Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Equivalences} | |
\label{cha:equivalences} | |
We now study in more detail the notion of \emph{equivalence of types} that was introduced briefly in \cref{sec:basics-equivalences}. | |
Specifically, we will give several different ways to define a type $\isequiv(f)$ having the properties mentioned there. | |
Recall that we wanted $\isequiv(f)$ to have the following properties, which we restate here: | |
\begin{enumerate} | |
\item $\qinv(f) \to \isequiv (f)$.\label{item:beb1} | |
\item $\isequiv (f) \to \qinv(f)$.\label{item:beb2} | |
\item $\isequiv(f)$ is a mere proposition.\label{item:beb3} | |
\end{enumerate} | |
Here $\qinv(f)$ denotes the type of quasi-inverses to $f$: | |
\begin{equation*} | |
\sm{g:B\to A} \big((f \circ g \htpy \idfunc[B]) \times (g\circ f \htpy \idfunc[A])\big). | |
\end{equation*} | |
By function extensionality, it follows that $\qinv(f)$ is equivalent to the type | |
\begin{equation*} | |
\sm{g:B\to A} \big((f \circ g = \idfunc[B]) \times (g\circ f = \idfunc[A])\big). | |
\end{equation*} | |
We will define three different types having properties~\ref{item:beb1}--\ref{item:beb3}, which we call | |
\begin{itemize} | |
\item half adjoint equivalences, | |
\item bi-invertible maps, | |
\index{function!bi-invertible} | |
and | |
\item contractible functions. | |
\end{itemize} | |
We will also show that all these types are equivalent. | |
These names are intentionally somewhat cumbersome, because after we know that they are all equivalent and have properties~\ref{item:beb1}--\ref{item:beb3}, we will revert to saying simply ``equivalence'' without needing to specify which particular definition we choose. | |
But for purposes of the comparisons in this chapter, we need different names for each definition. | |
Before we examine the different notions of equivalence, however, we give a little more explanation of why a different concept than quasi-invertibility is needed. | |
\section{Quasi-inverses} | |
\label{sec:quasi-inverses} | |
\index{quasi-inverse|(}% | |
We have said that $\qinv(f)$ is unsatisfactory because it is not a mere proposition, whereas we would rather that a given function could ``be an equivalence'' in at most one way. | |
However, we have given no evidence that $\qinv(f)$ is not a mere proposition. | |
In this section we exhibit a specific counterexample. | |
\begin{lem}\label{lem:qinv-autohtpy} | |
If $f:A\to B$ is such that $\qinv (f)$ is inhabited, then | |
\[\eqv{\qinv(f)}{\Parens{\prd{x:A}(x=x)}}.\] | |
\end{lem} | |
\begin{proof} | |
By assumption, $f$ is an equivalence; that is, we have $e:\isequiv(f)$ and so $(f,e):\eqv A B$. | |
By univalence, $\idtoeqv:(A=B) \to (\eqv A B)$ is an equivalence, so we may assume that $(f,e)$ is of the form $\idtoeqv(p)$ for some $p:A=B$. | |
Then by path induction, we may assume $p$ is $\refl{A}$, in which case $f$ is $\idfunc[A]$. | |
Thus we are reduced to proving $\eqv{\qinv(\idfunc[A])}{(\prd{x:A}(x=x))}$. | |
Now by definition we have | |
\[ \qinv(\idfunc[A]) \jdeq | |
\sm{g:A\to A} \big((g \htpy \idfunc[A]) \times (g \htpy \idfunc[A])\big). | |
\] | |
By function extensionality, this is equivalent to | |
\[ \sm{g:A\to A} \big((g = \idfunc[A]) \times (g = \idfunc[A])\big). | |
\] | |
And by \cref{ex:sigma-assoc}, this is equivalent to | |
\[ \sm{h:\sm{g:A\to A} (g = \idfunc[A])} (\proj1(h) = \idfunc[A]) | |
\] | |
However, by \cref{thm:contr-paths}, $\sm{g:A\to A} (g = \idfunc[A])$ is contractible with center $(\idfunc[A],\refl{\idfunc[A]})$; therefore by \cref{thm:omit-contr} this type is equivalent to $\idfunc[A] = \idfunc[A]$. | |
And by function extensionality, $\idfunc[A] = \idfunc[A]$ is equivalent to $\prd{x:A} x=x$. | |
\end{proof} | |
\noindent | |
We remark that \cref{ex:qinv-autohtpy-no-univalence} asks for a proof of the above lemma which avoids univalence. | |
Thus, what we need is some $A$ which admits a nontrivial element of $\prd{x:A}(x=x)$. | |
Thinking of $A$ as a higher groupoid, an inhabitant of $\prd{x:A}(x=x)$ is a natural transformation\index{natural!transformation} from the identity functor of $A$ to itself. | |
Such transformations are said to form the \define{center of a category}, | |
\index{center!of a category}% | |
\index{category!center of}% | |
since the naturality axiom requires that they commute with all morphisms. | |
Classically, if $A$ is simply a group regarded as a one-object groupoid, then this yields precisely its center in the usual group-theoretic sense. | |
This provides some motivation for the following. | |
\begin{lem}\label{lem:autohtpy} | |
Suppose we have a type $A$ with $a:A$ and $q:a=a$ such that | |
\begin{enumerate} | |
\item The type $a=a$ is a set.\label{item:autohtpy1} | |
\item For all $x:A$ we have $\brck{a=x}$.\label{item:autohtpy2} | |
\item For all $p:a=a$ we have $p\ct q = q \ct p$.\label{item:autohtpy3} | |
\end{enumerate} | |
Then there exists $f:\prd{x:A} (x=x)$ with $f(a)=q$. | |
\end{lem} | |
\begin{proof} | |
Let $g:\prd{x:A} \brck{a=x}$ be as given by~\ref{item:autohtpy2}. First we | |
observe that each type $\id[A]xy$ is a set. For since being a set is a mere | |
proposition, we may apply the induction principle of propositional truncation, and assume that $g(x)=\bproj | |
p$ and $g(y)=\bproj{p'}$ for $p:a=x$ and $p':a=y$. In this case, composing with | |
$p$ and $\opp{p'}$ yields an equivalence $\eqv{(x=y)}{(a=a)}$. But $(a=a)$ is | |
a set by~\ref{item:autohtpy1}, so $(x=y)$ is also a set. | |
Now, we would like to define $f$ by assigning to each $x$ the path $\opp{g(x)} | |
\ct q \ct g(x)$, but this does not work because $g(x)$ does not inhabit $a=x$ | |
but rather $\brck{a=x}$, and the type $(x=x)$ may not be a mere proposition, | |
so we cannot use induction on propositional truncation. Instead we can apply | |
the technique mentioned in \cref{sec:unique-choice}: we characterize | |
uniquely the object we wish to construct. Let us define, for each $x:A$, the | |
type | |
\[ B(x) \defeq \sm{r:x=x} \prd{s:a=x} (r = \opp s \ct q\ct s).\] | |
We claim that $B(x)$ is a mere proposition for each $x:A$. | |
Since this claim is itself a mere proposition, we may again apply induction on | |
truncation and assume that $g(x) = \bproj p$ for some $p:a=x$. | |
Now suppose given $(r,h)$ and $(r',h')$ in $B(x)$; then we have | |
\[ h(p) \ct \opp{h'(p)} : r = r'. \] | |
It remains to show that $h$ is identified with $h'$ when transported along this equality, which by transport in identity types and function types (\cref{sec:compute-paths,sec:compute-pi}), reduces to showing | |
\[ h(s) = h(p) \ct \opp{h'(p)} \ct h'(s) \] | |
for any $s:a=x$. | |
But each side of this is an equality between elements of $(x=x)$, so it follows from our above observation that $(x=x)$ is a set. | |
Thus, each $B(x)$ is a mere proposition; we claim that $\prd{x:A} B(x)$. | |
Given $x:A$, we may now invoke the induction principle of propositional truncation to assume that $g(x) = \bproj p$ for $p:a=x$. | |
We define $r \defeq \opp p \ct q \ct p$; to inhabit $B(x)$ it remains to show that for any $s:a=x$ we have | |
$r = \opp s \ct q \ct s$. | |
Manipulating paths, this reduces to showing that $q\ct (p\ct \opp s) = (p\ct \opp s) \ct q$. | |
But this is just an instance of~\ref{item:autohtpy3}. | |
\end{proof} | |
\begin{thm}\label{thm:qinv-notprop} | |
There exist types $A$ and $B$ and a function $f:A\to B$ such that $\qinv(f)$ is not a mere proposition. | |
\end{thm} | |
\begin{proof} | |
It suffices to exhibit a type $X$ such that $\prd{x:X} (x=x)$ is not a mere proposition. | |
Define $X\defeq \sm{A:\type} \brck{\bool=A}$, as in the proof of \cref{thm:no-higher-ac}. | |
It will suffice to exhibit an $f:\prd{x:X} (x=x)$ which is unequal to $\lam{x} \refl{x}$. | |
Let $a \defeq (\bool,\bproj{\refl{\bool}}) : X$, and let $q:a=a$ be the path corresponding to the nonidentity equivalence $e:\eqv\bool\bool$ defined by $e(\bfalse)\defeq\btrue$ and $e(\btrue)\defeq\bfalse$. | |
We would like to apply \cref{lem:autohtpy} to build an $f$. | |
By definition of $X$, equalities in subset types (\cref{subsec:prop-subsets}), and univalence, we have $\eqv{(a=a)}{(\eqv{\bool}{\bool})}$, which is a set, so~\ref{item:autohtpy1} holds. | |
Similarly, by definition of $X$ and equalities in subset types we have~\ref{item:autohtpy2}. | |
Finally, \cref{ex:eqvboolbool} implies that every equivalence $\eqv\bool\bool$ is equal to either $\idfunc[\bool]$ or $e$, so we can show~\ref{item:autohtpy3} by a four-way case analysis. | |
Thus, we have $f:\prd{x:X} (x=x)$ such that $f(a) = q$. | |
Since $e$ is not equal to $\idfunc[\bool]$, $q$ is not equal to $\refl{a}$, and thus $f$ is not equal to $\lam{x} \refl{x}$. | |
Therefore, $\prd{x:X} (x=x)$ is not a mere proposition. | |
\end{proof} | |
More generally, \cref{lem:autohtpy} implies that any ``Eilenberg--Mac Lane space'' $K(G,1)$, where $G$ is a nontrivial abelian\index{group!abelian} group, will provide a counterexample; see \cref{cha:homotopy}. | |
The type $X$ we used turns out to be equivalent to $K(\mathbb{Z}_2,1)$. | |
In \cref{cha:hits} we will see that the circle $\Sn^1 = K(\mathbb{Z},1)$ is another easy-to-describe example. | |
We now move on to describing better notions of equivalence. | |
\index{quasi-inverse|)}% | |
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% | |
\section{Half adjoint equivalences} | |
\label{sec:hae} | |
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% | |
\index{equivalence!half adjoint|(defstyle}% | |
\index{half adjoint equivalence|(defstyle}% | |
\index{adjoint!equivalence!of types, half|(defstyle}% | |
In \cref{sec:quasi-inverses} we concluded that $\qinv(f)$ is equivalent to $\prd{x:A} (x=x)$ by discarding a contractible type. | |
Roughly, the type $\qinv(f)$ contains three data $g$, $\eta$, and $\epsilon$, of which two ($g$ and $\eta$) could together be seen to be contractible when $f$ is an equivalence. | |
The problem is that removing these data left one remaining ($\epsilon$). | |
In order to solve this problem, the idea is to add one \emph{additional} datum which, together with $\epsilon$, forms a contractible type. | |
\begin{defn}\label{defn:ishae} | |
A function $f:A\to B$ is a \define{half adjoint equivalence} | |
if there are $g:B\to A$ and homotopies $\eta: g \circ f \htpy \idfunc[A]$ and $\epsilon:f \circ g \htpy \idfunc[B]$ such that there exists a homotopy | |
\[\tau : \prd{x:A} \map{f}{\eta x} = \epsilon(fx).\] | |
\end{defn} | |
Thus we have a type $\ishae(f)$, defined to be | |
\begin{equation*} | |
\sm{g:B\to A}{\eta: g \circ f \htpy \idfunc[A]}{\epsilon:f \circ g \htpy \idfunc[B]} \prd{x:A} \map{f}{\eta x} = \epsilon(fx). | |
\end{equation*} | |
Note that in the above definition, the coherence\index{coherence} condition relating $\eta$ and $\epsilon$ only involves $f$. | |
We might consider instead an analogous coherence condition involving $g$: | |
\[\upsilon : \prd{y:B} \map{g}{\epsilon y} = \eta(gy)\] | |
and a resulting analogous definition $\ishae'(f)$. | |
Fortunately, it turns out each of the conditions implies the other one: | |
\begin{lem}\label{lem:coh-equiv} | |
For functions $f : A \to B$ and $g:B\to A$ and homotopies $\eta: g \circ f \htpy \idfunc[A]$ and $\epsilon:f \circ g \htpy \idfunc[B]$, the following conditions are logically equivalent: | |
\begin{itemize} | |
\item $\prd{x:A} \map{f}{\eta x} = \epsilon(fx)$ | |
\item $\prd{y:B} \map{g}{\epsilon y} = \eta(gy)$ | |
\end{itemize} | |
\end{lem} | |
\begin{proof} | |
It suffices to show one direction; the other one is obtained by replacing $A$, $f$, and $\eta$ by $B$, $g$, and $\epsilon$ respectively. | |
Let $\tau : \prd{x:A}\;\map{f}{\eta x} = \epsilon(fx)$. | |
Fix $y : B$. | |
Using naturality of $\epsilon$ and applying $g$, we get the following commuting diagram of paths: | |
\[\uppercurveobject{{ }}\lowercurveobject{{ }}\twocellhead{{ }} | |
\xymatrix@C=3pc{gfgfgy \ar@{=}^-{gfg(\epsilon y)}[r] \ar@{=}_{g(\epsilon (fgy))}[d] & gfgy \ar@{=}^{g(\epsilon y)}[d] \\ gfgy \ar@{=}_{g(\epsilon y)}[r] & gy | |
}\] | |
Using $\tau(gy)$ on the left side of the diagram gives us | |
\[\uppercurveobject{{ }}\lowercurveobject{{ }}\twocellhead{{ }} | |
\xymatrix@C=3pc{gfgfgy \ar@{=}^-{gfg(\epsilon y)}[r] \ar@{=}_{gf(\eta (gy))}[d] & gfgy \ar@{=}^{g(\epsilon y)}[d] \\ gfgy \ar@{=}_{g(\epsilon y)}[r] & gy | |
}\] | |
Using the commutativity of $\eta$ with $g \circ f$ (\cref{cor:hom-fg}), we have | |
\[\uppercurveobject{{ }}\lowercurveobject{{ }}\twocellhead{{ }} | |
\xymatrix@C=3pc{gfgfgy \ar@{=}^-{gfg(\epsilon y)}[r] \ar@{=}_{\eta (gfgy)}[d] & gfgy \ar@{=}^{g(\epsilon y)}[d] \\ gfgy \ar@{=}_{g(\epsilon y)}[r] & gy | |
}\] | |
However, by naturality of $\eta$ we also have | |
\[\uppercurveobject{{ }}\lowercurveobject{{ }}\twocellhead{{ }} | |
\xymatrix@C=3pc{gfgfgy \ar@{=}^-{gfg(\epsilon y)}[r] \ar@{=}_{\eta (gfgy)}[d] & gfgy \ar@{=}^{\eta(gy)}[d] \\ gfgy \ar@{=}_{g(\epsilon y)}[r] & gy | |
}\] | |
Thus, canceling all but the right-hand homotopy, we have $g(\epsilon y) = \eta(g y)$ as desired. | |
\end{proof} | |
However, it is important that we do not include \emph{both} $\tau$ and $\upsilon$ in the definition of $\ishae (f)$ (whence the name ``\emph{half} adjoint equivalence''). | |
If we did, then after canceling contractible types we would still have one remaining datum --- unless we added another higher coherence condition. | |
In general, we expect to get a well-behaved type if we cut off after an odd number of coherences. | |
Of course, it is obvious that $\ishae(f) \to\qinv(f)$: simply forget the coherence datum. | |
The other direction is a version of a standard argument from homotopy theory and category theory. | |
\begin{thm}\label{thm:equiv-iso-adj} | |
For any $f:A\to B$ we have $\qinv(f)\to\ishae(f)$. | |
\end{thm} | |
\begin{proof} | |
Suppose that $(g,\eta,\epsilon)$ is a quasi-inverse for $f$. We have to provide | |
a quadruple $(g',\eta',\epsilon',\tau)$ witnessing that $f$ is a half adjoint equivalence. To | |
define $g'$ and $\eta'$, we can just make the obvious choice by setting $g' | |
\defeq g$ and $\eta'\defeq \eta$. However, in the definition of $\epsilon'$ we | |
need start worrying about the construction of $\tau$, so we cannot just follow our nose | |
and take $\epsilon'$ to be $\epsilon$. Instead, we take | |
\begin{equation*} | |
\epsilon'(b) \defeq \opp{\epsilon(f(g(b)))}\ct (\ap{f}{\eta(g(b))}\ct \epsilon(b)). | |
\end{equation*} | |
Now we need to find | |
\begin{equation*} | |
\tau(a): \ap{f}{\eta(a)}=\opp{\epsilon(f(g(f(a))))}\ct (\ap{f}{\eta(g(f(a)))}\ct \epsilon(f(a))). | |
\end{equation*} | |
Note first that by \cref{cor:hom-fg}, we have | |
%$\eta(g(f(a)))\ct\eta(a)=\ap{g}{\ap{f}{\eta(a)}}\ct\eta(a)$ and hence it follows that | |
$\eta(g(f(a)))=\ap{g}{\ap{f}{\eta(a)}}$. Therefore, we can apply | |
\cref{lem:htpy-natural} to compute | |
\begin{align*} | |
\ap{f}{\eta(g(f(a)))}\ct \epsilon(f(a)) | |
& = \ap{f}{\ap{g}{\ap{f}{\eta(a)}}}\ct \epsilon(f(a))\\ | |
& = \epsilon(f(g(f(a))))\ct \ap{f}{\eta(a)} | |
\end{align*} | |
from which we get the desired path $\tau(a)$. | |
\end{proof} | |
Combining this with \cref{lem:coh-equiv} (or symmetrizing the proof), we also have $\qinv(f)\to\ishae'(f)$. | |
It remains to show that $\ishae(f)$ is a mere proposition. | |
For this, we will need to know that the fibers of an equivalence are contractible. | |
\begin{defn}\label{defn:homotopy-fiber} | |
The \define{fiber} | |
\indexdef{fiber}% | |
\indexsee{function!fiber of}{fiber}% | |
of a map $f:A\to B$ over a point $y:B$ is | |
\[ \hfib f y \defeq \sm{x:A} (f(x) = y).\] | |
\end{defn} | |
In homotopy theory, this is what would be called the \emph{homotopy fiber} of $f$. | |
The path lemmas in \cref{sec:computational} yield the following characterization of paths in fibers: | |
\begin{lem}\label{lem:hfib} | |
For any $f : A \to B$, $y : B$, and $(x,p),(x',p') : \hfib{f}{y}$, we have | |
\[ \big((x,p) = (x',p')\big) \eqvsym \Parens{\sm{\gamma : x = x'} f(\gamma) \ct p' = p} \qedhere\] | |
\end{lem} | |
\begin{thm}\label{thm:contr-hae} | |
If $f:A\to B$ is a half adjoint equivalence, then for any $y:B$ the fiber $\hfib f y$ is contractible. | |
\end{thm} | |
\begin{proof} | |
Let $(g,\eta,\epsilon,\tau) : \ishae(f)$, and fix $y : B$. | |
As our center of contraction for $\hfib{f}{y}$ we choose $(gy, \epsilon y)$. | |
Now take any $(x,p) : \hfib{f}{y}$; we want to construct a path from $(gy, \epsilon y)$ to $(x,p)$. | |
By \cref{lem:hfib}, it suffices to give a path $\gamma : \id{gy}{x}$ such that $\ap f\gamma \ct p = \epsilon y$. | |
We put $\gamma \defeq \opp{g(p)} \ct \eta x$. | |
Then we have | |
\begin{align*} | |
f(\gamma) \ct p & = \opp{fg(p)} \ct f (\eta x) \ct p \\ | |
& = \opp{fg(p)} \ct \epsilon(fx) \ct p \\ | |
& = \epsilon y | |
\end{align*} | |
where the second equality follows by $\tau x$ and the third equality is naturality of $\epsilon$. | |
\end{proof} | |
We now define the types which encapsulate contractible pairs of data. | |
The following types put together the quasi-inverse $g$ with one of the homotopies. | |
\begin{defn}\label{defn:linv-rinv} | |
Given a function $f:A\to B$, we define the types | |
\begin{align*} | |
\linv(f) &\defeq \sm{g:B\to A} (g\circ f\htpy \idfunc[A])\\ | |
\rinv(f) &\defeq \sm{g:B\to A} (f\circ g\htpy \idfunc[B]) | |
\end{align*} | |
of \define{left inverses} | |
\indexdef{left!inverse}% | |
\indexdef{inverse!left}% | |
and \define{right inverses} | |
\indexdef{right!inverse}% | |
\indexdef{inverse!right}% | |
to $f$, respectively. | |
We call $f$ \define{left invertible} | |
\indexdef{function!left invertible}% | |
\indexdef{function!right invertible}% | |
if $\linv(f)$ is inhabited, and similarly \define{right invertible} | |
\indexdef{left!invertible function}% | |
\indexdef{right!invertible function}% | |
if $\rinv(f)$ is inhabited. | |
\end{defn} | |
\begin{lem}\label{thm:equiv-compose-equiv} | |
If $f:A\to B$ has a quasi-inverse, then so do | |
\begin{align*} | |
(f\circ \blank) &: (C\to A) \to (C\to B)\\ | |
(\blank\circ f) &: (B\to C) \to (A\to C). | |
\end{align*} | |
\end{lem} | |
\begin{proof} | |
If $g$ is a quasi-inverse of $f$, then $(g\circ \blank)$ and $(\blank\circ g)$ are quasi-inverses of $(f\circ \blank)$ and $(\blank\circ f)$ respectively. | |
\end{proof} | |
\begin{lem}\label{lem:inv-hprop} | |
If $f : A \to B$ has a quasi-inverse, then the types $\rinv(f)$ and $\linv(f)$ are contractible. | |
\end{lem} | |
\begin{proof} | |
By function extensionality, we have | |
\[\eqv{\linv(f)}{\sm{g:B\to A} (g\circ f = \idfunc[A])}.\] | |
But this is the fiber of $(\blank\circ f)$ over $\idfunc[A]$, and so | |
by \cref{thm:equiv-compose-equiv,thm:equiv-iso-adj,thm:contr-hae}, it is contractible. | |
Similarly, $\rinv(f)$ is equivalent to the fiber of $(f\circ \blank)$ over $\idfunc[B]$ and hence contractible. | |
\end{proof} | |
Next we define the types which put together the other homotopy with the additional coherence datum.\index{coherence}% | |
\begin{defn}\label{defn:lcoh-rcoh} | |
For $f : A \to B$, a left inverse $(g,\eta) : \linv(f)$, and a right inverse $(g,\epsilon) : \rinv(f)$, we denote | |
\begin{align*} | |
\lcoh{f}{g}{\eta} & \defeq \sm{\epsilon : f\circ g \htpy \idfunc[B]} \prd{y:B} g(\epsilon y) = \eta (gy), \\ | |
\rcoh{f}{g}{\epsilon} & \defeq \sm{\eta : g\circ f \htpy \idfunc[A]} \prd{x:A} f(\eta x) = \epsilon (fx). | |
\end{align*} | |
\end{defn} | |
\begin{lem}\label{lem:coh-hfib} | |
For any $f,g,\epsilon,\eta$, we have | |
\begin{align*} | |
\lcoh{f}{g}{\eta} & \eqvsym {\prd{y:B} \id[\hfib{g}{gy}]{(fgy,\eta(gy))}{(y,\refl{gy})}}, \\ | |
\rcoh{f}{g}{\epsilon} & \eqvsym {\prd{x:A} \id[\hfib{f}{fx}]{(gfx,\epsilon(fx))}{(x,\refl{fx})}}. | |
\end{align*} | |
\end{lem} | |
\begin{proof} | |
Using \cref{lem:hfib}. | |
\end{proof} | |
\begin{lem}\label{lem:coh-hprop} | |
If $f$ is a half adjoint equivalence, then for any $(g,\epsilon) : \rinv(f)$, the type $\rcoh{f}{g}{\epsilon}$ is contractible. | |
\end{lem} | |
\begin{proof} | |
By \cref{lem:coh-hfib} and the fact that dependent function types preserve contractible spaces, it suffices to show that for each $x:A$, the type $\id[\hfib{f}{fx}]{(gfx,\epsilon(fx))}{(x,\refl{fx})}$ is contractible. | |
But by \cref{thm:contr-hae}, $\hfib{f}{fx}$ is contractible, and any path space of a contractible space is itself contractible. | |
\end{proof} | |
\begin{thm}\label{thm:hae-hprop} | |
For any $f : A \to B$, the type $\ishae(f)$ is a mere proposition. | |
\end{thm} | |
\begin{proof} | |
By \cref{ex:prop-inhabcontr} it suffices to assume $f$ to be a half adjoint equivalence and show that $\ishae(f)$ is contractible. | |
Now by associativity of $\Sigma$ (\cref{ex:sigma-assoc}), the type $\ishae(f)$ is equivalent to | |
\[\sm{u : \rinv(f)} \rcoh{f}{\proj{1}(u)}{\proj{2}(u)}.\] | |
But by \cref{lem:inv-hprop,lem:coh-hprop} and the fact that $\Sigma$ preserves contractibility, the latter type is also contractible. | |
\end{proof} | |
Thus, we have shown that $\ishae(f)$ has all three desiderata for the type $\isequiv(f)$. | |
In the next two sections we consider a couple of other possibilities. | |
\index{equivalence!half adjoint|)}% | |
\index{half adjoint equivalence|)}% | |
\index{adjoint!equivalence!of types, half|)}% | |
\section{Bi-invertible maps} | |
\label{sec:biinv} | |
\index{function!bi-invertible|(defstyle}% | |
\index{bi-invertible function|(defstyle}% | |
\index{equivalence!as bi-invertible function|(defstyle}% | |
Using the language introduced in \cref{sec:hae}, we can restate the definition proposed in \cref{sec:basics-equivalences} as follows. | |
\begin{defn}\label{defn:biinv} | |
We say $f:A\to B$ is \define{bi-invertible} | |
if it has both a left inverse and a right inverse: | |
\[ \biinv (f) \defeq \linv(f) \times \rinv(f). \] | |
\end{defn} | |
In \cref{sec:basics-equivalences} we proved that $\qinv(f)\to\biinv(f)$ and $\biinv(f)\to\qinv(f)$. | |
What remains is the following. | |
\begin{thm}\label{thm:isprop-biinv} | |
For any $f:A\to B$, the type $\biinv(f)$ is a mere proposition. | |
\end{thm} | |
\begin{proof} | |
We may suppose $f$ to be bi-invertible and show that $\biinv(f)$ is contractible. | |
But since $\biinv(f)\to\qinv(f)$, by \cref{lem:inv-hprop} in this case both $\linv(f)$ and $\rinv(f)$ are contractible, and the product of contractible types is contractible. | |
\end{proof} | |
Note that this also fits the proposal made at the beginning of \cref{sec:hae}: we combine $g$ and $\eta$ into a contractible type and add an additional datum which combines with $\epsilon$ into a contractible type. | |
The difference is that instead of adding a \emph{higher} datum (a 2-dimensional path) to combine with $\epsilon$, we add a \emph{lower} one (a right inverse that is separate from the left inverse). | |
\begin{cor}\label{thm:equiv-biinv-isequiv} | |
For any $f:A\to B$ we have $\eqv{\biinv(f)}{\ishae(f)}$. | |
\end{cor} | |
\begin{proof} | |
We have $\biinv(f) \to \qinv(f) \to \ishae(f)$ and $\ishae(f) \to \qinv(f) \to \biinv(f)$. | |
Since both $\ishae(f)$ and $\biinv(f)$ are mere propositions, the equivalence follows from \cref{lem:equiv-iff-hprop}. | |
\end{proof} | |
\index{function!bi-invertible|)}% | |
\index{bi-invertible function|)}% | |
\index{equivalence!as bi-invertible function|)}% | |
\section{Contractible fibers} | |
\label{sec:contrf} | |
\index{function!contractible|(defstyle}% | |
\index{contractible!function|(defstyle}% | |
\index{equivalence!as contractible function|(defstyle}% | |
Note that our proofs about $\ishae(f)$ and $\biinv(f)$ made essential use of the fact that the fibers of an equivalence are contractible. | |
In fact, it turns out that this property is itself a sufficient definition of equivalence. | |
\begin{defn}[Contractible maps] \label{defn:equivalence} | |
A map $f:A\to B$ is \define{contractible} | |
if for all $y:B$, the fiber $\hfib f y$ is contractible. | |
\end{defn} | |
Thus, the type $\iscontr(f)$ is defined to be | |
\begin{align} | |
\iscontr(f) &\defeq \prd{y:B} \iscontr(\hfib f y)\label{eq:iscontrf} | |
% \\ | |
% &\defeq \prd{y:B} \iscontr (\setof{x:A | f(x) = y}). | |
\end{align} | |
Note that in \cref{sec:contractibility} we defined what it means for a \emph{type} to be contractible. | |
Here we are defining what it means for a \emph{map} to be contractible. | |
Our terminology follows the general homotopy-theoretic practice of saying that a map has a certain property if all of its (homotopy) fibers have that property. | |
Thus, a type $A$ is contractible just when the map $A\to\unit$ is contractible. | |
From \cref{cha:hlevels} onwards we will also call contractible maps and types \emph{$(-2)$-truncated}. | |
We have already shown in \cref{thm:contr-hae} that $\ishae(f) \to \iscontr(f)$. | |
Conversely: | |
\begin{thm}\label{thm:lequiv-contr-hae} | |
For any $f:A\to B$ we have ${\iscontr(f)} \to {\ishae(f)}$. | |
\end{thm} | |
\begin{proof} | |
Let $P : \iscontr(f)$. We define an inverse mapping $g : B \to A$ by sending each $y : B$ to the center of contraction of the fiber at $y$: | |
\[ g(y) \defeq \proj{1}(\proj{1}(Py)). \] | |
We can thus define the homotopy $\epsilon$ by mapping $y$ to the witness that $g(y)$ indeed belongs to the fiber at $y$: | |
\[ \epsilon(y) \defeq \proj{2}(\proj{1}(P y)). \] | |
It remains to define $\eta$ and $\tau$. This of course amounts to giving an element of $\rcoh{f}{g}{\epsilon}$. By \cref{lem:coh-hfib}, this is the same as giving for each $x:A$ a path from $(gfx,\epsilon(fx))$ to $(x,\refl{fx})$ in the fiber of $f$ over $fx$. But this is easy: for any $x : A$, the type $\hfib{f}{fx}$ | |
is contractible by assumption, hence such a path must exist. We can construct it explicitly as | |
\[\opp{\big(\proj{2}(P(fx))(gfx,\epsilon(fx))\big)} \ct \big(\proj{2}(P(fx)) (x,\refl{fx})\big). \qedhere \] | |
\end{proof} | |
It is also easy to see: | |
\begin{lem}\label{thm:contr-hprop} | |
For any $f$, the type $\iscontr(f)$ is a mere proposition. | |
\end{lem} | |
\begin{proof} | |
By \cref{thm:isprop-iscontr}, each type $\iscontr (\hfib f y)$ is a mere proposition. | |
Thus, by \cref{thm:isprop-forall}, so is~\eqref{eq:iscontrf}. | |
\end{proof} | |
\begin{thm}\label{thm:equiv-contr-hae} | |
For any $f:A\to B$ we have $\eqv{\iscontr(f)}{\ishae(f)}$. | |
\end{thm} | |
\begin{proof} | |
We have already established a logical equivalence ${\iscontr(f)} \Leftrightarrow {\ishae(f)}$, and both are mere propositions (\cref{thm:contr-hprop,thm:hae-hprop}). | |
Thus, \cref{lem:equiv-iff-hprop} applies. | |
\end{proof} | |
Usually, we prove that a function is an equivalence by exhibiting a quasi-inverse, but sometimes this definition is more convenient. | |
For instance, it implies that when proving a function to be an equivalence, we are free to assume that its codomain is inhabited. | |
\begin{cor}\label{thm:equiv-inhabcod} | |
If $f:A\to B$ is such that $B\to \isequiv(f)$, then $f$ is an equivalence. | |
\end{cor} | |
\begin{proof} | |
To show $f$ is an equivalence, it suffices to show that $\hfib f y$ is contractible for any $y:B$. | |
But if $e:B\to \isequiv(f)$, then given any such $y$ we have $e(y):\isequiv(f)$, so that $f$ is an equivalence and hence $\hfib f y$ is contractible, as desired. | |
\end{proof} | |
\index{function!contractible|)}% | |
\index{contractible!function|)}% | |
\index{equivalence!as contractible function|)}% | |
\section{On the definition of equivalences} | |
\label{sec:concluding-remarks} | |
\indexdef{equivalence} | |
We have shown that all three definitions of equivalence satisfy the three desirable properties and are pairwise equivalent: | |
\[ \iscontr(f) \eqvsym \ishae(f) \eqvsym \biinv(f). \] | |
(There are yet more possible definitions of equivalence, but we will stop with these three. | |
See \cref{ex:brck-qinv} and the exercises in this chapter for some more.) | |
Thus, we may choose any one of them as ``the'' definition of $\isequiv (f)$. | |
For definiteness, we choose to define | |
\[ \isequiv(f) \defeq \ishae(f).\] | |
\index{mathematics!formalized}% | |
This choice is advantageous for formalization, since $\ishae(f)$ contains the most directly useful data. | |
On the other hand, for other purposes, $\biinv(f)$ is often easier to deal with, since it contains no 2-dimensional paths and its two symmetrical halves can be treated independently. | |
However, for purposes of this book, the specific choice will make little difference. | |
In the rest of this chapter, we study some other properties and characterizations of equivalences. | |
\index{equivalence!properties of}% | |
\section{Surjections and embeddings} | |
\label{sec:mono-surj} | |
\index{set} | |
When $A$ and $B$ are sets and $f:A\to B$ is an equivalence, we also call it as \define{isomorphism} | |
\indexdef{isomorphism!of sets}% | |
or a \define{bijection}. | |
\indexdef{bijection}% | |
\indexsee{function!bijective}{bijection}% | |
(We avoid these words for types that are not sets, since in homotopy theory and higher category theory they often denote a stricter notion of ``sameness'' than homotopy equivalence.) | |
In set theory, a function is a bijection just when it is both injective and surjective. | |
The same is true in type theory, if we formulate these conditions appropriately. | |
For clarity, when dealing with types that are not sets, we will speak of \emph{embeddings} instead of injections. | |
\begin{defn}\label{defn:surj-emb} | |
Let $f:A\to B$. | |
\begin{enumerate} | |
\item We say $f$ is \define{surjective} | |
\indexsee{surjective!function}{function, surjective}% | |
\indexdef{function!surjective}% | |
(or a \define{surjection}) | |
\indexsee{surjection}{function, surjective}% | |
if for every $b:B$ we have $\brck{\hfib f b}$. | |
\item We say $f$ is an \define{embedding} | |
\indexdef{function!embedding}% | |
\indexsee{embedding}{function, embedding}% | |
if for every $x,y:A$ the function $\apfunc f : (\id[A]xy) \to (\id[B]{f(x)}{f(y)})$ is an equivalence. | |
\end{enumerate} | |
\end{defn} | |
In other words, $f$ is surjective if every fiber of $f$ is merely inhabited, or equivalently if for all $b:B$ there merely exists an $a:A$ such that $f(a)=b$. | |
In traditional logical notation, $f$ is surjective if $\fall{b:B}\exis{a:A} (f(a)=b)$. | |
This must be distinguished from the stronger assertion that $\prd{b:B}\sm{a:A} (f(a)=b)$; if this holds we say that $f$ is a \define{split surjection}. | |
\indexsee{split!surjection}{function, split surjective}% | |
\indexsee{surjection!split}{function, split surjective}% | |
\indexsee{surjective!function!split}{function, split surjective}% | |
\indexdef{function!split surjective}% | |
(Since this latter type is equivalent to $\sm{g:B\to A}\prd{b:B} (f(g(b))=b)$, being a split surjection is the same as being a \emph{retraction} as defined in \cref{sec:contractibility}.) | |
\index{retraction}% | |
\index{function!retraction}% | |
The axiom of choice from \cref{sec:axiom-choice} says exactly that every surjection \emph{between sets} is split. | |
However, in the presence of the univalence axiom, it is simply false that \emph{all} surjections are split. | |
In \cref{thm:no-higher-ac} we constructed a type family $Y:X\to \type$ such that $\prd{x:X} \brck{Y(x)}$ but $\neg \prd{x:X} Y(x)$; | |
for any such family, the first projection $(\sm{x:X} Y(x)) \to X$ is a surjection that is not split. | |
If $A$ and $B$ are sets, then by \cref{lem:equiv-iff-hprop}, $f$ is an embedding just when | |
\begin{equation} | |
\prd{x,y:A} (\id[B]{f(x)}{f(y)}) \to (\id[A]xy).\label{eq:injective} | |
\end{equation} | |
In this case we say that $f$ is \define{injective}, | |
\indexsee{injective function}{function, injective}% | |
\indexdef{function!injective}% | |
or an \define{injection}. | |
\indexsee{injection}{function, injective}% | |
We avoid these word for types that are not sets, because they might be interpreted as~\eqref{eq:injective}, which is an ill-behaved notion for non-sets. | |
It is also true that any function between sets is surjective if and only if it is an \emph{epimorphism} in a suitable sense, but this also fails for more general types, and surjectivity is generally the more important notion. | |
\begin{thm}\label{thm:mono-surj-equiv} | |
A function $f:A\to B$ is an equivalence if and only if it is both surjective and an embedding. | |
\end{thm} | |
\begin{proof} | |
If $f$ is an equivalence, then each $\hfib f b$ is contractible, hence so is $\brck{\hfib f b}$, so $f$ is surjective. | |
And we showed in \cref{thm:paths-respects-equiv} that any equivalence is an embedding. | |
Conversely, suppose $f$ is a surjective embedding. | |
Let $b:B$; we show that $\sm{x:A}(f(x)=b)$ is contractible. | |
Since $f$ is surjective, there merely exists an $a:A$ such that $f(a)=b$. | |
Thus, the fiber of $f$ over $b$ is inhabited; it remains to show it is a mere proposition. | |
For this, suppose given $x,y:A$ with $p:f(x)=b$ and $q:f(y)=b$. | |
Then since $\apfunc f$ is an equivalence, there exists $r:x=y$ with $\apfunc f (r) = p \ct \opp q$. | |
However, using the characterization of paths in $\Sigma$-types, the latter equality rearranges to $\trans{r}{p} = q$. | |
Thus, together with $r$ it exhibits $(x,p) = (y,q)$ in the fiber of $f$ over $b$. | |
\end{proof} | |
\begin{cor} | |
For any $f:A\to B$ we have | |
\[ \isequiv(f) \eqvsym (\mathsf{isEmbedding}(f) \times \mathsf{isSurjective}(f)).\] | |
\end{cor} | |
\begin{proof} | |
Being a surjection and an embedding are both mere propositions; now apply \cref{lem:equiv-iff-hprop}. | |
\end{proof} | |
Of course, this cannot be used as a definition of ``equivalence'', since the definition of embeddings refers to equivalences. | |
However, this characterization can still be useful; see \cref{sec:whitehead}. | |
We will generalize it in \cref{cha:hlevels}. | |
% \section{Fiberwise equivalences} | |
\section{Closure properties of equivalences} | |
\label{sec:equiv-closures} | |
\label{sec:fiberwise-equivalences} | |
\index{equivalence!properties of}% | |
% We end this chapter by observing some important closure properties of equivalences. | |
We have already seen in \cref{thm:equiv-eqrel} that equivalences are closed under composition. | |
Furthermore, we have: | |
\begin{thm}[The 2-out-of-3 property]\label{thm:two-out-of-three} | |
\index{2-out-of-3 property}% | |
Suppose $f:A\to B$ and $g:B\to C$. | |
If any two of $f$, $g$, and $g\circ f$ are equivalences, so is the third. | |
\end{thm} | |
\begin{proof} | |
If $g\circ f$ and $g$ are equivalences, then $\opp{(g\circ f)} \circ g$ is a quasi-inverse to $f$. | |
On the one hand, we have $\opp{(g\circ f)} \circ g \circ f \htpy \idfunc[A]$, while on the other we have | |
\begin{align*} | |
f \circ \opp{(g\circ f)} \circ g | |
&\htpy \opp g \circ g \circ f \circ \opp{(g\circ f)} \circ g\\ | |
&\htpy \opp g \circ g\\ | |
&\htpy \idfunc[B]. | |
\end{align*} | |
Similarly, if $g\circ f$ and $f$ are equivalences, then $f\circ \opp{(g\circ f)}$ is a quasi-inverse to $g$. | |
\end{proof} | |
This is a standard closure condition on equivalences from homotopy theory. | |
Also well-known is that they are closed under retracts, in the following sense. | |
\index{retract!of a function|(defstyle}% | |
\begin{defn}\label{defn:retract} | |
A function $g:A\to B$ is said to be a \define{retract} | |
of a function $f:X\to Y$ if there is a diagram | |
\begin{equation*} | |
\xymatrix{ | |
{A} \ar[r]^{s} \ar[d]_{g} | |
& | |
{X} \ar[r]^{r} \ar[d]_{f} | |
& | |
{A} \ar[d]^{g} | |
\\ | |
{B} \ar[r]_{s'} | |
& | |
{Y} \ar[r]_{r'} | |
& | |
{B} | |
} | |
\end{equation*} | |
for which there are | |
\begin{enumerate} | |
\item a homotopy $R:r\circ s \htpy \idfunc[A]$. | |
\item a homotopy $R':r'\circ s' \htpy\idfunc[B]$. | |
\item a homotopy $L:f\circ s\htpy s'\circ g$. | |
\item a homotopy $K:g\circ r\htpy r'\circ f$. | |
\item for every $a:A$, a path $H(a)$ witnessing the commutativity of the square | |
\begin{equation*} | |
\xymatrix@C=3pc{ | |
{g(r(s(a)))} \ar@{=}[r]^-{K(s(a))} \ar@{=}[d]_{\ap g{R(a)}} | |
& | |
{r'(f(s(a)))} \ar@{=}[d]^{\ap{r'}{L(a)}} | |
\\ | |
{g(a)} \ar@{=}[r]_-{\opp{R'(g(a))}} | |
& | |
{r'(s'(g(a)))} | |
} | |
\end{equation*} | |
\end{enumerate} | |
\end{defn} | |
Recall that in \cref{sec:contractibility} we defined what it means for a type to be a retract of another. | |
This is a special case of the above definition where $B$ and $Y$ are $\unit$. | |
Conversely, just as with contractibility, retractions of maps induce retractions of their fibers. | |
\begin{lem}\label{lem:func_retract_to_fiber_retract} | |
If a function $g:A\to B$ is a retract of a function $f:X\to Y$, then $\hfib{g}b$ is a retract of $\hfib{f}{s'(b)}$ | |
for every $b:B$, where $s':B\to Y$ is as in \cref{defn:retract}. | |
\end{lem} | |
\begin{proof} | |
Suppose that $g:A\to B$ is a retract of $f:X\to Y$. Then for any $b:B$ we have the functions | |
\begin{align*} | |
\varphi_b &:\hfiber{g}b\to\hfib{f}{s'(b)}, & | |
\varphi_b(a,p) & \defeq \pairr{s(a),L(a)\ct s'(p)},\\ | |
\psi_b &:\hfib{f}{s'(b)}\to\hfib{g}b, & | |
\psi_b(x,q) &\defeq \pairr{r(x),K(x)\ct r'(q)\ct R'(b)}. | |
\end{align*} | |
Then we have $\psi_b(\varphi_b({a,p}))\equiv\pairr{r(s(a)),K(s(a))\ct r'(L(a)\ct s'(p))\ct R'(b)}$. | |
We claim $\psi_b$ is a retraction with section $\varphi_b$ for all $b:B$, which is to say that for all $(a,p):\hfib g b$ we have $\psi_b(\varphi_b({a,p}))= \pairr{a,p}$. | |
In other words, we want to show | |
\begin{equation*} | |
\prd{b:B}{a:A}{p:g(a)=b} \psi_b(\varphi_b({a,p}))= \pairr{a,p}. | |
\end{equation*} | |
By reordering the first two $\Pi$s and applying a version of \cref{thm:omit-contr}, this is equivalent to | |
\begin{equation*} | |
\prd{a:A}\psi_{g(a)}(\varphi_{g(a)}({a,\refl{g(a)}}))=\pairr{a,\refl{g(a)}}. | |
\end{equation*} | |
For any $a$, by \cref{thm:path-sigma}, this equality of pairs is equivalent to a pair of equalities. The first components are equal by $R(a):r(s(a))= a$, so we need only show | |
\begin{equation*} | |
\trans{R(a)}{K(s(a))\ct r'(L(a))\ct R'(g(a))} = \refl{g(a)}. | |
\end{equation*} | |
But this transportation computes as $\opp{g(R(a))}\ct K(s(a))\ct r'(L(a))\ct R'(g(a))$, so the required path is given by $H(a)$. | |
\end{proof} | |
\begin{thm}\label{thm:retract-equiv} | |
If $g$ is a retract of an equivalence $f$, then $g$ is also an equivalence. | |
\end{thm} | |
\begin{proof} | |
By \cref{lem:func_retract_to_fiber_retract}, every fiber of $g$ is a retract of a fiber of $f$. | |
Thus, by \cref{thm:retract-contr}, if the latter are all contractible, so are the former. | |
\end{proof} | |
\index{retract!of a function|)}% | |
\index{fibration}% | |
\index{total!space}% | |
Finally, we show that fiberwise equivalences can be characterized in terms of equivalences of total spaces. | |
To explain the terminology, recall from \cref{sec:fibrations} that a type family $P:A\to\type$ can be viewed as a fibration over $A$ with total space $\sm{x:A} P(x)$, the fibration being the projection $\proj1:\sm{x:A} P(x) \to A$. | |
From this point of view, given two type families $P,Q:A\to\type$, we may refer to a function $f:\prd{x:A} (P(x)\to Q(x))$ as a \define{fiberwise map} or a \define{fiberwise transformation}. | |
\indexsee{transformation!fiberwise}{fiberwise transformation}% | |
\indexsee{function!fiberwise}{fiberwise transformation}% | |
\index{fiberwise!transformation|(defstyle}% | |
\indexsee{fiberwise!map}{fiberwise transformation}% | |
\indexsee{map!fiberwise}{fiberwise transformation} | |
Such a map induces a function on total spaces: | |
\begin{defn}\label{defn:total-map} | |
Given type families $P,Q:A\to\type$ and a map $f:\prd{x:A} P(x)\to Q(x)$, we define | |
\begin{equation*} | |
\total f \defeq \lam{w}\pairr{\proj{1}w,f(\proj{1}w,\proj{2}w)} : \sm{x:A}P(x)\to\sm{x:A}Q(x). | |
\end{equation*} | |
\end{defn} | |
\begin{thm}\label{fibwise-fiber-total-fiber-equiv} | |
Suppose that $f$ is a fiberwise transformation between families $P$ and | |
$Q$ over a type $A$ and let $x:A$ and $v:Q(x)$. Then we have an equivalence | |
\begin{equation*} | |
\eqv{\hfib{\total{f}}{\pairr{x,v}}}{\hfib{f(x)}{v}}. | |
\end{equation*} | |
\end{thm} | |
\begin{proof} | |
We calculate: | |
\begin{align} | |
\hfib{\total{f}}{\pairr{x,v}} | |
& \jdeq \sm{w:\sm{x:A}P(x)}\pairr{\proj{1}w,f(\proj{1}w,\proj{2}w)}=\pairr{x,v} | |
\notag \\ | |
& \eqv{}{} \sm{a:A}{u:P(a)}\pairr{a,f(a,u)}=\pairr{x,v} | |
\tag{by~\cref{ex:sigma-assoc}} \\ | |
& \eqv{}{} \sm{a:A}{u:P(a)}{p:a=x}\trans{p}{f(a,u)}=v | |
\tag{by \cref{thm:path-sigma}} \\ | |
& \eqv{}{} \sm{a:A}{p:a=x}{u:P(a)}\trans{p}{f(a,u)}=v | |
\notag \\ | |
& \eqv{}{} \sm{u:P(x)}f(x,u)=v | |
\tag{$*$}\label{eq:uses-sum-over-paths} \\ | |
& \jdeq \hfib{f(x)}{v}. \notag | |
\end{align} | |
The equivalence~\eqref{eq:uses-sum-over-paths} follows from \cref{thm:omit-contr,thm:contr-paths,ex:sigma-assoc}. | |
\end{proof} | |
We say that a fiberwise transformation $f:\prd{x:A} P(x)\to Q(x)$ is a \define{fiberwise equivalence}% | |
\indexdef{fiberwise!equivalence}% | |
\indexdef{equivalence!fiberwise} | |
if each $f(x):P(x) \to Q(x)$ is an equivalence. | |
\begin{thm}\label{thm:total-fiber-equiv} | |
Suppose that $f$ is a fiberwise transformation between families | |
$P$ and $Q$ over a type $A$. | |
Then $f$ is a fiberwise equivalence if and only if $\total{f}$ is an equivalence. | |
\end{thm} | |
\begin{proof} | |
Let $f$, $P$, $Q$ and $A$ be as in the statement of the theorem. | |
By \cref{fibwise-fiber-total-fiber-equiv} it follows for all | |
$x:A$ and $v:Q(x)$ that | |
$\hfib{\total{f}}{\pairr{x,v}}$ is contractible if and only if | |
$\hfib{f(x)}{v}$ is contractible. | |
Thus, $\hfib{\total{f}}{w}$ is contractible for all $w:\sm{x:A}Q(x)$ if and only if $\hfib{f(x)}{v}$ is contractible for all $x:A$ and $v:Q(x)$. | |
\end{proof} | |
\index{fiberwise!transformation|)}% | |
\section{The object classifier} | |
\label{sec:object-classification} | |
In type theory we have a basic notion of \emph{family of types}, namely a function $B:A\to\type$. | |
We have seen that such families behave somewhat like \emph{fibrations} in homotopy theory, with the fibration being the projection $\proj1:\sm{a:A} B(a) \to A$. | |
A basic fact in homotopy theory is that every map is equivalent to a fibration. | |
With univalence at our disposal, we can prove the same thing in type theory. | |
\begin{lem}\label{thm:fiber-of-a-fibration} | |
For any type family $B:A\to\type$, the fiber of $\proj1:\sm{x:A} B(x) \to A$ over $a:A$ is equivalent to $B(a)$: | |
\[ \eqv{\hfib{\proj1}{a}}{B(a)} \] | |
\end{lem} | |
\begin{proof} | |
We have | |
\begin{align*} | |
\hfib{\proj1}{a} &\defeq \sm{u:\sm{x:A} B(x)} \proj1(u)=a\\ | |
&\eqvsym \sm{x:A}{b:B(x)} (x=a)\\ | |
&\eqvsym \sm{x:A}{p:x=a} B(x)\\ | |
&\eqvsym B(a) | |
\end{align*} | |
using the left universal property of identity types. | |
\end{proof} | |
\begin{lem}\label{thm:total-space-of-the-fibers} | |
For any function $f:A\to B$, we have $\eqv{A}{\sm{b:B}\hfib{f}{b}}$. | |
\end{lem} | |
\begin{proof} | |
We have | |
\begin{align*} | |
\sm{b:B}\hfib{f}{b} &\defeq \sm{b:B}{a:A} (f(a)=b)\\ | |
&\eqvsym \sm{a:A}{b:B} (f(a)=b)\\ | |
&\eqvsym A | |
\end{align*} | |
using the fact that $\sm{b:B} (f(a)=b)$ is contractible. | |
\end{proof} | |
\begin{thm}\label{thm:nobject-classifier-appetizer} | |
For any type $B$ there is an equivalence | |
\begin{equation*} | |
\chi:\Parens{\sm{A:\type} (A\to B)}\eqvsym (B\to\type). | |
\end{equation*} | |
\end{thm} | |
\begin{proof} | |
We have to construct quasi-inverses | |
\begin{align*} | |
\chi & : \Parens{\sm{A:\type} (A\to B)}\to B\to\type\\ | |
\psi & : (B\to\type)\to\Parens{\sm{A:\type} (A\to B)}. | |
\end{align*} | |
We define $\chi$ by $\chi((A,f),b)\defeq\hfiber{f}b$, and $\psi$ by $\psi(P)\defeq\Pairr{(\sm{b:B} P(b)),\proj1}$. | |
Now we have to verify that $\chi\circ\psi\htpy\idfunc{}$ and that $\psi\circ\chi \htpy\idfunc{}$. | |
\begin{enumerate} | |
\item Let $P:B\to\type$. | |
By \cref{thm:fiber-of-a-fibration}, | |
$\hfiber{\proj1}{b}\eqvsym P(b)$ for any $b:B$, so it follows immediately | |
that $P\htpy\chi(\psi(P))$. | |
\item Let $f:A\to B$ be a function. We have to find a path | |
\begin{equation*} | |
\Pairr{\tsm{b:B} \hfiber{f}b,\,\proj1}=\pairr{A,f}. | |
\end{equation*} | |
First note that by \cref{thm:total-space-of-the-fibers}, we have | |
$e:\sm{b:B} \hfiber{f}b\eqvsym A$ with $e(b,a,p)\defeq a$ and $e^{-1}(a) | |
\defeq(f(a),a,\refl{f(a)})$. | |
By \cref{thm:path-sigma}, it remains to show $\trans{(\ua(e))}{\proj1} = f$. | |
But by the computation rule for univalence and~\eqref{eq:transport-arrow}, we have $\trans{(\ua(e))}{\proj1} = \proj1\circ e^{-1}$, and the definition of $e^{-1}$ immediately yields $\proj1 \circ e^{-1} \jdeq f$.\qedhere | |
\end{enumerate} | |
\end{proof} | |
\noindent | |
\indexdef{object!classifier}% | |
\indexdef{classifier!object}% | |
\index{.infinity1-topos@$(\infty,1)$-topos}% | |
In particular, this implies that we have an \emph{object classifier} in the sense of higher topos theory. | |
Recall from \cref{def:pointedtype} that $\pointed\type$ denotes the type $\sm{A:\type} A$ of pointed types. | |
\begin{thm}\label{thm:object-classifier} | |
Let $f:A\to B$ be a function. Then the diagram | |
\begin{equation*} | |
\vcenter{\xymatrix{ | |
A\ar[r]^-{\vartheta_f} \ar[d]_{f} & | |
\pointed{\type}\ar[d]^{\proj1}\\ | |
B\ar[r]_{\chi_f} & | |
\type | |
}} | |
\end{equation*} | |
is a pullback\index{pullback} square (see \cref{ex:pullback}). | |
Here the function $\vartheta_f$ is defined by | |
\begin{equation*} | |
\lam{a} \pairr{\hfiber{f}{f(a)},\pairr{a,\refl{f(a)}}}. | |
\end{equation*} | |
\end{thm} | |
\begin{proof} | |
Note that we have the equivalences | |
\begin{align*} | |
A & \eqvsym \sm{b:B} \hfiber{f}b\\ | |
& \eqvsym \sm{b:B}{X:\type}{p:\hfiber{f}b= X} X\\ | |
& \eqvsym \sm{b:B}{X:\type}{x:X} \hfiber{f}b= X\\ | |
& \eqvsym \sm{b:B}{Y:\pointed{\type}} \hfiber{f}b = \proj1 Y\\ | |
& \jdeq B\times_{\type}\pointed{\type} | |
\end{align*} | |
which gives us a composite equivalence $e:A\eqvsym B\times_\type\pointed{\type}$. | |
We may display the action of this composite equivalence step by step by | |
\begin{align*} | |
a & \mapsto \pairr{f(a),\; \pairr{a,\refl{f(a)}}}\\ | |
& \mapsto \pairr{f(a), \; \hfiber{f}{f(a)}, \; \refl{\hfiber{f}{f(a)}}, \; \pairr{a,\refl{f(a)}}}\\ | |
& \mapsto \pairr{f(a), \; \hfiber{f}{f(a)}, \; \pairr{a,\refl{f(a)}}, \; \refl{\hfiber{f}{f(a)}}}. | |
\end{align*} | |
Therefore, we get homotopies $f\htpy\proj1\circ e$ and $\vartheta_f\htpy \proj2\circ e$. | |
\end{proof} | |
\section{Univalence implies function extensionality} | |
\label{sec:univalence-implies-funext} | |
\index{function extensionality!proof from univalence}% | |
In the last section of this chapter we include a proof that the univalence axiom implies function | |
extensionality. Thus, in this section we work \emph{without} the function extensionality axiom. | |
The proof consists of two steps. First we show | |
in \cref{uatowfe} that the univalence | |
axiom implies a weak form of function extensionality, defined in \cref{weakfunext} below. The | |
principle of weak function extensionality in turn implies the usual function extensionality, | |
and it does so without the univalence axiom (\cref{wfetofe}). | |
\index{univalence axiom}% | |
Let $\type$ be a universe; we will explicitly indicate where we assume that it is univalent. | |
\begin{defn}\label{weakfunext} | |
The \define{weak function extensionality principle} | |
\indexdef{function extensionality!weak}% | |
asserts that there is a function | |
\begin{equation*} | |
\Parens{\prd{x:A}\iscontr(P(x))} \to\iscontr\Parens{\prd{x:A}P(x)} | |
\end{equation*} | |
for any family $P:A\to\type$ of types over any type $A$. | |
\end{defn} | |
The following lemma is easy to prove using function extensionality; the point here is that it also follows from univalence without assuming function extensionality separately. | |
\begin{lem} \label{UA-eqv-hom-eqv} | |
Assuming $\type$ is univalent, for any $A,B,X:\type$ and any $e:\eqv{A}{B}$, there is an equivalence | |
\begin{equation*} | |
\eqv{(X\to A)}{(X\to B)} | |
\end{equation*} | |
of which the underlying map is given by post-composition with the underlying function of $e$. | |
\end{lem} | |
\begin{proof} | |
% Immediate by induction on $\eqv{}{}$ (see \cref{thm:equiv-induction}). | |
As in the proof of \cref{lem:qinv-autohtpy}, we may assume that $e = \idtoeqv(p)$ for some $p:A=B$. | |
Then by path induction, we may assume $p$ is $\refl{A}$, so that $e = \idfunc[A]$. | |
But in this case, post-composition with $e$ is the identity, hence an equivalence. | |
\end{proof} | |
\begin{cor}\label{contrfamtotalpostcompequiv} | |
Let $P:A\to\type$ be a family of contractible types, i.e.\ \narrowequation{\prd{x:A}\iscontr(P(x)).} | |
Then the projection $\proj{1}:(\sm{x:A}P(x))\to A$ is an equivalence. Assuming $\type$ is univalent, it follows immediately that post-composition with $\proj{1}$ gives an equivalence | |
\begin{equation*} | |
\alpha : \eqv{\Parens{A\to\sm{x:A}P(x)}}{(A\to A)}. | |
\end{equation*} | |
\end{cor} | |
\begin{proof} | |
By \cref{thm:fiber-of-a-fibration}, for $\proj{1}:\sm{x:A}P(X)\to A$ and $x:A$ we have an equivalence | |
\begin{equation*} | |
\eqv{\hfiber{\proj{1}}{x}}{P(x)}. | |
\end{equation*} | |
Therefore $\proj{1}$ is an equivalence whenever each $P(x)$ is contractible. The assertion is now a consequence of \cref{UA-eqv-hom-eqv}. | |
\end{proof} | |
In particular, the homotopy fiber of the above equivalence at $\idfunc[A]$ is contractible. Therefore, we can show that univalence implies weak function extensionality by showing that the dependent function type $\prd{x:A}P(x)$ is a retract of $\hfiber{\alpha}{\idfunc[A]}$. | |
\begin{thm}\label{uatowfe} | |
In a univalent universe $\type$, suppose that $P:A\to\type$ is a family of contractible types | |
and let $\alpha$ be the function of \cref{contrfamtotalpostcompequiv}. | |
Then $\prd{x:A}P(x)$ is a retract of $\hfiber{\alpha}{\idfunc[A]}$. As a consequence, $\prd{x:A}P(x)$ is contractible. In other words, the univalence axiom implies the weak function extensionality principle. | |
\end{thm} | |
\begin{proof} | |
Define the functions | |
\begin{align*} | |
\varphi &: (\tprd{x:A}P(x))\to\hfiber{\alpha}{\idfunc[A]},\\ | |
\varphi(f) &\defeq (\lam{x} (x,f(x)),\refl{\idfunc[A]}), | |
\intertext{and} | |
\psi &: \hfiber{\alpha}{\idfunc[A]}\to \tprd{x:A}P(x), \\ | |
\psi(g,p) &\defeq \lam{x} \trans {\happly (p,x)}{\proj{2} (g(x))}. | |
\end{align*} | |
Then $\psi(\varphi(f))=\lam{x} f(x)$, which is $f$, by the uniqueness principle for dependent function types. | |
\end{proof} | |
We now show that weak function extensionality implies the usual function extensionality. | |
Recall from~\eqref{eq:happly} the function $\happly (f,g) : (f = g)\to(f\htpy g)$ which | |
converts equality of functions to homotopy. In the proof that follows, the univalence | |
axiom is not used. | |
\begin{thm}\label{wfetofe} | |
\index{function extensionality}% | |
Weak function extensionality implies the function extensionality \cref{axiom:funext}. | |
\end{thm} | |
\begin{proof} | |
We want to show that | |
\begin{equation*} | |
\prd{A:\type}{P:A\to\type}{f,g:\prd{x:A}P(x)}\isequiv(\happly (f,g)). | |
\end{equation*} | |
Since a fiberwise map induces an equivalence on total spaces if and only if it is fiberwise an equivalence by \cref{thm:total-fiber-equiv}, it suffices to show that the function of type | |
\begin{equation*} | |
\Parens{\sm{g:\prd{x:A}P(x)}(f= g)} \to \sm{g:\prd{x:A}P(x)}(f\htpy g) | |
\end{equation*} | |
induced by $\lam{g:\prd{x:A}P(x)} \happly (f,g)$ is an equivalence. | |
Since the type on the left is contractible by \cref{thm:contr-paths}, it suffices to show that the type on the right: | |
\begin{equation}\label{eq:uatofesp} | |
\sm{g:\prd{x:A}P(x)}\prd{x:A}f(x)= g(x) | |
\end{equation} | |
is contractible. | |
Now \cref{thm:ttac} says that this is equivalent to | |
\begin{equation}\label{eq:uatofeps} | |
\prd{x:A}\sm{u:P(x)}f(x)= u. | |
\end{equation} | |
The proof of \cref{thm:ttac} uses function extensionality, but only for one of the composites. | |
Thus, without assuming function extensionality, we can conclude that~\eqref{eq:uatofesp} is a retract\index{retract!of a type} of~\eqref{eq:uatofeps}. | |
And~\eqref{eq:uatofeps} is a product of contractible types, which is contractible by the weak function extensionality principle; hence~\eqref{eq:uatofesp} is also contractible. | |
\end{proof} | |
\sectionNotes | |
The fact that the space of continuous maps equipped with quasi-inverses has the wrong homotopy type to be the ``space of homotopy equivalences'' is well-known in algebraic topology. | |
In that context, the ``space of homotopy equivalences'' $(\eqv AB)$ is usually defined simply as the subspace of the function space $(A\to B)$ consisting of the functions that are homotopy equivalences. | |
In type theory, this would correspond most closely to $\sm{f:A\to B} \brck{\qinv(f)}$; see \cref{ex:brck-qinv}. | |
The first definition of equivalence given in homotopy type theory was the one that we have called $\iscontr(f)$, which was due to Voevodsky. | |
The possibility of the other definitions was subsequently observed by various people. | |
The basic theorems about adjoint equivalences\index{adjoint!equivalence} such as \cref{lem:coh-equiv,thm:equiv-iso-adj} are adaptations of standard facts in higher category theory and homotopy theory. | |
Using bi-invertibility as a definition of equivalences was suggested by Andr\'e Joyal. | |
The properties of equivalences discussed in \cref{sec:mono-surj,sec:equiv-closures} are well-known in homotopy theory. | |
Most of them were first proven in type theory by Voevodsky. | |
The fact that every function is equivalent to a fibration is a standard fact in homotopy theory. | |
The notion of object classifier | |
\index{object!classifier}% | |
\index{classifier!object}% | |
in $(\infty,1)$-category | |
\index{.infinity1-category@$(\infty,1)$-category}% | |
theory (the categorical analogue of \cref{thm:nobject-classifier-appetizer}) is due to Rezk (see~\cite{Rezk05,lurie:higher-topoi}). | |
Finally, the fact that univalence implies function extensionality (\cref{sec:univalence-implies-funext}) is due to Voevodsky. | |
Our proof is a simplification of his. | |
\cref{ex:funext-from-nondep} is also due to Voevodsky. | |
\sectionExercises | |
\begin{ex}\label{ex:two-sided-adjoint-equivalences} | |
Consider the type of ``two-sided adjoint equivalence\index{adjoint!equivalence} data'' for $f:A\to B$, | |
\begin{narrowmultline*} | |
\sm{g:B\to A}{\eta: g \circ f \htpy \idfunc[A]}{\epsilon:f \circ g \htpy \idfunc[B]} | |
\narrowbreak | |
\Parens{\prd{x:A} \map{f}{\eta x} = \epsilon(fx)} \times | |
\Parens{\prd{y:B} \map{g}{\epsilon y} = \eta(gy) }. | |
\end{narrowmultline*} | |
By \cref{lem:coh-equiv}, we know that if $f$ is an equivalence, then this type is inhabited. | |
Give a characterization of this type analogous to \cref{lem:qinv-autohtpy}. | |
Can you give an example showing that this type is not generally a mere proposition? | |
(This will be easier after \cref{cha:hits}.) | |
\end{ex} | |
\begin{ex}\label{ex:symmetric-equiv} | |
Show that for any $A,B:\UU$, the following type is equivalent to $\eqv A B$. | |
\begin{equation*} | |
\sm{R:A\to B\to \type} | |
\Parens{\prd{a:A} \iscontr\Parens{\sm{b:B} R(a,b)}} \times | |
\Parens{\prd{b:B} \iscontr\Parens{\sm{a:A} R(a,b)}}. | |
\end{equation*} | |
Can you extract from this a definition of a type satisfying the three desiderata of $\isequiv(f)$? | |
\end{ex} | |
\begin{ex} \label{ex:qinv-autohtpy-no-univalence} | |
Reformulate the proof of \cref{lem:qinv-autohtpy} without using univalence. | |
\end{ex} | |
\begin{ex}[The unstable octahedral axiom]\label{ex:unstable-octahedron} | |
\index{axiom!unstable octahedral}% | |
\index{octahedral axiom, unstable}% | |
Suppose $f:A\to B$ and $g:B\to C$ and $b:B$. | |
\begin{enumerate} | |
\item Show that there is a natural map $\hfib{g\circ f}{g(b)} \to \hfib{g}{g(b)}$ whose fiber over $(b,\refl{g(b)})$ is equivalent to $\hfib f b$. | |
\item Show that $\eqv{\hfib{g\circ f}{c}}{\sm{w:\hfib{g}{c}} \hfib f {\proj1 w}}$. | |
\end{enumerate} | |
\end{ex} | |
\begin{ex}\label{ex:2-out-of-6} | |
\index{2-out-of-6 property}% | |
Prove that equivalences satisfy the \emph{2-out-of-6 property}: given $f:A\to B$ and $g:B\to C$ and $h:C\to D$, if $g\circ f$ and $h\circ g$ are equivalences, so are $f$, $g$, $h$, and $h\circ g\circ f$. | |
Use this to give a higher-level proof of \cref{thm:paths-respects-equiv}. | |
\end{ex} | |
\begin{ex}\label{ex:qinv-univalence} | |
For $A,B:\UU$, define | |
\[ \mathsf{idtoqinv}_{A,B} :(A=B) \to \sm{f:A\to B}\qinv(f) \] | |
by path induction in the obvious way. | |
Let \textbf{\textsf{qinv}-univalence} denote the modified form of the univalence axiom which asserts that for all $A,B:\UU$ the function $\mathsf{idtoqinv}_{A,B}$ has a quasi-inverse. | |
\begin{enumerate} | |
\item Show that \qinv-univalence can be used instead of univalence in the proof of function extensionality in \cref{sec:univalence-implies-funext}. | |
\item Show that \qinv-univalence can be used instead of univalence in the proof of \cref{thm:qinv-notprop}. | |
\item Show that \qinv-univalence is inconsistent (i.e.\ allows construction of an inhabitant of $\emptyt$). | |
Thus, the use of a ``good'' version of $\isequiv$ is essential in the statement of univalence. | |
\end{enumerate} | |
\end{ex} | |
\begin{ex}\label{ex:embedding-cancellable} | |
Show that a function $f:A\to B$ is an embedding if and only if the following two conditions hold: | |
\begin{enumerate} | |
\item $f$ is \emph{left cancellable}, i.e.\ for any $x,y:A$, if $f(x)=f(y)$ then $x=y$.\label{item:ex:ec1} | |
\item For any $x:A$, the map $\apfunc f: \Omega(A,x) \to \Omega(B,f(x))$ is an equivalence.\label{item:ex:ec2} | |
\end{enumerate} | |
(In particular, if $A$ is a set, then $f$ is an embedding if and only if it is left-cancellable and $\Omega(B,f(x))$ is contractible for all $x:A$.) | |
Give examples to show that neither of~\ref{item:ex:ec1} or~\ref{item:ex:ec2} implies the other. | |
\end{ex} | |
\begin{ex}\label{ex:cancellable-from-bool} | |
Show that the type of left-cancellable functions $\bool\to B$ (see \cref{ex:embedding-cancellable}) is equivalent to $\sm{x,y:B}(x\neq y)$. | |
Give a similar explicit characterization of the type of embeddings $\bool\to B$. | |
\end{ex} | |
\begin{ex}\label{ex:funext-from-nondep} | |
The \textbf{na\"{i}ve non-dependent function extensionality axiom} says that for $A,B:\type$ and $f,g:A\to B$ there is a function $(\prd{x:A} f(x)=g(x)) \to (f=g)$. | |
\indexdef{function extensionality!non-dependent}% | |
Modify the argument of \cref{sec:univalence-implies-funext} to show that this axiom implies the full function extensionality axiom (\cref{axiom:funext}). | |
\end{ex} | |
% Local Variables: | |
% TeX-master: "hott-online" | |
% End: | |