Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Singular cohomology} | |
Here's one way to motivate this chapter. It turns out that: | |
\begin{itemize} | |
\ii $H_n(\CP^2) \cong H_n(S^2 \vee S^4)$ for every $n$. | |
\ii $H_n(\CP^3) \cong H_n(S^2 \times S^4)$ for every $n$. | |
\end{itemize} | |
This is unfortunate, because if possible we would like | |
to be able to tell these spaces apart (as they are | |
in fact not homotopy equivalent), but the homology groups | |
cannot tell the difference between them. | |
In this chapter, we'll define a \emph{cohomology group} $H^n(X)$ and $H^n(Y)$. | |
In fact, the $H^n$'s are completely determined by the $H_n$'s | |
by the so-called \emph{universal coefficient theorem}. | |
However, it turns out that one can take all the cohomology groups and put | |
them together to form a \emph{cohomology ring} $H^\bullet$. | |
We will then see that $H^\bullet(X) \not\cong H^\bullet(Y)$ as rings. | |
\section{Cochain complexes} | |
\begin{definition} | |
A \vocab{cochain complex} $A^\bullet$ is algebraically the same as a chain complex, except that the indices increase. | |
So it is a sequence of abelian groups | |
\[ \dots \taking{\delta} A^{n-1} \taking\delta A^n \taking\delta A^{n+1} \taking\delta \dots. \] | |
such that $\delta^2 = 0$. | |
Notation-wise, we're now using subscripts, and use $\delta$ rather $\partial$. | |
We define the \vocab{cohomology groups} by | |
\[ H^n(A^\bullet) = \ker\left( A^n \taking\delta A^{n+1} \right) | |
/ \img\left( A^{n-1} \taking\delta A^n \right). \] | |
\end{definition} | |
\begin{example}[de Rham cohomology] | |
We have already met one example of a cochain complex: | |
let $M$ be a smooth manifold and $\Omega^k(M)$ be the | |
additive group of $k$-forms on $M$. | |
Then we have a cochain complex | |
\[ 0 \taking d \Omega^0(M) | |
\taking d \Omega^1(M) \taking d \Omega^2(M) | |
\taking d \dots. \] | |
The resulting cohomology is called \vocab{de Rham cohomology}, | |
described later. | |
\end{example} | |
Aside from de Rham's cochain complex, | |
\textbf{the most common way to get a cochain complex | |
is to \emph{dualize} a chain complex.} | |
Specifically, pick an abelian group $G$; | |
note that $\Hom(-, G)$ is a contravariant functor, | |
and thus takes every chain complex | |
\[ \dots \taking\partial A_{n+1} \taking\partial | |
A_n \taking\partial A_{n-1} \taking\partial \dots \] | |
into a cochain complex: letting $A^n = \Hom(A_n, G)$ we obtain | |
\[ \dots \taking\delta A^{n-1} \taking\delta | |
A^n \taking\delta A^{n+1} \taking\delta \dots. \] | |
where $\delta(A_n \taking{f} G) = A_{n+1} \taking\partial A \taking{f} G$. | |
These are the cohomology groups we study most in algebraic topology, | |
so we give a special notation to them. | |
\begin{definition} | |
Given a chain complex $A_\bullet$ of abelian groups and another group $G$, | |
we let \[ H^n(A_\bullet; G) \] denote the cohomology groups | |
of the dual cochain complex $A^\bullet$ obtained by applying $\Hom(-,G)$. | |
In other words, $H^n(A_\bullet; G) = H^n(A^\bullet)$. | |
\end{definition} | |
\section{Cohomology of spaces} | |
\prototype{$C^0(X;G)$ all functions $X \to G$ while $H^0(X)$ are those functions $X \to G$ | |
constant on path components.} | |
The case of interest is our usual geometric situation, with $C_\bullet(X)$. | |
\begin{definition} | |
For a space $X$ and abelian group $G$, | |
we define $C^\bullet(X;G)$ to be the dual to the | |
singular chain complex $C_\bullet(X)$, | |
called the \vocab{singular cochain complex} of $X$; | |
its elements are called \vocab{cochains}. | |
Then we define the \vocab{cohomology groups} | |
of the space $X$ as | |
\[ H^n(X; G) \defeq H^n(C_\bullet(X); G) = H_n(C^\bullet(X;G)). \] | |
\end{definition} | |
\begin{remark} | |
Note that if $G$ is also a ring (like $\ZZ$ or $\RR$), | |
then $H^n(X; G)$ is not only an abelian group but actually a $G$-module. | |
\end{remark} | |
\begin{example} | |
[$C^0(X; G)$, $C^1(X; G)$, and $H^0(X;G)$] | |
Let $X$ be a topological space and consider $C^\bullet(X)$. | |
\begin{itemize} | |
\ii $C_0(X)$ is the free abelian group on $X$, | |
and $C^0(X) = \Hom(C_0(X), G)$. | |
So a $0$-cochain is a function that | |
takes every point of $X$ to an element of $G$. | |
\ii $C_1(X)$ is the free abelian group on $1$-simplices in $X$. | |
So $C^1(X)$ needs to take every $1$-simplex to an element of $G$. | |
\end{itemize} | |
Let's now try to understand $\delta : C^0(X) \to C^1(X)$. | |
Given a $0$-cochain $\phi \in C^0(X)$, | |
i.e.\ a homomorphism $\phi : C^0(X) \to G$, | |
what is $\delta\phi : C^1(X) \to G$? | |
Answer: | |
\[ \delta\phi : [v_0, v_1] \mapsto \phi([v_0]) - \phi([v_1]). \] | |
Hence, elements of | |
$\ker(C^0 \taking\delta C^1) \cong H^0(X;G)$ | |
are those cochains | |
that are \emph{constant on path-connected components}. | |
\end{example} | |
In particular, much like $H_0(X)$, we have \[ H^0(X) \cong G^{\oplus r} \] | |
if $X$ has $r$ path-connected components (where $r$ is finite\footnote{% | |
Something funny happens if $X$ has \emph{infinitely} many path-connected components: | |
say $X = \coprod_\alpha X_\alpha$ over an infinite indexing set. | |
In this case we have | |
$H_0(X) = \bigoplus_\alpha G$ while $H^0(X) = \prod_\alpha G$. | |
For homology we get a \emph{direct sum} while | |
for cohomology we get a \emph{direct product}. | |
These are actually different for infinite indexing sets. | |
For general modules $\bigoplus_\alpha M_\alpha$ is \emph{defined} to only allow | |
to have \emph{finitely many} nonzero terms. | |
(This was never mentioned earlier in the Napkin, | |
since I only ever defined $M \oplus N$ and extended it to finite direct sums.) | |
No such restriction holds for $\prod_\alpha G_\alpha$ a product of groups. | |
This corresponds to the fact that $C_0(X)$ is formal linear sums of $0$-chains | |
(which, like all formal sums, are finite) | |
from the path-connected components of $G$. | |
But a cochain of $C^0(X)$ is a \emph{function} | |
from each path-connected component of $X$ to $G$, | |
where there is no restriction. | |
}). | |
To the best of my knowledge, the higher cohomology groups $H^n(X; G)$ | |
(or even the cochain groups $C^n(X; G) = \Hom(C_n(X), G)$) are harder to describe concretely. | |
\begin{abuse} | |
In this chapter the only cochain complexes | |
we will consider are dual complexes as above. | |
So, any time we write a chain complex $A^\bullet$ it is implicitly given | |
by applying $\Hom(-,G)$ to $A_\bullet$. | |
\end{abuse} | |
\section{Cohomology of spaces is functorial} | |
We now check that the cohomology groups still exhibit the same nice functorial behavior. | |
First, let's categorize the previous results we had: | |
\begin{ques} | |
Define $\catname{CoCmplx}$ | |
the category of cochain complexes. | |
\end{ques} | |
\begin{exercise} | |
Interpret $\Hom(-,G)$ as a contravariant functor | |
from \[ \Hom(-,G) : \catname{Cmplx}\op \to \catname{CoCmplx}. \] | |
This means in particular that given a chain map $f : A_\bullet \to B_\bullet$, | |
we naturally obtain a dual map $f^\vee : B^\bullet \to A^\bullet$. | |
\end{exercise} | |
\begin{ques} | |
Interpret $H^n : \catname{CoCmplx} \to \catname{Grp}$ as a functor. | |
Compose these to get a contravariant functor | |
$H^n(-;G) : \catname{Cmplx}\op \to \catname{Grp}$. | |
\end{ques} | |
Then in exact analog to our result that $H_n : \catname{hTop} \to \catname{Grp}$ we have: | |
\begin{theorem}[$H^n (-;G): \catname{hTop}\op \to \catname{Grp}$] | |
For every $n$, $H^n(-;G)$ is a contravariant functor | |
from $\catname{hTop}\op$ to $\catname{Grp}$. | |
\end{theorem} | |
\begin{proof} | |
The idea is to leverage the work we already did in constructing | |
the prism operator earlier. | |
First, we construct the entire sequence of functors | |
from $\catname{Top}\op \to \catname{Grp}$: | |
\begin{center} | |
\begin{tikzcd} | |
\catname{Top}\op \ar[rr, "C_\bullet"] | |
&& \catname{Cmplx}\op \ar[rr, "\Hom(-;G)"] | |
&& \catname{CoCmplx} \ar[rr, "H^n"] | |
&& \catname{Grp} \\ | |
X \ar[dd, "f"', near start] | |
&& C_\bullet(X) \ar[dd, "f_\sharp"', near start] | |
&& C^\bullet(X;G) | |
&& H^n(X;G) \\ | |
\quad \ar[rr, mapsto] && | |
\quad \ar[rr, mapsto] && | |
\quad \ar[rr, mapsto] && | |
\quad \\ | |
Y | |
&& C_\bullet(Y) | |
&& C^\bullet(Y;G) \ar[uu, "f_\sharp"', near start] | |
&& H^n(Y;G). \ar[uu, "f^\ast"', near start] | |
\end{tikzcd} | |
\end{center} | |
Here $f^\sharp = (f_\sharp)^\vee$, and $f^\ast$ | |
is the resulting induced map on homology groups of the cochain complex. | |
So as before all we have to show is that $f \simeq g$, | |
then $f^\ast = g^\ast$. | |
Recall now that there is a prism operator such that | |
$f_\sharp - g_\sharp = P \partial + \partial P$. | |
If we apply the entire functor $\Hom(-;G)$ we get that | |
$f^\sharp - g^\sharp = \delta P^\vee + P^\vee \delta$ | |
where $P^\vee : C^{n+1}(Y;G) \to C^n(X;G)$. | |
So $f^\sharp$ and $g^\sharp$ are chain homotopic thus $f^\ast = g^\ast$. | |
\end{proof} | |
\section{Universal coefficient theorem} | |
We now wish to show that the cohomology groups are determined up to isomorphism | |
by the homology groups: given $H_n(A_\bullet)$, we can extract $H^n(A_\bullet; G)$. | |
This is achieved by the \emph{universal coefficient theorem}. | |
\begin{theorem} | |
[Universal coefficient theorem] | |
Let $A_\bullet$ be a chain complex of \emph{free} abelian groups, | |
and let $G$ be another abelian group. | |
Then there is a natural short exact sequence | |
\[ | |
0 \to \Ext(H_{n-1}(A_\bullet), G) \to H^n(A_\bullet; G) | |
\taking{h} \Hom(H_n(A_\bullet), G) \to 0. \] | |
In addition, this exact sequence is \emph{split} | |
so in particular | |
\[ H^n(C_\bullet; G) \cong \Ext(H_{n-1}(A_\bullet, G)) | |
\oplus \Hom(H_n(A_\bullet), G). \] | |
\end{theorem} | |
Fortunately, in our case of interest, $A_\bullet$ is $C_\bullet(X)$ | |
which is by definition free. | |
There are two things we need to explain, what the map $h$ is and the map $\Ext$ is. | |
It's not too hard to guess how \[ h : H^n(A_\bullet; G) \to \Hom(H_n(A_\bullet), G) \] is defined. | |
An element of $H^n(A_\bullet;G)$ is represented by a function which sends a cycle | |
in $A_n$ to an element of $G$. | |
The content of the theorem is to show that $h$ is surjective with kernel $\Ext(H_{n-1}(A_\bullet), G)$. | |
What about $\Ext$? | |
It turns out that $\Ext(-,G)$ is the so-called \vocab{Ext functor}, defined as follows. | |
Let $H$ be an abelian group, and consider a \vocab{free resolution} of $H$, | |
by which we mean an exact sequence | |
\[ \dots \taking{f_2} F_1 \taking{f_1} F_0 \taking{f_0} H \to 0 \] | |
with each $F_i$ free. | |
Then we can apply $\Hom(-,G)$ to get a cochain complex | |
\[ \dots \xleftarrow{f_2^\vee} \Hom(F_1, G) \xleftarrow{f_1^\vee} | |
\Hom(F_0, G) \xleftarrow{f_0^\vee} \Hom(H,G) \leftarrow 0. \] | |
but \emph{this cochain complex need not be exact} | |
(in categorical terms, $\Hom(-,G)$ does not preserve exactness). | |
We define \[ \Ext(H,G) \defeq \ker(f_2^\vee) / \img(f_1^\vee) \] | |
and it's a theorem that this doesn't depend on the choice of the free resolution. | |
There's a lot of homological algebra that goes into this, | |
which I won't take the time to discuss; | |
but the upshot of the little bit that I did include is that the $\Ext$ | |
functor is very easy to compute in practice, since | |
you can pick any free resolution you want and compute the above. | |
%By ``natural'', we mean that if $f : A_\bullet \to B_\bullet$ is a chain map, | |
%then we obtain a commutative diagram | |
%\begin{diagram} | |
% 0 & \rTo & \Ext(H_{n-1}(A_\bullet), G) & \rTo | |
% & H^n(A_\bullet;G) & \rTo & \Hom(H_n(A_\bullet), G) & \rTo & 0 \\ | |
% & & \uTo^{ \Ext(f_\ast, G) } & & \uTo^{f^\ast} & & \uTo^{\Hom(f_\ast, G)} & & \\ | |
% 0 & \rTo & \Ext(H_{n-1}(B_\bullet), G) & \rTo | |
% & H^n(A_\bullet;G) & \rTo & \Hom(H_n(B_\bullet), G) & \rTo & 0 \\ | |
%\end{diagram} | |
%where $f_\ast$ is the induced arrow $H_n(A_\bullet) \to H_n(B_\bullet)$. | |
\begin{lemma} | |
[Computing the $\Ext$ functor] | |
For any abelian groups $G$, $H$, $H'$ we have | |
\begin{enumerate}[(a)] | |
\ii $\Ext(H \oplus H', G) = \Ext(H, G) \oplus \Ext(H', G)$. | |
\ii $\Ext(H,G) = 0$ for $H$ free, and | |
\ii $\Ext(\Zc n, G) = G / nG$. | |
\end{enumerate} | |
\end{lemma} | |
\begin{proof} | |
For (a), note that if $\dots \to F_1 \to F_0 \to H \to 0$ | |
and $\dots \to F_1' \to F_0' \to F_0' \to H' \to 0$ are free resolutions, | |
then so is $F_1 \oplus F_1' \to F_0 \oplus F_0' \to H \oplus H' \to 0$. | |
For (b), note that $0 \to H \to H \to 0$ is a free resolution. | |
Part (c) follows by taking the free resolution | |
\[ 0 \to \ZZ \taking{\times n} \ZZ \to \Zc n \to 0 \] | |
and applying $\Hom(-,G)$ to it. | |
\begin{ques} | |
Finish the proof of (c) from here. \qedhere | |
\end{ques} | |
\end{proof} | |
\begin{ques} | |
Some $\Ext$ practice: compute | |
$\Ext(\ZZ^{\oplus 2015}, G)$ and $\Ext(\Zc{30}, \Zc 4)$. | |
\end{ques} | |
\section{Example computation of cohomology groups} | |
\prototype{Possibly $H^n(S^m)$.} | |
The universal coefficient theorem gives us a direct way to compute | |
any cohomology groups, provided we know the homology ones. | |
\begin{example} | |
[Cohomolgy groups of $S^m$] | |
It is straightforward to compute $H^n(S^m)$ now: | |
all the $\Ext$ terms vanish since $H_n(S^m)$ is always free, | |
and hence we obtain that | |
\[ H^n(S^m) \cong \Hom(H_n(S^m), G) \cong | |
\begin{cases} | |
G & n=m, n=0 \\ | |
0 & \text{otherwise}. | |
\end{cases} | |
\] | |
% By UCT for reduced groups, we also have | |
% \[ \wt H^n(S^m) \cong \Hom(\wt H_n(S^m), G) \cong | |
% \begin{cases} | |
% G & n=m \\ | |
% 0 & \text{otherwise}. | |
% \end{cases} | |
% \] | |
% since $\Hom(\ZZ, G)$. | |
\end{example} | |
\begin{example} | |
[Cohomolgy groups of torus] | |
This example has no nonzero $\Ext$ terms either, | |
since this time $H^n(S^1 \times S^1)$ is always free. | |
So we obtain | |
\[ H^n(S^1 \times S^1) \cong \Hom(H_n(S^1 \times S^1), G). \] | |
Since $H_n(S^1 \times S^1)$ is $\ZZ$, $\ZZ^{\oplus 2}$, $\ZZ$ | |
in dimensions $n=1,2,1$ we derive that | |
\[ | |
H^n(S^1 \times S^1) | |
\cong | |
\begin{cases} | |
G & n = 0,2 \\ | |
G^{\oplus 2} & n = 1. | |
\end{cases} | |
\] | |
\end{example} | |
From these examples one might notice that: | |
\begin{lemma} | |
[$0$th homology groups are just duals] | |
For $n = 0$ and $n = 1$, we have | |
\[ H^n(X;G) \cong \Hom(H_n(X), G). \] | |
\end{lemma} | |
\begin{proof} | |
It's already been shown for $n=0$. | |
For $n=1$, notice that $H_0(X)$ is free, | |
so the $\Ext$ term vanishes. | |
\end{proof} | |
\begin{example} | |
[Cohomolgy groups of Klein bottle] | |
This example will actually have $\Ext$ term. | |
Recall that if $K$ is a Klein Bottle then its homology groups are | |
$\ZZ$ in dimension $n=0$ and $\ZZ \oplus \Zc 2$ in $n=1$, and $0$ elsewhere. | |
For $n=0$, we again just have $H^0(K;G) \cong \Hom(\ZZ, G) \cong G$. | |
For $n=1$, the $\Ext$ term is $\Ext(H_0(K), G) \cong \Ext(\ZZ, G) = 0$ | |
so \[ H^1(K;G) \cong \Hom(\ZZ \oplus \Zc2, G) \cong G \oplus \Hom(\Zc2, G). \] | |
We have that $\Hom(\Zc2,G)$ is the subgroup | |
of elements of order $2$ in $G$ (and $0 \in G$). | |
But for $n=2$, we have our first interesting $\Ext$ group: | |
the exact sequence is | |
\[ 0 \to \Ext(\ZZ \oplus \Zc 2, G) \to H^2(X;G) \to \underbrace{H_2(X)}_{=0} \to 0. \] | |
Thus, we have | |
\[ H^2(X;G) \cong \left( \Ext(\ZZ,G) \oplus \Ext(\Zc2,G) \right) \oplus 0 | |
\cong G/2G. \] | |
All the higher groups vanish. | |
In summary: | |
\[ | |
H^n(X;G) \cong | |
\begin{cases} | |
G & n = 0 \\ | |
G \oplus \Hom(\Zc2, G) & n = 1 \\ | |
G/2G & n = 2 \\ | |
0 & n \ge 3. | |
\end{cases} | |
\] | |
\end{example} | |
\section{Relative cohomology groups} | |
One can also define relative cohomology groups in the obvious way: | |
dualize the chain complex | |
\[ \dots \taking\partial C_1(X,A) \taking\partial C_0(X,A) \to 0 \] | |
to obtain a cochain complex | |
\[ | |
\dots \xleftarrow\delta C^1(X,A;G) \xleftarrow\delta C^0(X,A;G) | |
\leftarrow 0. | |
\] | |
We can take the cohomology groups ofthis. | |
\begin{definition} | |
The groups thus obtained are the \vocab{relative cohomology groups} | |
are denoted $H^n(X,A;G)$. | |
\end{definition} | |
In addition, we can define reduced cohomology groups as well. | |
One way to do it is to take the augmented singular chain complex | |
\[ \dots \taking\partial C_1(X) \taking\partial C_0(X) \taking\eps \ZZ \to 0 \] | |
and dualize it to obtain | |
\[ | |
\dots \xleftarrow\delta C^1(X;G) \xleftarrow\delta C^0(X;G) | |
\xleftarrow{\eps^\vee} \underbrace{\Hom(\ZZ, G)}_{\cong G} | |
\leftarrow 0. | |
\] | |
Since the $\ZZ$ we add is also free, | |
the universal coefficient theorem still applies. | |
So this will give us reduced cohomology groups. | |
However, since we already defined the relative cohomology groups, | |
it is easiest to simply define: | |
\begin{definition} | |
The \vocab{reduced cohomology groups} of a nonempty space $X$, | |
denoted $\wt H^n(X; G)$, | |
are defined to be $H^n(X, \{\ast\} ; G)$ | |
for some point $\ast \in X$. | |
\end{definition} | |
\section\problemhead | |
\begin{sproblem} | |
[Wedge product cohomology] | |
For any $G$ and $n$ we have | |
\[ | |
\wt H^n(X \vee Y; G) | |
\cong | |
\wt H^n(X; G) \oplus \wt H^n(Y; G). | |
\] | |
\end{sproblem} | |
\begin{dproblem} | |
Prove that for a field $F$ of characteristic zero and a space $X$ | |
with finitely generated homology groups: | |
\[ H^k(X, F) \cong \left( H_k(X) \right)^\vee. \] | |
Thus over fields cohomology is the dual of homology. | |
\end{dproblem} | |
\begin{problem}[$\Zc2$-cohomology of $\RP^n$] | |
Prove that | |
\[ | |
H^m(\RP^n, \Zc2) | |
\cong | |
\begin{cases} | |
\ZZ & \text{$m=0$, or $m$ is odd and $m=n$} \\ | |
\Zc2 & \text{$0 < m < n$ and $m$ is odd} \\ | |
0 & \text{otherwise}. | |
\end{cases} | |
\] | |
\end{problem} | |