Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Bonus: Cellular homology} | |
We now introduce cellular homology, which essentially lets us compute | |
the homology groups of any CW complex we like. | |
\section{Degrees} | |
\prototype{$z \mapsto z^d$ has degree $d$.} | |
For any $n > 0$ and map $f : S^n \to S^n$, consider | |
\[ f_\ast : \underbrace{H_n(S^n)}_{\cong \ZZ} \to \underbrace{H_n(S^n)}_{\cong \ZZ} \] | |
which must be multiplication by some constant $d$. | |
This $d$ is called the \vocab{degree} of $f$, denoted $\deg f$. | |
\begin{ques} | |
Show that $\deg(f \circ g) = \deg(f) \deg(g)$. | |
\end{ques} | |
\begin{example} | |
[Degree] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii For $n=1$, the map $z \mapsto z^k$ (viewing $S^1 \subseteq \CC$) | |
has degree $k$. | |
\ii A reflection map $(x_0, x_1, \dots, x_n) \mapsto (-x_0, x_1, \dots, x_n)$ | |
has degree $-1$; we won't prove this, but geometrically this should be clear. | |
\ii The antipodal map $x \mapsto -x$ has degree $(-1)^{n+1}$ | |
since it's the composition of $n+1$ reflections as above. | |
We denote this map by $-\id$. | |
\end{enumerate} | |
\end{example} | |
Obviously, if $f$ and $g$ are homotopic, then $\deg f = \deg g$. | |
In fact, a theorem of Hopf says that this is a classifying invariant: | |
anytime $\deg f = \deg g$, we have that $f$ and $g$ are homotopic. | |
One nice application of this: | |
\begin{theorem} | |
[Hairy ball theorem] | |
If $n > 0$ is even, then $S^n$ doesn't have a continuous field | |
of nonzero tangent vectors. | |
\end{theorem} | |
\begin{proof} | |
If the vectors are nonzero then WLOG they have norm $1$; | |
that is for every $x$ we have an orthogonal unit vector $v(x)$. | |
Then we can construct a homotopy map $F : S^n \times [0,1] \to S^n$ by | |
\[ (x,t) \mapsto (\cos \pi t)x + (\sin \pi t) v(x). \] | |
which gives a homotopy from $\id$ to $-\id$. | |
So $\deg(\id) = \deg(-\id)$, which means $1 = (-1)^{n+1}$ | |
so $n$ must be odd. | |
\end{proof} | |
Of course, the one can construct such a vector field whenever $n$ is odd. | |
For example, when $n=1$ such a vector field is drawn below. | |
\begin{center} | |
\begin{asy} | |
size(5cm); | |
draw(unitcircle, blue+1); | |
label("$S^1$", dir(100), dir(100), blue); | |
void arrow(real theta) { | |
pair P = dir(theta); | |
dot(P); | |
pair delta = 0.8*P*dir(90); | |
draw( P--(P+delta), EndArrow ); | |
} | |
arrow(0); | |
arrow(50); | |
arrow(140); | |
arrow(210); | |
arrow(300); | |
\end{asy} | |
\end{center} | |
\section{Cellular chain complex} | |
Before starting, we state: | |
\begin{lemma} | |
[CW homology groups] | |
Let $X$ be a CW complex. Then | |
\begin{align*} | |
H_k(X^n, X^{n-1}) &\cong | |
\begin{cases} | |
\ZZ^{\oplus\text{\#$n$-cells of $X$}} & k = n \\ | |
0 & \text{otherwise}. | |
\end{cases} \\ | |
\intertext{and} | |
H_k(X^n) &\cong | |
\begin{cases} | |
H_k(X) & k \le n-1 \\ | |
0 & k \ge n+1. | |
\end{cases} | |
\end{align*} | |
\end{lemma} | |
\begin{proof} | |
% I'll prove just the case where $X$ is finite-dimensional as usual. | |
The first part is immediate by noting that $(X^n, X^{n-1})$ is a good pair | |
and $X^n/X^{n-1}$ is a wedge sum of two spheres. | |
For the second part, fix $k$ and note that, as long as $n \le k-1$ or $n \ge k+2$, | |
\[ | |
\underbrace{H_{k+1}(X^n, X^{n-1})}_{=0} | |
\to H_k(X^{n-1}) | |
\to H_k(X^n) | |
\to \underbrace{H_{k}(X^n, X^{n-1})}_{=0}. | |
\] | |
So we have isomorphisms | |
\[ H_k(X^{k-1}) \cong H_k(X^{k-2}) \cong \dots \cong H_k(X^0) = 0 \] | |
and | |
\[ H_k(X^{k+1}) \cong H_k(X^{k+2}) \cong \dots \cong H_k(X). \qedhere \] | |
\end{proof} | |
So, we know that the groups $H_k(X^k, X^{k-1})$ are super nice: | |
they are free abelian with basis given by the cells of $X$. | |
So, we give them a name: | |
\begin{definition} | |
For a CW complex $X$, we define | |
\[ \Cells_k(X) = H_k(X^k, X^{k-1}) \] | |
where $\Cells_0(X) = H_0(X^0, \varnothing) = H_0(X^0)$ by convention. | |
So $\Cells_k(X)$ is an abelian group with basis given by | |
the $k$-cells of $X$. | |
\end{definition} | |
Now, using $\Cells_k = H_k(X^k, X^{k-1})$ let's use | |
our long exact sequence and try to string together maps between these. | |
Consider the following diagram. | |
\begin{center} | |
\newcommand{\CX}[1]{\boxed{\color{blue}\Cells_{#1}(X)}} | |
\begin{tikzcd}[column sep=tiny] | |
& \underbrace{H_3(X^2)}_{=0} \ar[d, "0"] \\ | |
\CX{4} \ar[r, "\partial_4"] \ar[rd, "d_4", blue] | |
& H_3(X^3) \ar[r, two heads] \ar[d, "0"] | |
& \underbrace{H_3(X^4)}_{\cong H_3(X)} \ar[r, "0"] | |
& \underbrace{H_3(X^4, X^3)}_{= 0} \\ | |
& \CX{3} \ar[d, "\partial_3"] \ar[rd, "d_3", blue] | |
&& \underbrace{H_1(X^0)}_{=0} \ar[d, "0"] \\ | |
\underbrace{H_2(X^1)}_{=0} \ar[r, "0"] | |
& H_2(X^2) \ar[r, hook] \ar[d, two heads] | |
& \CX{2} \ar[r, "\partial_2"] \ar[rd, "d_2", blue] | |
& H_1(X^1) \ar[r, two heads] | |
& \underbrace{H_1(X^2)}_{\cong H_1(X)} \ar[r, "0"] | |
& \underbrace{H_1(X^2, X^1)}_{=0} \\ | |
& \underbrace{H_2(X^3)}_{\cong H_2(X)} \ar[d, "0"] | |
&& \CX{1} \ar[d, "\partial_1"] \ar[rd, "d_1", blue] \\ | |
& \underbrace{H_2(X^3, X^2)}_{=0} | |
& \underbrace{H_0(\varnothing)}_{=0} \ar[r, "0"] | |
& H_0(X^0) \ar[r, hook] \ar[d, two heads] | |
& \CX{0} \ar[r, "\partial_0"] | |
& \dots \\ | |
&&& \underbrace{H_0(X^1)}_{\cong H_0(X)} \ar[d, "0"] \\ | |
&&& \underbrace{H_0(X^1, X^0)}_{=0} | |
\end{tikzcd} | |
\end{center} | |
The idea is that we have taken all the exact sequences generated by adjacent | |
skeletons, and strung them together at the groups $H_k(X^k)$, | |
with half the exact sequences being laid out vertically | |
and the other half horizontally. | |
In that case, composition generates a sequence of dotted maps | |
between the $H_k(X^k, X^{k-1})$ as shown. | |
\begin{ques} | |
Show that the composition of two adjacent dotted arrows is zero. | |
\end{ques} | |
So from the diagram above, we can read off a sequence of arrows | |
\[ | |
\dots \taking{d_5} \Cells_4(X) \taking{d_4} \Cells_3(X) | |
\taking{d_3} \Cells_2(X) \taking{d_2} \Cells_1(X) | |
\taking{d_1} \Cells_0(X) \taking{d_0} 0. | |
\] | |
This is a chain complex, called the \vocab{cellular chain complex}; | |
as mentioned before before all the homology groups are free, | |
but these ones are especially nice because for most reasonable CW complexes, | |
they are also finitely generated | |
(unlike the massive $C_\bullet(X)$ that we had earlier). | |
In other words, the $H_k(X^k, X^{k-1})$ are especially nice ``concrete'' free groups | |
that one can actually work with. | |
The other reason we care is that in fact: | |
\begin{theorem}[Cellular chain complex gives $H_n(X)$] | |
\label{thm:cellular_chase} | |
The $k$th homology group of the cellular chain complex | |
is isomorphic to $H_k(X)$. | |
\end{theorem} | |
\begin{proof} | |
Follows from the diagram; \Cref{prob:diagram_chase}. | |
\end{proof} | |
A nice application of this is to define | |
the \vocab{Euler characteristic} of a finite CW complex $X$. | |
Of course we can write | |
\[ \chi(X) = \sum_n (-1)^n \cdot \#(\text{$n$-cells of $X$}) \] | |
which generalizes the familiar $V-E+F$ formula. | |
However, this definition is unsatisfactory because it | |
depends on the choice of CW complex, while we actually | |
want $\chi(X)$ to only depend on the space $X$ itself | |
(and not how it was built). In light of this, we prove that: | |
\begin{theorem} | |
[Euler characteristic via Betti numbers] | |
For any finite CW complex $X$ we have | |
\[ \chi(X) = \sum_n (-1)^n \rank H_n(X). \] | |
\end{theorem} | |
Thus $\chi(X)$ does not depend on the choice of CW decomposition. | |
The numbers | |
\[ b_n = \rank H_n(X) \] | |
are called the \vocab{Betti numbers} of $X$. | |
In fact, we can use this to define $\chi(X)$ for any reasonable space; | |
we are happy because in the (frequent) case that $X$ is a CW complex, | |
\begin{proof} | |
We quote the fact that if $0 \to A \to B \to C \to D \to 0$ | |
is exact then $\rank B + \rank D = \rank A + \rank C$. | |
Then for example the row | |
\begin{center} | |
\begin{tikzcd} | |
\underbrace{H_2(X^1)}_{=0} \ar[r, "0"] & H_2(X^2) \ar[r, hook] & | |
H_2(X^2, X^1) \ar[r, "\partial_2"] & H_1(X^1) \ar[r, two heads] & | |
\underbrace{H_1(X^2)}_{\cong H_1(X)} \ar[r, "0"] | |
& \underbrace{H_1(X^2, X^1)}_{=0} | |
\end{tikzcd} | |
\end{center} | |
from the cellular diagram gives | |
\[ \#(\text{$2$-cells}) + \rank H_1(X) | |
= \rank H_2(X^2) + \rank H_1(X^1). \] | |
More generally, | |
\[ \#(\text{$k$-cells}) + \rank H_{k-1}(X) | |
= \rank H_k(X^k) + \rank H_{k-1}(X^{k-1}) \] | |
which holds also for $k=0$ if we drop the $H_{-1}$ terms | |
(since $\#\text{$0$-cells} = \rank H_0(X^0)$ is obvious). | |
Multiplying this by $(-1)^k$ and summing across $k \ge 0$ | |
gives the conclusion. | |
\end{proof} | |
\begin{example} | |
[Examples of Betti numbers] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii The Betti numbers of $S^n$ are $b_0 = b_n = 1$, | |
and zero elsewhere. The Euler characteristic is $1 + (-1)^n$. | |
\ii The Betti numbers of a torus $S^1 \times S^1$ | |
are $b_0 = 1$, $b_1 = 2$, $b_2 = 1$, and zero elsewhere. | |
Thus the Euler characteristic is $0$. | |
\ii The Betti numbers of $\CP^n$ are $b_0 = b_2 = \dots = b_{2n} = 1$, | |
and zero elsewhere. Thus the Euler characteristic is $n+1$. | |
\ii The Betti numbers of the Klein bottle | |
are $b_0 = 1$, $b_1 = 1$ and zero elsewhere. | |
Thus the Euler characteristic is $0$, the same as the sphere | |
(also since their CW structures use the same number of cells). | |
\end{enumerate} | |
One notices that in the ``nice'' spaces $S^n$, $S^1 \times S^1$ and $\CP^n$ | |
there is a nice symmetry in the Betti numbers, namely $b_k = b_{n-k}$. | |
This is true more generally; see Poincar\'e duality and \Cref{prob:betti}. | |
\end{example} | |
\section{The cellular boundary formula} | |
In fact, one can describe explicitly what the maps $d_n$ are. | |
Recalling that $H_k(X^k, X^{k-1})$ has a basis the $k$-cells of $X$, we obtain: | |
\begin{theorem} | |
[Cellular boundary formula for $k=1$] | |
For $k=1$, \[ d_1 : \Cells_1(X) \to \Cells_0(X) \] is just the boundary map. | |
\end{theorem} | |
\begin{theorem} | |
[Cellular boundary for $k > 1$] | |
Let $k > 1$ be a positive integer. | |
Let $e^k$ be an $k$-cell, and let $\{e_\beta^{k-1}\}_\beta$ | |
denote all $(k-1)$-cells of $X$. | |
Then \[ d_k : \Cells_k(X) \to \Cells_{k-1}(X) \] | |
is given on basis elements by | |
\[ d_k(e^k) = \sum_\beta d_\beta e_\beta^{k-1} \] | |
where $d_\beta$ is be the degree of the composed map | |
\[ S^{k-1} = \partial D_\beta^k \xrightarrow{\text{attach}} | |
X^{k-1} \surjto S_\beta^{k-1}. \] | |
Here the first arrow is the attaching map for $e^k$ | |
and the second arrow is the quotient of collapsing | |
$X^{k-1} \setminus e^{k-1}_\beta$ to a point. | |
\end{theorem} | |
This gives us an algorithm for computing homology groups of a CW complex: | |
\begin{itemize} | |
\ii Construct the cellular chain complex, | |
where $\Cells_k(X)$ is $\ZZ^{\oplus \# \text{$k$-cells}}$. | |
\ii $d_1 : \Cells_1(X) \to \Cells_0(X)$ is just the boundary map | |
(so $d_1(e^1)$ is the difference of the two endpoints). | |
\ii For any $k > 1$, we compute $d_k : \Cells_k(X) \to \Cells_{k-1}(X)$ | |
on basis elements as follows. | |
Repeat the following for each $k$-cell $e^k$: | |
\begin{itemize} | |
\ii For every $k-1$ cell $e^{k-1}_\beta$, | |
compute the degree of the boundary of $e^k$ welded onto | |
the boundary of $e^{k-1}_\beta$, say $d_\beta$. | |
\ii Then $d_k(e^k) = \sum_\beta d_\beta e^{k-1}_\beta$. | |
\end{itemize} | |
\ii Now we have the maps of the cellular chain complex, | |
so we can compute the homologies directly | |
(by taking the quotient of the kernel by the image). | |
\end{itemize} | |
We can use this for example to compute the homology groups of the torus again, | |
as well as the Klein bottle and other spaces. | |
\begin{example} | |
[Cellular homology of a torus] | |
Consider the torus built from $e^0$, $e^1_a$, $e^1_b$ and $e^2$ as before, | |
where $e^2$ is attached via the word $aba\inv b\inv$. | |
For example, $X^1$ is | |
\begin{center} | |
\begin{asy} | |
unitsize(0.8cm); | |
draw(shift(-1,0)*unitcircle, blue, MidArrow); | |
draw(shift(1,0)*rotate(180)*unitcircle, red, MidArrow); | |
label("$e^1_a$", 2*dir(180), dir(180), blue); | |
label("$e^1_b$", 2*dir(0), dir(0), red); | |
dotfactor *= 1.4; | |
dot("$e^0$", origin, dir(0)); | |
\end{asy} | |
\end{center} | |
The cellular chain complex is | |
\begin{center} | |
\begin{tikzcd} | |
0 \ar[r] | |
& \ZZ e^2 \ar[r, "d_2"] | |
& \ZZ e^1_a \oplus \ZZ e^1_b \ar[r, "d_1"] | |
& \ZZ e^0 \ar[r, "d_0"] | |
& 0 | |
\end{tikzcd} | |
\end{center} | |
Now apply the cellular boundary formulas: | |
\begin{itemize} | |
\ii Recall that $d_1$ was the boundary formula. | |
We have $d_1(e^1_a) = e_0 - e_0 = 0$ and similarly $d_1(e^1_b) = 0$. | |
So $d_1 = 0$. | |
\ii For $d_2$, consider the image of the boundary $e^2$ on $e^1_a$. | |
Around $X^1$, it wraps once around $e^1_a$, once around $e^1_b$, | |
again around $e^1_a$ (in the opposite direction), | |
and again around $e^1_b$. | |
Once we collapse the entire $e^1_b$ to a point, | |
we see that the degree of the map is $0$. | |
So $d_2(e^2)$ has no $e^1_a$ coefficient. | |
Similarly, it has no $e^1_b$ coefficient, hence $d_2 = 0$. | |
\end{itemize} | |
Thus \[ d_1=d_2=0. \] | |
So at every map in the complex, the kernel of the map | |
is the whole space while the image is $\{0\}$. | |
So the homology groups are $\ZZ$, $\ZZ^{\oplus 2}$, $\ZZ$. | |
\end{example} | |
\begin{example} | |
[Cellular homology of the Klein bottle] | |
Let $X$ be a Klein bottle. | |
Consider cells $e^0$, $e^1_a$, $e^1_b$ and $e^2$ as before, | |
but this time $e^2$ is attached via the word $abab\inv$. | |
So $d_1$ is still zero, but this time we have | |
$d_2(e^2) = 2e^1_a$ instead (why?). | |
So our diagram looks like | |
\begin{center} | |
\begin{tikzcd}[row sep = tiny] | |
0 \ar[r, "0"] | |
& \ZZ e^2 \ar[r, "d_2"] | |
& \ZZ e^1_a \oplus \ZZ e^1_b \ar[r, "d_1"] | |
& \ZZ e^0 \ar[r, "d_0"] & 0 \\ | |
& e^2 \ar[r, mapsto] & 2e^1_a \\ | |
&& e_1^a \ar[r, mapsto] & 0 && \\ | |
&& e_1^b \ar[r, mapsto] & 0 | |
\end{tikzcd} | |
\end{center} | |
So we get that $H_0(X) \cong \ZZ$, | |
but \[ H_1(X) \cong \ZZ \oplus \Zc2 \] this time | |
(it is $\ZZ^{\oplus 2}$ modulo a copy of $2\ZZ$). | |
Also, $\ker d_2 = 0$, and so now $H_2(X) = 0$. | |
\end{example} | |
\section\problemhead | |
\begin{dproblem} | |
Let $n$ be a positive integer. | |
Show that | |
\[ | |
H_k(\CP^n) \cong | |
\begin{cases} | |
\ZZ & k=0,2,4,\dots,2n \\ | |
0 & \text{otherwise}. | |
\end{cases} | |
\] | |
\begin{hint} | |
$\CP^n$ has no cells in adjacent dimensions, | |
so all $d_k$ maps must be zero. | |
\end{hint} | |
\end{dproblem} | |
\begin{problem} | |
Show that a non-surjective map $f : S^n \to S^n$ has degree zero. | |
\begin{hint} | |
The space $S^n - \{x_0\}$ is contractible. | |
\end{hint} | |
\end{problem} | |
\begin{problem}[Moore spaces] | |
\gim | |
Let $G_1$, $G_2$, \dots, $G_N$ be a sequence of | |
finitely generated abelian groups. | |
Construct a space $X$ such that | |
\[ | |
\wt H_n(X) | |
\cong | |
\begin{cases} | |
G_n & 1 \le n \le N \\ | |
0 & \text{otherwise}. | |
\end{cases} | |
\] | |
\end{problem} | |
\begin{problem} | |
\label{prob:diagram_chase} | |
Prove \Cref{thm:cellular_chase}, | |
showing that the homology groups of $X$ | |
coincide with the homology groups of the cellular chain complex. | |
\begin{hint} | |
You won't need to refer to any elements. | |
Start with \[ H_2(X) \cong H_2(X^3) \cong | |
H_2(X^2) / \ker \left[ H_2(X^2) \surjto H_2(X^3) \right], \] say. | |
Take note of the marked injective and surjective arrows. | |
\end{hint} | |
\begin{sol} | |
For concreteness, let's just look at the homology at $H_2(X^2, X^1)$ | |
and show it's isomorphic to $H_2(X)$. | |
According to the diagram | |
\begin{align*} | |
H_2(X) &\cong H_2(X^3) \\ | |
&\cong H_2(X^2) / \ker \left[ H_2(X^2) \surjto H_2(X^3) \right] \\ | |
&\cong H_2(X^2) / \img \partial_3 \\ | |
&\cong \img\left[ H_2(X^2) \injto H_2(X^2, X^1) \right] / \img \partial_3 \\ | |
&\cong \ker(\partial_2) / \img\partial_3 \\ | |
&\cong \ker d_2 / \img d_3. \qedhere | |
\end{align*} | |
\end{sol} | |
\end{problem} | |
\begin{dproblem} | |
\gim | |
Let $n$ be a positive integer. Show that | |
\[ | |
H_k(\RP^n) | |
\cong | |
\begin{cases} | |
\ZZ & \text{if $k=0$ or $k=n\equiv 1 \pmod 2$} \\ | |
\Zc2 & \text{if $k$ is odd and $0 < k < n$} \\ | |
0 & \text{otherwise}. | |
\end{cases} | |
\] | |
\begin{hint} | |
There is one cell of each dimension. | |
Show that the degree of $d_k$ is $\deg(\id)+\deg(-\id)$, | |
hence $d_k$ is zero or $\cdot 2$ depending | |
on whether $k$ is even or odd. | |
\end{hint} | |
\end{dproblem} | |