-
Notifications
You must be signed in to change notification settings - Fork 0
/
locallyconst.tex
196 lines (172 loc) · 19.3 KB
/
locallyconst.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
188
189
190
191
192
193
194
195
196
%!TEX root = main.tex
%
% locallyconst.tex
%
\chapter{Locally Constant Objects}
% \begin{definition}
% Let $\mathscr{E}$ be a topos. An object $X \in \mathscr{E}$ is called \emph{constant} if $X \cong \Delta S$ for some set $S \in \mathbf{Set}$.
% \end{definition}
% \begin{definition}
% Let $\mathscr{E}$ be a topos. An object $X \in \mathscr{E}$ is called \emph{locally constant} if there is an object $U \in \mathscr{E}$ such that $U \to 1$ is epi and a set $S \in \mathbf{Set}$ such that $X \times U \cong \Delta S \times U$ in the slice topos $\mathscr{E}/U$.
% \end{definition}
% \begin{definition}
% A geometric morphism $f : \mathscr{F} \to \mathscr{E}$ between toposes is called \emph{connected} if the inverse image part $f^*$ is fully faithful.
% \end{definition}
% Recall that if $\mathscr{E}$ is a Grothendieck topos, then it has a unique structure morphism $\gamma : \mathscr{E} \to \mathbf{Set}$ whose direct image part are the global sections $g^* = \Hom_\mathscr{E}(1,-)$ and whose inverse image part is the constant sheaf $\gamma^* = \Delta(-)$.
% \begin{definition}
% Let $\mathscr{E}$ be a Grothendieck topos. Let $\gamma : \mathscr{E} \to \mathbf{Set}$ be the structure morphism. Then $\mathscr{E}$ is called \emph{connected} if $\gamma$ is connected.
% \end{definition}
% \begin{proposition}
% Let $\mathscr{E}$ be a Grothendieck topos. Then the following are equivalent.
% \begin{enumerate}
% \item $\mathscr{E}$ is connected.
% \item If the terminal object $1 \in \mathscr{E}$ is expressed as a coproduct $1 = A + B$ for some objects $A,B \in \mathscr{E}$, then $A = 0$ or $B = 0$ (i.e. initial).
% \end{enumerate}
% \end{proposition}
\begin{proposition}
\label{prop:locally constant iff every restriction map is a bijection}
Let $\mathbf{C}$ be a category and $P \in \mathbf{Set}^{\mathbf{C}^{op}}$. If the topos $\mathbf{Set}^{\mathbf{C}^{op}}$ is connected, then the following are equivalent.
\begin{enumerate}
\item $P$ is locally constant.
\item Every restriction map of $P$ is a bijection.
\end{enumerate}
\end{proposition}
\begin{proof}
$(1) \implies (2)$. There exists a presheaf $U$ with $U \to 1$ epi, a set $S \in \mathbf{Set}$ and an isomorphism $\varphi : P \times U \to \Delta S \times U$ in $\mathbf{Set}^{\mathbf{C}^{op}}/U$. So for every morphism $f : x \to y$ in $\mathbf{C}$ we have a commutative diagram
\[ \begin{tikzcd}
Py \times Uy \arrow{r}{\left(\varphi_{Py}, \id_{Uy} \right)} \arrow[swap]{d}{Pf \times Uf} & S \times Uy \arrow{d}{\id_S \times Uf} \\
Px \times Ux \arrow[swap]{r}{\left(\varphi_{Px}, \id_{Ux} \right)} & S \times Ux
\end{tikzcd} \]
We'll show that $Pf$ is a bijection. So let $c \in Px$. Take any $b \in Uy$. Let $s = \varphi_{Px}\left( c, (Uf)(b) \right) \in S$. Then
\[ \begin{tikzcd}
\; & \left(s, b \right) \arrow[maps to]{d} \\
\left( c, (Uf)(b) \right) \arrow[maps to]{r} & \left( s, (Uf)(b) \right)
\end{tikzcd} \]
The map $\varphi_y$ is a bijection, so there exists a unique $(d,b) \in Py \times Uy$ such that $\varphi_y(d,b) = (s,b)$. Write $\left(Pf \times Uf\right) \left(d,b \right) = \left(c', (Uf)(b) \right)$. The diagram commutes, so $(c', (Uf)(b)) \mapsto (s, (Uf)(b))$. The map $\varphi_x$ is a bijection, so $c = c'$. Hence $Pf$ is a bijection. \\
$(2) \implies (1)$. Suppose $P$ is a presheaf with the property that all of its restriction maps are bijections. Take any object $x_0 \in \mathbf{C}$. Let
\[ S = \bigsqcup_{a \in Px_0} *. \]
Observe that for all objects $x \in \mathbf{C}$ we now have a bijection of sets $Px \cong S$ because the topos is assumed to be connected.
Define the presheaf $U$ as follows. For every object $x \in \mathbf{C}$, we set
\[ U(x) = \Iso(S, P(x)). \]
Given a morphism $f : x \to y$ in $\mathbf{C}$, the restriction map $Uf : Uy \to Ux$ is defined by
\[ \Iso(S, Py) \to \Iso(S, Px), \qquad g \mapsto (Pf) \circ g. \]
This is well-defined, because $Pf$ is a bijection for each morphism $f$ by assumption. The restriction maps are compatible with composition of morphisms because $P$ is a presheaf. The unique morphism $U \to 1$ is epi since the set $\Iso(S, Px)$ is non-empty for all $x \in \mathbf{C}$. Define the natural transformation $\varphi : P \times U \to \Delta S \times U$ as follows. Given an object $x \in \mathbf{C}$, the component $\varphi_x$ is defined as
\[ \varphi_x : Px \times \Iso(S, Px) \to S \times \Iso(S, Px), \qquad (a,g) \mapsto \left(g^{-1}(a), g \right). \]
The inverse of $\varphi_x$ is then given by $\varphi_x^{-1} \left( s, g \right) = (g(s), g)$. Clearly, $\varphi_x$ respects the projection onto $Ux$ for every $x \in \mathbf{C}$. It remains to check that for every $f : x \to y$ in $\mathbf{C}$ the usual diagram commutes. So let $f : x \to y$ be any morphism in $\mathbf{C}$. Let $(a,g) \in Py \times \Iso(S, Py)$. Then on the one hand,
\[ \left(\left(\id_S \times Uf\right) \circ \varphi_y\right) \left( a, g \right) = \left( \id_S \times Uf \right)\left(g^{-1}(a), g \right) = \left(g^{-1}(a), \left(Pf \right) \circ g \right). \]
On the other hand,
\[ \left( \varphi_x \circ \left( Pf \times Uf \right) \right)\left(a, g \right) = \varphi_x \left((Pf)(a), (Pf) \circ g \right) = \left( g^{-1}(a), (Pf) \circ g \right). \]
Hence $\varphi$ is a natural isomorphism.
\end{proof}
The proof shows that when $\mathbf{Set}^{\mathbf{C}^{op}}$ is not connected it is still true that $(1) \implies (2)$.
An $n$-simplex $\sigma$ of a simplicial set $S$ is called \emph{degenerate} if there exists an $m < n$ and an order-preserving surjection $f : \mathbf{n} \to \mathbf{m}$ and an $m$-simplex $\tau$ of $S$ such that $\sigma = (Sf)(\tau)$. Note that $0$-simplices (vertices) are never degenerate. An $n$-simplex $\sigma$ is called \emph{non-degenerate} if it is not degenerate.
\begin{lemma}[Eilenberg-Zilber]
\label{lem:eilenberg-zilber} Let $S \in \mathbf{SSets}$. For each $n$-simplex $\sigma \in S_n$ there exists a unique order-preserving surjection $f : \mathbf{n} \to \mathbf{m}$ and a unique non-degenerate $m$-simplex $\tau \in S_m$ such that $\sigma = (Sf)(\tau)$.
\end{lemma}
\begin{proof}
\begin{color}{red} Put a reference here... \end{color}
If $\sigma$ is already non-degenerate, take $m=n$ and $f = \id$. If $\sigma$ is degenerate, then there exists some surjection $f_1 : [n] \to [m_1]$ with $m_1 < n$ and an $m_1$-simplex $\sigma_1$ such that $\sigma = (Sf_1)(\sigma_1)$. If $\sigma_1$ is non-degenerate, we are done. Otherwise, continue in this way until we reach a $\sigma_j$ that is non-degenerate. This process ends eventually since with each iteration, $m_i < m_{i-1}$, and vertices are always non-degenerate.
So the existence of such a pair $(f,\tau)$ with $\sigma = (Sf)(\tau)$ is established. Suppose then that $(f',\tau')$ is another pair with $\sigma = (Sf')(\tau')$. Consider the pushout
\[ \begin{tikzcd}
\left[n\right] \arrow{r}{f} \arrow[swap]{d}{f'} & \left[m\right] \arrow{d}{\pi} \\
\left[m'\right] \arrow[swap]{r}{\pi'} & \left[p\right]
\end{tikzcd} \]
in $\mathbf{\Delta}$. We can apply the Yoneda functor to get a commutative diagram
\[ \begin{tikzcd}
\mathbf{\Delta}\left(-,\left[n\right]\right) \arrow{r}{\mathbf{\Delta}\left(-,f\right)} \arrow[swap]{d}{\mathbf{\Delta}\left(-,f'\right)} & \mathbf{\Delta}\left(-,\left[m\right]\right) \arrow{d}{\mathbf{\Delta}\left(-,\pi\right)} \\
\mathbf{\Delta}\left(-,\left[m'\right]\right) \arrow[swap]{r}{\mathbf{\Delta}\left(-,\pi'\right)} & \mathbf{\Delta}\left(-,\left[p\right]\right)
\end{tikzcd} \]
in $\mathbf{SSet}$. This diagram is still a pushout. While the $n$-simplex $\sigma$ is an element of the set $S_n$, it may equivalently be regarded as a functor $\sigma : \mathbf{\Delta}(-,[n]) \to S$ describing how the simplex is laid into the simplicial set. Similarly for $\tau$ and $\tau'$. Hence we have a commutative diagram
\[ \begin{tikzcd}
\mathbf{\Delta}\left(-,\left[n\right]\right) \arrow{r}{\mathbf{\Delta}\left(-,f\right)} \arrow[swap]{d}{\mathbf{\Delta}\left(-,f'\right)} & \mathbf{\Delta}\left(-,\left[m\right]\right) \arrow{d}{\mathbf{\Delta}\left(-,\pi\right)} \arrow[bend left]{ddr}{\tau} \\
\mathbf{\Delta}\left(-,\left[m'\right]\right) \arrow[swap]{r}{\mathbf{\Delta}\left(-,\pi'\right)} \arrow[swap, bend right]{rrd}{\tau'} & \mathbf{\Delta}\left(-,\left[p\right]\right) \arrow[dotted]{dr}{\exists ! \rho} \\
\; & \; & S
\end{tikzcd} \]
Since $\tau \circ \mathbf{\Delta}(-,f) = \sigma = \tau' \circ\mathbf{\Delta}(-,f')$, there exists a unique mediating $\rho$ such that $\mathbf{\Delta}(-,\pi) \circ \rho = \tau$ and $\mathbf{\Delta}(-,\pi') \circ \rho = \tau'$ as indicated by the dotted arrow. We can regard $\rho$ equivalently as a $p$-simplex in the set $S_p$, while $\pi$ and $\pi'$ are order-preserving surjections. So we see that $\tau = (S \pi)(\rho)$ and $\tau' = (S \pi')(\rho)$. But $\tau$ and $\tau'$ are both non-degenerate. So we must have $\pi = \pi' = \id$, $m=p=m'$, and hence $\tau = \tau'$.
\end{proof}
% \begin{lemma}
% Let $S \in \mathbf{SSets}$ be a simplicial set. For every $p \in |S|$ there exists a unique non-degenerate simplex $\sigma \in S$ such that $p \in |\sigma| \subseteq |S|$ and if $v_1,\ldots,v_n \in S_0$ are the vertices of $\sigma$, then for the barycentric coordinates $p = \sum_{i=1}^n \lambda_i |v_i|$, we have $\lambda_i > 0$ for all $1 \leq i \leq n$.
% \end{lemma}
% \begin{proof}
% Let $p \in |S|$. By \cref{lem:eilenberg-zilber}, we may assume $p$ is contained in some non-degenerate $n$-simplex $\sigma : \mathbf{\Delta}(-,[n]) \to S$, so that $p \in |\sigma| \subset |S|$. Write $p = \sum_{i=0}^n \lambda_i |v_i|$, where $v_i \in S_0$ is $\sigma_{[0]}(i)$, $\sum_{i=0}^n \lambda_i = 1$ and $\lambda_i \geq 0$. If $\lambda_i = 0$ for some $i$, observe then that $p \in |d_i \sigma|$, where $d_i \sigma$ is non-degenerate. Continue in this way until for each remaining indices $i$ we have $\lambda_i > 0$. During each application of a face map $d_i$, the resulting simplex stays non-degenerate.
% \end{proof}
% \begin{proposition}
% Let $X$ be a finite $T_0$-space. Then there exists an equivalence of categories
% \[ \Sh(\mathcal{O}(X)) \cong \mathbf{Set}^{X^{op}}. \]
% \end{proposition}
% \begin{proof}
% Define $\varphi : \Sh(\mathcal{O}(X)) \to \mathbf{Set}^{X^{op}}$ by sending a sheaf $F$ on $\mathcal{O}X$ to the presheaf defined as sending an object $x \in X$ to $FU_x$. If $x \leq y$ in $X$, then $U_x \subseteq U_y$, so we get restriction maps $FU_y \to FU_x$ for free. Thus $\varphi(F)$ is a presheaf on the poset $X$. Given a morphism $f : F \to F'$ in $\Sh(\mathcal{O}X)$, The morphism $\varphi(f) : \varphi(F) \to \varphi(F')$ is defined on components by $\varphi(f)_x = f_{U_x} : FU_x \to F'U_x$. Hence $\varphi$ is a functor. We'll construct a quasi-inverse $\psi : \mathbf{Set}^{X^{op}} \to \Sh(\mathcal{O}X)$ for $\varphi$. Let $P \in \mathbf{Set}^{X^{op}}$. By \cite[Comparison Lemma, paragraph 4 of the appendix]{MacLaneMoerdijk91}, it suffices to define $\psi(P)$ on a basis and checking the sheaf condition on that basis. Of course, as basis for the topological space $X$ we take we minimal opens $\{U_x : x \in X\}$. Hence $\varphi(P)$ is defined on basis-opens by $\varphi(P)(U_x) = P(x)$. The sheaf condition for basis-opens is automatic, since it is minimal. It's straightforward to see that $\varphi$ and $\psi$ are each other's quasi-inverse. The $T_0$ condition is used to show that $(\varphi \circ \psi)(P) = P$.
% \end{proof}
% \begin{corollary}
% Let $X$ be a finite $T_0$-space and let $|NX|$ be the geometric realization of the nerve of $X$ regarded as a poset. Then there is a geometric morphism
% \[ f : \Sh(\mathcal{O}|NX|) \to \mathbf{Set}^{X^{op}} \]
% whose left-adjoint $f^*$ induces an equivalence of categories
% \[ \mathbf{Set}^{X^{op}}_{lcf} \cong \Sh(\mathcal{O}|NX|)_{lcf}. \]
% \end{corollary}
% \begin{proof}
% Let $\mu : |NX| \to X$ be the McCord map sending a point $p = \sum_{i=0}^n \lambda_i |v_i| \in |NX|$ to the minimum of its support $\min \{v_i : \lambda_i > 0\}$. Then the continuous map $\mu$ induces two functors
% \[ f_* : \Sh(\mathcal{O}|NX|) \to \mathbf{Set}^{X^{op}}, \qquad F \mapsto F \circ \mu^{-1} \]
% and
% \[ f^* : \mathbf{Set}^{X^{op}} \to \Sh(\mathcal{O}|NX|), \qquad P \mapsto \left( U \mapsto \colim_{V \supseteq \mu(U)} \lim_{x \in V} P(x) \right)^{+} \]
% which are adjoint, $f^* \dashv f_*$, and $f^*$ is left-exact. Hence $f = (f^* \dashv f_*)$ is a geometric morphism. Since $\mu$ is a weak homotopy equivalence by \cite[Theorem 1.4.6]{barmak11} it is immediate that $\mathbf{Set}^{X^{op}}_{lcf} \cong \Sh(\mathcal{O}|NX|)_{lcf}$.
% \end{proof}
% \begin{example}
% We saw that every finite $G$-set, for $G$ a discrete group, is in fact locally constant finite. This is not so for monoids. For instance, consider the finite set $\Z/n\Z$ whereupon $\N$ acts as follows. For $(x,k) \in \Z/n\Z \times \N$, we set $x * k = 0$ if $k > 0$ and $x * k = x$ if $k=0$. Then $*$ is a monoid action, but the map $ - *1 : \Z/n\Z \to \Z/n\Z$ given by $x \mapsto x * 1$ is not injective. So $\Z/n\Z$ with the given action is not locally constant by \cref{prop:locally constant iff every restriction map is a bijection}.
% \end{example}
% \begin{lemma}
% \label{lem:every lcf N set is also an lcf Z set}
% Every locally constant finite $\N$-set is also a locally constant finite $\Z$-set.
% \end{lemma}
% \begin{proof}
% Let $X \in (\mathbf{B}\N)_{lcf}$. Then $1 : X \to X$, $x \mapsto x * 1$ is bijective by \cref{prop:locally constant iff every restriction map is a bijection}. Let $(-1) : X \to X$ be its inverse. Then the map $-n : X \to X$ for $n > 0$ is defined as applying $(-1)$ an $n$ number of times. Thus $X \times \Z \to X_0 \times \Z$, $(x,n) \mapsto (x * (-n), n)$ is a $\Z$-trivialization, so $X \in (\mathbf{B}\Z)_{lcf}$.
% \end{proof}
% \begin{corollary}
% Let $p : \mathbf{Set} \to \mathbf{B}\N$ be the point whose inverse image part forgets the $\N$-action. Then
% \[ \pi_1\left( \mathbf{B}\N, p \right) \cong \widehat{\Z}. \]
% \end{corollary}
% \begin{proof}
% Every locally constant finite $\Z$-set is a locally constant finite $\N$-set since $\Z$ has a canonical $\N$-action. Moreover, by \cref{lem:every lcf N set is also an lcf Z set} the converse is also true. So $\left(\mathbf{B}\N \right)_{lcf} \cong \left( \mathbf{B} \Z \right)_{lcf}$. Consequently, by \cref{thm:the fundamental group of BG is the profinite completion of G} the result follows.
% \end{proof}
% The inclusion functor $i : \mathbf{Grp} \to \mathbf{Mon}$ from the category of groups to the category of monoids has both a left and right adjoint.
% The right adjoint sends a monoid $M$ to its group of units $M^\times$. Indeed, given a homomorphism $f : i(G) \to M$ of monoids, $f$ must map units into units.
% But everything in $i(G)$ is a unit since $G$ is a group. So the codomain of $f$ can be restricted to $M^\times$.
% On the other hand, given a homomorphism of groups $f : G \to M^\times$, we clearly have a homomorphism of monoids $f : iG \to M$ simply by extending the codomain. Hence we have set up a bijection between $\Hom_{\mathbf{Mon}} \left( iG, M \right)$ and $\Hom_{\mathbf{Grp}} \left( G, M^\times \right)$.
% The left adjoint of $i$ is the free group functor $M^{grp}$ for a monoid $M$.
% It is defined as the free group on the set of symbols $\{x_m : m \in M\}$ modulo the normal subgroup generated by $x_{1}$ and elements of the form $x_m \cdot x_{m'}$ and $x_{m \cdot m'}^{-1}$ with $m,m' \in M$. For instance, $\N^{grp} = \Z$, $\left( \Z / 3 \Z, + \right)^{grp} = \left( \Z / 3\Z, + \right)$, $ \left( \Z / 3\Z, \times \right)^{grp} = 0$.
% If $G$ is already a group, then $G^{grp} \cong G$.
% Given a homomorphism $f : M \to N$ of monoids, we get a group homomorphism $f^{grp} : M^{grp} \to N^{grp}$ defined by $f(x_{m}) = x_{f(m)}$.
% The bijection between $\Hom_{\mathbf{Grp}} \left( M^{grp}, G\right)$ and $\Hom_{\mathbf{Mon}} \left(M, iG \right)$ is then clear.
% Summarized, we have three functors
% \[ \begin{tikzcd} \mathbf{Grp} \arrow{r}{i} & \mathbf{Mon} \arrow[bend left]{l}{\left(-\right)^{grp}} \arrow[swap, bend right]{l}{\left( - \right)^{\times}} \end{tikzcd}, \qquad \left( - \right)^{grp} \dashv i \dashv \left(- \right)^{\times}. \]
% \begin{corollary}
% Let $M$ be a monoid and let $p : \mathbf{Set} \to \mathbf{B}M$ be the point whose inverse image part forgets the $M$-action. Then
% \[ \pi_1\left( \mathbf{B}M, p \right) \cong \widehat{M^{grp}}. \]
% \end{corollary}
% \begin{proof}
% \end{proof}
% \begin{definition}
% \label{def:monoid ideals, prime ideals, faces}
% Let $M$ be a monoid. A subset $I \subseteq M$ is called an \emph{ideal} if $x \in I$ and $y \in M$ imply that $xy \in I$. An ideal $I$ is called \emph{prime} if $I \neq M$ and $xy \in I$ implies that $x \in I$ or $y \in I$. A submonoid $F \subseteq M$ is called a \emph{face} if $xy \in F$ implies that both $x \in F$ as well as $y \in F$.
% \end{definition}
% Observe that if $\mathfrak{p}$ is a prime ideal of $M$, then $M \setminus \mathfrak{p}$ is a face of $M$. Convserely, if $F$ is a face of $M$ then $M \setminus F$ is a prime ideal of $M$. So we have found an inclusion-reversing bijection between the faces of $M$ and the prime ideals of $M$.
% \begin{definition}
% \label{def: spectrum of a monoid}
% Let $M$ be a monoid. The \emph{spectrum} of $M$ is the set
% \[ \Spec M := \left\{ \mathfrak{p} \mid \mathfrak{p} \text{ is a prime ideal of } M \right\}. \]
% \end{definition}
% \begin{example}
% \label{ex: spectrum of N and spectrum of Z/4Z}
% The ideals of $\N$ are the empty set $\emptyset$ and subsets of the form $\N_{\geq k}$ for any $k \in \N$.
% Of these, the prime ideals are $\emptyset$ and $\N_{\geq 1}$. So $\Spec \N = \{\emptyset, \N_{\geq 1}\}$. The ideals of the monoid $\Z/4\Z$ with multiplication are $\emptyset$, $\{0\}$, $\{0,2\}$ and $\Z/4\Z$. Of these, the prime ideals are $\emptyset$, $\{0\}$ and $\{0,2\}$. So $\Spec \Z/4\Z =\left\{ \emptyset, \{0\}, \{0,2\} \right\}$.
% \end{example}
% In \cref{ex: spectrum of N and spectrum of Z/4Z}, notice that there is always a maximal prime ideal (in the sense of inclusion). This is not a coincidence. A maximal ideal is a prime ideal that is maximal with respect to inclusion.
% \begin{lemma}
% Let $M$ be a monoid. Then $\Spec M$ has a unique maximal ideal, namely the complement of the units $M^\times$.
% \end{lemma}
% \begin{proof}
% First we show that $M^\times$ is a face of $M$. Suppose that $xy \in M^\times$. If $x$ is not a unit, then it follows that $y$ is also not a unit. For if $y$ was a unit, then $(xy) \cdot y^{-1} = x \in M^\times$, since $M^\times$ is a submonoid. But if both $x$ and $y$ are not units, then their product $xy$ is also not a unit. Contradiction. Hence both $x$ and $y$ are units. So $M^\times$ is a face of $M$. Now let $F$ be an arbitrary face of $M$. Let $x \in M^\times$. The claim is that $x \in F$. We have $1 \in F$ because $F$ is a submonoid. Hence $xx^{-1} = x^{-1}x = 1 \in F$ implies that both $x$ and $x^{-1}$ are in $F$. So $M^\times \subseteq F$.
% \end{proof}
% Hence all the primes in $\Spec M$ are contained between the primes $\emptyset$ and $M \setminus M^\times$. Another easy corollary is that if $M = M^\times$, i.e. $M$ is a group, then $\Spec M = \{\emptyset\}$. Of course, $\Spec M$ can be given the structure of a topological space by declaring that sets of the form $Z(I) = \left\{ \mathfrak{p} \in \Spec M \mid I \subseteq \mathfrak{p} \right\}$, with $I$ an ideal of $M$, are closed, and then checking the usual axioms for a topology. We will not explore this direction.
% \begin{remark}
% Let $I$ be the unit interval of $\R$. Then $\mathcal{O}(I)$ is a locale and a category at the same time. So if a geometric morphism $\Sh(|N\mathcal{O}(I)|) \to \mathbf{Set}^{\mathcal{O}(I)^{op}}$ were to exist,
% \end{remark}