\chapter{Excision and relative homology} We have already seen how to use the Mayer-Vietoris sequence: we started with a sequence \[ \dots \to H_n(U \cap V) \to H_n(U) \oplus H_n(V) \to H_n(U+V) \to H_{n-1}(U \cap V) \to \dots \] and its reduced version, then appealed to the geometric fact that $H_n(U+V) \cong H_n(X)$. This allowed us to algebraically make computations on $H_n(X)$. In this chapter, we turn our attention to the long exact sequence associated to the chain complex \[ 0 \to C_n(A) \injto C_n(X) \surjto C_n(X,A) \to 0. \] The setup will look a lot like the previous two chapters, except in addition to $H_n : \catname{hTop} \to \catname{Grp}$ we will have a functor $H_n : \catname{hPairTop} \to \catname{Grp}$ which takes a pair $(X,A)$ to $H_n(X,A)$. Then, we state (again without proof) the key geometric result, and use this to make deductions. \section{The long exact sequences} Recall \Cref{thm:long_exact_rel}, which says that the sequences \[ \dots \to H_n(A) \to H_n(X) \to H_n(X,A) \to H_{n-1}(A) \to \dots. \] and \[ \dots \to \wt H_n(A) \to \wt H_n(X) \to H_n(X,A) \to \wt H_{n-1}(A) \to \dots \] are long exact. By \Cref{prob:triple_long_exact} we even have a long exact sequence \[ \dots \to H_n(B,A) \to H_n(X,A) \to H_n(X,B) \to H_{n-1}(B,A) \to \dots. \] for $A \subseteq B \subseteq X$. An application of the second long exact sequence above gives: \begin{lemma} [Homology relative to contractible spaces] \label{lem:rel_contractible} Let $X$ be a topological space, and let $A \subseteq X$ be contractible. For all $n$, \[ H_n(X, A) \cong \wt H_n(X). \] \end{lemma} \begin{proof} Since $A$ is contractible, we have $\wt H_n(A) = 0$ for every $n$. For each $n$ there's a segment of the long exact sequence given by \[ \dots \to \underbrace{\wt H_n(A)}_{=0} \to \wt H_n(X) \to H_n(X,A) \to \underbrace{\wt H_{n-1}(A)}_{=0} \to \dots. \] So since $0 \to \wt H_n(X) \to H_n(X,A) \to 0$ is exact, this means $H_n(X,A) \cong \wt H_n(X)$. \end{proof} In particular, the theorem applies if $A$ is a single point. The case $A = \varnothing$ is also worth noting. We compile these results into a lemma: \begin{lemma} [Relative homology generalizes absolute homology] Let $X$ be any space, and $\ast \in X$ a point. Then for all $n$, \[ H_n(X, \{\ast\}) \cong \wt H_n(X) \qquad\text{and}\qquad H_n(X, \varnothing) = H_n(X). \] \end{lemma} \section{The category of pairs} Since we now have an $H_n(X,A)$ instead of just $H_n(X)$, a natural next step is to create a suitable category of \emph{pairs} and give ourselves the same functorial setup as before. \begin{definition} Let $\varnothing \neq A \subseteq X$ and $\varnothing \neq B \subseteq X$ be subspaces, and consider a map $f : X \to Y$. If $f\im(A) \subseteq B$ we write \[ f : (X,A) \to (Y,B). \] We say $f$ is a \vocab{map of pairs}, between the pairs $(X,A)$ and $(Y,B)$. \end{definition} \begin{definition} We say that $f,g : (X,A) \to (Y,B)$ are \vocab{pair-homotopic} if they are ``homotopic through maps of pairs''. More formally, a \vocab{pair-homotopy} $f, g : (X,A) \to (Y,B)$ is a map $F : [0,1] \times X \to Y$, which we'll write as $F_t(X)$, such that $F$ is a homotopy of the maps $f,g : X \to Y$ and each $F_t$ is itself a map of pairs. \end{definition} Thus, we naturally arrive at two categories: \begin{itemize} \ii $\catname{PairTop}$, the category of \emph{pairs} of topological spaces, and \ii $\catname{hPairTop}$, the same category except with maps only equivalent up to homotopy. \end{itemize} \begin{definition} As before, we say pairs $(X,A)$ and $(Y,B)$ are \vocab{pair-homotopy equivalent} if they are isomorphic in $\catname{hPairTop}$. An isomorphism of $\catname{hPairTop}$ is a \vocab{pair-homotopy equivalence}. \end{definition} We can do the same song and dance as before with the prism operator to obtain: \begin{lemma}[Induced maps of relative homology] We have a functor \[ H_n : \catname{hPairTop} \to \catname{Grp}. \] \end{lemma} That is, if $f : (X,A) \to (Y,B)$ then we obtain an induced map \[ f_\ast : H_n(X,A) \to H_n(Y,B). \] and if two such $f$ and $g$ are pair-homotopic then $f_\ast = g_\ast$. Now, we want an analog of contractible spaces for our pairs: i.e.\ pairs of spaces $(X,A)$ such that $H_n(X,A) = 0$. The correct definition is: \begin{definition} Let $A \subseteq X$. We say that $A$ is a \vocab{deformation retract} of $X$ if there is a map of pairs $r : (X, A) \to (A, A)$ which is a pair homotopy equivalence. \end{definition} \begin{example} [Examples of deformation retracts] \listhack \begin{enumerate}[(a)] \ii If a single point $p$ is a deformation retract of a space $X$, then $X$ is contractible, since the retraction $r : X \to \{\ast\}$ (when viewed as a map $X \to X$) is homotopic to the identity map $\id_X : X \to X$. \ii The punctured disk $D^2 \setminus \{0\}$ deformation retracts onto its boundary $S^1$. \ii More generally, $D^{n} \setminus \{0\}$ deformation retracts onto its boundary $S^{n-1}$. \ii Similarly, $\RR^n \setminus \{0\}$ deformation retracts onto a sphere $S^{n-1}$. \end{enumerate} \end{example} Of course in this situation we have that \[ H_n(X,A) \cong H_n(A,A) = 0. \] \begin{exercise} Show that if $A \subseteq V \subseteq X$, and $A$ is a deformation retract of $V$, then $H_n(X,A) \cong H_n(X,V)$ for all $n$. (Use \Cref{prob:triple_long_exact}. Solution in next section.) \end{exercise} \section{Excision} Now for the key geometric result, which is the analog of \Cref{thm:open_cover_homology} for our relative homology groups. \begin{theorem} [Excision] Let $Z \subseteq A \subseteq X$ be subspaces such that the closure of $Z$ is contained in the interior of $A$. Then the inclusion $\iota (X \setminus Z, A \setminus Z) \injto (X,A)$ (viewed as a map of pairs) induces an isomorphism of relative homology groups \[ H_n(X \setminus Z, A \setminus Z) \cong H_n(X,A). \] \end{theorem} This means we can \emph{excise} (delete) a subset $Z$ of $A$ in computing the relative homology groups $H_n(X,A)$. This should intuitively make sense: since we are ``modding out by points in $A$'', the internals of the point $A$ should not matter so much. The main application of excision is to decide when $H_n(X,A) \cong \wt H_n(X/A)$. Answer: \begin{theorem} [Relative homology $\implies$ quotient space] \label{thm:good_pair} Let $X$ be a space and $A$ be a subspace such that $A$ is a deformation retract of some open set $V \subseteq X$. Then the quotient map $q : X \to X/A$ induces an isomorphism \[ H_n(X,A) \cong H_n(X/A, A/A) \cong \wt H_n(X/A). \] \end{theorem} \begin{proof} By hypothesis, we can consider the following maps of pairs: \begin{align*} r & : (V,A) \to (A,A) \\ q & : (X,A) \to (X/A, A/A) \\ \widehat q &: (X-A, V-A) \to (X/A-A/A, V/A-A/A). \end{align*} Moreover, $r$ is a pair-homotopy equivalence. Considering the long exact sequence of a triple (which was \Cref{prob:triple_long_exact}) we have a diagram \begin{center} \begin{tikzcd}[row sep=huge] H_n(V,A) \ar[r] \ar[d, "\cong"', "r"] & H_n(X,A) \ar["f", r] & H_n(X, V) \ar[r] & H_{n-1}(V,A) \ar[d, "\cong"', "r"] \\ \underbrace{H_n(A,A)}_{=0} & & & \underbrace{H_{n-1}(A,A)}_{=0} \end{tikzcd} \end{center} where the isomorphisms arise since $r$ is a pair-homotopy equivalence. So $f$ is an isomorphism. Similarly the map \[ g : H_n(X/A, A/A) \to H_n(X/A, V/A) \] is an isomorphism. Now, consider the commutative diagram \begin{center} \begin{tikzcd}[sep=huge] H_n(X,A) \ar[r, "f"] \ar[d, "q_\ast"'] & H_n(X,V) & H_n(X-A, V-A) \ar[l, "\text{Excise}"'] \ar[d, "\widehat{q}_\ast", "\cong"'] \\ H_n(X/A,A/A) \ar[r, "g"'] & H_n(X/A,V/A) & \ar["\text{Excise}"', l] H_n(X/A-A/A, V/A-A/A) \end{tikzcd} \end{center} and observe that the rightmost arrow $\widehat{q}_\ast$ is an isomorphism, because outside of $A$ the map $\widehat q$ is the identity. We know $f$ and $g$ are isomorphisms, as are the two arrows marked with ``Excise'' (by excision). From this we conclude that $q_\ast$ is an isomorphism. Of course we already know that homology relative to a point is just the relative homology groups (this is the important case of \Cref{lem:rel_contractible}). \end{proof} \section{Some applications} One nice application of excision is to compute $\wt H_n(X \vee Y)$. \begin{theorem}[Homology of wedge sums] Let $X$ and $Y$ be spaces with basepoints $x_0 \in X$ and $y_0 \in Y$, and assuming each point is a deformation retract of some open neighborhood. Then for every $n$ we have \[ \wt H_n(X \vee Y) = \wt H_n(X) \oplus \wt H_n(Y). \] \end{theorem} \begin{proof} Apply \Cref{thm:good_pair} with the subset $\{x_0, y_0\}$ of $X \amalg Y$, \begin{align*} \wt H_n (X \vee Y) \cong \wt H_n( (X \amalg Y) / \{x_0, y_0\} ) &\cong H_n(X \amalg Y, \{x_0,y_0\}) \\ &\cong H_n(X, \{x_0\}) \oplus H_n(Y, \{y_0\}) \\ &\cong\wt H_n(X) \oplus \wt H_n(Y). \qedhere \end{align*} \end{proof} Another application is to give a second method of computing $H_n(S^m)$. To do this, we will prove that \[ \wt H_n(S^m) \cong \wt H_{n-1}(S^{m-1}) \] for any $n,m > 1$. However, \begin{itemize} \ii $\wt H_0(S^n)$ is $\ZZ$ for $n=0$ and $0$ otherwise. \ii $\wt H_n(S^0)$ is $\ZZ$ for $m=0$ and $0$ otherwise. \end{itemize} So by induction on $\min \{m,n\}$ we directly obtain that \[ \wt H_n(S^m) \cong \begin{cases} \ZZ & m=n \\ 0 & \text{otherwise} \end{cases} \] which is what we wanted. To prove the claim, let's consider the exact sequence formed by the pair $X = D^2$ and $A = S^1$. \begin{example}[The long exact sequence for $(X,A) = (D^2, S^1)$] Consider $D^2$ (which is contractible) with boundary $S^1$. Clearly $S^1$ is a deformation retraction of $D^2 \setminus \{0\}$, and if we fuse all points on the boundary together we get $D^2 / S^1 \cong S^2$. So we have a long exact sequence \begin{center} \begin{tikzcd} \wt H_2(S^1) \ar[r] & \underbrace{\wt H_2(D^2)}_{=0} \ar[r] & \wt H_2(S^2) \ar[lld] \\ \wt H_1(S^1) \ar[r] & \underbrace{\wt H_1(D^2)}_{=0} \ar[r] & \wt H_1(S^2) \ar[lld] \\ \wt H_0(S^1) \ar[r] & \underbrace{\wt H_0(D^2)}_{=0} \ar[r] & \underbrace{\wt H_0(S^2)}_{=0} \end{tikzcd} \end{center} From this diagram we read that \[ \dots, \quad \wt H_3(S^2) = \wt H_2(S^1), \quad \wt H_2(S^2) = \wt H_1(S^1), \quad \wt H_1(S^2) = \wt H_0(S^1). \] \end{example} More generally, the exact sequence for the pair $(X,A) = (D^m, S^{m-1})$ shows that $\wt H_n(S^m) \cong \wt H_{n-1}(S^{m-1})$, which is the desired conclusion. \section{Invariance of dimension} Here is one last example of an application of excision. \begin{definition} Let $X$ be a space and $p \in X$ a point. The $k$th \vocab{local homology group} of $p$ at $X$ is defined as \[ H_k(X, X \setminus \{p\}). \] \end{definition} Note that for any open neighborhood $U$ of $p$, we have by excision that \[ H_k(X, X \setminus \{p\}) \cong H_k(U, U \setminus \{p\}). \] Thus this local homology group only depends on the space near $p$. \begin{theorem} [Invariance of dimension, Brouwer 1910] Let $U \subseteq \RR^n$ and $V \subseteq \RR^m$ be nonempty open sets. If $U$ and $V$ are homeomorphic, then $m = n$. \end{theorem} \begin{proof} Consider a point $x \in U$ and its local homology groups. By excision, \[ H_k(\RR^n, \RR^n \setminus \{x\}) \cong H_k(U, U \setminus \{x\}). \] But since $\RR^n \setminus \{x\}$ is homotopic to $S^{n-1}$, the long exact sequence of \Cref{thm:long_exact_rel} tells us that \[ H_k(\RR^n, \RR^n \setminus \{x\}) \cong \begin{cases} \ZZ & k = n \\ 0 & \text{otherwise}. \end{cases} \] Analogously, given $y \in V$ we have \[ H_k(\RR^m, \RR^m \setminus\{y\}) \cong H_k(V, V\setminus\{y\}). \] If $U \cong V$, we thus deduce that \[ H_k(\RR^n, \RR^n\setminus\{x\}) \cong H_k(\RR^m, \RR^m\setminus\{y\}) \] for all $k$. This of course can only happen if $m=n$. \end{proof} \section\problemhead \begin{problem} Let $X = S^1 \times S^1$ and $Y = S^1 \vee S^1 \vee S^2$. Show that \[ H_n(X) \cong H_n(Y) \] for every integer $n$. \end{problem} \begin{problem}[Hatcher \S2.1 exercise 18] Consider $\QQ \subset \RR$. Compute $\wt H_1(\RR, \QQ)$. \begin{hint} Use \Cref{thm:long_exact_rel}. \end{hint} \begin{sol} We have an exact sequence \[ \underbrace{\wt H_1(\RR)}_{=0} \to \wt H_1(\RR, \QQ) \to \wt H_0(\QQ) \to \underbrace{\wt H_0(\RR)}_{=0}. \] Now, since $\QQ$ is path-disconnected (i.e.\ no two of its points are path-connected) it follows that $\wt H_0(\QQ)$ consists of countably infinitely many copies of $\ZZ$. \end{sol} \end{problem} \begin{sproblem} What are the local homology groups of a topological $n$-manifold? \end{sproblem} \begin{problem} Let \[ X = \{(x,y) \mid x \ge 0\} \subseteq \RR^2 \] denote the half-plane. What are the local homology groups of points in $X$? % http://math.stackexchange.com/questions/350667/local-homology-group-a-homeomorphism-takes-the-boundary-to-the-boundary \end{problem} \begin{problem} [Brouwer-Jordan separation theorem, generalizing Jordan curve theorem] \yod Let $X \subseteq \RR^n$ be a subset which is homeomorphic to $S^{n-1}$. Prove that $\RR^n \setminus X$ has exactly two path-connected components. \begin{hint} For any $n$, prove by induction for $k=1,\dots,n-1$ that (a) if $X$ is a subset of $S^n$ homeomorphic to $D^k$ then $\wt H_i(S^n \setminus X) = 0$; (b) if $X$ is a subset of $S^n$ homeomorphic to $S^k$ then $\wt H_i(S^n \setminus X) = \ZZ$ for $i=n-k-1$ and $0$ otherwise. \end{hint} \begin{sol} This is shown in detail in Section 2.B of Hatcher. \end{sol} \end{problem}