]{p-1}u_{n}\rightarrow\sin_{p}\text{\textit{ \ \ uniformly in} }I_{p}. \] \end{description} \noindent\textbf{Proof.} In $I_{p}$ we hav \begin{align*} \frac{u_{n}}{u_{n+1}} & =\dfrac{\phi_{n}}{\phi_{n+1}}\left( \dfrac {\left\Vert \phi_{n}\right\Vert _{\infty}}{\left\Vert \phi_{n+1}\right\Vert _{\infty}}\right) ^{-1}\\ & \geqslant\left( \inf\limits_{I_{p}}\dfrac{\phi_{n}}{\phi_{n+1}}\right) \left( \dfrac{\left\Vert \phi_{n}\right\Vert _{\infty}}{\left\Vert \phi _{n+1}\right\Vert _{\infty}}\right) ^{-1}\\ & =\left( \dfrac{\left\Vert \phi_{n}\right\Vert _{\infty}}{\left\Vert \phi_{n+1}\right\Vert _{\infty}}\right) \left( \dfrac{\left\Vert \phi _{n}\right\Vert _{\infty}}{\left\Vert \phi_{n+1}\right\Vert _{\infty}}\right) ^{-1}\\ & =1, \end{align*} that is, $\left\{ u_{n}\left( x\right) \right\} _{n\geqslant1}$ is decreasing for each $x\in I_{p},$ and the whole sequence is bounded below by $u_{1}$. Thus, there exist \[ u:=\lim u_{n}. \] We have $\left\Vert u_{n}\right\Vert _{\infty}=1$ for each $n$. Moreover, sinc \[ \frac{\left\Vert \phi_{n}\right\Vert _{\infty}}{\left\Vert \phi_{n+1 \right\Vert _{\infty}}=\inf\limits_{I_{p}}\dfrac{\phi_{n}}{\phi_{n+1 }\leqslant\left\Vert \frac{\phi_{n}}{\phi_{n+1}}\right\Vert _{\infty \leqslant\left\Vert \frac{\phi_{1}}{\phi_{2}}\right\Vert _{\infty}=:C, \] we also have, for every $x\in I_{p}, \begin{align*} \left\vert u_{n}^{\prime}\left( x\right) \right\vert & =\frac {1}{\left\Vert \phi_{n}\right\Vert _{\infty}}\psi_{p^{\prime}}\left( \int _{x}^{\pi_{p}/2}\psi_{p}\left( \phi_{n-1}\left( s\right) \right) ds\right) \\ & =\frac{\left\Vert \phi_{n-1}\right\Vert _{\infty}}{\left\Vert \phi _{n}\right\Vert _{\infty}}\psi_{p^{\prime}}\left( \int_{x}^{\pi_{p}/2 \psi_{p}\left( \frac{\phi_{n-1}\left( s\right) }{\left\Vert \phi _{n-1}\right\Vert _{\infty}}\right) ds\right) \\ & \leqslant C\psi_{p^{\prime}}\left( \int_{0}^{\pi_{p}/2}\psi_{p}\left( u_{n-1}\right) ds\right) \\ & \leqslant C\psi_{p^{\prime}}\left( \int_{0}^{\pi_{p}/2}\psi_{p}\left( 1\right) ds\right) \\ & =\frac{C\pi_{p}}{2}. \end{align*} It follows from Arzela-Ascoli's theorem that $u_{n}\rightarrow u\in C\left( I_{p}\right) $, uniformly. In order to conclude the proof, we need just to show that \begin{equation} u=\dfrac{\sin_{p}}{\sqrt[p]{p-1}}. \label{12 \end{equation} From (\ref{11}) we can write the following expression: \[ u_{n+1}\left( x\right) =\gamma_{n}\int_{0}^{x}\psi_{p^{\prime}}\left( \int_{\theta}^{\pi_{p}/2}\psi_{p}\left( u_{n}\left( s\right) \right) ds\right) d\theta, \] where \[ \gamma_{n}:=\frac{\left\Vert \phi_{n}\right\Vert _{\infty}}{\left\Vert \phi_{n+1}\right\Vert _{\infty}}. \] In view of the boundedness of $\left\{ \gamma_{n}\right\} $, there exists $\gamma:=\lim\gamma_{n_{k}}$ for some subsequence $\left\{ \gamma_{n_{k }\right\} .$ Thus, letting $k\rightarrow\infty$ in \[ u_{n_{k}+1}\left( x\right) =\gamma_{n_{k}}\int_{0}^{x}\psi_{p^{\prime }\left( \int_{\theta}^{\pi_{p}/2}\psi_{p}\left( u_{n_{k}}\left( s\right) \right) ds\right) d\theta, \] we ge \[ u\left( x\right) =\gamma\int_{0}^{x}\psi_{p^{\prime}}\left( \int_{\theta }^{\pi_{p}/2}\psi_{p}\left( u\left( s\right) \right) ds\right) d\theta\in C^{1}\left( I_{p}\right) , \] which means that $u$ is a positive solution to the following proble \[ \left\{ \begin{array} [c]{ll \psi_{p}\left( u^{\prime}\right) ^{\prime}=-\gamma\psi_{p}\left( u\right) & \text{ \ \ if }x\in I_{p},\\ u\left( 0\right) =u^{\prime}\left( \pi_{p}/2\right) =0. & \end{array} \right. \] In view of the positivity of $u$, we can integrate the equation above multiplied by $u^{\prime}$ and proceed as in the derivation of (\ref{09}) to find $\gamma=1$. From this we conclude that in fact $\lim\gamma_{n}=1$ (the whole sequence converges to the eingenvalue $1$) and that $u=\lim u_{n}$ satisfies the same boundary value problem that $\sin_{p}/\left\Vert \sin _{p}\right\Vert $ does. Since both $u$ and $\sin_{p}$ are positive and $\left\Vert u\right\Vert _{\infty}=\left\Vert \sin_{p}/\left\Vert \sin _{p}\right\Vert \right\Vert _{\infty}=1$, we must have \[ u=\dfrac{\sin_{p}}{\left\Vert \sin_{p}\right\Vert \] whence (\ref{12}) follows. $\blacksquare$ \section{Numerical Results} Next we examine the computational time of each method. Computations were performed on a WindowsXP/Pentium 4-2.8GHz platform, using the GCC compiler. Although the method of computing $\sin_{p}$ by solving an ODE suggested in \cite{BR2} (which we implemented by means of a standard Runge-Kutta fourth power method) is by far the fastest, the computational times of the other two methods are competitive, the inverse power method being on average more than twice as fast as the power series method of \cite{Lindqvist} for values of $p$ greater than 2. Also, the average number of $8$ iterations that the inverse power method uses to obtain the same (and sometimes better; see Table 2) accuracy of the differential equation method of \cite{BR2} is quite remarkable, specially taking into account that the functions $\phi_{n}$ converge to $0$ rather rapidly. We emphasize that the computational time of the inverse power method is not the main subject of this presentation. The method demands the computation of double integrals at each iteration for each grid point. We opted for a classical, computationally easy to implement and reasonably fast method to compute these integrals, namely, the Simpson composite method. However, a greater effort spent in lessening the computational time of the numerical integrations certainly would be reflected in a substantial decrease in the time spent computing $\sin_{p}$ overall. Nevertheless, by considering the accuracy and the comparison scale among the three methods (on the range of miliseconds) we may say that the results presented in this paper validate the inverse power method as an effective and reasonably fast method for numerically obtaining $\sin_{p}$. Below we present the average time spent in computing $\sin_{p}$ on the whole interval $I_{p}$ divided in $101$ grid points by each method for six values of $p$ (the average was taken out of five computer runs); the stop criterion in each method was an error tolerance of $10^{-8}$ between successive iterations and less than 500 terms in the power series \ \begin{tabular} [c]{|l|c|c|c|c|c|c|}\hline $p$ & $1.1$ & $1.5$ & $2.0$ & $2.5$ & $3.0$ & $3.5$\\\hline Inverse power method & $21.5$ & $32.1$ & $1.1$ & $37.7$ & $37.8$ & $31.7$\\\hline Differential equation method & $1.9$ & $1.8$ & $1.1$ & $1.5$ & $1.5$ & $1.5$\\\hline Power series & $92.9$ & $2.2$ & $2.0$ & $79.6$ & $79.3$ & $73.3$\\\hline \end{tabular} \ \ \] \begin{center} {\small Table 1: Average time (in miliseconds) for the computation of $\sin_{p}$ on $I_{p}$ for each method.}\vspace{0.2cm} \end{center} Besides the trivial point $0$, the only point where the value of $\sin_{p}$ is exactly known is $\pi_{p}/2$, with $\sin_{p}\left( \pi_{p}/2\right) =\sqrt[p]{p-1}$. In the next table we present the computed value for $\sin _{p}\left( \pi_{p}/2\right) $ obtained using each method \ \begin{tabular} [c]{|l|c|c|c|c|c|c|}\hline $p$ & $1.1$ & $1.5$ & $2.0$ & $2.5$ & $3.0$ & $3.5$\\\hline $\sqrt[p]{p-1}$ & $0.123285$ & $0.629961$ & $1$ & $1.17608$ & $1.25992$ & $1.29926$\\\hline Inverse power method & $0.123285$ & $0.629961$ & $1$ & $1.17608$ & $1.25992$ & $1.29926$\\\hline Differential equation method & $0.123285$ & $0.629966$ & $1.00017$ & $1.17647$ & $1.26044$ & $1.29983$\\\hline Power series & $5.3\times10^{128}$ & $0.629961$ & $1$ & $1.17608$ & $1.25993$ & $1.29928$\\\hline \end{tabular} \ \ \ \] \begin{center} {\small Table 2: Value of $\sin_{p}\left( \pi_{p}/2\right) =\sqrt[p]{p-1}$ obtained independently using each method.}\vspace{0.2cm} \end{center} \noindent Notice that the inverse power method appears to be more accurate when computing $\sin_{p}$ at values close to $\pi_{p}/2$. Indeed, in order to obtain a good approximation close to this point, it was necessary to allow for a greater number of terms in the power series than would be necessary for points far from $\pi_{p}/2$ \ \begin{tabular} [c]{|c|c|c|c|c|c|c|}\hline $p$ & $1.1$ & $1.5$ & $2.0$ & $2.5$ & $3.0$ & $3.5$\\\hline Inverse power method & $5$ & $8$ & $9$ & $8$ & $8$ & $8$\\\hline Power Series & $501$ & $13$ & $8$ & $470$ & $501$ & $501$\\\hline \end{tabular} \ \] \begin{center} {\small Table 3: Number of iterations.}\vspace{0.2cm}\vspace{0.2cm} \end{center} \noindent We see that the number of iterations used by the inverse power method is remarkably low. Below, we present the graphics of $\sin_{p}$ for the same values of $p$ computed using the three methods (except for $p=1.1$, since the power series appears to diverge in this case). Notice that all three methods agree very well with each other, being virtually indistinguishable. \[ \includegraphics[scale=0.4] {p_1.1.eps}\ \ \ \ \includegraphics[scale=0.4] {p_1.5.eps} \ \[ \includegraphics[scale=0.4] {p_2.0.eps}\ \ \ \ \includegraphics[scale=0.4] {p_2.5.eps} \ \[ \includegraphics[scale=0.4] {p_3.0.eps}\ \ \ \ \includegraphics[scale=0.4] {p_3.5.eps} \] \section*{Acknowledgments} The second author would like to thank the support of FAPEMIG and CNPq.
{ "timestamp": "2010-11-16T02:04:25", "yymm": "1011", "arxiv_id": "1011.3486", "language": "en", "url": "https://arxiv.org/abs/1011.3486", "abstract": "In this paper, we discuss a new iterative method for computing $\\sin_{p}$. This function was introduced by Lindqvist in connection with the unidimensional nonlinear Dirichlet eigenvalue problem for the $p$-Laplacian. The iterative technique was inspired by the inverse power method in finite dimensional linear algebra and is competitive with other methods available in the literature.", "subjects": "Classical Analysis and ODEs (math.CA); Numerical Analysis (math.NA)", "title": "Computing the $\\sin_{p}$ function via the inverse power method", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.976310532836284, "lm_q2_score": 0.8198933403143929, "lm_q1q2_score": 0.8004705039512656 }
https://arxiv.org/abs/2209.02588
A generalization of Chu-Vandermonde's Identity
We present and prove a general form of Vandermonde's identity and use it as an alternative solution to a classic probability problem.
\section{Introduction} Vandermonde's identity, its diverse proofs, and applications have been the spot of research throughout the centuries. Nowadays, it is a classic identity whose proof is available in most combinatorics-related books~\cite{knuth94}. However, there are still some efforts to give new proofs for this well-established identity, e.g.,~\cite{spivey16}. In addition to its original form proof, some specific types of its generalization have been proposed. Some works concentrated on its higher dimension~\cite{yaacov17}. Some others put their efforts into its complex~\cite{mestrovic18} or coefficients of its numbers~\cite{knuth94}. However, our focus was not generalizing this identity directly. We were doing some exercises in Stochastic Process and reached a classic problem during its teaching, which we put in this paper with an alternative solution in section~\ref{seq:app}. To have a formal proof of it, we reached an equation that we later realized could be considered a generalization of \textit{Vandermonde's identity}. This article gives a general form of Vandermonde's identity, including first order and higher orders of this identity. Then, we proof the mentioned classic problem, borrowed from Ross book~\cite{ross10} using the first order of general Vandermonde's identity. Lastly, we briefly discuss some of our thoughts about this proof. \section{The Vandermonde's identity General form.} At first, we encountered the first order of Vandermonde's identity. We prove it as follows. \begin{theorem}[First Order of Vandermonde's Identity General Form.]\label{1storder} Let assume $k\leq m$ and $r-k\leq n$, then $\sum_{k=0}^{r}{k{m \choose k}{n \choose r-k}} = m {m+n-1 \choose r-1} \label{eq:mainFormula}$, where ${x \choose y}=\frac{x!}{(x-y)!y!}$, and $x,y\in \mathbb{N}$ . \end{theorem} \begin{proof} Let start with \begin{equation*} \begin{aligned} m \left( x+1 \right)^{m+n-1} & = m \left( x+1 \right)^{m-1} \left( x+1 \right)^{n}\\ \end{aligned} \end{equation*} By $m \left({1+x}\right) ^{m-1} = \sum_{i=0}^{m}{i{m \choose i} x^{i-1}}$~\cite{Koh} we have, \begin{equation*} \begin{aligned} & = \sum_{i=0}^{m}{i{m \choose i} x^{i-1}} \sum_{j=0}^{n}{{n \choose j} x^{j}}.\\ \end{aligned} \end{equation*} We get factor of $\frac{1}{x}$ from $x^{i-1}$ in the above equation; thus, we have \begin{equation*} \begin{aligned} & = \frac{1}{x}\sum_{i=0}^{m}{i{m \choose i} x^{i}} \sum_{j=0}^{n}{{n \choose j} x^{j}}.\\ \end{aligned} \end{equation*} Using \textit{polynomial ring}, $\sum_{i=0}^{m}{a_{i}{x^{i}}} \sum_{j=0}^{n}{b_{j}x^{j}} = \sum_{r=0}^{m+n}{{ \left( \sum_{k=0}^{r}{a_{k}b_{r-k}} \right) }x^{r}}$ ~\cite{whitelow}, we convert the above formula as follows. \begin{equation*} \begin{aligned} & = \frac{1}{x}\sum_{r=0}^{m+n}{{ \left( \sum_{k=0}^{r}{k{m \choose k}{n \choose r-k}} \right) }x^{r}}\\ & = \sum_{r=0}^{m+n}{{ \left( \sum_{k=0}^{r}{k{m \choose k}{n \choose r-k}} \right) }x^{r-1}}\\ & = 0 + \sum_{r=1}^{m+n}{{ \left( \sum_{k=0}^{r}{k{m \choose k}{n \choose r-k}} \right) }x^{r-1}} \end{aligned} \end{equation*} , or \begin{equation} m \left( x+1 \right)^{m+n-1} = \sum_{r=1}^{m+n}{{ \left( \sum_{k=0}^{r}{k{m \choose k}{n \choose r-k}} \right) }x^{r-1}}. \label{eq:firstExpansion} \end{equation} The equation~\ref{eq:firstExpansion} is our first expansion. On the other hand, we consider \textit{binomial expansion}, \begin{equation*} \left( x+1 \right)^{m+n-1} = \sum_{r=0}^{m+n-1}{{m+n-1 \choose r}}x^{r}. \end{equation*} By changing the index variable, the equation is \begin{equation*} = \sum_{r=1}^{m+n}{{m+n-1 \choose r-1}}x^{r-1}. \end{equation*} Multiplying both sides with $m$ \begin{equation} m \left( x+1 \right)^{m+n-1} = \sum_{r=1}^{m+n}{m{m+n-1 \choose r-1}}x^{r-1} \label{eq:binomialexp} \end{equation} The left sides of equation~\ref{eq:firstExpansion} and equation~\ref{eq:binomialexp} are equal. Thus, \begin{equation*} \begin{aligned} \sum_{r=1}^{m+n}{{ \left( \sum_{k=0}^{r}{k{m \choose k}{n \choose r-k}} \right) }x^{r-1}}= \sum_{r=1}^{m+n}{m{m+n-1 \choose r-1}}x^{r-1} \end{aligned} \end{equation*} , and results in \begin{equation} \begin{aligned} \sum_{k=0}^{r}{k{m \choose k}{n \choose r-k}} = m {m+n-1 \choose r-1}. \end{aligned} \label{eq:rightSidesAreEqual} \end{equation} \end{proof} \textbf{Theorem}~\ref{1storder} is a leading part of an alternative solution to the problem in section~\ref{seq:app}. However, going further, we realized that the \textbf{Theorem}~\ref{1storder} is a particular case of a more general form. Thus, we introduce the general form of \textit{Vandermonde's identity} as the below theorem. \begin{theorem}[General Form of Vandermonde's Identity] \label{generalVI} Let assume $l\leq k$, $k\leq m$ and $r-k\leq n$ then, \begin{equation} \sum_{k=0}^{r}{{k \brack l}{m \choose k}{n \choose r-k}} = {m \brack l} {m+n-l \choose r-l} \label{eq:generalizedFormula} \end{equation} , where ${x \brack y}=\frac{x!}{(x-y)!}$ and ${x \choose y}=\frac{x!}{(x-y)!y!}$, and $x,y\in \mathbb{N}$. \end{theorem} \begin{proof} Let $f(x)= (1 + x)^{m}$. its binomial expansion is \begin{equation*} f(x) = (1 + x)^{m} = \sum_{i=0}^{m}{{m \choose i}x^{i}}. \end{equation*} Then, utilizing induction, its $l$-th derivative is \begin{equation} f^{\left( l \right)}(x) = {m \brack l} (1+x)^{m-l} = \sum_{i=0}^{m}{{i \brack l}{m \choose i}x^{i-l}}. \label{eq:mthderivative} \end{equation} Now, we start with ${m \brack l} \left( x+1\right)^{m+n-l}$. We have \begin{equation*} \begin{aligned} {m \brack l} \left( x+1 \right)^{m+n-l} & = {m \brack l} \left( x+1 \right)^{m-l} \left( x+1 \right)^{n}\\ \end{aligned} \end{equation*} using equation~\ref{eq:mthderivative} and in a similar way to the corresponding steps of \textbf{Theorem}~\ref{1storder}, we utilize the \textit{polynomial ring} and the equation is as follows. \begin{equation*} \begin{aligned} & = \sum_{i=0}^{m}{{i \brack l}{m \choose i} x^{i-l}} \sum_{j=0}^{n}{{n \choose j} x^{j}}\\ & = \sum_{r=0}^{m+n}{{ \left( \sum_{k=0}^{r}{{k \brack l}{m \choose k}{n \choose r-k}} \right) }x^{r-l}}\\ & = 0 + \sum_{r=l}^{m+n}{{ \left( \sum_{k=0}^{r}{{k \brack l}{m \choose k}{n \choose r-k}} \right) }x^{r-l}} \end{aligned} \end{equation*} , or \begin{equation} \begin{aligned} {m \brack l} \left( x+1 \right)^{m+n-l} & = \sum_{r=l}^{m+n}{{ \left( \sum_{k=0}^{r}{{k \brack l}{m \choose k}{n \choose r-k}} \right) }x^{r-l}} \end{aligned} \label{eq:gen1stExpansion} \end{equation} On the other hand, we use binomial expansion and have \begin{equation*} \left( x+1 \right)^{m+n-l} = \sum_{r=0}^{m+n-l}{{m+n-l \choose r}}x^{r} \label{eq:gen2ndExpansion} \end{equation*} changing the index variable and multiplying both sides with ${m \brack l}$ \begin{equation} {m \brack l} \left( x+1 \right)^{m+n-l} = \sum_{r=l}^{m+n}{{m \brack l}{m+n-l \choose r-l}}x^{r-l} \label{eq:gen2ndExpansion} \end{equation} The left-hand sides of both equations~\ref{eq:gen1stExpansion} and~equation \ref{eq:gen2ndExpansion} are the same, so the right-hand side of both equations are the same as well, or simply \begin{equation} \begin{aligned} \sum_{k=0}^{r}{{k \brack l}{m \choose k}{n \choose r-k}} = {m \brack l} {m+n-l \choose r-l}. \end{aligned} \end{equation} \end{proof} \begin{remark} In \textbf{Theorem}~\ref{generalVI}, if \begin{enumerate} \item $l=0$, then it is equivalent with Vandermonde's identity, or \begin{equation} \begin{aligned} \sum_{k=0}^{r}{{m \choose k}{n \choose r-k}} = {m+n \choose r}; \end{aligned} \end{equation} \item $l=1$, then it is equivalent with first order of general order of Vandermonde's identity; \item $l>1$, then it is equivalent to higher order of general form of Vandermonde's identity. \end{enumerate} \end{remark} \section{Application.}~\label{seq:app} In this section, we present the query which led us to face the first order of Vandermonde's identity general form. We prove the problem using \textbf{Theorem}~\ref{1storder}. \subsection{Statement of the Problem.} ``An urn has $r$ red and $w$ white balls that are randomly removed one at a time. Let $R_{i}$ be the event that the $i$th ball removed is red. Find $P\left( R_{i} \right)$''~\cite{ross10}. \textbf{Solution.} The answer is $\frac{r}{r+w}$. \textit{Also sprach Ross} ``... each of the $r+w$ balls is equally to be the $i$th ball removed''. In addition to the intuition introduced by Sheldon Ross, we would like to have a direct alternative formal proof of the answer. Thus, we tried to prove it, and here it is. \begin{proof} Let assume $r_{j}$ denotes choosing of $j$ red balls in $i-1$ previous steps. Then, \begin{equation*} P\left(R_{i}\right) = \sum_{j=0}^{i-1}{P\left( R_{i} \cap r_{j} \right)}\\ \end{equation*} By applying \textit{Law of total probability} we have, \begin{equation*} = \sum_{j=0}^{i-1}{P\left( R_{i} | r_{j} \right) P\left(r_{j}\right) } \end{equation*} $P\left(r_{j}\right)$, as mentioned above, is equal to choosing $j$ red balls among the $i-1$ balls that we have already picked. Thus, its value is equal to $\frac{{ r \choose j}{ w \choose i-1-j}}{{r+w \choose i-1}}$. In addition, $P\left( R_{i} | r_{j}\right)$ denotes choosing a red ball in the $i$th step. Because, we have already picked $j$ red balls in the previous steps, its value is equal to $\frac{{r-j \choose 1}}{{r+w-\left( i-1 \right) \choose 1}}$. Consequently, we have, \begin{equation*} \begin{aligned} & = \sum_{j=0}^{i-1}{\left(\frac{{r-j \choose 1}}{{r+w-\left( i-1 \right) \choose 1}} \frac{{ r \choose j}{ w \choose i-1-j}}{{r+w \choose i-1}}\right)} \\ &= \sum_{j=0}^{i-1}{\left(\frac{(r-j)}{(r+w-i+1)} \frac{{ r \choose j}{ w \choose i-1-j}}{{r+w \choose i-1}}\right)}\\ & = \frac{1}{ \left(r+w-i+1\right) {r+w \choose i-1}} \sum_{j=0}^{i-1}{\left((r-j) { r \choose j}{ w \choose i-1-j}\right)} \\ & = \frac{1}{\left(r+w-i+1\right) {r+w \choose i-1}} \left(\sum_{j=0}^{i-1}{r{ r \choose j}{ w \choose i-1-j}} - \sum_{j=0}^{i-1}{j{ r \choose j}{ w \choose i-1-j}}\right)\\ \end{aligned} \end{equation*} The $\sum_{j=0}^{i-1}{j{ r \choose j}{ w \choose i-1-j}}$ term in the above formula is in accordance with \textbf{Theorem}~\ref{1storder}. Thus, we proceed with it as follows, \begin{equation*} \begin{aligned} & = \frac{1}{\left(r+w-i+1\right) {r+w \choose i-1}} \left(r{r+w \choose i-1} - r {r+w-1 \choose i-2 } \right)\\ & = \frac{r}{\left(r+w-i+1\right) {r+w \choose i-1}} \left({r+w \choose i-1} - {r+w-1 \choose i-2 } \right)\\ & = \frac{r}{\left(r+w-i+1\right) {r+w \choose i-1}} \left({r+w \choose i-1} - \frac{i-1}{r+w}{r+w \choose i-1 } \right)\\ & = \frac{r{r+w \choose i-1 }}{\left(r+w-i+1\right) {r+w \choose i-1}} \left( 1 - \frac{i-1}{r+w} \right) \\ & = \frac{r}{\left(r+w-i+1\right)} \left( \frac{\left(r+w-i+1\right)}{r+w} \right)\\ & = \frac{r}{r+w}. \end{aligned} \label{eq:ithRedBallProbability} \end{equation*} \end{proof} \section{Final Thoughts} There have been many efforts to expand or extend Vandermonde's identity or Vandermonde's convolution. For example, Graham et al.~\cite{knuth94} defined some general forms of identities, including Vandermonde's identity, in chapter~5. But, that generalization is different from our point of view. They are a generalization of Vandermonde's identity, discussing the coefficients of $m$ or $n$ in the equation. The same happens in works such as~\cite{fang07, mestrovic18 }. Yaccov~\cite{yaacov17} investigated the determinant of the Vandermonde in a higher dimension. However, to our knowledge, we proposed this type of generalization of Vandermonde's identity for the first time, or at least we have given a new way of this generalization proof. In our proposal, we consider the general form of Vandermonde's identity from the point of view of Combinatorics' coefficients. It is worth mentioning that it is possible to observe our proposed generalization as a higher derivative of Vandermonde's identity. \bibliographystyle{amsplain}
{ "timestamp": "2022-09-07T02:51:52", "yymm": "2209", "arxiv_id": "2209.02588", "language": "en", "url": "https://arxiv.org/abs/2209.02588", "abstract": "We present and prove a general form of Vandermonde's identity and use it as an alternative solution to a classic probability problem.", "subjects": "General Mathematics (math.GM)", "title": "A generalization of Chu-Vandermonde's Identity", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9763105273220727, "lm_q2_score": 0.8198933425148213, "lm_q1q2_score": 0.8004705015785019 }
https://arxiv.org/abs/1508.04721
Every natural number is the sum of forty-nine palindromes
It is shown that the set of decimal palindromes is an additive basis for the natural numbers. Specifically, we prove that every natural number can be expressed as the sum of forty-nine (possibly zero) decimal palindromes.
\section{Statement of result} Let $\NN\defeq\{0,1,2,\ldots\}$ denote the set of natural numbers (including zero). Every number $n\in\NN$ has a unique decimal representation of the form \begin{equation} \label{eq:representation} n=\sum_{j=0}^{L-1}10^j\delta_j, \end{equation} where each digit $\delta_j$ belongs to the digit set $$ \cD\defeq\{0,1,2,\ldots,9\}, $$ and the leading digit $\delta_{L-1}$ is nonzero whenever $L\ge 2$. In what follows, we use diagrams to illustrate the ideas; for example, $$ n=\begin{tabular}{|c|c|c|c|} \hline $\delta_{L-1}$&$\cdots$&$\delta_1$&$\delta_0$\\ \hline \end{tabular} $$ represents the relation \eqref{eq:representation}. The integer $n$ is said to be a \emph{palindrome} if its digits satisfy the symmetry condition $$ \delta_j=\delta_{L-1-j}\qquad(0\le j<L). $$ Denoting by $\cP$ the collection of all palindromes in $\NN$, the aim of this note is to show that $\cP$ is an additive basis for $\NN$. \begin{theorem} \label{thm:main} The set $\cP$ of decimal palindromes is an additive basis for the natural numbers $\NN$. Every natural number is the sum of forty-nine (possibly zero) decimal palindromes. \end{theorem} The proof is given in the next section. It is unlikely that the second statement is optimal; a refinement of our method may yield an improvement. No attempt has been made to generalize this theorem to bases other than ten; for large bases, this should be straightforward, but small bases may present new obstacles (for example, obtaining the correct analogue of Lemma~\ref{lem:passage} may be challenging in the binary case, where the only nonzero digit is the digit one). We remark that arithmetic properties of palindromes (in various bases) have been previously investigated by many authors; see\cite{AllSha,BanHarSak,BanShp1,BanShp2,BerZie,CilLucShp,CilTesLuc,Col, CooGab,Goi,HerLuc,Kor,Luc,Sim} and the references therein. \section{The proof} \label{sec:proof} \subsection{Notation} \label{sec:notate} For every $n\in\NN$, let $L(n)$ (the ``length'' of $n$) denote the number of decimal digits $L$ in the expansion \eqref{eq:representation}; in particular, $L(0)\defeq 1$. For any $\ell\in\NN$ and $d\in\cD$, we denote \begin{equation} \label{eq:plddefn} p_\ell(d)\defeq\begin{cases} 0&\quad\hbox{if $\ell=0$};\\ d&\quad\hbox{if $\ell=1$};\\ 10^{\ell-1}d+d&\quad\hbox{if $\ell\ge 2$}. \end{cases} \end{equation} Note that $p_\ell(d)$ is a palindrome, and $L(p_\ell(d))=\ell$ if $d\ne 0$. If $\ell\ge 2$, then the decimal expansion of $p_\ell(d)$ has the form $$ p_\ell(d)=\begin{tabular}{|c|c|c|} \hline $d$&$0\cdots 0$&$d$\\ \hline \end{tabular} $$ with $\ell-2$ zeros nested between two copies of the digit $d$. More generally, for any integers $\ell\ge k\ge 0$ and $d\in\cD$, let $$ p_{\ell,k}(d)\defeq 10^k p_{\ell-k}(d)=\begin{cases} 0&\quad\hbox{if $\ell=k$};\\ 10^k d&\quad\hbox{if $\ell=k+1$};\\ 10^{\ell-1}d+10^k d&\quad\hbox{if $\ell\ge k+2$}. \end{cases} $$ If $\ell\ge k+2$, then the decimal expansion of $p_{\ell,k}(d)$ has the form $$ p_{\ell,k}(d)=\begin{tabular}{|c|c|c|c|} \hline $d$&$0\cdots 0$&$d$&$0\cdots 0$\\ \hline \end{tabular} $$ with $\ell-k-2$ zeros nested between two copies of the digit $d$, followed by $k$ copies of the digit zero. Next, for any integers $\ell,k\in\NN$, $\ell\ge k+4$, and digits $a,b\in\cD$, we denote \begin{equation} \label{eq:qlkabdefn} q_{\ell,k}(a,b)\defeq p_{\ell,k}(a)+p_{\ell-1,k}(b)= 10^{\ell-1}a+10^{\ell-2}b+10^k(a+b). \end{equation} Taking into account that the relation $n=10\cdot\fl{n/10}+\delta_0(n)$ holds for every natural number $n<100$, where $\fl{\cdot}$ is the floor function and $\delta_0(r)$ denotes the one's digit of any natural number $r$, one sees that the decimal expansion of $q_{\ell,k}(a,b)$ has the form $$ q_{\ell,k}(a,b)=\begin{tabular}{|c|c|c|c|c|c|} \hline $a$&$b$&$0\cdots 0$&$\fl{(a+b)/10}$&$\delta_0(a+b)$&$0\cdots 0$\\ \hline \end{tabular} $$ with $\ell-k-4$ zeros nested between the digits $a,b$ and the digits of $a+b$, followed by $k$ copies of the digit zero. For example, $q_{10,2}(7,8)=7800001500$. Finally, for any integers $\ell,k\in\NN$, $\ell\ge k+4$, and digits $a,b,c\in\cD$, $a\ne 0$, we denote by $\NN_{\ell,k}(a,b;c)$ the set of natural numbers described as follows. Given $n\in\NN$, let $L$ and $\{\delta_j\}_{j=0}^{L-1}$ be defined as in \eqref{eq:representation}. Then $\NN_{\ell,k}(a,b;c)$ consists of those integers $n$ for which $L=\ell$, $\delta_{\ell-1}=a$, $\delta_{\ell-2}=b$, $\delta_k=c$, and $10^k\mid n$. In other words, $\NN_{\ell,k}(a,b;c)$ is the set of natural numbers $n$ that have a decimal expansion of the form $$ n=\begin{tabular}{|c|c|c|c|c|} \hline $a$&$b$&$\bigast\cdots\bigast$&$c$&$0\cdots 0$\\ \hline \end{tabular} $$ with $\ell-k-3$ arbitrary digits nested between the digits $a,b$ and the digit $c$, followed by $k$ copies of the digit zero. We reiterate that $a\ne 0$. \subsection{Handling small integers} \label{sec:handling} Let $f:\cD\to\cD$ be the function whose values are provided by the following table: $$ \begin{tabular}{|c||c|c|c|c|c|c|c|c|c|c|} \hline $d$&$0$&$1$&$2$&$3$&$4$&$5$&$6$&$7$&$8$&$9$\\ \hline $f(d)$&$0$&$1$&$9$&$8$&$7$&$6$&$5$&$4$&$3$&$2$\\ \hline \end{tabular} $$ We begin our proof of Theorem~\ref{thm:main} with the following observation. \begin{lemma} \label{lem:10^jd-f(d)} Every number $f(d)$ is a palindrome, and $10^jd-f(d)$ is a palindrome for every integer $j\ge 1$. \end{lemma} \begin{proof} This is easily seen if $d=0$ or $j=1$. For $j\ge 2$ and $d=1$, the number $10^jd-f(d)=10^j-1$ is a repunit of the form $9\cdots 9$, hence a palindrome. Finally, for $j\ge 2$ and $2\le d\le 9$, the number $10^jd-f(d)$ is a palindrome that has a decimal expansion of the form $$ 10^jd-f(d)=\begin{tabular}{|c|c|c|} \hline $d-1$&$9\cdots 9$&$d-1$\\ \hline \end{tabular} $$ with $j-2$ nines nested between two copies of the digit $d-1$. \end{proof} \begin{lemma} \label{lem:small} If $n$ is a natural number with at most $K$ nonzero decimal digits, then $n$ is the sum of $2K+1$ palindromes. \end{lemma} \begin{proof} Starting with the expansion \eqref{eq:representation} we write $$ n=\delta_0+\sum_{j\in\cJ}10^j\delta_j, $$ where $$ \cJ\defeq\big\{1\le j<L:\delta_j\ne 0\big\}. $$ Since \begin{equation} \label{eq:nexpr} n=\delta_0 +\sum_{j\in\cJ}f(\delta_j) +\sum_{j\in\cJ}\bigl(10^j\delta_j-f(\delta_j)\bigr), \end{equation} Lemma~\ref{lem:10^jd-f(d)} implies that $n$ is the sum of $2|\cJ|+1$ palindromes. Since zero is a palindrome, we obtain the stated result by adding $2K-2|\cJ|$ additional zeros on the right side of \eqref{eq:nexpr}. \end{proof} Lemma~\ref{lem:small} implies, in particular, that $n\in\NN$ is a sum of 49 palindromes whenever $L(n)\le 24$. Therefore, we can assume that $L(n)\ge 25$ in the sequel. \subsection{Reduction to $\NN_{\ell,0}(5^+;c)$} Recall the definition of $\NN_{\ell,k}(a,b;c)$ given in \S\ref{sec:notate}. For any given integers $\ell,k\in\NN$, $\ell\ge k+4$, and a digit $c\in\cD$, we now denote $$ \NN_{\ell,k}(5^+;c)\defeq\bigcup_{\substack{a,b\in\cD\\a\ge 5}} \NN_{\ell,k}(a,b;c). $$ The set $\NN_{\ell,k}(5^+;c)$ can be described as follows. For each $n\in\NN$, let $L$ and $\{\delta_j\}_{j=0}^{L-1}$ be defined as in \eqref{eq:representation}. The set $\NN_{\ell,k}(5^+;c)$ consists of those integers $n$ for which $L=\ell$, $\delta_{\ell-1}\ge 5$, $\delta_k=c$, and $10^k\mid n$. In other words, $\NN_{\ell,k}(5^+;c)$ is the set of natural numbers $n$ that have a decimal expansion of the form $$ n=\begin{tabular}{|c|c|c|c|c|} \hline $a$&$\bigast\cdots\bigast$&$c$&$0\cdots 0$\\ \hline \end{tabular} $$ with $\ell-k-2$ arbitrary digits nested between the digit $a$ ($\ge 5$) and the digit $c$, followed by $k$ copies of the digit zero. \begin{lemma} \label{lem:1streduction} Let $n\in\NN$, and put $L\defeq L(n)$ as in \eqref{eq:representation}. If $L\ge 5$, then $n$ is the sum of two palindromes and an element of $\NN_{\ell,0}(5^+;c)$ with some $\ell\in\{L-1,L\}$ and $c\in\cD$. \end{lemma} \begin{proof} Let $\{\delta_j\}_{j=0}^{L-1}$ be defined as in \eqref{eq:representation}. If the leading digit $\delta_{L-1}$ exceeds four, then $n\in\NN_{L,0}(5^+;\delta_0)$, and there is nothing to prove (since zero is a palindrome). Now suppose that $\delta_{L-1}\le 4$. Put $m\defeq 10\delta_{L-1}+\delta_{L-2}-6$, and observe that $4\le m\le 43$. If $4\le m\le 9$, then using \eqref{eq:plddefn} we see that \begin{align*} n-p_{L-1}(m) &=n-(10^{L-2}m+m)\\ &=\sum_{j=0}^{L-1}10^j\delta_j-10^{L-2}(10\delta_{L-1}+\delta_{L-2}-6)-m\\ &=6\cdot 10^{L-2}+\sum_{j=0}^{L-3}10^j\delta_j-m, \end{align*} and the latter number evidently lies in $\NN_{L-1,0}(5^+;c)$, where $c\equiv (\delta_0-m)\bmod 10$. Since $p_{L-1}(m)$ is a palindrome, this yields the desired result for $4\le m\le 9$. In the case that $10\le m\le 43$, we write $m=10a+b$ with digits $a,b\in\cD$, $a\ne 0$. Using \eqref{eq:qlkabdefn} we have \begin{align*} n-q_{L,0}(a,b) &=n-(10^{L-1}a+10^{L-2}b+a+b)\\ &=n-(10^{L-2}m+a+b)\\ &=\sum_{j=0}^{L-1}10^j\delta_j-10^{L-2}(10\delta_{L-1}+\delta_{L-2}-6)-a-b\\ &=6\cdot 10^{L-2}+\sum_{j=0}^{L-3}10^j\delta_j-a-b, \end{align*} and the latter number lies in $\NN_{L-1,0}(5^+;c)$, where $c\equiv (\delta_0-a-b)\bmod 10$. Since $q_{L,0}(a,b)$ is the sum of two palindromes, we are done in this case as well. \end{proof} \subsection{Inductive passage from $\NN_{\ell,k}(5^+;c_1)$ to $\NN_{\ell-1,k+1}(5^+;c_2)$} \begin{lemma} \label{lem:passage} Let $\ell,k\in\NN$, $\ell\ge k+6$, and $c_\ell\in\cD$ be given. Given $n\in\NN_{\ell,k}(5^+;c_1)$, one can find digits $a_1,\ldots,a_{18},b_1,\ldots,b_{18}\in\cD\setminus\{0\}$ and $c_2\in\cD$ such that the number $$ n-\sum_{j=1}^{18}q_{\ell-1,k}(a_j,b_j) $$ lies in the set $\NN_{\ell-1,k+1}(5^+;c_2)$. \end{lemma} \begin{proof} Fix $n\in\NN_{\ell,k}(5^+;c_1)$, and let $\{\delta_j\}_{j=0}^{\ell-1}$ be defined as in \eqref{eq:representation} (with $L\defeq\ell$). Let $m$ be the three-digit integer formed by the first three digits of $n$; that is, $$ m\defeq 100\delta_{\ell-1}+10\delta_{\ell-2}+\delta_{\ell-3}. $$ Clearly, $m$ is an integer in the range $500\le m\le 999$, and we have \begin{equation} \label{eq:expandit} n=\sum_{j=k}^{\ell-1}10^j\delta_j=10^{\ell-3}m+\sum_{j=k}^{\ell-4}10^j\delta_j. \end{equation} Let us denote $$ \cS\defeq\{19,29,39,49,59\}. $$ In view of the fact that $$ 9\cS\defeq \mathop{\underbracket{\hskip3pt\cS+\cdots+\cS\hskip1pt}}\limits_{\text{nine copies}} =\{171,181,191,\ldots,531\}, $$ it is possible to find an element $h\in 9\cS$ for which $m-80<2h\le m-60$. With $h$ fixed, let $s_1,\ldots,s_9$ be elements of $\cS$ such that $$ s_1+\cdots+s_9=h. $$ Finally, let $\eps_1,\ldots,\eps_9$ be natural numbers, each equal to zero or two: $\eps_j\in\{0,2\}$ for $j=1,\ldots,9$. A specific choice of these numbers is given below. We now put $$ t_j\defeq s_j+\eps_j\mand t_{j+9}\defeq s_j-\eps_j\qquad(j=1,\ldots,9), $$ and let $a_1,\ldots,a_{18},b_1,\ldots,b_{18}\in\cD$ be determined from the digits of $t_1,\ldots,t_{18}$, respectively, via the relations $$ 10a_j+b_j=t_j\qquad (j=1,\ldots,18). $$ Since $$ \cS+2=\{21,31,41,51,61\}\mand \cS-2=\{17,27,37,47,57\}, $$ all of the digits $a_1,\ldots,a_{18},b_1,\ldots,b_{18}$ are \emph{nonzero}, as required. Using \eqref{eq:qlkabdefn} we compute \begin{align*} \sum_{j=1}^{18}q_{\ell-1,k}(a_j,b_j) &=\sum_{j=1}^{18}\bigl(10^{\ell-2}a_j+10^{\ell-3}b_j+10^k(a_j+b_j)\bigr)\\ &=10^{\ell-3}\sum_{j=1}^{18}t_j+10^k\sum_{j=1}^{18}(a_j+b_j)\\ &=2h\cdot 10^{\ell-3}+10^k\sum_{j=1}^{18}(a_j+b_j) \end{align*} since $$ t_1+\cdots+t_{18}=2(s_1+\cdots+s_9)=2h $$ regardless of the choice of the $\eps_j$'s. Taking \eqref{eq:expandit} into account, we have \begin{equation} \label{eq:eureka} n-\sum_{j=1}^{18}q_{\ell-1,k}(a_j,b_j) =10^{\ell-3}(m-2h)+\sum_{j=k}^{\ell-4}10^j\delta_j -10^k\sum_{j=1}^{18}(a_j+b_j), \end{equation} and since $60\le m-2h<80$ it follows that the number defined by either side of \eqref{eq:eureka} lies in the set $\NN_{\ell-1,k}(5^+;c)$, where $c$ is the unique digit in $\cD$ determined by the congruence \begin{equation} \label{eq:congone} \delta_k-\sum_{j=1}^{18}(a_j+b_j)\equiv c\bmod 10. \end{equation} To complete the proof, it suffices to show that for an appropriate choice of the $\eps_j$'s we have $c=0$, for this implies that $n\in\NN_{\ell-1,k+1}(5^+;c_2)$ for some $c_2\in\cD$. To do this, let $g(r)$ denote the sum of the decimal digits of any $r\in\NN$. Then $$ \sum_{j=1}^{18}(a_j+b_j)=\sum_{j=1}^{18}g(t_j) =\sum_{j=1}^9g(s_j+\eps_j)+\sum_{j=1}^9g(s_j-\eps_j). $$ For every number $s\in\cS$, one readily verifies that $$ g(s+2)+g(s-2)=2\,g(s)-9. $$ Therefore, \eqref{eq:congone} is equivalent to the congruence condition $$ \delta_k-\sum_{j=1}^{18}g(s_j)+9E\equiv c\bmod 10, $$ where $E$ is the number of integers $j\in\{1,\ldots,9\}$ such that $\eps_j\defeq 2$. As we can clearly choose the $\eps_j$'s so the latter congruence is satisfied with $c=0$, the proof of the lemma is complete. \end{proof} \subsection{Proof of Theorem~\ref{thm:main}} Let $n$ be an arbitrary natural number. To show that $n$ is the sum of 49 palindromes, we can assume that $L:=L(n)$ is at least $25$, as mentioned in \S\ref{sec:handling}. By Lemma~\ref{lem:1streduction} we can find two palindromes ${\widetilde p}_1,{\widetilde p}_2$ such that the number \begin{equation} \label{eq:n1reln} n_1\defeq n-{\widetilde p}_1-{\widetilde p}_2 \end{equation} belongs to $\NN_{\ell,0}(5^+;c_1)$ for some $\ell\in\{L-1,L\}$ and $c_1\in\cD$. Since $\ell\ge 24$, by Lemma~\ref{lem:passage} we can find digits $a_1^{(1)},\ldots,a_{18}^{(1)},b_1^{(1)},\ldots,b_{18}^{(1)}\in\cD\setminus\{0\}$ and $c_2\in\cD$ such that the number $$ n_2\defeq n_1-\sum_{j=1}^{18}q_{\ell-1,0}\bigl(a_j^{(1)},b_j^{(1)}\bigr) $$ lies in the set $\NN_{\ell-1,1}(5^+;c_2)$. Similarly, using Lemma~\ref{lem:passage} again we can find digits $a_1^{(2)},\ldots,a_{18}^{(2)},b_1^{(2)},\ldots,b_{18}^{(2)}\in\cD\setminus\{0\}$ and $c_3\in\cD$ such that $$ n_3\defeq n_2-\sum_{j=1}^{18}q_{\ell-2,1}\bigl(a_j^{(2)},b_j^{(2)}\bigr) $$ belongs to the set $\NN_{\ell-2,2}(5^+;c_3)$. Proceeding inductively in this manner, we continue to construct the sequence $n_1,n_2,n_3,\ldots$, where each number \begin{equation} \label{eq:ninduct} n_i\defeq n_{i-1}-\sum_{j=1}^{18}q_{\ell-i+1,i-2}\bigl(a_j^{(i-1)},b_j^{(i-1)}\bigr) \end{equation} lies in the set $\NN_{\ell-i+1,i-1}(5^+;c_i)$. The method works until we reach a specific value of $i$, say $i\defeq\nu$, where $\ell-\nu+1<(\nu-1)+6$; at this point, Lemma~\ref{lem:passage} can no longer be applied. Notice that, since $\ell-\nu+1\le (\nu-1)+5$, every element of $\NN_{\ell-\nu+1,\nu-1}(5^+;c_\nu)$ has at most five nonzero digits. Therefore, by Lemma~\ref{lem:small} we can find eleven palindromes ${\widetilde p}_3,{\widetilde p}_4,\ldots,{\widetilde p}_{13}$ such that \begin{equation} \label{eq:ni0reln} n_{\nu}={\widetilde p}_3+{\widetilde p}_4+\cdots+{\widetilde p}_{13}. \end{equation} Now, combining \eqref{eq:n1reln}, \eqref{eq:ninduct} with $i=2,3,\ldots,\nu$, and \eqref{eq:ni0reln}, we see that $$ n=\sum_{i=1}^{13}{\widetilde p}_j+\sum_{j=1}^{18}N_j, $$ where $$ N_j\defeq\sum_{i=2}^\nu q_{\ell-i+1,i-2}\bigl(a_j^{(i-1)},b_j^{(i-1)}\bigr) \qquad(j=1,\ldots,18). $$ To complete the proof of the theorem, it remains to verify that every integer $N_j$ is the sum of two palindromes. Indeed, by \eqref{eq:qlkabdefn} we have $$ N_j=\sum_{i=2}^\nu p_{\ell-i+1,i-2}\bigl(a_j^{(i-1)}\bigr) +\sum_{i=2}^\nu p_{\ell-i,i-2}\bigl(b_j^{(i-1)}\bigr). $$ Considering the form of the decimal expansions, for each $j$ we see that $$ \sum_{i=2}^\nu p_{\ell-i+1,i-2}\bigl(a_j^{(i-1)}\bigr) =\begin{tabular}{|c|c|c|c|c|c|c|} \hline $\vphantom{\Big|}a_j^{(1)}$&$\cdots$&$a_j^{(\nu-1)}$&$0\cdots 0$&$a_j^{(\nu-1)}$&$\cdots$&$a_j^{(1)}$\\ \hline \end{tabular} $$ which is a palindrome of length $\ell-1$ (since $a_j^{(1)}\ne 0$) having precisely $2(\nu-1)$ nonzero entries, and $$ \sum_{i=2}^\nu p_{\ell-i,i-2}\bigl(b_j^{(i-1)}\bigr) =\begin{tabular}{|c|c|c|c|c|c|c|} \hline $\vphantom{\Big|}b_j^{(1)}$&$\cdots$&$b_j^{(\nu-1)}$&$0\cdots 0$&$b_j^{(\nu-1)}$&$\cdots$&$b_j^{(1)}$\\ \hline \end{tabular} $$ which is a palindrome of length $\ell-2$ (since $b_j^{(1)}\ne 0$), also having precisely $2(\nu-1)$ nonzero entries.
{ "timestamp": "2015-08-20T02:11:43", "yymm": "1508", "arxiv_id": "1508.04721", "language": "en", "url": "https://arxiv.org/abs/1508.04721", "abstract": "It is shown that the set of decimal palindromes is an additive basis for the natural numbers. Specifically, we prove that every natural number can be expressed as the sum of forty-nine (possibly zero) decimal palindromes.", "subjects": "Number Theory (math.NT)", "title": "Every natural number is the sum of forty-nine palindromes", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9905874115238966, "lm_q2_score": 0.8080672204860316, "lm_q1q2_score": 0.8004612162785679 }
https://arxiv.org/abs/1811.03008
Limit points of normalized prime gaps
We show that at least 1/3 of positive real numbers are in the set of limit points of normalized prime gaps. More precisely, if $p_n$ denotes the $n$th prime and $\mathbb{L}$ is the set of limit points of the sequence $\{(p_{n+1}-p_n)/\log p_n\}_{n=1}^\infty,$ then for all $T\geq 0$ the Lebesque measure of $\mathbb{L} \cap [0,T]$ is at least $T/3.$ This improves the result of Pintz (2015) that the Lebesque measure of $\mathbb{L} \cap [0,T]$ is at least $(1/4-o(1))T,$ which was obtained by a refinement of the previous ideas of Banks, Freiberg, and Maynard (2015). Our improvement comes from using Chen's sieve to give, for a certain sum over prime pairs, a better upper bound than what can be obtained using Selberg's sieve. Even though this improvement is small, a modification of the arguments Pintz and Banks, Freiberg, and Maynard shows that this is sufficient. In addition, we show that there exists a constant $C$ such that for all $T \geq 0$ we have $\mathbb{L} \cap [T,T+C] \neq \emptyset,$ that is, gaps between limit points are bounded by an absolute constant.
\section{Introduction and main results} The Prime Number Theorem tells us that the gap $p_{n+1}-p_n$ between consecutive primes is asymptotically $\log p_n$ on average ($p_n$ denotes the $n$th prime). It is therefore reasonable to consider the distribution of the normalized prime gaps $(p_{n+1}-p_n)/\log p_n;$ by heuristics given by Cram\'er's model we expect that for all $b>a \geq0$ \begin{align} \label{heur} \frac{1}{N} \bigg\{ n \leq N: \, (p_{n+1}-p_n)/\log p_n \in [a,b] \bigg\} \sim \int_a^b e^{-u} \, du, \quad \quad N \to \infty. \end{align} That is, we expect the sequence of normalized prime gaps to satisfy a Poisson distribution (cf. Soundararajan's account \cite{Sound} for details). Gallagher \cite{Gal} has shown this to be true assuming a sufficiently uniform version of the Hardy-Littlewood conjecture. To approach (\ref{heur}), consider the following conjecture of Erd\"os \cite{Erdos}: if $\mathbb{L}$ denotes the set limit points of the sequence $\{(p_{n+1}-p_n)/\log p_n\}_{n=1}^\infty,$ then $\mathbb{L} = [0,\infty].$ By the 1931 result of Westzynthius \cite{West} we know that $\infty \in \mathbb{L},$ and from the seminal work of Goldston, Pintz and Y\i ld\i r\i m \cite{GPY} it follows that $0 \in \mathbb{L}.$ Besides $0$ and $\infty$ no other real number is known to be in $\mathbb{L}$. It is known that $\mathbb{L}$ has a positive Lebesque measure (Erd\"os \cite{Erdos} and Ricci \cite{Ricci}). Goldston and Ledoan \cite{GL} extended the method of Erd\"os to show that intervals of certain specific form, e.g. $[1/8,2]$, contain limit points. In addition, Pintz \cite{Pintz2} has shown that there is an ineffective constant $c$ such that $[0,c] \subseteq \mathbb{L}$ (by applying the ground-breaking work of Zhang \cite{Zhang} on bounded gaps between primes). Note that $\mathbb{L}$ is Lebesque-measurable since it is a closed set. Hildebrand and Maier \cite{HM} showed that there exists a positive constant $c$ such that the Lebesque measure of $\mathbb{L} \cap [0,T] $ is at least $ cT$ for all sufficiently large $T$. Following the breakthrough of Maynard \cite{May} on bounded gaps between primes, it was proved by Banks, Freiberg and Maynard \cite{BFM} that this holds with $c=1/8 -o(1),$ that is, asymptotically at least 1/8 of positive real numbers are limit points. Pintz \cite{Pintz} improved this to $c=1/4 -o(1)$ by modifying the argument of \cite{BFM}; this was shown by Pintz for more general normalizations also. This was then extended to more general and especially larger normalizing factors than $\log p_n$ by Freiberg and Baker \cite{BF}, by combining the arguments with the work of Ford, Green, Konyagin, Maynard, and Tao \cite{Ford} on long prime gaps. For clarity we only consider the set of limit points $\mathbb{L}$ with the logarithmic normalization as defined above. Our main results are deduced from the following \begin{theorem} \label{maint} Let $\beta_1 \leq \beta_2 \leq \beta_3 \leq \beta_4$ be any real numbers. Then \begin{align*} \mathbb{L} \cap \{\beta_j -\beta_i: \, \, 1 \leq i < j \leq 4\} \neq \emptyset. \end{align*} \end{theorem} The proof of this will be given in Section \ref{mainsec}. We note that \cite[Theorem 1.1]{BFM} gives this for nine real numbers in place of four, and \cite[Theorem 1]{Pintz} is the same but for five real numbers. Using the same argument as in the proof of \cite[Corollary 1.2]{BFM} this implies that the Lebesque measure of $\mathbb{L} \cap [0,T]$ is $\geq (1/3-o(1))T$ as $T \to \infty$, where the $o(1)$ is ineffective. Using a more elaborate construction based on similar ideas we will show below \begin{cor} \label{cor1} For all $T > 0$ we have \begin{align*} \mu (\mathbb{L} \cap [0,T]) \geq T/3, \end{align*} where $\mu$ denotes the Lebesque measure on $\mathbb{R}$. \end{cor} Another way to approach the conjecture that $\mathbb{L} = [0,\infty]$ would be to show that for any given positive real $x$ we can find a limit point close to $x$; using Theorem \ref{maint}, we will show below that gaps between limit points are bounded by an absolute (ineffective) constant (note that this actually follows already from \cite[Theorem 1.1]{BFM}, as is evident from the proof): \begin{cor} \label{cor2} There exists a constant $C \geq 0$ such that for all $T\geq 0$ we have \begin{align*} \mathbb{L} \cap [T,T+C] \neq \emptyset. \end{align*} \end{cor} In the language of combinatorics, a set $A \subseteq [0,\infty]$ is called syndetic if there is a constant $C$ such that every interval of length $C$ intersects with $A$ (cf. \cite{BFW}, for example). Thus, Corollary \ref{cor2} can be rephrased as saying that the set of limit points $\mathbb{L}$ is syndetic. \begin{remark} By similar ideas as in the work of Baker and Freiberg \cite{BF}, one can extend our results to other normalizations of prime gaps, replacing $\log p_n$ by a function which can grow somewhat quicker than the logarithm (cf. \cite[Theorem 6.2]{BF} for what normalizations are allowed). We have restricted our attention to the logarithmic normalization to avoid having to define cumbersome notation, with the hope that this makes the article more accessible. \end{remark} \subsection{Proof of Corollary \ref{cor1}} Corollary \ref{cor1} follows from combining Theorem \ref{maint} with the following general proposition: \begin{prop} Let $k \geq 2$ and let $\mathbb{B} \subseteq [0,\infty)$ be any Lebesque-measurable set satisfying the following property: for any real numbers $\beta_1 \leq \beta_2 \leq \cdots \leq \beta_k$ we have \begin{align*} \mathbb{B} \cap \{\beta_j -\beta_i: \, \, 1 \leq i < j \leq k\} \neq \emptyset. \end{align*} Then for any $T >0$ we have \begin{align*} \mu (\mathbb{B} \cap [0,T]) \geq T/(k-1). \end{align*} \end{prop} \begin{proof} For any $m \geq 2$ and for any real numbers $\beta_1,\dots,\beta_m$, define the set of differences \begin{align*} \Delta(\beta_1,\dots,\beta_m) := \{\beta_j -\beta_i: \, \, 1 \leq i < j \leq m\}. \end{align*} For any $\epsilon > 0,$ let us inductively define increasing sequences of real numbers $r_j$ and $s_j$ as follows: set $r_0=s_0=0,$ and for $j >0$, having defined $r_0,\dots, r_{j-1}$ and $s_0,\dots,s_{j-1},$ let \begin{align*} S_{j}=S_j(r_0,\dots,r_{j-1}) := \{ s > r_{j-1}: \, \Delta(r_0,r_1,\dots, r_{j-1}, s) \cap \mathbb{B} = \emptyset \}, \quad \quad s_{j} := \inf S_j, \end{align*} and pick any $r_j \in S_j$ with $r_j \in [s_j,s_j+\epsilon).$ Then by the assumption on $\mathbb{B}$ for some $\ell \leq k-1$ we have $s_\ell = \infty$ (i.e. $S_\ell = \emptyset$), and we stop there and set $r_\ell = \infty$. We now note the following property which holds for all $j \in \{0,1,\dots,\ell-1\}:$ since $s_{j+1}$ is the infimum of $S_{j+1}$, for any $t \in [r_j,s_{j+1})$ we have $\Delta(r_0,r_1,\dots, r_{j}, t) \cap \mathbb{B} \neq \emptyset$. Since $\Delta(r_0,r_1,\dots, r_{j}) \cap \mathbb{B} = \emptyset,$ this implies that \begin{align*} [r_j,s_{j+1}) \subseteq \bigcup_{i=0}^j (\mathbb{B}+r_i). \end{align*} Hence, for all $t \in (r_j,s_{j+1}]$ we have by sub-additivity \begin{align} \label{tin} \mu([r_j,t)) =\mu \bigg([r_j,t)\cap \bigcup_{i=0}^j (\mathbb{B}+r_i)\bigg) \leq \sum_{i=0}^j \mu\left( [r_j,t)\cap (\mathbb{B}+r_i)\right). \end{align} Let $T > 0.$ Then there is some $\lambda \leq \ell -1$ with $T \in (r_\lambda, r_{\lambda+1}]$ (since $r_\ell = \infty$). Denote $\tilde{T}= \min \{T,s_{\lambda+1}\}$. Then, by using $r_j < s_j+\epsilon$, we have \begin{align*} \mu([0,T)) = \sum_{j=0}^{\lambda-1}\mu([r_j,r_{j+1})) + \mu([r_\lambda,T)) \leq (\lambda+1)\epsilon + \sum_{j=0}^{\lambda-1}\mu([r_j,s_{j+1})) + \mu([r_\lambda,\tilde{T})) . \end{align*} By using (\ref{tin}) for all of the summands we get \begin{align*} T=\mu([0,T)) & \leq (\lambda+1)\epsilon + \sum_{j=0}^{\lambda-1}\sum_{i=0}^j \mu( [r_j,s_{j+1})\cap (\mathbb{B}+r_i)) + \sum_{i=0}^\lambda \mu( [r_\lambda,\tilde{T})\cap (\mathbb{B}+r_i)) \\ &\leq (\lambda+1)\epsilon + \sum_{j=0}^{\lambda-1}\sum_{i=0}^j \mu( [r_j,r_{j+1})\cap (\mathbb{B}+r_i)) + \sum_{i=0}^\lambda \mu( [r_\lambda,T)\cap (\mathbb{B}+r_i)) \\ &= (\lambda+1)\epsilon + \sum_{i=0}^\lambda \bigg( \sum_{j=i}^{\lambda-1} \mu( [r_j,r_{j+1})\cap (\mathbb{B}+r_i)) + \mu( [r_\lambda,T)\cap (\mathbb{B}+r_i)) \bigg) \\ &= (\lambda+1)\epsilon + \sum_{i=0}^\lambda \mu( [r_i,T)\cap (\mathbb{B}+r_i)) \\ &= (\lambda+1)\epsilon + \sum_{i=0}^\lambda \mu( [0,T-r_i)\cap \mathbb{B}) \leq (\lambda+1)\epsilon +(\lambda+1) \mu( [0,T)\cap \mathbb{B}). \end{align*} Hence, for any $T > 0$ we have $ \mu( [0,T)\cap \mathbb{B}) \geq T/(\lambda+1) - \epsilon \geq T/(k-1) - \epsilon$ by using $\lambda+1 \leq \ell \leq k-1$. Since $\epsilon>0$ can be made arbitrarily small, we have $\mu( [0,T)\cap \mathbb{B}) \geq T/(k-1).$ \end{proof} \subsection{Proof of Corollary \ref{cor2}} Corollary \ref{cor2} follows from Theorem \ref{maint} using the following general proposition. This is also proved in the work of Bergelson, Furstenberg, and Weiss \cite[Section 1, second paragraph]{BFW} but we give our own different proof of this. \begin{prop} \label{gen2} Let $\mathbb{B} \subseteq [0,\infty)$ be any set satisfying the following property: there exists an integer $k \geq 2$ such that for any real numbers $\beta_1 \leq \beta_2 \leq \cdots \leq \beta_k$ we have \begin{align*} \mathbb{B} \cap \{\beta_j -\beta_i: \, \, 1 \leq i < j \leq k\} \neq \emptyset. \end{align*} Then there exists a constant $C\geq 0$ (ineffective) such that for all $T\geq 0$ we have \begin{align*} \mathbb{B} \cap [T,T+C] \neq \emptyset. \end{align*} \end{prop} To prove this proposition we first prove the following weaker version: \begin{lemma} \label{boundlemma} Let $\mathbb{B} \subseteq [0,\infty)$ satisfy the assumptions of Proposition \ref{gen2}. Let $w$ be any given function such that $w(T) \to \infty$ as $T \to \infty,$ and $w(T) >0$ for $T>0.$ Then there exists a constant $C,$ depending only on the choice of $w,$ such that for all $T > C$ we have \begin{align*} \mathbb{B} \cap [T-w(T),T] \neq \emptyset. \end{align*} \end{lemma} \begin{proof} Define \begin{align*} \mathcal{A}:=\{A > 0: \, \, \mathbb{B} \cap [A-w(A),A] = \emptyset \}. \end{align*} Suppose that the conclusion of the lemma is not true, so that $\mathcal{A}$ is unbounded. Then we can choose $A_1, A_2, \dots, A_{k-1} \in \mathcal{A}$ such that \begin{align} & A_1 < A_2 < \cdots < A_{k-1}, \\ & w(A_1) < w(A_2) < \cdots < w(A_{k-1}) \quad \text{and} \label{2}\\ & A_j < w(A_{j+1}) \quad \text{for} \quad j=1,2, \dots, k-1. \label{3} \end{align} Define $k$ real numbers by $\beta_0:=0$ and $\beta_j:=A_{j}$ if $j=1,2, \dots k-1.$ Then by (\ref{2}) and (\ref{3}) we have $\beta_i < w(A_j)$ if $1\leq i<j \leq k-1.$ Hence, \begin{align*} \{\beta_j-\beta_i: \, \, 0 \leq i < j \leq k-1\} \subseteq \bigcup_{j=1}^{k-1} [A_j-w(A_j),A_j]. \end{align*} But by assumption we also have \begin{align*} \mathbb{B} \cap \{\beta_j-\beta_i: \, \, 0 \leq i < j \leq k-1\} \neq \emptyset, \end{align*} which gives a contradiction. \end{proof} \emph{Proof of Proposition \ref{gen2}.} Suppose that no such constant $C$ exists. This implies that for every $C$ there are arbitrarily large $A$ such that $\mathbb{B} \cap [A-C,A] = \emptyset$. Hence, it is possible to find a strictly increasing sequence of positive real numbers $A_n \to \infty$ as $n \to \infty,$ such that \begin{align*} \mathbb{B} \cap [A_n -n, A_n ] = \emptyset \end{align*} for all $n \geq 1.$ Fix any such sequence $A_n$ and define a step function $w$ by setting (with $A_0=0$) \begin{align*} w(A) = n \quad \text{for} \quad A \in (A_{n-1},A_n] \quad \text{for any} \, \, n \geq 1. \end{align*} Then $w(A) \to \infty$ as $A \to \infty$, and there are arbitrarily large $A$ such that $\mathbb{B} \cap [A-w(A),A] = \emptyset,$ namely $A = A_n$ for any $n\geq 1$. This is a contradiction with Lemma \ref{boundlemma}. \qed \subsection{Outline of the proof of Theorem \ref{maint}} The proof of Theorem \ref{maint} will occupy us for the remainder of the article; our proof builds heavily on the earlier work of Banks, Freiberg and Maynard \cite{BFM}, and the refinement of Pintz \cite{Pintz} to their argument. We now give an informal outline of the basic ideas and indicate our modifications to them. A finite set of integers $\mathcal{H}$ is said to be admissible if for every prime $p$ the set $\mathcal{H}$ avoids at least one residue class modulo $p$, that is, if \begin{align*} \bigg| \bigg\{ n \,\, (p): \, \prod_{h \in \mathcal{H}} (n+h) \equiv 0 \quad (p) \bigg\} \bigg| < p. \end{align*} Let $N$ be large and suppose we are given an admissible $K$-tuple $\mathcal{H} = \{h_1,\dots, h_K\}$ with $h_j \leq C \log N$ for all $j$ for some large $C.$ Then by a variant of the Erd\"os-Rankin construction (cf. \cite[Section 5]{BFM}), one can show that there is an integer $b$ and a smooth modulus $W < N^\epsilon$ such that for any $N < n \leq 2N$ with $n \equiv b \,(W),$ if there are prime numbers in the interval $[n,n+C\log N]$, then they must belong to the set $n+\mathcal{H}$. By using the Maynard-Tao sieve, we can show that there exists $N < n \leq 2N$ with $n \equiv b \,(W)$ such that $n+\mathcal{H}$ contains prime numbers once $K=|\mathcal{H}|$ is large enough. Furthermore, suppose that we have a partition $\mathcal{H} = \mathcal{H}_1 \cup \mathcal{H}_2 \cup \cdots \cup \mathcal{H}_{M}$ into $M$ sets of equal size. Then we can show that there exists a constant $A$ such that for any integer $a \geq 1,$ if $M = \lceil Aa \rceil +1,$ then for at least $a+1$ distinct indices $j$ the set $n+\mathcal{H}_j$ contains a prime number. That is, the prime numbers that we find by the Maynard-Tao sieve are not too much concentrated on any particular set $n+\mathcal{H}_j.$ The constant $A$ is determined by how well we can control sums over prime pairs; more precisely, it is the best constant so that for all distinct $h,h' \in \mathcal{H}$ we can show the bound \begin{align} \label{pairsout} \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W)}} 1_{\mathbb{P}}(n+h)1_{\mathbb{P}}(n+h')\bigg( \sum_{\substack{d_1, \dots, d_K \\ d_i | n+h_i}} \lambda_{d_1, \dots, d_K} \bigg)^2 \leq (A+o(1)) X, \end{align} where $X$ is the expected main term and $\lambda_{d_1, \dots, d_K}$ are sieve weights of Maynard-Tao type supported on $d_1\cdots d_K \leq N^\delta$ for some small $\delta >0$. In \cite[Section 4]{BFM}, Selberg's upper bound sieve is used to show this for $A=4.$ We improve this to $A=3.99$ by using Chen's sieve \cite{Chen}, \cite{Pan} (cf. Proposition \ref{pairs} below). The reason why this small improvement is sufficient is as follows: we choose $a=100$ so that $\lceil 3.99a \rceil +1 = 4a,$ and partition our tuple \begin{align*} \mathcal{H} = \mathcal{H}_1 \cup \mathcal{H}_2 \cup \mathcal{H}_3 \cup \mathcal{H}_4, \quad \quad \quad \quad \mathcal{H}_i = \bigcup_{j=1}^a \mathcal{H}_{ij}, \quad i \in \{1,2,3,4\}. \end{align*} Then we find $N < n \leq 2N$ with $n \equiv b \,(W)$ such that for at least $a+1$ distinct $(i,j)$ the set $n+\mathcal{H}_{ij}$ contains a prime number. Thus, by the pigeon-hole principle we must have at least two indices $i \neq i'$ such that both $n+\mathcal{H}_i, n+\mathcal{H}_{i'}$ contain primes. By the restriction $n\equiv b \, (W)$ given by the modified Erd\"os-Rankin construction, we then know that there are two consecutive primes, one in $n+\mathcal{H}_i$ and one in $n+\mathcal{H}_{i'},$ for some $i \neq i'.$ For $\beta_1 \leq \beta_2 \leq \beta_3 \leq \beta_4$ as in Theorem \ref{maint}, it is then enough to choose $\mathcal{H}_i$ so that for all $h \in \mathcal{H}_i$ we have $h = (\beta_i + o(1)) \log N.$ From this argument we see that the exact numerical value of $A=3.99$ is not important, what matters is that $A$ is strictly less than $4.$ To show the bound (\ref{pairsout}) with $A=3.99,$ we require a Bombieri-Vinogradov type equidistribution result for primes, where the moduli run over multiples of $W < N^\epsilon.$ The possibility of exceptional zeros of $L$-functions causes some technical problems, but the result \cite[Theorem 4.2]{BFM} turns out to be sufficient. Since we are using Chen's sieve, we also need to extend this to almost-primes; this is done in Section \ref{bvsec}. In Section \ref{chensec} we apply Chen's sieve to obtain the required bound (\ref{pairsout}) for prime pairs (Proposition \ref{pairs}). We then state and prove in Section \ref{maysec} the precise version of the Maynard-Tao sieve which we will use (Proposition \ref{strong}), and in Section \ref{mainsec} we prove our main result Theorem \ref{maint}. \begin{remark} By the same argument, if we could show the bound (\ref{pairsout}) with any constant $A <3$ in place of $3.99,$ we would obtain Theorem \ref{maint} with sequence of four real numbers replaced by three. This in turn would give that $\mu(\mathbb{L} \cap [0,T]) \geq T/2.$ Similarly, if we had (\ref{pairsout}) with any constant $A < 2$ in place of $3.99,$ we could show that $\mathbb{L} =[0,\infty],$ which is the conjecture of Erd\"os. However, by the parity principle this should be just as hard as obtaining a lower bound for such a sum over prime pairs, which would immediately imply $\mathbb{L} =[0,\infty]$ (cf. \cite[Chapter 16]{FI} for a quantitative version due to Bombieri of the parity principle). \end{remark} \subsection{Notations} We use the following asymptotic notations: for positive functions $f,g,$ we write $f \ll g$ or $f= \mathcal{O}(g)$ if there is a constant $C$ such that $f \leq C g.$ $f \asymp g$ means $g \ll f \ll g.$ The constant may depend on some parameter, which is indicated in the subscript (e.g. $\ll_{\epsilon}$). We write $f=o(g)$ if $f/g \to 0$ for large values of the variable. In general, $C$ stands for some large constant, which may not be the same from place to place. For variables we write $n \sim N$ meaning $N<n \leq eN$ (an $e$-adic interval), and $n \asymp N$ meaning $N/C < n < CN$ (a $C^2$-adic interval) for some constant $C>1$ which is large enough depending on the situation. If not otherwise stated the symbols $p,q,r$ denote primes and $d,k,\ell,m,n$ denote integers. For a statement $E$ we denote by $1_E$ the characteristic function of that statement. For a set $A$ we use $1_A$ to denote the characteristic function of $A,$ so that $1_\mathbb{P}$ will denote the characteristic function of primes. We define $P(w):= \prod_{p\leq w} p,$ and for any integer $d$ we write $P^-(d):= \min \{p: \, p | d\},$ $P^+(d):= \max \{p: \, p | d\}.$ The $k$-fold divisor function is denoted by $\tau_k(d).$ We denote the ceiling function by $\lceil \cdot \rceil$, that is, $\lceil x \rceil $ is the smallest integer $n \geq x.$ Overall we use similar notations as in \cite{BFM}, especially when we use the Maynard-Tao sieve; these are recalled in the text as needed. \subsection*{Acknowledgements} I am grateful to my supervisor Kaisa Matom\"aki for support and comments. I also express my gratitude to Emmanuel Kowalski for helpful comments as well as for hospitality during my visit to ETH Z\"urich. I wish to thank James Maynard for bringing the article \cite{BFM} to my attention. I also wish to thank Pavel Zorin-Kranich for useful suggestions and the anonymous referee for comments. During the work the author was supported by a grant from the Magnus Ehrnrooth Foundation. \section{Modified Bombieri-Vinogradov Theorem} \label{bvsec} As was outlined above, we need to show an upper bound of type (\ref{pairsout}) for prime pairs, where the modulus $W$ can be as large as $N^\epsilon.$ For this purpose we require a modified version of the Bombieri-Vinogradov Theorem. Before stating this we need the following lemma on exceptional zeros of Dirichlet $L$-functions (this is \cite[Lemma 4.1]{BFM}): \begin{lemma}Let $T \geq 3$ and $P \geq T^{1/\log_2 T}.$ For a sufficiently small constant $c > 0,$ there is at most one modulus $q \leq T$ with $P^+(q) \leq P$ and one primitive character $\chi$ modulo $q$ such that the function $L(s,\chi)$ has a zero in the region \begin{align*} \Re(s) \geq 1-\frac{c}{\log P}, \quad \quad |\Im(s)| \leq \exp \bigg( \log P / \sqrt{\log T}\bigg). \end{align*} If such a character $\chi$ mod $q$ exists, then it is real, $L(s,\chi)$ has at most one zero in the above region, which is then real and simple, and \begin{align*} P^+(q) \gg \log q \gg \log_2 T. \end{align*} \end{lemma} Fix a constant $c>0$ for which the above lemma holds. Similarly as in \cite{BFM}, if such an exceptional modulus $q \leq T$ exists with $P = T^{1/\log_2 T}$, we define \begin{align} \label{Z} Z_T = P^+(q), \end{align} and we set $Z_T =1$ if no such modulus exists. We then have the following variant of the Bombieri-Vinogradov Theorem (this is \cite[Theorem 4.2]{BFM}): \begin{prop} \label{bv} \emph{\textbf{(Modified Bombieri-Vinogradov).}} Let $N > 2$ and fix constants $C > 0,$ $\epsilon > 0,$ and $\delta > 0$. Let $q_0 < N^{\epsilon}$ be a square-free integer with $P^+(q_0) < N^{\epsilon/ \log_2 N}.$ Then for $\epsilon$ small enough we have \begin{align*} \sum_{\substack{q \leq N^{1/2-\delta} \\ q_0 | q \\ (q, Z_{N^{2\epsilon}})=1}} \max_{(a,q) = 1} \bigg | \sum_{\substack{ n \leq N \\ n \equiv a \, (q)}} \Lambda (n) \, - \frac{1}{\phi(q)} \sum_{\substack{ n \leq N}} \Lambda (n)\bigg | \, \ll_{\delta, C} \, \frac{N}{\phi(q_0) \log^C N}. \end{align*} \end{prop} From the proof of \cite[Theorem 4.2]{BFM} we obtain the following lemma, which we require for the proof of Proposition \ref{bv2} below: \begin{lemma} \label{char} With the same notations and assumptions as in Proposition \ref{bv} we have \begin{align} \sup_{\substack{A,B \\ AB \leq N^{1/2-\delta} \\ A \leq q_0}} \sum_{\substack{A < a \leq 2A \\ a | q_0}} \sum_{\substack{B \leq b \leq 2B \\ (b, q_0 Z_{N^{2\epsilon}})=1}} \frac{1}{\phi (b)} \sideset{}{'}\sum_{\chi \, \, (ab)} \bigg | \sum_{n \leq N} \Lambda(n) \chi(n) \bigg | \, \ll_C \frac{N}{\log^C N}, \end{align} where $\Sigma '$ denotes the sum over primitive characters modulo $ab.$ \end{lemma} Since we plan to apply Chen's sieve, we also require a similar equidistribution result for almost-primes. To prove such a result we require the large sieve for multiplicative characters, which follows from Theorem 9.10 of \cite{FI}: \begin{lemma} \label{large}\emph{\textbf{(Large sieve for multiplicative characters).}} For any sequence $c_n$ of complex numbers and for any $M,N \geq 1$ we have \begin{align*} \sum_{q \leq Q} \frac{q}{\phi(q)} \sideset{}{'} \sum_{\chi \, \, (q)} \bigg | \sum_{M< n \leq M+N} c_n \chi(n)\bigg |^2 \, \leq \, (Q^2+N) \sum_n |c_n|^2. \end{align*} \end{lemma} To state the equidistribution result for almost-primes, we need to set up some notation: fix $0 < \alpha < 1/2,$ and let $N^\alpha \ll A_1 \ll N^{1-\alpha}$, for sufficiently large $N.$ Define \begin{align} \label{p0} \Lambda_0(n) := (f \ast g)(n), \end{align} where $f(m) = 1_{\mathbb{P}}(m)(\log m)1_{m \leq A_1},$ and $g$ is any function such that $|g(n)| \, \ll 1,$ and $g(n) \neq 0$ only if $P^-(n) \geq N^{\alpha}$ and $n \asymp N/A_1$. Note that then $\Lambda_0(n)$ is supported on almost-primes $n \ll N$. We then have that Proposition \ref{bv} holds also with $\Lambda(n)$ replaced by $\Lambda_0 (n)$: \begin{prop} \label{bv2}\emph{\textbf{(Modified Bombieri-Vinogradov for almost-primes).}} Let $N > 2$ and fix constants $C > 0,$ $\epsilon > 0,$ and $\delta > 0$. Let $q_0 < N^{\epsilon}$ be a square-free integer with $P^+(q_0) < N^{\epsilon/ \log_2 N}.$ Let $\Lambda_0(n)$ be as in (\ref{p0}). Then for all small enough $\epsilon$ we have \begin{align*} \sum_{\substack{q \leq N^{1/2-\delta} \\ q_0 | q \\ (q, Z_{N^{2\epsilon}})=1}} \max_{(a,q) = 1} \bigg | \sum_{\substack{ n \equiv a \, (q)}} \Lambda_0 (n) \, - \frac{1}{\phi(q)} \sum_{\substack{ n }} \Lambda_0 (n)\bigg| \, \ll_{\delta, C} \, \frac{N}{\phi(q_0) \log^C N}. \end{align*} \end{prop} \begin{proof} The basic idea is to use the large sieve inequality for large moduli and for small moduli use Lemma \ref{char}. For convenience we set $D:=N^{1/2-\delta}.$ Using the expansion \begin{align*} \sum_{\substack{ n \equiv a \, (q)}} \Lambda_0 (n) \, - \frac{1}{\phi(q)} \sum_{\substack{ n }} \Lambda_0 (n) = \frac{1}{\phi(q)}\sum_{\substack{\chi \, \, (q) \\ \chi \neq \chi_0}} \overline{\chi}(a) \sum_{n } \Lambda_0 (n) \chi (n), \end{align*} we are reduced to obtaining the bound \begin{align} \label{ave} \sum_{\substack{q \leq D \\ q_0 | q \\ (q, Z_{N^{2\epsilon}})=1}} \frac{1}{\phi(q)} \sum_{\substack{\chi \, \, (q) \\ \chi \neq \chi_0}} \bigg | \sum_{n } \Lambda_0 (n) \chi (n) \bigg| \, \ll_{\delta, C} \, \frac{N}{\phi(q_0) \log^C N}. \end{align} We then replace the character $\chi$ modulo $q$ by the primitive character $\chi'$ modulo $q'$ which induces $\chi$; we have \begin{align*} \chi(n) = \chi'(n) - \chi'(n)1_{(n,q/q') > 1}. \end{align*} Hence, the left-hand side of (\ref{ave}) is bounded by \begin{align} \label{ave2} \sum_{\substack{q \leq D \\ q_0 | q \\ (q, Z_{N^{2\epsilon}})=1}} \frac{1}{\phi(q)} \sum_{\substack{\chi \, \, (q) \\ \chi \neq \chi_0}} \bigg | \sum_{n } \Lambda_0 (n) \chi' (n) \bigg| + \sum_{\substack{q \leq D \\ q_0 | q \\ (q, Z_{N^{2\epsilon}})=1}} \frac{1}{\phi(q)} \sum_{\substack{\chi \, \, (q) \\ \chi \neq \chi_0}} \bigg | \sum_{m,n } f(m) g(n) \chi' (mn) 1_{(mn,q/q') > 1} \bigg|. \end{align} We have \begin{align*} 1_{(mn,q/q') > 1} = 1_{(n,q/q') > 1} + 1_{(m,q/q') > 1} -1_{(m,q/q') > 1}1_{(n,q/q') > 1}. \end{align*} Define $h_1(n,d) := 1$ and $h_2(n,d) := 1_{(n,d)>1}.$ Then (\ref{ave2}) is bounded by \begin{align} \label{ave3} \sum_{i,j=1}^2 \sum_{\substack{q \leq D \\ q_0 | q \\ (q, Z_{N^{2\epsilon}})=1}} \frac{1}{\phi(q)} \sum_{\substack{\chi \, \, (q) \\ \chi \neq \chi_0}} \bigg | \sum_{m,n } f(m)h_i(m,q/q')\chi'(m) g(n)h_j(n,q/q') \chi' (n) \bigg|. \end{align} If $i =2$ or $j=2,$ we remove the additional conditions for $q$, write $q=dq'$, and bound the sum by \begin{align*} \sum_{\substack{q \leq D}} \frac{1}{\phi(q)} \sum_{\substack{\chi \, \, (q) \\ \chi \neq \chi_0}} &\bigg | \sum_{m,n } f(m)h_i(m,q/q')\chi'(m) g(n)h_j(n,q/q') \chi' (n) \bigg| \\ & \ll \sum_{d \leq D} \frac{1}{\phi(d)} \sum_{\substack{q' \leq D}} \frac{1}{\phi(q')} \sideset{}{'}\sum_{\substack{\chi \, \, (q')}} \bigg | \sum_{m,n } f(m)h_i(m,d)\chi(m) g(n)h_j(n,d) \chi (n) \bigg| \\ & \ll (\log N) \sum_{d \leq D} \frac{1}{\phi(d)} \sup_{E \leq D} \frac{1}{E} \bigg( \sum_{\substack{q' \sim E}} \frac{q'}{\phi(q')} \sideset{}{'} \sum_{\substack{\chi \, \, (q')}} \bigg | \sum_{m } f(m)h_i(m,d)\chi(m) \bigg|^2 \bigg)^{1/2} \\ & \hspace{80pt} \cdot\bigg( \sum_{\substack{q' \sim E}} \frac{q'}{\phi(q')} \sideset{}{'}\sum_{\substack{\chi \, \, (q')}} \bigg | \sum_{n } g(n)h_j(n,d) \chi (n) \bigg|^2 \bigg)^{1/2}, \end{align*} where in the last bound we have split the sum over $q'$ dyadically and applied Cauchy-Schwarz. By Lemma \ref{large} and by the assumptions on $f$ and $g,$ the last expression is bounded by \begin{align} \label{j2} (\log N)\sum_{d \leq D} \frac{1}{\phi(d)} \sup_{E \leq D} \frac{1}{E} &\bigg( \bigg(E^2 + A_1 \bigg) \sum_{m } |f(m)h_i(m,d)|^2 \bigg)^{1/2} \\ \nonumber & \hspace{20pt} \cdot\ \bigg( \bigg(E^2 + N/A_1 \bigg) \sum_{n } |g(n)h_j(n,d)|^2 \bigg)^{1/2} \end{align} Suppose at first that $j=2$ so that $h_j(n,d)= 1_{(n,d)>1}.$ Since $g(n)$ is supported on $P^-(n) \geq N^\alpha,$ this means that $(n,d) \geq N^\alpha.$ We obtain that (\ref{j2}) is bounded by \begin{align*} (\log N)&\sum_{d \leq D} \frac{1}{\phi(d)} \sup_{E \leq D} \frac{1}{E} \bigg( \bigg(E^2 + A_1 \bigg) \sum_{m } |f(m)|^2 \bigg)^{1/2} \\ & \hspace{150pt} \cdot\ \bigg( \bigg(E^2 + N/A_1 \bigg) \sum_{\substack{k| d \\ N^{\alpha} \leq k \leq D}} \sum_{n \asymp N/(A_1k)} |g(kn)|^2 \bigg)^{1/2} \\ & \ll (\log^2 N) \sum_{d \leq D} \frac{\tau(d)^{1/2}}{\phi(d)} \sup_{E \leq D} \frac{1}{E} \bigg( \bigg(E^2 + A_1 \bigg) A_1 \bigg)^{1/2} \bigg( \bigg(E^2 + N/A_1 \bigg) N^{1-\alpha}/A_1 \bigg)^{1/2} \\ & \leq (\log^4 N) \sup_{E \leq D} ( E N^{(1-\alpha)/2} + N^{1-\alpha/2}/A_1 + A_1 + N^{1-\alpha/2} /E) \ll N^{1- \alpha/3}, \end{align*} which is sufficient. For $i=2, j=1,$ since $f(m)=1_{\mathbb{P}}(m)(\log m)1_{n \leq A_1},$ we have that if $(m,d)>1,$ then $m$ is a prime dividing $d$. Hence, by a similar argument as above we get a bound $\ll N^{1- \alpha/3}$. For $i=j=1$ we have to estimate \begin{align} \label{j1} \sum_{\substack{q \leq D \\ q_0 | q \\ (q, Z_{N^{2\epsilon}})=1}} \frac{1}{\phi(q)} \sum_{\substack{\chi \, \, (q) \\ \chi \neq \chi_0}} \bigg | \sum_{m,n } f(m)\chi'(m) g(n) \chi' (n) \bigg|. \end{align} We begin by extracting a factor of $1/\phi(q_0)$ similarly as in the proof of \cite[Theorem 4.2]{BFM}: if $q'$ denotes the modulus of $\chi'$, then (\ref{j1}) is bounded by (writing $q'=ab,$ where $a | q_0$ and $(b,q_0)=1$; recall that $q_0$ is square-free) \begin{align*} \sum_{\substack{q' \leq D \\ (q', Z_{N^{2\epsilon}})=1}} \sideset{}{'}\sum_{\substack{\chi \, \, (q') }} &\bigg | \sum_{m,n } f(m)\chi(m) g(n) \chi(n) \bigg| \sum_{\substack{q \leq D \\ [q',q_0] | q \\ (q, Z_{N^{2\epsilon}})=1}} \frac{1}{\phi(q)} \\ & \ll \frac{\log N}{\phi (q_0)} \sum_{a | q_0} \sum_{\substack{b \leq D/a \\ (b, q_0Z_{N^{2\epsilon}})=1}} \frac{1}{\phi (b)} \sideset{}{'}\sum_{\substack{\chi \, \, (ab) }} \bigg | \sum_{m,n } f(m)\chi(m) g(n) \chi (n) \bigg| \\ & \ll \frac{\log^3 N}{\phi (q_0)} \sup_{\substack{A,B \\ AB \leq D \\ A \leq q_0}} \sum_{\substack{A < a \leq 2A \\ a | q_0}} \sum_{\substack{B \leq b \leq 2B \\ (b, q_0 Z_{N^{2\epsilon}})=1}} \frac{1}{\phi (b)} \sideset{}{'}\sum_{\substack{\chi \, \, (ab) }} \bigg | \sum_{m,n } f(m)\chi(m) g(n) \chi (n) \bigg| \\ \end{align*} Hence, it remains to show that \begin{align*} \sup_{\substack{A,B \\ AB \leq D \\ A \leq q_0}} \sum_{\substack{A < a \leq 2A \\ a | q_0}} \sum_{\substack{B \leq b \leq 2B \\ (b, q_0 Z_{N^{2\epsilon}})=1}} \frac{1}{\phi (b)} \sideset{}{'}\sum_{\substack{\chi \, \, (ab) }} \bigg | \sum_{m,n } f(m)\chi(m) g(n) \chi (n) \bigg| \, \ll_C \frac{N}{\log^C N} \end{align*} For $B \geq N^{\epsilon}$ we have by Cauchy-Schwarz and Lemma \ref{large} \begin{align*} & \sum_{\substack{A < a \leq 2A \\ a | q_0}} \sum_{\substack{B \leq b \leq 2B \\ (b, q_0 Z_{N^{2\epsilon}})=1}} \frac{1}{\phi (b)} \sideset{}{'}\sum_{\substack{\chi \, \, (ab) }} \bigg | \sum_{m,n } f(m)\chi(m) g(n) \chi (n) \bigg| \\ & \ll \frac{1}{B} \bigg( \sum_{q \ll AB} \frac{q}{\phi(q)} \sideset{}{'}\sum_{\substack{\chi \, \, (q) }} \bigg | \sum_{m } f(m)\chi(m) \bigg|^2\bigg)^{1/2} \bigg(\sum_{q \ll AB} \frac{q}{\phi(q)} \sideset{}{'}\sum_{\substack{\chi \, \, (q) }} \bigg | \sum_{n } g(n) \chi (n) \bigg|^2 \bigg)^{1/2} \\ & \ll \frac{\log N}{B} \bigg((AB)^2 A_1 + A_1^2 \bigg)^{1/2} \bigg( (AB)^2 N/A_1 + (N/A_1)^2 \bigg)^{1/2} \\ & \leq (\log N) ( A^2 B N^{1/2} + A N / A_1^{1/2} + A_1^{1/2} A N^{1/2} + N/B ) \ll N^{1-\epsilon}, \end{align*} if $\epsilon$ is small enough in terms of $\delta$ and $\alpha.$ For $B < N^{\epsilon}$ we replace $f(m)$ by $\Lambda(m)1_{m \leq A_1},$ which causes an error term bounded by using a trivial bound \begin{align*} \sum_{\substack{A < a \leq 2A \\ a | q_0}} \sum_{\substack{B \leq b \leq 2B \\ (b, q_0 Z_{N^{2\epsilon}})=1}} \frac{1}{\phi (b)} \sideset{}{'}\sum_{\substack{\chi \, \, (ab) }} &\bigg | \sum_{\substack{p^k \leq A_1 \\ k \geq 2}} \sum_{n } \log(p) \chi(p^k) g(n) \chi (n) \bigg| \\ & \hspace{20pt} \ll (AB)^2 N^{1-\alpha/2} \log N \ll N^{1-\alpha/3} \end{align*} if $\epsilon$ is sufficiently small. We then use Cauchy-Schwarz to get \begin{align} \nonumber \sum_{\substack{A < a \leq 2A \\ a | q_0}} \sum_{\substack{B \leq b \leq 2B \\ (b, q_0 Z_{N^{2\epsilon}})=1}} & \frac{1}{\phi (b)} \sideset{}{'}\sum_{\substack{\chi \, \, (ab) }} \bigg | \sum_{m,n } \Lambda(m)1_{m \leq A_1} \chi(m) g(n) \chi (n) \bigg| \\ \label{smallb} & \ll \bigg( \sum_{\substack{A < a \leq 2A \\ a | q_0}} \sum_{\substack{B \leq b \leq 2B \\ (b, q_0 Z_{N^{2\epsilon}})=1}} \frac{1}{\phi (b)} \sideset{}{'}\sum_{\substack{\chi \, \, (ab) }} \bigg | \sum_{m \leq A_1 } \Lambda (m) \chi(m) \bigg|^2\bigg)^{1/2} \\ \nonumber & \hspace{140pt} \cdot \bigg(\frac{1}{B}\sum_{q \ll AB} \frac{q}{\phi(q)} \sideset{}{'}\sum_{\substack{\chi \, \, (q) }} \bigg | \sum_{n } g(n) \chi (n) \bigg|^2 \bigg)^{1/2} \end{align} Since $AB < N^{2\epsilon} < A_1^{1/2-\delta},$ we may use the bound Lemma \ref{char} with $A_1$ in place of $N$ (decreasing $\epsilon$ also if necessary), which yields \begin{align*} \sum_{\substack{A < a \leq 2A \\ a | q_0}}& \sum_{\substack{B \leq b \leq 2B \\ (b, q_0 Z_{N^{2\epsilon}})=1}} \frac{1}{\phi (b)} \sideset{}{'}\sum_{\substack{\chi \, \, (ab) }} \bigg | \sum_{ m \leq A_1 } \Lambda(m)\chi(m) \bigg|^2 \\ & \ll A_1 \sum_{\substack{A < a \leq 2A \\ a | q_0}} \sum_{\substack{B \leq b \leq 2B \\ (b, q_0 Z_{N^{2\epsilon}})=1}} \frac{1}{\phi (b)} \sideset{}{'}\sum_{\substack{\chi \, \, (ab) }} \bigg | \sum_{m \leq A_1 } \Lambda(m)\chi(m) \bigg| \ll_C \frac{A_1^2}{\log^{2(C+5)} N}. \end{align*} Using Lemma \ref{large} to bound the sum with $g(n) \chi (n)$ in (\ref{smallb}) we get that \begin{align*} \sum_{\substack{A < a \leq 2A \\ a | q_0}} & \sum_{\substack{B \leq b \leq 2B \\ (b, q_0 Z_{N^{2\epsilon}})=1}} \frac{1}{\phi (b)} \sideset{}{'}\sum_{\substack{\chi \, \, (ab) }} \bigg | \sum_{m,n } f(m)\chi(m) g(n) \chi (n) \bigg| \\ & \ll_C \frac{A_1}{\log^{C+5} N} \bigg( A^2 B N/A_1 + (N/A_1)^2/B \bigg)^{1/2} \ll_C \frac{N}{\log^{C+5} N}. \end{align*} \end{proof} \section{Chen's sieve upper bound for prime pairs} \label{chensec} In this section we will apply Chen's sieve to obtain an upper bound for prime pairs, which is 3.99 times the expected main term. As will become apparent in the next section, the exact numerical value of this constant does not matter, only that it is stricly less than four. To state the result, we first need to set up some notation from \cite{BFM}. Let $K>1,$ $N > 3,$ and define the Maynard-Tao sieve weights (recall the definition of $Z_T$ from (\ref{Z})) \begin{align} \label{l1} \lambda_{d_1, \dots, d_K} = \begin{cases} \bigg(\prod_{i=1}^K \mu (d_i)\bigg) \sum_{j=1}^{J} \prod_{\ell=1}^K F_{\ell,j}\bigg( \frac{\log d_\ell}{\log N}\bigg), & \text{if} \, \, (d_1\cdots d_K, Z_{N^{4 \epsilon}}) =1, \\ 0, & \text{otherwise,} \end{cases} \end{align} for some fixed $J$, where $F_{\ell,j}:[0,\infty) \to \mathbb{R}$ are smooth compactly supported functions, not identically zero, satisfying a support condition \begin{align} \label{l2} \sup \bigg \{\sum_{\ell=1}^K t_l: \, \, \prod_{\ell=1}^K F_{\ell,j}(t_\ell) \neq 0 \bigg \} \leq \delta \end{align} for all $j=1,2,\dots,J$ for some small $\delta >0.$ Note that this implies that $\lambda_{d_1,\dots, d_K}$ are supported on $d_1 \cdots d_K \leq N^{\delta}.$ Define \begin{align*} F(t_1, \dots, t_K) := \sum_{j=1}^J \prod_{\ell=1}^K F_{\ell,j}'\bigg( t_\ell \bigg), \end{align*} where $F_{\ell,j}'$ is the derivative of $F_{\ell,j}.$ Set \begin{align} \label{Lint} L_K(F)& := \int_0^\infty \cdots \int_0^\infty \bigg( \int_0^\infty \int_0^\infty F(t_1,\dots t_K) dt_{K-1} dt_K\bigg)^2 dt_1 \cdots dt_{K-2} \\ \nonumber & = \sum_{j,j'=1}^J F_{K-1,j}(0)F_{K-1,j'}(0)F_{K,j}(0)F_{K,j'}(0) \prod_{\ell=1}^{K-2} \int_0^\infty F'_{\ell,j}(t_\ell)F'_{\ell,j'}(t_\ell) dt_\ell. \end{align} We note here that $F_{\ell,j}$ will be chosen so that $F(t_1,\dots,t_K)$ is symmetric with respect to permutations of the variables (cf. \cite{BFM}). Let $Z_{N^{4\epsilon}}$ be as in (\ref{Z}) and define \begin{align*} W := \prod_{\substack{p \leq \epsilon \log N \\ p \nmid Z_{N^{4\epsilon}}}} p, \quad \quad \quad \quad B := \frac{\phi(W)}{W} \log N. \end{align*} Using the above notation, we have that \cite[Lemma 4.6 (iii)]{BFM} holds with the constant $4$ replaced by $3.99:$ \begin{prop} \label{pairs} For all sufficiently large $N$ the following holds: Let $\mathcal{H} = \{h_1, \dots, h_K\} \subseteq [0,N]$ be an admissible $K$-tuple such that \begin{align} \label{smooth} P^+\bigg (\prod_{1 \leq i < j\leq K} (h_j-h_i) \bigg) \leq \epsilon \log N. \end{align} Let $b$ be an integer such that \begin{align*} \bigg( \prod_{j=1}^K (b+h_j) , W\bigg) =1. \end{align*} Then for all distinct $h_j,h_\ell \in \mathcal{H}$ we have \begin{align*} S:=\sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W)}} 1_{\mathbb{P}}(n+h_j)1_{\mathbb{P}}(n+h_\ell)\bigg( \sum_{\substack{d_1, \dots, d_K \\ d_i | n+h_i}} \lambda_{d_1, \dots, d_K} \bigg)^2 \leq (3.99 + \mathcal{O}(\delta)) \frac{N}{W} B^{-K}L_K(F). \end{align*} \end{prop} The proof in \cite[Lemma 4.6 (iii)]{BFM} uses Selberg's sieve combined with the Modified Bombieri-Vinogradov Theorem. Our improvement comes from using Chen's sieve instead of Selberg's sieve. Similarly as in \cite[Lemma 4.6 (iii)]{BFM}, we first note that we may replace \begin{align*} \bigg( \sum_{\substack{d_1, \dots d_K \\ d_i | n+h_i}} \lambda_{d_1, \dots, d_K} \bigg)^2 \quad \text{by} \quad \nu_{\mathcal{H},j,\ell} (n) := \bigg( \sum_{\substack{d_1, \dots, d_K \\ d_i | n+h_i \\ d_j=d_\ell=1}} \lambda_{d_1, \dots, d_K} \bigg)^2 1_{((n+h_j)(n+h_\ell),Z_{N^{4\epsilon}})=1} \end{align*} in the sum $S.$ We then require the following weighted sieve inequality of Chen type (this is essentially Lemma 4.1 of \cite{Wu}, which is in there attributed to Chen \cite{Chen}; according to Wu, the idea that this simple sieve inequality is sufficient is due to Pan \cite{Pan}). \begin{lemma} \label{chen} Let $0 < \alpha < \beta < 1/4,$ $Y:= N^\alpha,$ and $Z:=N^{\beta}.$ Then $S \leq S_1 -S_2 /2 + S_3 /2,$ where \begin{align*} S_1 & := \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W)}} 1_{\mathbb{P}}(n+h_j)1_{(n+h_\ell,P(Y))=1}\nu_{\mathcal{H},j,\ell} (n) \\ S_2 & := \sum_{Y < p \leq Z} \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W) \\ p | n+h_\ell}} 1_{\mathbb{P}}(n+h_j)1_{(n+h_\ell,P(Y))=1}\nu_{\mathcal{H},j,\ell} (n) , \quad \quad \text{and} \\ S_3 &:= \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W)}} 1_{\mathbb{P}}(n+h_j) \sum_{Y < p < q < r \leq Z} \sum_{(s,P(q))=1} 1_{n+h_\ell=pqrs}\nu_{\mathcal{H},j,\ell} (n) . \end{align*} \end{lemma} \begin{proof} By positivity of $\nu_{\mathcal{H},j,\ell} (n)$ it suffices to show that for any $n \in (N+h_\ell,2N+h_\ell]$ \begin{align} \label{sieve} 1_{(n,P(Z))=1} \leq 1_{(n,P(Y)) = 1} - \frac{1}{2} \sum_{Y < p \leq Z} 1_{p|n} 1_{(n,P(Y)) = 1} +\frac{1}{2} \sum_{Y < p < q < r \leq Z} \sum_{(s,P(q))=1} 1_{n=pqrs}. \end{align} For $(n,P(Y))>1$ this is obvious, so let $(n,P(Y))=1$ and denote $k=\sum_{Y < p \leq Z} 1_{p|n}.$ If $k=0,$ then both sides of (\ref{sieve}) are equal to one. For $k \geq 1$ the left-hand side is zero. If $k=1,$ then the right-hand side is $1-1/2+0 = 1/2 > 0.$ For $k \geq 2$ the right-hand side is $1-k/2 + (k-2)/2=0,$ since in the last sum $p$ and $q$ are fixed and there are $k-2$ ways to choose $r$. \end{proof} \begin{remark} Note that $\beta < 1/4$ implies that in the sum $S_3$ we have $s \gg N/(pqr) > N^{1/4} > q.$ The above lemma holds also for $\beta \geq 1/4,$ but then we sometimes may have $s=1$ in the sum $S_3.$ \end{remark} We now proceed to estimate $S_1$, $S_2$ and $S_3$ separately by applying the linear sieve. For this we use similar notations as in \cite[Chapters 11 and 12]{FI} (using the subscript `lin' for clarity): we let $F_{\text{lin}}(s),f_{\text{lin}}(s)$ be the continuous solution to the system of delay-differential equations \begin{align*} \begin{cases} (sF_{\text{lin}}(s))' = f_{\text{lin}}(s-1) \\ (sf_{\text{lin}}(s))' = F_{\text{lin}}(s-1) \end{cases} \end{align*} with the condition \begin{align*} \begin{cases} sF_{\text{lin}}(s) = 2e^{\gamma}, & \text{if} \, \, 1 \leq s \leq 3 \\ sf_{\text{lin}}(s) = 0, & \text{if} \, \, s\leq 2. \end{cases}. \end{align*} Here $\gamma$ is the Euler-Mascheroni constant. We record here that for $2 \leq s \leq 4$ \begin{align*} f_{\text{lin}}(s) = \frac{2 e^\gamma \log (s-1)}{s}. \end{align*} By \cite[Chapters 11 and 12]{FI} we then have \begin{lemma}\label{linear} \emph{\textbf{(Linear sieve).}} Let $(a_n)_{n \geq 1}$ be a sequence of non-negative real numbers. For some fixed $X$ depending only on the sequence $(a_n)_{n \geq 1}$, define $r_d$ for all square-free $d \geq 1$ by \begin{align*} \sum_{n \equiv 0 \, (d)} a_n = g(d) X + r_d, \end{align*} where $g(d)$ is a multiplicative function, depending only on the sequence $(a_n)_{n \geq 1}$, satisfying $0 \leq g(p) < 1$ for all primes $p.$ Let $D\geq 2$ (the level of distribution), and let $z=D^{1/s}$ for some $s\geq 1.$ Suppose that there exists a constant $L >0$ that for any $2 \leq w < z$ we have \begin{align*} \prod_{w \leq p < z} (1-g(p))^{-1} \leq \frac{\log z}{\log w} \bigg(1+\frac{L}{\log w}\bigg). \end{align*} Then \begin{align*} \sum_{n} a_n 1_{(n,P(z))=1} &\leq (F_{\text{\emph{lin}}}(s) + \mathcal{O}(\log^{-1/6} D)) X \prod_{p\leq z} (1-g(p)) + \sum_{\substack{d \leq D \\ d \, \, \text{\emph{squarefree}}}} |r_d|, \\ \sum_{n} a_n 1_{(n,P(z))=1} &\geq (f_{\text{\emph{lin}}}(s) - \mathcal{O}(\log^{-1/6} D)) X \prod_{p\leq z} (1-g(p)) - \sum_{\substack{d \leq D \\ d \, \, \text{\emph{squarefree}}}} |r_d|. \end{align*} \end{lemma} We now estimate the sums $S_1,S_2$ and $S_3$ in the following three lemmata. \begin{lemma} \label{s1} We have \begin{align*} S_1 \leq \frac{F_{\text{\emph{lin}}}(1/(2\alpha)) + \mathcal{O}(\delta)}{\alpha e^\gamma} \frac{N}{W} B^{-K}L_K(F) \end{align*} \end{lemma} \begin{proof} Define $r_d$ by the equation \begin{align} \label{rd} \sum_{\substack{N < n \leq 2N \\ n \equiv -h_\ell \, (d)}} 1_{\mathbb{P}}(n+h_j) 1_{n \equiv b \, (W)}\nu_{\mathcal{H},j,\ell} (n) = g(d) \sum_{\substack{N < n \leq 2N}} 1_{\mathbb{P}}(n+h_j) 1_{n \equiv b \, (W)} \nu_{\mathcal{H},j,\ell} (n) +r_d, \end{align} where $g(d)$ is a multiplicative function, supported on square-free integers, defined by \begin{align*} g(p) := \begin{cases} \frac{1}{p-1}, & \text{if} \, p \, \nmid W Z_{N^{4\epsilon}} \\ 0, & \text{if} \, p \, \mid W Z_{N^{4\epsilon}}. \end{cases} \end{align*} We note that by the same argument as in the proof of \cite[Lemma 4.6]{BFM} (recall that $d_j=d_\ell=1$ in $\nu_{\mathcal{H},j,\ell} (n) $), the sum on the right-hand side in (\ref{rd}) is \begin{align} \nonumber \sum_{\substack{N < n \leq 2N}} 1_{\mathbb{P}}(n+h_j) 1_{n \equiv b \, (W)} \nu_{\mathcal{H},j,\ell} (n) &= (1+o(1)) \frac{N }{\phi(W) \log N} B^{-K+2}L_K(F) \\ \label{main} &= (1+o(1)) \frac{N}{W} B^{-K+1}L_K(F). \end{align} (to show this we just expand the square in $\nu_{\mathcal{H},j,\ell} (n)$, swap the order of summation, and use the Proposition \ref{bv} with moduli $[d_1,d_1'] \cdots [d_K,d_K']W \leq N^{3\delta}$ similarly as in \cite[Lemma 4.6]{BFM}). Hence, by the upper bound of the linear sieve (Lemma \ref{linear} with level of distribution $D=N^{1/2-4\delta}$, sifting up to $Y=N^\alpha$) we get \begin{align*} S_1 \leq (F_{\text{lin}}(1/(2\alpha)) + \mathcal{O}(\delta))\bigg( \prod_{p \leq Y} (1- g(p)) \bigg) \frac{N}{W} B^{-K+1 }L_K(F) + \sum_{\substack {d \leq N^{1/2-4\delta} \\ d \, \, \text{squarefree}} } |r_d|. \end{align*} By Merten's Theorem \begin{align*} \prod_{p \leq Y} (1- g(p)) &= \prod_{W < p < Y}\bigg( 1- \frac{1}{p-1} \bigg) = \prod_{W < p < Y}\bigg( 1- \frac{1+ \mathcal{O}(1/p)}{p} \bigg) \\ & = (1 + o(1))\frac{W}{\phi(W)} \prod_{ p < Y}\bigg( 1- \frac{1}{p} \bigg) = (1+o(1))\frac{W}{\phi(W) e^{\gamma} \log Y} , \end{align*} so that \begin{align*} S_1 \leq \frac{F_{\text{lin}}(1/(2\alpha)) + \mathcal{O}(\delta)}{\alpha e^\gamma} \frac{N}{W} B^{-K}L_K(F)+ \sum_{\substack {d \leq N^{1/2-4\delta} \\ d \, \, \text{squarefree}} } |r_d|. \end{align*} For the error term we expand the square in $\nu_{\mathcal{H},j,\ell} (n) $ and swap the order of summation to get \begin{align*} r_d &= \sum_{\substack{N < n \leq 2N \\ n \equiv -h_\ell \, (d)}} 1_{\mathbb{P}}(n+h_j) 1_{n \equiv b \, (W)} \nu_{\mathcal{H},j,\ell} (n) - g(d) \sum_{\substack{N < n \leq 2N}} 1_{\mathbb{P}}(n+h_j) 1_{n \equiv b \, (W)} \nu_{\mathcal{H},j,\ell} (n) \\ &= \sum_{\substack{d_1, \dots, d_K \\ d'_1,\dots d_K' \\ d_j=d_j'=d_\ell=d_\ell'=1}} \lambda_{d_1,\dots, d_K} \lambda_{d_1',\dots, d_K'} \bigg( \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W) \\n \equiv -h_\ell \, (d) \\ n \equiv -h_i \, ([d_i,d_i'])}} 1_{\mathbb{P}}(n+h_j) - g(d) \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W) \\ n \equiv -h_i \, ([d_i,d_i'])}} 1_{\mathbb{P}}(n+h_j) \bigg). \end{align*} Similarly as in the proof of \cite[Lemma 4.6]{BFM}, we note that since $h'-h$ is $\epsilon \log N$-smooth for all distinct $h,h' \in \mathcal{H}$ by (\ref{smooth}), and by the support conditions (\ref{l1}), (\ref{l2}) of $\lambda_{d_1,\dots,d_k},$ we may assume that $d,$ $[d_1,d_1'],\dots, [d_K,d_K'],$ $W Z_{N^{4\epsilon}}$ are pairwise coprime. In that case we have $g(d)=1/\phi(d)$, \begin{align*} \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W) \\n \equiv -h_\ell \, (d) \\ n \equiv -h_i \, ([d_i,d_i'])}} 1_{\mathbb{P}}(n+h_j) = \frac{\pi(2N+h_j) - \pi (N+h_j)}{\phi(d) \phi(W) \prod_{i=1}^K \phi([d_i,d_i'])} + \mathcal{O} \bigg( E(N, d [d_1,d_1'] \cdots [d_K, d_K'] W) \bigg), \end{align*} and \begin{align*} g(d) \hspace{-5pt}\sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W) \\ n \equiv -h_i \, ([d_i,d_i'])}} 1_{\mathbb{P}}(n+h_j) = \frac{\pi(2N+h_j) - \pi (N+h_j)}{\phi(d) \phi(W) \prod_{i=1}^K \phi([d_i,d_i'])} + \mathcal{O} \bigg( E(N, [d_1,d_1'] \cdots [d_K, d_K'] W) \bigg) \end{align*} where \begin{align*} E(N,q) = \max_{(a,q)=1 } \bigg | \pi (2N+h_j; q, a ) - \pi(N+h_j;q,a) - \frac{\pi(2N+h_j) - \pi (N+h_j)}{\phi(q)} \bigg |, \end{align*} if $(q, Z_{N^{4\epsilon}}) =1 $ and we set $E(N,q) = 0$ if $(q, Z_{N^{4\epsilon}}) >1.$ Hence, by the triangle inequality \begin{align*} \sum_{\substack {d \leq N^{1/2-4\delta} \\ d \,\, \text{squarefree}} } |r_d| \, \ll \sum_{\substack {d \leq N^{1/2-4\delta} \\ d \,\, \text{squarefree} \\ (d,W)=1} } \sum_{\substack{d_1, \dots, d_K \\ d'_1,\dots d_K' \\ d_j=d_j'=d_\ell=d_\ell'=1 }} | \lambda_{d_1,\dots, d_K} \lambda_{d_1',\dots, d_K'} | E(N, d [d_1,d_1'] \cdots [d_K, d_K'] W) \\ + \sum_{\substack {d \leq N^{1/2-4\delta} \\ d \,\, \text{squarefree} \\ (d,W)=1}} \frac{1}{\phi(d)}\sum_{\substack{d_1, \dots, d_K \\ d'_1,\dots d_K' \\ d_j=d_j'=d_\ell=d_\ell'=1}} | \lambda_{d_1,\dots, d_K} \lambda_{d_1',\dots, d_K'} | E(N, [d_1,d_1'] \cdots [d_K, d_K'] W) \end{align*} The second sum on the right-hand side is bounded by $\log N$ times the first sum. We have the trivial bounds $|\lambda_{d_1,\dots, d_K}| \, \ll 1$ and $E(N,q) \ll 1 +N/\phi(q).$ Hence, using Cauchy-Schwarz and Proposition \ref{bv} the first sum is bounded by \begin{align*} \sum_{\substack{q \leq N^{1/2-2\delta} \\ (q, W Z_{N^{4\epsilon}}) =1 }} & \tau_{3K}(q) E(N, qW) \\ &\leq \bigg( \sum_{\substack{q \leq N^{1/2-2\delta} \\ (q, W Z_{N^{4\epsilon}}) =1 }}\tau_{3K}(q)^2 (1+N /\phi(qW))\bigg)^{1/2}\bigg( \sum_{\substack{q \leq N^{1/2-2\delta} \\ (q, W Z_{N^{4\epsilon}}) =1 }} E(N, qW) \bigg)^{1/2} \\ & \, \ll_{K,C} \frac{N}{W \log^C N}, \end{align*} which is sufficient. \end{proof} \begin{lemma} \label{s2} We have \begin{align*} S_2 \geq \frac{1-\mathcal{O}(\delta)}{\alpha e^{\gamma}} \int_\alpha ^\beta f_{\text{\emph{lin}}} \bigg( \frac{1/2 -t}{\alpha} \bigg) \frac{dt}{t} \frac{N}{W} B^{-K}L_K(F). \end{align*} \end{lemma} \begin{proof} Set \begin{align*} S_{2,p} := \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W) \\ p | n+h_\ell}} 1_{\mathbb{P}}(n+h_j)1_{(n+h_\ell,P(Y))=1}\nu_{\mathcal{H},j,\ell} (n), \end{align*} so that $S_2 = \sum_{Y < p \leq Z} S_{2,p}.$ We will apply the lower bound of the linear sieve to each of the sums $S_{2,p}:$ for $(d,p)=1,$ let $r_{dp}$ be defined by \begin{align*} \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W) \\ p | n+h_\ell \\ n \equiv - h_\ell \, (d)}} 1_{\mathbb{P}}(n+h_j) 1_{n \equiv b \, (W)}\nu_{\mathcal{H},j,\ell} (n) = \frac{g(d)}{p-1} \sum_{\substack{N < n \leq 2N}} 1_{\mathbb{P}}(n+h_j) 1_{n \equiv b \, (W)} \nu_{\mathcal{H},j,\ell} (n) +r_{dp}, \end{align*} where $g(d)$ is as in the proof of Lemma \ref{s1}, that is, a multiplicative function, supported on square-free integers, defined by \begin{align*} g(q) := \begin{cases} \frac{1}{q-1}, & \text{if} \, q \, \nmid W Z_{N^{4\epsilon}} \\ 0, & \text{if} \, q \, \mid W Z_{N^{4\epsilon}}. \end{cases} \end{align*} Applying the lower bound of the linear sieve (Lemma \ref{linear} with level of distribution $D=N^{1/2-4\delta}/p$ and shifting up to $Y=N^\alpha$), using (\ref{main}) and Merten's Theorem similarly as in the proof of Lemma \ref{s1}, we find that \begin{align*} S_{2,p} & \geq \bigg( f_{\text{lin}} \bigg( \frac{\log N^{1/2}/p}{\log Y} \bigg) - \mathcal{O}(\delta)\bigg)\frac{1}{p-1} \bigg( \prod_{q \leq Y} (1- g(q)) \bigg) \frac{N}{W} B^{-K+1}L_K(F) - \sum_{\substack {d \leq N^{1/2-4\delta}/p \\ d \, \, \text{squarefree}} } |r_{dp}| \\ & \geq \frac{1}{\alpha e^\gamma} \bigg( f_{\text{lin}} \bigg( \frac{\log N^{1/2}/p}{\log Y} \bigg) - \mathcal{O}(\delta)\bigg)\frac{1}{ p} \frac{N}{W} B^{-K}L_K(F) - \sum_{\substack {d \leq N^{1/2-4\delta}/p \\ d \, \, \text{squarefree}} } |r_{dp}|. \end{align*} Summing over $p$ we get, by a similar argument as in the proof of Lemma \ref{s1}, a sufficient bound for the error term \begin{align*} \sum_{Y < p \leq Z} \sum_{\substack {d \leq N^{1/2-4\delta}/p \\ d \, \, \text{squarefree}} } |r_{dp}| \, \ll_{C,K} \frac{N}{W\log^C N}. \end{align*} Hence, we have \begin{align*} S_{2} & \geq \frac{1- \mathcal{O}(\delta)}{\alpha e^\gamma} \bigg( \sum_{Y< p \leq Z} \frac{1}{p} f_{\text{lin}} \bigg( \frac{\log N^{1/2}/p}{\log Y} \bigg) \bigg) \frac{N}{W} B^{-K}L_K(F) \\ & \geq \frac{1- \mathcal{O}(\delta)}{\alpha e^\gamma} \bigg( \int_{Y < z \leq Z} f_{\text{lin}} \bigg( \frac{\log N^{1/2}/z}{\log Y} \bigg) \frac{dz}{z \log z} \bigg) \frac{N}{W} B^{-K}L_K(F) \\ & \geq \frac{1-\mathcal{O}(\delta)}{\alpha e^{\gamma}} \int_\alpha ^\beta f_{\text{lin}} \bigg( \frac{1/2 -t}{\alpha} \bigg) \frac{dt}{t} \frac{N}{W} B^{-K}L_K(F) \end{align*} by the change of variables $z=N^t$. \end{proof} For the next Lemma we need the Buchstab function, defined as the continuous solution to the delay-differential equation \begin{align*} \begin{cases} s \omega(s) = 1, & \text{if} \, \, 1 \leq s \leq 2,\\ (s \omega (s))' = \omega(s-1), & \text{if} \, \, s > 2. \end{cases} \end{align*} Then by \cite[Lemma 12.1]{FI} for any $N^\epsilon < z < N$ we have \begin{align} \label{buchstabfun} \sum_{N< n \leq 2N} 1_{(n,P(z))=1} = (1+o(1)) \omega(\log N / \log z) \frac{N}{\log z}, \quad \quad N \to \infty. \end{align} \begin{lemma} \label{s3} We have \begin{align*} S_3 \leq (4 + \mathcal{O}(\delta))\int_{\alpha < u_1 < u_2 < u_3 < \beta} \omega \bigg(\frac{1-u_1-u_2-u_3}{u_2} \bigg)\frac{du_1 d u_2 du_3 }{u_1 u_2^2 u_3} \frac{N}{W} B^{-K}L_K(F). \end{align*} \end{lemma} \begin{proof} Here we apply the switching, to sieve out the prime divisors of $n+h_j$ rather than $n+h_\ell$; define \begin{align*} a_n := \sum_{Y < p< q < r \leq Z} \sum_{(s,P(q)) = 1} 1_{n=pqrs} \end{align*} so that \begin{align*} S_3 = \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W)}} 1_\mathbb{P}(n+h_j) a_{n+h_\ell} \nu_{\mathcal{H},j,\ell} (n). \end{align*} We use a similar Selberg upper bound sieve as in \cite[Lemma 4.6]{BFM} (we could just as well use the linear sieve upper bound as in the above but the argument is slightly simpler this way); let $G:[0, \infty) \to \mathbb{R}$ be a smooth function supported on $[0,1/4-2\delta]$ with $G(0)=1.$ Then \begin{align*} S_3 &\leq \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W)}} a_{n+h_\ell} \bigg( \sum_{e \,| n+ h_j}\mu(e) G \bigg( \frac{\log e}{\log N}\bigg)\bigg)^2\nu_{\mathcal{H},j,\ell} (n) \\ &\leq \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W)}} a_{n+h_\ell} \bigg( \sum_{\substack{e \,| n+ h_j\\ (e, Z_{N^{4\epsilon}})=1}}\mu(e) G \bigg( \frac{\log e}{\log N}\bigg)\bigg)^2 \bigg( \sum_{\substack{d_1, \dots, d_K \\ d_i | n+h_i \\ d_j=d_\ell=1}} \lambda_{d_1, \dots, d_K} \bigg)^2. \end{align*} We then expand the squares and rearrange the sum to get \begin{align*} \sum_{\substack{d_1, \dots, d_K \\ d'_1,\dots d_K' \\ d_j=d_j'=d_\ell=d_\ell'=1 }} \lambda_{d_1,\dots, d_K} \lambda_{d_1',\dots, d_K'} \sum_{\substack{e,e' \\ (ee', Z_{N^{4\epsilon}})=1}} \mu(e) \mu(e') G \bigg( \frac{\log e}{\log N}\bigg) G \bigg( \frac{\log e'}{\log N}\bigg) \sum_{\substack{N < n \leq 2N \\ n \equiv b \, (W) \\ [d_i,d_i'] | n+h_i \\ [e,e'] | n+h_j }} a_{n+h_\ell} \end{align*} In the innermost sum, we may again assume that $[d_1,d_1'],\dots, [d_K,d_K'],$ $[e,e'],$ $W Z_{N^{4\epsilon}}$ are pairwise coprime, and insert the estimates (for $d=[d_1,d'_1]\cdots [d_K,d_K'][e,e'] W$) \begin{align*} \sum_{\substack{N < n \leq 2N \\ n \equiv a \, (d) }} a_{n+h_\ell} = \frac{1}{\phi(d)} \sum_{\substack{N < n \leq 2N }} a_{n+h_\ell} + \tilde{r}_d. \end{align*} By essentially the same argument as in the proof of \cite[Lemma 4.6 (iii)]{BFM}, choosing the function $G$ optimally gives \begin{align} \label{third} S_3 &\leq (4+ \mathcal{O}(\delta)) \frac{\log N}{N} \bigg(\sum_{N < n \leq 2N} a_n \bigg) \frac{N}{W} B^{-K}L_K(F) + \mathcal{O}(R), \end{align} where \begin{align*} R = \sum_{\substack{d_1, \dots, d_K \\ d'_1,\dots d_K' \\ d_j=d_j'=d_\ell=d_\ell'=1 }} | \lambda_{d_1,\dots, d_K} \lambda_{d_1',\dots, d_K'} | \sum_{\substack{e,e' \leq N^{1/4-2\delta} \\ (ee',Z_{N^{4\epsilon}}) = 1}}E_0(N, [d_1,d_1'] \cdots [d_K, d_K'] [e,e'] W) \end{align*} with \begin{align*} E_0(N,d) := \max_{(a,d)=1} \bigg | \sum_{\substack{N +h_\ell< n \leq 2N+h_\ell \\ n \equiv a \, (d)}} a_n -\frac{1}{\phi(d)} \sum_{\substack{N+h_\ell < n \leq 2N+h_\ell }} a_n \bigg |. \end{align*} Note that the condition $e,e' \leq N^{1/4-2\delta}$ comes from the support restriction of the function $G$. Using Cauchy-Schwarz and the trivial bound $ | \lambda_{d_1,\dots, d_K}| \ll 1$ similarly as in the proof of Lemma \ref{s1}, the error term $R$ has a sufficient bound if we can show that \begin{align*} \sum_{\substack{d \leq N^{1/2-2\delta} \\ (d, W Z_{N^{4\epsilon}}) =1 }} |E_0(N,dW)| \, \ll_C \frac{N}{W \log^C N}. \end{align*} To show this we use finer-than-dyadic decomposition to write $a_n 1_{N +h_\ell < n \leq 2N + h_\ell}$ as a sum of terms of the form \begin{align*} \sum_{\substack{Y < p < q < r \leq Z \\ p \in I_1, \, \, q \in I_2}}\, \, \sum_{\substack{ (N+h_\ell)/(pqr) < s \leq (2N+h_\ell)/(pqr) \\ (s,P(q))=1}} 1_{n=pqrs}, \end{align*} where each $I_j$ is of the form $(A_j, \lambda A_j ]$ for $\lambda= 1 + \log^{-2C} N$. We remove the cross-conditions $Y < p < q;$ this causes an error bounded using triangle inequality by the sum of (\ref{error1}) and (\ref{error2}), which are given by \begin{align} \label{error1} \sum_{\substack{d \leq N^{1/2-2\delta} \\ (d, W Z_{N^{4\epsilon}}) =1 }} & \max_{(a,d)=1} \sum_{\substack{Y < p < q < r \leq Z \\ p \in [\lambda^{-2}Y, \lambda^2 Y] \cup [\lambda^{-2}q, \lambda^2 q] \\ (pq,d) =1 }} \, \, \sum_{\substack{s \asymp N/(pqr) \\ (s,(P(q)))=1 \\ rs \equiv a \overline{pq} \, (dW)}} 1 \, \\ \nonumber & \ll \sum_{\substack{d \leq N^{1/2-2\delta} \\ (d, W Z_{N^{4\epsilon}}) =1 }} \max_{(a,d)=1} \sum_{\substack{Y < p < q \leq Z \\ p \in [\lambda^{-2}Y, \lambda^2 Y] \cup [\lambda^{-2}q, \lambda^2 q] \\ (pq,d) =1 }} \, \, \sum_{\substack{m \asymp N/(pq) \\ m \equiv a \overline{pq} \, (dW)}} 1 \ll_C \frac{N}{W \log^C N} \end{align} (since $m=rs \gg N/pq > N^{1/2}$ by using $\beta < 1/4$), and \begin{align}\label{error2} \sum_{\substack{d \leq N^{1/2-2\delta} \\ (d, W Z_{N^{4\epsilon}}) =1 }} \frac{1}{\phi(dW)} \sum_{\substack{Y < p < q < r \leq Z \\ p \in [\lambda^{-2}Y, \lambda^2 Y] \cup [\lambda^{-2}q, \lambda^2 q]}} \, \, \sum_{\substack{s \asymp N/(pqr) \\ (s,(P(q)))=1}} 1 \, \ll_C \frac{N}{W \log^C N}, \end{align} which is sufficient. Similarly, if we replace the condition $N+h_\ell < pqrs \leq 2N+h_\ell$ by $ (N+h_\ell)/(A_1qr) < s \leq (2N+h_\ell)/(A_1qr),$ then we get a sufficient bound for the contribution of the part where $pqrs \notin (N+h_\ell,2N+h_\ell].$ Thus, we can replace $a_n 1_{N < n \leq 2N}$ by a sum of $\mathcal{O}(\log^{4C+2} N)$ functions of the form $(P\ast g)(n), $ where for $Y \ll A_1, A_2 \ll Z$ \begin{align*} P(m)= 1_\mathbb{P}(m)1_{m \in (A_1,\lambda A_1]} \quad \text{and} \quad g(n)= \sum_{\substack{ q < r \leq Z \\ q \in (A_2,\lambda A_2] }} \sum_{\substack{ (N+h_\ell)/(A_1qr) < s \leq (2N+h_\ell)/(A_1qr)\\ (s,P(q))=1}} 1_{n=qrs}. \end{align*} We can then replace $P(m)$ by $f(m)/\log A_1,$ where $f(m) := P(m)\log m$; this is because for all $m \in (A_1,\lambda A_1]$ we have \begin{align*} \log m = \log A_1 + \mathcal{O}\bigg( \log^{-2C} N \bigg), \end{align*} so that the error term from this has a sufficient bound by trivial estimates. Finally, writing $f(m) = 1_\mathbb{P}(m)(\log m )1_{m \leq \lambda A_1} - 1_\mathbb{P}(m)(\log m) 1_{m \leq A_1}$ and using triangle inequality, we obtain by Proposition \ref{bv2} that \begin{align*} \sum_{\substack{d \leq N^{1/2-2\delta} \\ (d, W Z_{N^{4\epsilon}}) =1 }} |E_0(N,dW)| \, \ll_C \frac{N}{W \log^C N}, \end{align*} which suffices by the previous remarks to bound the error term $R$ in (\ref{third}). To compute the main term in (\ref{third}) we write by using (\ref{buchstabfun}) \begin{align*} \sum_{N < n \leq 2N} a_n &= \sum_{Y < p< q < r \leq Z} \, \,\sum_{ \substack{ N/(pqr) < s \leq 2N/(pqr) \\ (s,P(q)) = 1}} 1 \\ & = (1+o(1)) N \sum_{Y < p< q < r \leq Z} \frac{\omega \bigg( \frac{\log(N/(pqr))}{\log q}\bigg)}{pqr \log q} \\ &= ( 1+o(1)) N \int_{Y<z_1 < z_2 < z_3 \leq Z} \omega \bigg( \frac{\log(N/(z_1z_2z_3))}{\log z_2}\bigg)\frac{dz_1 dz_2 dz_3}{z_1 z_2 z_3 (\log z_1)( \log^2 z_2) \log z_3} \\ &= ( 1+o(1)) \frac{N}{\log N} \int_{\alpha < u_1 < u_2 < u_3 < \beta} \omega \bigg(\frac{1-u_1-u_2-u_3}{u_2} \bigg)\frac{du_1 d u_2 du_3 }{u_1 u_2^2 u_3} \end{align*} after the change of variables $z_j=N^{u_j}.$ \end{proof} \emph{Proof of Proposition \ref{pairs}.} Combining Lemmata \ref{chen}, \ref{s1}, \ref{s2} and \ref{s3} we obtain \begin{align*} S \leq (\Omega_1 - \Omega_2 + \Omega_3 + \mathcal{O}(\delta)) \frac{N}{W} B^{-K}L_K(F), \end{align*} where \begin{align*} \Omega_1 &= \frac{F_{\text{lin}}(1/(2\alpha))}{\alpha e^\gamma}, \quad \quad \quad \Omega_2 = \frac{1}{2\alpha e^{\gamma}} \int_\alpha ^\beta f_{\text{lin}} \bigg( \frac{1/2 -t}{\alpha} \bigg) \frac{dt}{t}, \quad \quad \text{and} \\ \Omega_3 &= 2 \int_{\alpha < u_1 < u_2 < u_3 < \beta} \omega \bigg(\frac{1-u_1-u_2-u_3}{u_2} \bigg)\frac{du_1 d u_2 du_3 }{u_1 u_2^2 u_3}. \end{align*} We choose $\alpha = 1/7$ and $\beta= 3/14$ (so that $(1/2-t)/\alpha \geq 2$ in the integral defining $\Omega_2$). For this choice we get \begin{align*} \Omega_1 = \frac{7F_{\text{lin}}(7/2)}{e^\gamma} = 2\bigg( \frac{3F_{\text{lin}}(3)}{e^\gamma} + \int_3^{7/2} \frac{f_{\text{lin}}(s-1)}{e^\gamma} ds \bigg) \\ = 4+ 4\int_3^{7/2} \frac{\log (s-2)}{s-1} ds \leq 4.19, \end{align*} \begin{align*} \Omega_2 = \frac{7}{2e^{\gamma}} \int_{1/7}^{3/14} f_{\text{lin}} \bigg( 7/2-7t \bigg) \frac{dt}{t} = 7 \int_{1/7}^{3/14}\frac{\log(7/2-7t-1)}{7/2-7t}\frac{dt}{t} \geq 0.279, \end{align*} and \begin{align*} \Omega_3 = 2 \int_{1/7 < u_1 < u_2 < u_3 < 3/14} \omega \bigg(\frac{1-u_1-u_2-u_3}{u_2} \bigg)\frac{du_1 d u_2 du_3 }{u_1 u_2^2 u_3} \leq 0.076. \end{align*} Hence, $\Omega_1 - \Omega_2 + \Omega_3 < 3.99.$ \qed \begin{remark} The upper bound for the integral in $\Omega_3$ was computed using Python 7.3; the code is available at \url{http://codepad.org/2emT1dHN}. The choice of exponents $\alpha=1/7$ and $\beta=3/14$ has not been optimized since this is not relevant to our application. \end{remark} \section{Modified Maynard-Tao sieve} \label{maysec} We are now ready to prove the following version of the Maynard-Tao sieve, which is modelled after \cite[Theorem 4.3]{BFM}: \begin{prop} \label{strong} \emph{\textbf{(Modified Maynard-Tao sieve).}} Let $K$ be a sufficiently large multiple of $4.$ Let $\epsilon > 0$ be sufficiently small. Then for all sufficiently large $N$ the following holds: Let $Z_{N^{4\epsilon}}$ be as in (\ref{Z}) and define \begin{align*} W := \prod_{\substack{p \leq \epsilon \log N \\ p \, \nmid Z_{N^{4\epsilon}}}} p; \end{align*} Let $\mathcal{H} = \{h_1, \dots, h_K\} \subseteq [0,N]$ be an admissible $K$-tuple such that \begin{align*} P^+ \bigg( \prod_{1 \leq i < j\leq K} (h_j-h_i) \bigg) \leq \epsilon \log N \end{align*} Let $b$ be an integer such that \begin{align*} \bigg( \prod_{j=1}^K (b+h_j) , W\bigg) =1. \end{align*} Let \begin{align*} \mathcal{H} = \mathcal{H}_1 \cup \mathcal{H}_2 \cup \mathcal{H}_3 \cup \mathcal{H}_{4} \end{align*} be a partition of $\mathcal{H}$ into four sets of equal size. Then there is an integer $n \in [N,2N]$ with $n \equiv b \, (W)$ such that $n+\mathcal{H}_i$ contains a prime number for at least two distinct indices $i \in \{1,2,3,4\}.$ \end{prop} To prove the above proposition we will show that it suffices to prove the following seemingly weaker \begin{prop} \label{weak} Let $a \geq 1$ be an integer and let $K$ be a sufficiently large multiple of $\lceil 3.99a \rceil +1.$ Let $\epsilon > 0$ be sufficiently small. Then for all sufficiently large $N$ the following holds: Let $Z_{N^{4\epsilon}}$ be as in (\ref{Z}) and define \begin{align*} W := \prod_{\substack{p \leq \epsilon \log N \\ p \nmid Z_{N^{4\epsilon}}}} p. \end{align*} Let $\mathcal{H} = \{h_1, \dots, h_K\} \subseteq [0,N]$ be an admissible $K$-tuple such that \begin{align*} P^+ \bigg( \prod_{1 \leq i < j\leq K} (h_j-h_i) \bigg) \leq \epsilon \log N \end{align*} Let $b$ be an integer such that \begin{align*} \bigg( \prod_{j=1}^K (b+h_j) , W\bigg) =1. \end{align*} Let \begin{align*} \mathcal{H} = \mathcal{H}_1 \cup \mathcal{H}_2 \cup \cdots \cup \mathcal{H}_{\lceil 3.99a \rceil+1} \end{align*} be a partition of $\mathcal{H}$ into $\lceil 3.99a \rceil+1$ sets of equal size. Then there is an integer $n \in [N,2N]$ with $n \equiv b \, (W)$ and a set of $a+1$ distinct indices $\{j_1,j_2,\dots, j_{a+1}\} \subseteq \{1,2,\dots,\lceil 3.99a \rceil+1\}$ such that $n+\mathcal{H}_j$ contains a prime number for every $j \in \{j_1,j_2,\dots, j_{a+1}\}.$ \end{prop} \emph{Proof of Proposition \ref{strong} using Proposition \ref{weak}.} We take $a=100$ so that $\lceil 3.99a \rceil+1 =4a.$ By taking a larger $K$ if necessary, we may suppose that $K$ is a sufficiently large multiple of $4a$. Given a partition $\mathcal{H} = \mathcal{H}_1 \cup \mathcal{H}_2 \cup \mathcal{H}_3 \cup \mathcal{H}_{4}$ as in Proposition \ref{strong}, we take a further partition \begin{align*} \mathcal{H}_i = \mathcal{H}_{i1} \cup \mathcal{H}_{i2} \cup \cdots \cup \mathcal{H}_{ia} \end{align*} into sets of equal sizes for all $i \in \{1,2,3,4\}.$ Then by Proposition \ref{weak} there is an integer $n \in [N,2N]$ with $n \equiv b \, (W)$ so that for at least $a+1$ distinct sets $\mathcal{H}_{ij}$ the set $n+ \mathcal{H}_{ij}$ contains a prime number. By the pigeon-hole principle this implies that $n+\mathcal{H}_i$ contains a prime number for at least two distinct indices $i \in \{1,2,3,4\}.$ \qed \vspace{7pt} \emph{Proof of Proposition \ref{weak}.} We use Pintz's refined version of the argument in \cite{BFM} (cf. proof of \cite[Theorem 3]{Pintz} and especially \cite[Theorem 5.4]{BF}): using the notations of \cite{BF}, let us denote $M:=\lceil 3.99a \rceil+1$, and let $\mu, \mu'$ be positive real numbers with (defining $\binom{1}{2}=0$) \begin{align} \label{mu} \mu' = \max_{v \in \mathbb{N}} \bigg(v- \mu \binom{v}{2} \bigg). \end{align} For any integer $n$ consider \begin{align} \label{quant} \sum_{j=1}^M \bigg( \sum_{h \in \mathcal{H}_j} 1_\mathbb{P}(n+h) - \mu \sum_{\substack{\{h,h'\} \subseteq \mathcal{H}_j \\ h \neq h'}} 1_\mathbb{P}(n+h)1_\mathbb{P}(n+h') \bigg). \end{align} If there are at most $a$ indices $j$ such that $n+\mathcal{H}_j$ contains a prime, then the sum (\ref{quant}) is at most $\mu' a$. Hence, if \begin{align*} \sum_{h \in \mathcal{H}} 1_\mathbb{P}(n+h)-\mu' a - \mu \sum_{j=1}^M \sum_{\substack{\{h,h'\} \subseteq \mathcal{H}_j \\ h \neq h'}} 1_\mathbb{P}(n+h)1_\mathbb{P}(n+h') \, > \,0, \end{align*} then there are at least $a+1$ distinct indices $j$ such that $n+\mathcal{H}_j$ contains a prime. Therefore, the proposition follows once we show that \begin{align*} \sum_{\substack{N<n\leq 2N \\ n \equiv b \, (W)}}\bigg( \sum_{h \in \mathcal{H}} 1_\mathbb{P}(n+h)-\mu' a - \mu \sum_{j=1}^M \sum_{\substack{\{h,h'\} \subseteq \mathcal{H}_j \\ h \neq h'}} 1_\mathbb{P}(n+h)1_\mathbb{P}(n+h')\bigg) \bigg(\sum_{\substack{d_1, \dots, d_K \\ d_i | n+h_i}} \lambda_{d_1, \dots, d_K} \bigg)^2 \, > \,0. \end{align*} Let $\Sigma$ denote the above sum. Using \cite[Lemma 4.6 (i),(ii)]{BFM} to evaluate the first two sums, and Proposition \ref{pairs} to bound the third, we obtain that $\Sigma$ is bounded from below by \begin{align*} (1+\mathcal{O}(\delta))\frac{N}{WB^K} \bigg( K J_K(F) - \mu' a I_K(F) - 3.99 \mu M \binom{K/M}{2} L_K(F) \bigg), \end{align*} where $I_K(F)$, $J_K(F)$ and $L_K(F)$ are the integrals in \cite[Lemma 4.6]{BFM} ($L_K(F)$ is the same as in (\ref{Lint}) above). By \cite[Lemma 4.7]{BFM}, for any given $\rho \in (0,1)$ there is a choice of $F$ such that \begin{align*} J_K(F) &\geq (1+ \mathcal{O}(\log^{-1/2} K)) \frac{\rho \delta\log K}{K} I_K(F), \\ L_K(F) & \leq (1+ \mathcal{O}(\log^{-1/2} K)) \bigg(\frac{\rho \delta\log K}{K} \bigg)^2 I_K(F). \end{align*} Thus, we have \begin{align} \label{Sigma} \Sigma \geq \mathfrak{S}(1+\mathcal{O}(\delta)) N W^{-1}B^{-K} I_K(F), \end{align} where \begin{align*} \mathfrak{S} := \rho \delta \log K - \mu' a - 3.99 \mu M \binom{K/M}{2}\bigg(\frac{\rho\delta \log K}{K} \bigg)^2, \end{align*} if we pick $K$ large enough so that $\log^{-1/2} K < \delta.$ Choosing $\mu=1/L$ for some positive integer $L$ we observe that $\mu' = (1+L)/2,$ the maximum (\ref{mu}) being obtained at $v= L$ and $v=1+L$. Define the quantity $X$ by $XM: = \rho \delta \log K.$ Then by using $3.99 \leq (M-1)/a$ we obtain \begin{align*} \mathfrak{S} &= XM - \frac{1+L}{2} a - 3.99 \frac{M}{L}\binom{K/M}{2}\bigg(\frac{XM}{K} \bigg)^2 \\ & \geq XM - \frac{1+L}{2} a - \frac{M-1}{a} \frac{M}{L} \frac{K^2}{2M^2} \bigg(\frac{XM}{K} \bigg)^2 \\ & = XM - \frac{1+L}{2}a - \frac{M-1}{a}\frac{X^2 M}{2L} = \frac{a}{2(M-1)}>0, \end{align*} for $X=aL/(M-1)$ and $L=M,$ requiring that $K$ is large enough so that $\rho < 1$ for this choice of $X$. \qed \section{Proof of Theorem \ref{maint}} \label{mainsec} Theorem \ref{maint} now follows by the same argument as in \cite[Section 6]{BFM}, using our Proposition \ref{strong} in place of \cite[Theorem 4.3]{BFM}; for this we need the modified Erd\"os-Rankin construction given by \cite[Lemma 5.2]{BFM} which states: \begin{lemma} \label{erlemma} Let $K \geq 1$ and $\beta_K \geq \beta_{K-1} \geq \cdots \geq \beta_1 \geq 0.$ Then there is a real number $y(\bm{\beta},K)$ such that the following holds: Let $x,y,z$ be any real numbers such that $x \geq 1$, $y \geq y(\bm{\beta},K)$, and \begin{align*} 2y(1+(1+\beta_K)x)\leq 2z \leq y (\log_2 y)(\log_3 y)^{-1}. \end{align*} Let $\mathcal{Z}$ be any (possibly empty) set of primes such that for any $q \in \mathcal{Z}$ we have \begin{align*} \sum_{p\in \mathcal{Z}, \, p \geq q} 1/p \ll 1/q \ll 1/\log z. \end{align*} Then there is a set of integers $\{a_p:p\leq y, \, \, p \notin \mathcal{Z}\}$ and an admissible $K$-tuple $\{h_1,h_2,\dots,h_K\}$ such that \begin{align*} \{h_1,h_2,\dots,h_K\} &= ((0,z] \cap \mathbb{Z}) \setminus \bigcup_{p \leq y, \, p \notin \mathcal{Z}} \{m: m \equiv a_p \quad (p) \}, \\ P^+ \bigg( \prod_{1 \leq i < j\leq K} (h_j-h_i) \bigg) & \leq y, \end{align*} and for all $i=1,2,\dots,K$ \begin{align*} h_i = \beta_i xy +y + \mathcal{O}\left( y e^{-\log^{1/4} y}\right). \end{align*} \end{lemma} Given $\beta_1 \leq \beta_2 \leq \beta_3 \leq \beta_4$ as in Theorem \ref{maint} and any sufficiently large $N$, we will apply the above lemma with \begin{align*} x&:= 1/\epsilon, \quad \quad y:= \epsilon \log N, \quad \quad z:=y (\log_2 y)(2\log_3 y)^{-1}, \\ \bm{\beta}&:= \{\beta_1,\dots,\beta_1, \beta_2,\dots,\beta_2,\beta_3,\dots,\beta_3,\beta_4,\dots,\beta_4,\}, \end{align*} where $\epsilon>0$ is sufficiently small and each $\beta_i$ is repeated $K/4$ times for some sufficiently large $K \equiv 0 \,(4)$; by translation we may assume $\beta_1\geq 0.$ We let $\mathcal{Z}:= \{Z_{N^{4\epsilon}}\}$ if $Z_{N^{4\epsilon}} > 1,$ and $\mathcal{Z}=\emptyset$ otherwise (recall (\ref{Z}) for the definition of $Z_T$). The conditions of Lemma \ref{erlemma} are satisfied, so we get a set of integers $\{a_p: p\leq y, \, \, p \neq Z_{N^{4\epsilon}}\}$ and an admissible $K$-tuple $\mathcal{H}$ such that \begin{align} \label{er}&\mathcal{H} = ((0,z] \cap \mathbb{Z}) \setminus \bigcup_{p \leq \epsilon \log N, \, p \neq Z_{N^{4\epsilon}}} \{m: m \equiv a_p \quad (p) \}, \\ \nonumber & P^+ \bigg( \prod_{1 \leq i < j\leq K} (h_j-h_i) \bigg) \leq \epsilon \log N, \end{align} such that there is a partition $\mathcal{H}=\mathcal{H}_1 \cup \mathcal{H}_2 \cup\mathcal{H}_3 \cup\mathcal{H}_4$ into sets of equal sizes so that for all $i=1,2,3,4$ and for all $h \in \mathcal{H}_i$ \begin{align*} h = (\beta_i+\epsilon + o(1)) \log N. \end{align*} Let $b$ be an integer satisfying \begin{align*} b \equiv -a_p \quad (p) \quad \quad \text{for all} \quad \quad p \leq \epsilon \log N, \, p \neq Z_{N^{4\epsilon}}. \end{align*} Then the assumptions of Proposition \ref{strong} are satisfied, so that the proposition yields two indices $1\leq i<j \leq 4$ and an integer $n \in [N,2N]$ with $n \equiv b \, (W)$ such that both $n+\mathcal{H}_i$ and $n+\mathcal{H}_j$ contain a prime number. Furthermore, since $n \equiv b \, (W),$ by (\ref{er}) we have \begin{align*} \mathbb{P} \cap (n,n+z] \subseteq n + \mathcal{H}. \end{align*} Thus, for some $1\leq i<j \leq 4$, there are consecutive primes $p,q \in n+ \mathcal{H}$ such that \begin{align*} p =(\beta_i+\epsilon + o(1)) \log N, \quad \quad \text{and} \quad \quad q = (\beta_j+\epsilon + o(1)) \log N. \end{align*} Since this holds for all sufficiently large $N$, we obtain that for some $1\leq i<j \leq 4$ we have $\beta_j-\beta_i \in \mathbb{L}$. \qed
{ "timestamp": "2020-11-03T02:39:38", "yymm": "1811", "arxiv_id": "1811.03008", "language": "en", "url": "https://arxiv.org/abs/1811.03008", "abstract": "We show that at least 1/3 of positive real numbers are in the set of limit points of normalized prime gaps. More precisely, if $p_n$ denotes the $n$th prime and $\\mathbb{L}$ is the set of limit points of the sequence $\\{(p_{n+1}-p_n)/\\log p_n\\}_{n=1}^\\infty,$ then for all $T\\geq 0$ the Lebesque measure of $\\mathbb{L} \\cap [0,T]$ is at least $T/3.$ This improves the result of Pintz (2015) that the Lebesque measure of $\\mathbb{L} \\cap [0,T]$ is at least $(1/4-o(1))T,$ which was obtained by a refinement of the previous ideas of Banks, Freiberg, and Maynard (2015). Our improvement comes from using Chen's sieve to give, for a certain sum over prime pairs, a better upper bound than what can be obtained using Selberg's sieve. Even though this improvement is small, a modification of the arguments Pintz and Banks, Freiberg, and Maynard shows that this is sufficient. In addition, we show that there exists a constant $C$ such that for all $T \\geq 0$ we have $\\mathbb{L} \\cap [T,T+C] \\neq \\emptyset,$ that is, gaps between limit points are bounded by an absolute constant.", "subjects": "Number Theory (math.NT)", "title": "Limit points of normalized prime gaps", "lm_name": "Qwen/Qwen-72B", "lm_label": "1. YES\n2. YES", "lm_q1_score": 0.9905874098566368, "lm_q2_score": 0.8080672158638527, "lm_q1q2_score": 0.8004612103526376 }