Skip to content

Commit

Permalink
Various changes
Browse files Browse the repository at this point in the history
* Make notation consistent: use `|` instead of `:` for set comprehension, and `#` instead of `*` for
  induced map between fundamental group
* Category theory: Add explicit example for natural transformation
* Category theory: Add some intuition for additive category and short exact sequence
* Algebraic topology: Add some explanation for Ext functor
* Minor grammar fixes
  • Loading branch information
user202729 committed Apr 8, 2024
1 parent 8a733c5 commit 8832fcc
Show file tree
Hide file tree
Showing 14 changed files with 202 additions and 31 deletions.
2 changes: 1 addition & 1 deletion tex/alg-geom/sheaves.tex
Original file line number Diff line number Diff line change
Expand Up @@ -524,7 +524,7 @@ \section{Sheaves}

\ii Let $X = \RR$, and define the presheaf of rings $\SF$ by:
\[
\SF(U) = \{ f \colon U \to \RR: \text{there exists continuous $g
\SF(U) = \{ f \colon U \to \RR \mid \text{there exists continuous $g
\colon \RR \to \RR$ such that $g\restrict{U} = f$} \}.
\]
Then $\SF$ is not a sheaf. Indeed, $s_1(x) = 0$ in $\SF((-1, 0))$
Expand Down
53 changes: 53 additions & 0 deletions tex/cats/abelian.tex
Original file line number Diff line number Diff line change
Expand Up @@ -104,6 +104,12 @@ \section{Additive and abelian categories}
The zero map serves as the identity element for each group.
\end{itemize}
\end{definition}
In short:
\begin{moral}
In an additive category, you can add two morphisms.
\end{moral}
Which is the only definition that makes sense anyway, we cannot talk about elements.

\begin{definition}
An \vocab{abelian category} $\AA$ is one with the additional properties that
for any morphism $A \taking f B$,
Expand Down Expand Up @@ -136,6 +142,9 @@ \section{Additive and abelian categories}
\end{enumerate}
\end{example}

From now on, you can basically forget about additive category, we will be working in abelian
category.

In general, once you assume a category is abelian, all the properties you would want
of these kernels, cokernels, \dots\ that you would guess hold true.
For example,
Expand Down Expand Up @@ -181,6 +190,7 @@ \section{Additive and abelian categories}
\end{proof}

\section{Exact sequences}
\label{sec:exact_sequences}
\prototype{$0 \to G \to G \times H \to H \to 0$ is exact.}
Exact sequences will seem exceedingly unmotivated until you learn about homology groups,
which is one of the most natural places that exact sequences appear.
Expand Down Expand Up @@ -209,6 +219,31 @@ \section{Exact sequences}
\end{enumerate}
\end{example}

If you look at the prototypical example, actually, a \vocab{short exact sequence} (an exact sequence
of the form $0 \to A \to B \to C \to 0$) is the most natural things ever:
\begin{moral}
It's basically just an equation $C = B/A$.
\end{moral}
Whenever you see ``there is a short exact sequence $0 \to A \to B \to C \to 0$'', you
can mentally translate it to ``$C \cong B/A$''; but there's a slight difference:
A group has more structures than a number, so the sequence also contains the information of the
maps --- the map that identifies $A$ with a subgroup of $B$, and the map that identifies $C$
with the quotient group $B/A$.
\begin{example}[More exact sequences]
\listhack
\begin{enumerate}[(a)]
\ii The sequence
\[ 0 \to \ZZ \taking{\times 3} \ZZ \to \Zc{3} \to 0 \]
is short exact.
\ii So is
\[ 0 \to \ZZ \taking{\times 5} \ZZ \to \Zc{5} \to 0. \]
\end{enumerate}
As you can see, the written equation ``$C \cong B/A$'' is not completely accurate, the map $A
\to B$ also matters in determining what $C$ is. This also explains the common notation: the
image of the map $\ZZ \taking{\times 3} \ZZ$ is usually written $3 \ZZ$, thus $\Zc{3} =
\frac{\ZZ}{3 \ZZ}$.
\end{example}

Now, we want to mimic this definition in a general \emph{abelian} category $\AA$.
So, let's write down a criterion for when $A \taking f B \taking g C$ is exact.
First, we had better have that $g \circ f = 0$,
Expand Down Expand Up @@ -450,3 +485,21 @@ \section{Breaking long exact sequences}
Prove that there is an exact sequence
\[ \Ker a \to \Ker b \to \Ker c \to \Coker a \to \Coker b \to \Coker c. \]
\end{sproblem}

\begin{problem}
[An additive category that is not abelian]
Consider a category, where:
\begin{itemize}
\ii the objects are pairs of abelian groups $(B, A)$ where $A$ is a subgroup of $B$.
\ii the morphisms $(B, A) \to (B', A')$ are maps $f \colon B \to B'$ where
$f\im(A) \subseteq A'$.
\end{itemize}
(You can think of this similar to the $\catname{PairTop}$ category, seen in
\Cref{ch:relative_hom}. We use abelian groups here to make the category additive.)

This category can be equivalently viewed as the category of short exact sequences
$0 \to A \to B \to B/A \to 0$ of abelian groups.

Show that the arrow $(X, 0) \to (X, X)$ is monic and epic, but not an
isomorphism. Conclude that the category is not abelian.
\end{problem}
6 changes: 5 additions & 1 deletion tex/cats/functors.tex
Original file line number Diff line number Diff line change
Expand Up @@ -506,6 +506,8 @@ \section{(Optional) Natural transformations}
the fact that it is an isomorphism follows from the fact that $V$ and $(V^\vee)^\vee$
have equal dimensions and $\eps_V$ is injective).

Another example can be found in \Cref{remark:hurewicz}.

\section{(Optional) The Yoneda lemma}
Now that I have natural transformations, I can define:
\begin{definition}
Expand Down Expand Up @@ -603,7 +605,9 @@ \section{(Optional) The Yoneda lemma}
Each $A$ gives us some information on how $X$ and $Y$ are similar,
but the whole natural isomorphism is strong enough to imply $X \cong Y$.

\ii Consider the functor $U \colon \catname{Grp} \to \catname{Set}$.
\ii Consider the covariant forgetful functor $U \colon \catname{Grp} \to
\catname{Set}$.\footnote{Actually, you need to apply a dual version.
\Cref{thm:yoneda} uses contravariant functor.}
It can be represented by $H^\ZZ$, in the sense that
\[ \Hom_{\catname{Grp}}(\ZZ, G) \cong U(G)
\qquad\text{ by }\qquad \phi \mapsto \phi(1). \]
Expand Down
4 changes: 2 additions & 2 deletions tex/diffgeo/manifolds.tex
Original file line number Diff line number Diff line change
Expand Up @@ -547,8 +547,8 @@ \subsection{Sanity check}
Then
\[
\begin{aligned}
\km_0 &= \{ \text{smooth functions } f \colon f(0) = 0 \} \\
\km_0^2 &= \{ \text{smooth functions } f \colon f(0) = 0, (\nabla f)_0 = 0 \}.
\km_0 &= \{ \text{smooth functions } f \mid f(0) = 0 \} \\
\km_0^2 &= \{ \text{smooth functions } f \mid f(0) = 0, (\nabla f)_0 = 0 \}.
\end{aligned}
\]
In other words $\km_0^2$ is the set of functions which vanish at $0$
Expand Down
71 changes: 67 additions & 4 deletions tex/homology/cohomology.tex
Original file line number Diff line number Diff line change
Expand Up @@ -308,6 +308,67 @@ \section{Universal coefficient theorem}
$\Ext(\ZZ^{\oplus 2015}, G)$ and $\Ext(\Zc{30}, \Zc 4)$.
\end{ques}

\section{Explanation for universal coefficient theorem}

There is so much unexplained symbols and formulas in the previous chapter that may make you scream:
\begin{quote}
I don't care if $\CP^2$ and $S^2 \vee S^4$ are distinct anymore! What are these spaces anyway?
\end{quote}
Nevertheless, it is not all that difficult. There are two key points to be read from the theorem:
\begin{itemize}
\ii Even though $H_n(A_\bullet) = 0$, it is still possible for $H^n(A_\bullet; G) \neq 0$
if $\Ext(H_{n-1}(A_\bullet), G) \neq 0$.

In low-dimensional cases, we can actually visualize it --- \Cref{sec:visual_cohom_groups}
does that for the Klein bottle.
\ii $H^n(A_\bullet; G)$ is uniquely determined by $H_n(A_\bullet)$ and $G$, regardless of what
$A_\bullet$ is, as long as each $A_n$ is free.
\end{itemize}

Which means: if you wish, you can forget about the formula in the universal coefficient theorem,
and use the cellular chain complex $\Cells_\bullet(X)$ to compute cohomology by:
\[ H^n(X; G) = \frac{\ker(\Hom(\Cells_n(X), G) \to \Hom(\Cells_{n+1}(X), G))}
{\img(\Hom(\Cells_{n-1}(X), G) \to \Hom(\Cells_{n}(X), G))}. \]
After all, the cellular chain complex and the singular chain complex are both free and have the same
homology groups, so by the universal coefficient theorem they must have the same cohomology groups.

Nevertheless, the formula of the universal coefficient theorem is desirable because, more often than
not, the chain complex $A_\bullet$ is more complicated than $H_\bullet(A_\bullet)$.

\begin{example}
The Klein bottle's cellular chain complex has the following form:
\[ \cdots \to \ZZ \taking{1 \mapsto (0, 2)} \ZZ^2 \taking {(a, b) \mapsto 0} \ZZ. \]
The homology groups is:
\[ H_2 = 0, H_1 = \ZZ \oplus \Zc{2}, H_0 = \ZZ. \]
It's indeed simpler, but only marginally (there are $3$ generators instead of $4$, and we don't
need to keep track of the maps)
because cellular homology is already so efficient.
\end{example}

Where does the formula come from, again? You can think of it like this.
Because the universal coefficient theorem tells us that $H^\bullet(A_\bullet; G)$ only depends on
$H_\bullet(A_\bullet)$, if we're given $H_\bullet$, we just construct \emph{any} chain complex of
free abelian groups $A_\bullet$ and dualize it.

Assume $H_k = 0$ for every terms, except $H_{n-1}\neq 0$. Then, tautologically,
$H^n \cong \Ext(H_{n-1}; G)$ --- a free resolution \emph{is} a chain complex!

\begin{exercise}
Verify this. (Hint: Starting from the exact sequence $Z_{n-1} \to H_{n-1} \to 0$.
Can you extend it to a free resolution of $H_{n-1}$?)
\end{exercise}

Assume $H_k = 0$ for every terms, except $H_n \neq 0$. Then we can see $H^n \cong \Hom(H_n, G)$.

The universal coefficient theorem simply states that the choice of free resolution doesn't matter,
and that if the other terms can be nonzero, $H^n$ is the direct sum of the two groups in the two
cases above.

If you want, you can even prove the fact that the choice of free resolution does not matter yourself
--- it's a bit tricky, but not all that difficult. It boils down to the construction of maps
between the chain complexes (it's not difficult to ensure the diagram commutes, the groups are free
so we can send the basis wherever we want), and show the two free resolutions are chain homotopic.

\section{Example computation of cohomology groups}
\prototype{Possibly $H^n(S^m)$.}

Expand Down Expand Up @@ -397,6 +458,7 @@ \section{Example computation of cohomology groups}
\end{example}

\section{Visualization of cohomology groups}
\label{sec:visual_cohom_groups}

We try to make sense of $C^n(X; G)$ and $H^n(X; G)$, for higher values of $n$.

Expand Down Expand Up @@ -461,10 +523,11 @@ \section{Visualization of cohomology groups}
\begin{moral}
Homotopic chains with the same boundary are mapped to the same value by cocycles.
\end{moral}
We defined what it means for two $k$-simplex to be homotopic in \Cref{sec:higher_homotopy_grps} ---
in the current situation, we require in addition that the boundaries are always fixed.
We defined what it means for two $k$-simplices to be homotopic in \Cref{sec:higher_homotopy_grps}
--- in the current situation, we require in addition that the boundaries are always fixed.

For instance, the blue and the orange $1$-simplex below are homotopic, but not the red $1$-simplex.
For instance, the blue and the orange $1$-simplices below are homotopic, but not the red
$1$-simplex.
\begin{center}
\begin{asy}
size(4cm);
Expand All @@ -481,7 +544,7 @@ \section{Visualization of cohomology groups}
\end{center}

Proof is not difficult --- you just need to show that the difference between two homotopic
$k$-simplex is the boundary of something (their interior!), and write the interior as the
$k$-simplices is the boundary of something (their interior!), and write the interior as the
sum of some $k+1$-simplices. (Hint: The easiest way is actually to write the interior as the
difference of two $k+1$-simplices instead, and be careful of vertex ordering issues.)
\begin{exercise}
Expand Down
7 changes: 3 additions & 4 deletions tex/homology/excision.tex
Original file line number Diff line number Diff line change
Expand Up @@ -21,8 +21,7 @@ \section{Motivation}

The main motivation is that:
\begin{moral}
Relative homology is the algebraic topology analog of quotient homology.
% (or analogue?)
Relative homology is the algebraic analog of quotient space.
\end{moral}
So, for instance, when you see a map of pairs
$f \colon (X, A) \to (Y, B)$, you should think of $X/A \to Y/B$.
Expand Down Expand Up @@ -104,8 +103,8 @@ \section{Motivation}
``finitely many circles'' (or all but finitely many).

We haven't had enough tools to formalize all these yet.
Formally speaking, the quotient map $q \colon X \to X/A$ and $q \colon A \to A/A$ induces $q_*
\colon H(X, A) \to H(X/A, A/A)$, and $q_*$ is not injective.
Formally speaking, the quotient maps $q \colon X \to X/A$ and $q \colon A \to A/A$ induces $q_*
\colon H_1(X, A) \to H_1(X/A, A/A)$, and $q_*$ is not injective.
\end{example}

Regardless, for nice spaces $A \subseteq X$ such that $H_n(X, A) \cong \wt H_n(X/A)$, we would be
Expand Down
6 changes: 6 additions & 0 deletions tex/homology/long-exact.tex
Original file line number Diff line number Diff line change
Expand Up @@ -46,6 +46,12 @@ \section{Short exact sequences and four examples}
\end{center}
\end{definition}

\begin{moral}
This basically means $C_\bullet = B_\bullet / A_\bullet$, for suitable definition of
$/$ on chain complexes.
\end{moral}
This agrees with the definition in \Cref{sec:exact_sequences}.

\begin{example}
[Mayer-Vietoris short exact sequence and its augmentation]
\label{ex:mayer_short_exact}
Expand Down
32 changes: 32 additions & 0 deletions tex/homology/singular.tex
Original file line number Diff line number Diff line change
Expand Up @@ -252,6 +252,38 @@ \section{The singular homology groups}
are exactly the nulhomotopic loops in $\pi_1(X, x_0)$.
The difference is that $H_1(X)$ allows loops to commute,
whereas $\pi_1(X, x_0)$ does not.

\begin{remark}[Digression: category theory interpretation]
\label{remark:hurewicz}
From this, you can say that there is a Hurewicz map $\pi_1(X, x_0) \taking\phi H_1(X)$
for each $(X, x_0)$.

But there is more than that: this map is \emph{natural}, in the sense that for $h \colon (X,
x_0) \to (Y, y_0)$ map of pointed spaces, then
\begin{center}
\begin{tikzcd}
\pi_1(X, x_0) \rar["h_\sharp"] \dar["\phi"] & \pi_1(Y, y_0) \dar["\phi"] \\
H_1(X) \rar["h_\ast"] & H_1(Y) \\
\end{tikzcd}
\end{center}
commutes.

In category theory terms, we say that $\phi$ is a \emph{natural transformation}
from $\pi_1$ to $H_1$.

Another way to say this is: we have families of groups
\[ \{ \pi_1(X, x_0) \mid (X, x_0)\text{ pointed space} \} \]
and
\[ \{ H_1(X) \mid (X, x_0)\text{ pointed space} \} \]
then the natural transformation $\phi$ can be seen as a family of homomorphisms
\[ \{ \phi \colon \pi_1(X, x_0) \to H_1(X) \mid (X, x_0)\text{ pointed space} \} \]
satisfying the naturality conditions.

Of course, the fact that $\pi_1$ is a functor means
$\{ \pi_1(X, x_0) \mid (X, x_0)\text{ pointed space} \}$ is a lot more than a family of groups
indexed by pointed spaces, as explained in \Cref{thm:fundgrp_functor}.
\end{remark}

\begin{example}
[The first homology group of the annulus]
To give a concrete example, consider the annulus $X$ above.
Expand Down
6 changes: 3 additions & 3 deletions tex/measure/martingale.tex
Original file line number Diff line number Diff line change
Expand Up @@ -98,9 +98,9 @@ \section{Sub-$\sigma$-algebras and filtrations}
variable taking on one of three values.
If we're interested in $X$ then we could define
\begin{align*}
A &= \{\omega \colon X(\omega) = 1\} \\
B &= \{\omega \colon X(\omega) = 2\} \\
C &= \{\omega \colon X(\omega) = 3\}
A &= \{\omega \mid X(\omega) = 1\} \\
B &= \{\omega \mid X(\omega) = 2\} \\
C &= \{\omega \mid X(\omega) = 3\}
\end{align*}
then we could write
\[ \SF = \left\{ \varnothing, \; A, \; B, \; C, \;
Expand Down
4 changes: 2 additions & 2 deletions tex/measure/pontryagin.tex
Original file line number Diff line number Diff line change
Expand Up @@ -62,8 +62,8 @@ \section{LCA groups}
That means that $\mu$ is ``translation-invariant''
under translation by $G$.
\ii $\mu(K)$ is finite for any compact set $K$.
\ii if $S$ is measurable, then $\mu(S) = \inf\left\{ \mu(U) \colon U \supseteq S \text{ open} \right\}$.
\ii if $U$ is open, then $\mu(U) = \sup\left\{ \mu(S) \colon S \supseteq U \text{ measurable} \right\}$.
\ii if $S$ is measurable, then $\mu(S) = \inf\left\{ \mu(U) \mid U \supseteq S \text{ open} \right\}$.
\ii if $U$ is open, then $\mu(U) = \sup\left\{ \mu(S) \mid S \supseteq U \text{ measurable} \right\}$.
\end{itemize}
Moreover, it is unique up to scaling by a positive constant.
\end{theorem}
Expand Down
2 changes: 1 addition & 1 deletion tex/riemann-surface/complex-structure.tex
Original file line number Diff line number Diff line change
Expand Up @@ -209,7 +209,7 @@ \section{Complex structures}
equivalent to stating that a Riemann surface carries a complex-analytic structure.

\section{Riemann surface}
\prototype{The Riemann sphere, or any open subset of $\CC$ such as $\{ z \in \CC: |z| < 1 \}$.}
\prototype{The Riemann sphere, or any open subset of $\CC$ such as $\{ z \in \CC \mid |z| < 1 \}$.}

From the motivation above, the definition of a Riemann surface naturally falls out:
\begin{definition}[Riemann surface]
Expand Down
6 changes: 3 additions & 3 deletions tex/set-theory/CH.tex
Original file line number Diff line number Diff line change
Expand Up @@ -308,12 +308,12 @@ \section{Infinite combinatorics}
\end{lemma}
\begin{proof}
Assume not. Let
\[ \left\{ p_\alpha \colon \alpha < \omega_1 \right\} \]
\[ \left\{ p_\alpha \mid \alpha < \omega_1 \right\} \]
be a strong antichain. Let
\[ C = \left\{ \dom(p_\alpha) \colon \alpha < \omega_1 \right\}. \]
\[ C = \left\{ \dom(p_\alpha) \mid \alpha < \omega_1 \right\}. \]
Let $\ol C \subseteq C$ be such that $\ol C$ is uncountable, and $\ol C$ is a $\Delta$-system with root $R$.
Then let
\[ B = \left\{ p_\alpha \colon \dom(p_\alpha) \in R \right\}. \]
\[ B = \left\{ p_\alpha \mid \dom(p_\alpha) \in R \right\}. \]
Each $p_\alpha \in B$ is a function $p_\alpha \colon R \to \{0,1\}$,
so there are two that are the same.
\end{proof}
Expand Down
20 changes: 16 additions & 4 deletions tex/topology/cover-project.tex
Original file line number Diff line number Diff line change
Expand Up @@ -61,7 +61,19 @@ \section{Even coverings and covering projections}
This is essentially wrapping the real line
into a single helix and projecting it down.
\end{example}
\missingfigure{helix}
\begin{center}
\begin{asy}
var flatten = yscale(0.4);
guide g;
for (real a=90+360*5; a>=90; a-=20){
g = g .. flatten*dir(a)+(0, a/360/3);
}
draw(g, Arrows);
draw((0, -0.4)--(0, -1.4), Arrow, L=Label("$p$"));
draw(shift(0, -2)*flatten*unitcircle, L=Label("$S^1$", Relative(0.1)));
\end{asy}
\end{center}


We claim this is a covering projection.
Indeed, consider the point $1 \in S^1$
Expand Down Expand Up @@ -254,13 +266,13 @@ \section{Lifting theorem}
and consider a covering projection $p \colon (E, e_0) \to (B, b_0)$.
(As usual, $Y$, $B$, $E$ are path-connected.)
Then a lifting $\tilde f$ with $\tilde f(y_0) = e_0$ exists if and only if
\[ f_\ast\im(\pi_1(Y, y_0)) \subseteq p_\ast\im(\pi_1(E, e_0)), \]
\[ f_\sharp\im(\pi_1(Y, y_0)) \subseteq p_\sharp\im(\pi_1(E, e_0)), \]
i.e.\ the image of $\pi_1(Y, y_0)$ under $f$ is contained in
the image of $\pi_1(E, e_0)$ under $p$ (both viewed as subgroups of $\pi_1(B, b_0)$).
If this lifting exists, it is unique.
\end{theorem}
As $p_\ast$ is injective,
we actually have $p_\ast\im(\pi_1(E, e_0)) \cong \pi_1(E, e_0)$.
As $p_\sharp$ is injective,
we actually have $p_\sharp\im(\pi_1(E, e_0)) \cong \pi_1(E, e_0)$.
But in this case we are interested in the actual elements, not just the isomorphism classes of the groups.
\begin{ques}
What happens if we put $Y= [0,1]$?
Expand Down
Loading

0 comments on commit 8832fcc

Please sign in to comment.