-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathappendixB.tex
187 lines (171 loc) · 11.8 KB
/
appendixB.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
%!TEX root = main.tex
%
% appendixB.tex
%
\chapter{Category Theory}
Most can be found in \cite{MacLaneMoerdijk91}.
\begin{lemma}
\label{lem:mono iff pullback}
Let $\mathcal{C}$ be a category with pullbacks. If $f : A \to B$ is a morphism in $\mathcal{C}$, then $f$ is a monomorphism if and only if the diagram
\[
\begin{tikzcd}
A \arrow[r, equal] \arrow[d, equal] & A \arrow[d, "f"] \\
A \arrow[r, "f"] & B
\end{tikzcd}
\]
is a pullback diagram.
\end{lemma}
\begin{proof}
Suppose $f$ is mono. Let $P$ be the pullback of $f$ with itself. There exists a unique $\phi : A \to P$ such that
\[
\begin{tikzcd}
A \arrow[dr, dotted, "\exists ! \phi"] \arrow[drr, "\id_A", bend left=20] \arrow[ddr, "\id_A", bend right=20] \\
& P \arrow[r, "p_1"] \arrow[d, "p_2"] & A \arrow[d, "f"] \\
& A \arrow[r, "f"] & B
\end{tikzcd}
\]
Notice first that $p_1 = p_2$, using that $f$ is mono. Next, we have $f p_1 \phi = f \id_A$, so $p_1 \phi = \id_A$. Monomorphisms are stable under pullbacks, so $p_1$ is also mono. Hence, we have $p_1 \phi p_1 = p_1 \id_P \Rightarrow \phi p_1 = \id_P$. We conclude that $P \cong A$.
On the other hand, Let $\begin{tikzcd} C \arrow[r, shift left=.50ex, "g"] \arrow[r, shift right=.50ex, swap, "h"] & A \arrow[r, "f"] & B \end{tikzcd}$ be morphisms such that $fg = fh$. Then there exists a unique morphism $\phi : C \to A$, such that the diagram
\[
\begin{tikzcd}
C \arrow[dr, dotted, "\exists ! \phi"] \arrow[drr, "g", bend left=20] \arrow[ddr, "h", bend right=20] \\
& A \arrow[r, equal] \arrow[d, equal] & A \arrow[d, "f"] \\
& A \arrow[r, hook, "f"] & B
\end{tikzcd}
\]
commutes. In other words, $g = \phi = h$.
\end{proof}
\begin{corollary}
Let $\mathcal{C}$ be a category with pushouts. If $f : A \to B$ is a morphism in $\mathcal{C}$, then $f$ is an epimorphism if and only if the diagram
\[
\begin{tikzcd}
A \arrow[r, "f"] \arrow[d, "f"] & B \arrow[d, equal] \\
A \arrow[r, equal] & B
\end{tikzcd}
\]
is a pushout.
\end{corollary}
\begin{proof}
Follows from duality of Lemma \ref{lem:mono iff pullback}.
\end{proof}
\begin{lemma}
Every equalizer is a monomorphism.
\end{lemma}
\begin{proof}
Let $\mathcal{C}$ be a category with equalizers. Suppose that
\[ \begin{tikzcd} E \arrow[r, "e"] & A \arrow[r, shift left=0.50ex, "f"] \arrow[r, swap, shift right=0.50ex, "g"] & B \end{tikzcd} \]
is an equalizer diagram in $\mathcal{C}$. Suppose that there are morphisms
\[ \begin{tikzcd} F
\arrow[r, shift left=0.50ex, "x"]
\arrow[r, swap, shift right=0.50ex, "y"] & E \arrow[r, "e"] & A \end{tikzcd} \]
such that $ex = ey$. Clearly, $fex = gex$. It follows that, as shown in the diagram below, there exists a unique commuting $u : F \to E$ because $e$ is an equalizer.
\[ \begin{tikzcd} E \arrow[r, "e"] & A
\arrow[r, shift left=0.50ex, "f"]
\arrow[r, swap, shift right=0.50ex, "g"] & B \\
F
\arrow[ur, swap, "ex = ey"]
\arrow[u, dotted, "\exists ! u"] \end{tikzcd} \]
By uniqueness of this $u$, we may conclude that $x=u=y$.
\end{proof}
\begin{proposition}
In the category $\mathbf{Sets}$, monomorphisms are in bijection with injective maps, and epimorphisms are in bijection with surjective maps.
\end{proposition}
\begin{proof}
Suppose $f : A \to B$ is a monomorphism. We want to show that $f$ is injective. So let $a,a' \in A$ be such that $f(a) = f(a')$. Consider the diagram
\[ \begin{tikzcd} *
\arrow[r, shift left=0.50ex, "a"]
\arrow[r, swap, shift right=0.50ex, "a'"]& A \arrow[r, "f"] & B \end{tikzcd} \]
where $a(*) = a \in A$ and $a'(*) = a' \in A$. Then $f a = f a'$, so $a = a'$. We have shown that $f$ is injective. Assume now instead that $f$ is injective. We want to show that it is mono. So let
\[ \begin{tikzcd} C
\arrow[r, shift left=0.50ex, "g"]
\arrow[r, swap, shift right=0.50ex, "h"]
& A \arrow[r, "f"] & B \end{tikzcd} \]
be an arbitrary diagram such that $fg = fh$. This means that for all $c \in C$, we have $f(g(c)) = f(h(c))$. Since $f$ is injective, this implies that for all $c \in C$ we have $g(c) = h(c)$, in other words that $g = h$. We have shown that $f$ is mono. Suppose now that $f$ is an epimorphism. We want to show that $f$ is surjective. Let $b \in B$. Consider the diagram
\[ \begin{tikzcd} A \arrow[r, "f"] & B \arrow[r, two heads] & \{b\} \end{tikzcd} \]
\end{proof}
\begin{definition}
Let $\mathcal{C}$ be a locally small category and let $C \in \mathcal{C}$. A \emph{sieve on $C$} is a subfunctor $S \subseteq \Hom(\cdot,C)$.
\end{definition}
Note that a subfunctor just means that there is a monic natural transformation from $S$ to $\Hom(\cdot,C)$.
If $S$ is a sieve on $C$ and $h : D \to C$ a morphism, then the pullback sieve $h^*S$ is a sieve on $D$; just compose with $h$. We call a sieve $S$ a \emph{maximal} sieve if $S = \Hom(\cdot, C)$.
\begin{lemma}
Let $S$ be a sieve on $C$. Then $S$ is a maximal sieve if and only if $\id_C \in S(C)$.
\end{lemma}
\begin{proof}
Suppose that $\id_C \in S(C)$. Let $f : D \to C$ be a morphism. The natural inclusion $i : S \to \Hom(\cdot, C)$ induces a commutative diagram
\[
\begin{tikzcd}
S(C) \arrow[r, "f^*"] \arrow[d, swap, "i_C"] & S(D) \arrow[d, "i_D"] & \id_C \arrow[r, mapsto] \arrow[d, mapsto] & f \arrow[d, mapsto] \\
\Hom(C,C) \arrow[r, "f^*"] & \Hom(D,C) & \id_C \arrow[r, mapsto] & f
\end{tikzcd}
\]
Since $f$ was arbitrary, $S = \Hom(\cdot, C)$.
\end{proof}
If $\mathcal{C}$ and $\mathcal{D}$ are categories, we denote by $[\mathcal{C},\mathcal{D}]$ the category of functors $\mathcal{C} \to \mathcal{D}$, with as morphisms the natural transformations between them.
\begin{definition}
Let $\mathcal{C}$ be a locally small category. We set
\[ \widehat{\mathcal{C}} := [\mathcal{C}^{\text{op}}, \mathbf{Sets}].\]
Objects of this category are called \emph{presheaves}.
\end{definition}
The motivation for presheaves stems from topological spaces. If $X$ is a topological space, then we have a category $\mathcal{O}(X)$ where the objects are the open sets of $X$ and the morphisms are inclusions $U \subseteq V$ of open sets. A presheaf $F$ on $\mathcal{O}(X)$ is then a rule which associates to every open $U$ a set $FU$ and whenever we have an inclusion $U \subseteq V$, we require a map $FV \to FU$, called a \emph{restriction} map. For instance, in order to study continuous maps $X \to \mathbb{R}$, one could restrict attention to the presheaf $F$ which assigns to every open $U$ the set of all continous maps $U \to \mathbb{R}$. The restriction map $FV \to FU$ can then be defined as actual restriction of continuous functions.
The term ``presheaf'' suggests that there is also a definition of a sheaf. This is indeed the case. So what's wrong with presheaves themselves? Suppose that $X,Y$ are topological spaces and $A,B \subseteq X$ open sets. We do not require that $A \cap B \neq \emptyset$, as this is a perfectly possible option. Suppose that we have two continuous maps $f : A \to Y$ and $g : B \to Y$. From the gluing lemma of topology, we know that if $f|_{A \cap B} = g|_{A \cap B}$, then there exists a unique continuous map $h : A \cup B \to Y$ such that $h|_{A} = f$ and $h|_{B} = g$.
Presheaves do not guarantee the existence of such $h$. To give an example, let $X = \{a,b\}$ be the discrete space with two points. For our presheaf $F \in \widehat{\mathcal{O}(X)}$, we set $F\emptyset = FX = F\{a\} = F\{b\} = \Z$. We can regard the elements of $FU$ as constant functions $U \to \Z$. Then $0 \in F\{a\}$ and $1 \in F\{b\}$, and the fact that they agree on the intersection $\{a\} \cap \{b\}$ is vacuously true, yet we cannot ``glue'' the constant functions $0$ and $1$ to some constant function $n \in \Z$.
The notion of a sheaf fixes this problem by forcing this gluing property on it.
\begin{definition}
Let $F \in \widehat{\mathcal{O}(X)}$. Then $F$ is called a \emph{sheaf} if for every open $U \in \mathcal{O}(X)$ and for every open cover $\{U_i\}_{i \in I}$ the diagram
\[
\begin{tikzcd}
FU \arrow[r, dotted, "e"] & \prod_{i \in I} FU_i \arrow[r, shift left=.50ex, "g"] \arrow[r, shift right=.50ex, swap, "h"] & \prod_{i,j \in I} F(U_i \cap U_j)
\end{tikzcd}
\]
is an equalizer, where $g\{s_i\} := \{s_i|_{U_i \cap U_j}\}$ and $h\{s_j\} := \{s_j|_{U_i \cap U_j}\}$.
\end{definition}
The natural numbers $\N$ include $0$, so $\N = \{0,1,2,\ldots\}$.
\begin{definition}
Let $n \in \N$ By $\underline{n}$ we denote the subset $\{0,1,\ldots,n-1,n\} \subset \N$. If $m,n \in \N$, then a map $f : \underline{m} \to \underline{n}$ is called an \emph{order-preserving} map if $x \leq y \Rightarrow f(x) \leq f(y)$ for all $x,y \in \underline{m}$. By $\Delta$ we denote the category whose objects are the $\underline{n} \subset \N$ and whose morphisms are the order-preserving maps. The objects of the presheaf category $\widehat{\Delta}$ are called \emph{simplicial sets}.
\end{definition}
Specifying a simplicial set $X \in \widehat{\Delta}$ amounts to giving a set $X_n$ for each $n \in \N$, and for each order-preserving map $f : \underline{m} \to \underline{n}$ a map of sets $Xf : X_n \to X_m$ with the usual requirements of $X$ being a functor. A map between simplicial sets $X,Y \in \widehat{\Delta}$ is then, of course, a natural transformation of functors $X \to Y$.
\begin{definition}
For each $n \in \N$, the \emph{standard geometric $n$-simplex} is the topological space
\[ \Delta_n := \left\{ x \in \mathbb{R}^{n+1} : \sum_{i=0}^n x_i = 1,\; x_i \geq 0 \right\} \]
\end{definition}
\begin{definition}
For each $n \in \N$, the \emph{standard $n$-simplex} is the (representable) simplicial set $\Hom_{\Delta}(\cdot, \underline{n})$.
\end{definition}
Hence, we see that standard $n$-simplices are exactly the maximal sieves.
\begin{lemma}
There exists a functor $F: \Delta \to \mathbf{Top}$.
\end{lemma}
\begin{proof}
Given $\underline{n} \in \Delta$, we send it to the geometric $n$-simplex $ F \underline{n} := \Delta_n$. Given an order-preserving map $f : \underline{n} \to \underline{m}$, we send it to the map $Ff : \Delta_n \to \Delta_m$ defined as follows:
\[ \left( x_0,\ldots,x_n \right) \mapsto \left( y_0, \ldots, y_m \right), \qquad y_j := \sum_{f(i) = j}x_i. \]
It remains to be seen why this assignment gives us a continuous map $Ff : \Delta_n \to \Delta_m$.
% We do this by induction on $m$. For $m=0$, the object $\underline{m} \in \Delta$ is terminal, so there is for each $n$ only one morphism $\underline{n} \to \underline{0}$. Then $y_0 = \sum_{i} x_i = 1 \in \Delta_0$, so this is a continuous map. Suppose now that the statement is proven for all $0 \leq m < M$. Consider $m=M$. Since $f$ is order-preserving, $m-1 < M \Rightarrow f(m-1) \leq f(M)$.
The first thing to see is that $Ff$ is well-defined. Clearly, $y_j \geq 0$ for all $j \in \underline{m}$ since $x_i \geq 0$ for all $i \in \underline{n}$. Also
\[ \sum_{j=0}^m y_j = \sum_{j=0}^m \sum_{f(i) = j} x_i = \sum_{f(i)=0}x_i + \sum_{f(i)=1}x_i + \ldots + \sum_{f(i)=m}x_i \]
\[ = \sum_{i \in f^{-1}(0)}x_i + \ldots + \sum_{i \in f^{-1}(m)}x_i = \sum_{i \in \underline{n}}x_i = 1. \]
In the last equation, we use that $f$ is order preserving: Whenever $0 \leq a < b \leq m$, this implies that for all $x \in f^{-1}(a)$ and all $y \in f^{-1}(b)$, we have $x < y$. Hence $Ff$ is well-defined. Since $Ff$ is linear, it is continuous.
\end{proof}
\begin{lemma}
Let $X \in \widehat{\Delta}$ be a simplicial set. Then
\[ X_n \cong \Hom_{\widehat{\Delta}}\left( \Hom_{\Delta}\left( \cdot, \underline{n} \right), X \right), \qquad n \in \N. \]
That is, any simplicial set can be seen as mapping the standard $n$-simplices into it.
\end{lemma}
\begin{proof}
Yoneda. (TODO?)
\end{proof}
\begin{theorem}
Every simplical set $X \in \widehat{\Delta}$ is a colimit of standard $n$-simplices.
\end{theorem}
\begin{proof}
TODO.
\end{proof}
\begin{theorem}
\end{theorem}
\begin{theorem}
There exists an adjoint pair of functors
\[ \begin{tikzcd}
\widehat{\Delta} \arrow[r, shift right=.50ex, swap, "|\cdot|"] & \arrow[l, shift right=.50ex, swap, "\Sing"] \mathbf{Top}
\end{tikzcd} \]
where $\Sing_n X := \Hom_{\mathbf{Top}}(\Delta_n, X)$ and $|X| := ...$
\end{theorem}