-
Notifications
You must be signed in to change notification settings - Fork 10
/
Copy path2_21_wirthmuller.tex
174 lines (160 loc) · 10.2 KB
/
2_21_wirthmuller.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
\begin{quote}
\textit{``What do homotopy theorists do when changing planes?''\\
``Look for a transfer map.''}
\end{quote}
In this section, we'll conver the Wirthmüller isomorphism, the equivariant analogue of the nonequivariant statement
that in the stable homotopy category $\Ho(\Spc)$, finite sums and products are equivalent:
\begin{equation}
\label{additivespec}
\bigvee_{i=1}^n X_i\stackrel\cong\longrightarrow \prod_{i=1}^n X_i.
\end{equation}
This is a backbone of the stable category: among other things, it ensures $\Ho(\Spc)$ is additive. We'll assume $G$
is a finite group; there's a statement for compact Lie groups, but it's more complicated.
Tom Dieck splitting is another important structural result about $\pi_G^*(\Sigma^\infty X)$. When $X = S^0$, this
is used to compute the equivariant stable homotopy groups of the spheres (which is, of course, not completely
known). This can be used to recover the Burnside category.\index{Burnside category}
Transfers play a key role in both of these results, somewhat implicitly in the Wirthmüller isomorphism, but very
explicitly in tom Dieck splitting.
\begin{thm}[Wirthmüller isomorphism]
\label{Wirthiso}
\index{Wirthmüller isomorphism}\index{pistar-isomorphism@$\pi_*$-isomorphism}
Let $H$ be a subgroup of $G$ and $X$ be an object of $\Sp^H$.\footnote{Here, we're implicitly using the complete
universe $U$ to define $\Spc^G$ and $\Spc^H$.\index{complete universe}} Then, there is a $\pi_*$-isomorphism
$G\wedge_H X\simeqto F_H(G,X)$.
\end{thm}
\begin{ex}
Adapt the proof of~\eqref{additivespec} for nonequivariant spectra to show that the natural map from finite sums to
finite products in $\Spc^G$ is also an isomorphism.
\end{ex}
However, if you only have~\eqref{additivespec}, you've described the category of $G$-spectra structured by a
universe with only trivial representations!
Applying \cref{Wirthiso} to $X = S^0$ computes the Spanier-Whitehead duals of orbits.
\begin{cor}
\label{wirthcor}
In the situation of \cref{Wirthiso}, $\Sigma^\infty(G/H)_+\cong F_H(G,S^0)\cong F((G/H)_+, S^0)$.
\end{cor}
That is, $\Sigma^\infty (G/H)_+$ is its own Spanier-Whitehead dual. The slogan is that orbits are self-dual in the
equivariant stable category when $G$ is finite.\footnote{When $G$ is a compact Lie group, there's a degree shift
arising from the tangent representation of $G$ on the Lie algebra of $H$. \TODO: double-check
this.}\index{Spanier-Whitehead duality}
Recall that the forgetful map $i_H^*\colon \Spc^G\to\Spc^H$ has a left and a right adjoint, respectively
$G\wedge_H\bl$ and $F_H(G,\bl)$. These spectrum-level constructions are induced by applying the space-level
functors levelwise.
\begin{ex}
Show that these definitions of $G\wedge_H\bl$ and $F_H(G,\bl)$ are consistent with the structure maps.
\end{ex}
There are three approaches to proving \cref{Wirthiso}.
\begin{enumerate}
\item One is a connectivity argument: show that on the space level, the maps get more and more highly
connected.
\item\label{transferproof} Another approach is to construct an explicit inverse, using a transfer map.
\item One can also set up a Grothendieck six-functor formalism to prove it, which sets up a general theory for
when a lax monoidal functor's left and right adjoints coincide. See~\cite{FHM, WirthRevisited, BDS} for a proof
using this approach; the first and the third papers set up general theory, and the second shows that it applies
to the Wirthmüller map. This makes interesting contact with base-change theory in algebraic
geometry.\index{six-functor formalism}
\end{enumerate}
\begin{warn}
The proof of \cref{Wirthiso} given in~\cite{LMS} is incorrect, and the fix is nontrivial. It takes
approach~\eqref{transferproof}.
\end{warn}
We'll use a connectivity argument, which will introduce some tools we'll find useful later.
\begin{proof}[Proof of \cref{Wirthiso} ($X$ connected)]
Though we assume $X$ is connected, the same proof can be adapted when $X$ is bounded-below. The theorem is true for
general $X$, but this proof may not work in that case.
Let $X\in H\Top_*$. Then, there's a map $\theta\colon G\wedge_H X\to F_H(G,X)$ defined by
\[\theta(g_1,x)(g_2) = \begin{cases}
g_2g_1x, &g_2g_1\in H\\
*, &\text{otherwise.}
\end{cases}\]
\begin{ex}
Show that $\theta$ is a $G$-map.
\end{ex}
$\theta$ induces a map $\overline\theta\colon G\wedge_H X\to F_H(G,X)$ in $\Spc^G$. We'll show that it's an
equivalence by computing the connectivity of $\theta$.
Let $K\subseteq G$, $\rho$ be the regular representation of $G$, and $m\in\N$. Then, we'll compute the connectivity
of\index{regular representation}
\[\theta_{m,K}\colon \paren{G\wedge_H \Sigma^{m\rho}X}^K\longrightarrow \paren{G_H(G, \Sigma^{m\rho} X)}^K.\]
When $G$ is finite, the sequence $(\rho, 2\rho, 3\rho,\dotsc)$ is cofinal (i.e.\ the colimits are the same) in the
filtered diagram of finite-dimensional representations of the complete universe $U$. Thus, to understand
$\colim_{V\subset U} \Omega^V\Sigma^V X$, it suffices to understand what happens when $V = m\rho$.
\begin{ex}
The reason this is true is that when $G$ is finite, the regular representation $\rho$ contains a copy of every
irreducible as a summand. Prove this by doing a character computation.\index{regular representation}
\end{ex}
We'll show the connectivity of $\theta_{m,K}$ is increasing in $m$, which means that
$\pi^K_*\overline\theta\colon\pi_*^K(G\wedge_H X)\to\pi_*^K(F_H(G,X))$ is an isomorphism.
The calculation itself will use the $K, H$ double coset decomposition of $G$ to identify the $K$-fixed points as a
sum and as a product, and we understand the connectivity of both of these from nonequivariant homotopy
theory.\index{double cosets}
Let $S\subset G$ be a set of
representatives for the double coset partition of $G$, i.e.\ under the $K\times H$-action $(k,h)\cdot g =
k^{-1}gh$. Then,
\begin{equation}
\label{GZKfixed}
\paren{G\wedge_H Z}^K = \paren{\coprod_{g\in S} (KgH)_+\wedge_H Z}^K.
\end{equation}
Either $g^{-1}Kg\subseteq H$ or it isn't.
\begin{itemize}
\item If $g^{-1}Kg\subseteq H$, then the action is only in $Z$, so $((KgH)_+\wedge_H Z)^K = Z^{g^{-1}Kg}$.
\item If $g^{-1}Kg\not\subseteq H$, then the action is free, so $((KgH)_+\wedge_H Z)^K = *$.
\end{itemize}
In particular~\eqref{GZKfixed} simplifies to
\[\paren{G\wedge_H Z}^K\cong \bigvee_{\substack{g\in S\\g^{-1}Kg\subseteq H}} Z^{g^{-1}Kg}.\]
Next we look at $(F_H(G,Z))^K$. If $S$ is a set of representatives of the double cosets, so is
$S^{-1}\coloneqq \set{g^{-1}\mid g\in S}$, so we can decompose
\begin{align*}
\paren{F_H(G,Z)}^K &\cong \paren{F_H\paren{\coprod_{g\in S} Kg^{-1}H, Z}}^K\\
&\cong \prod_{g\in S} F_H(Kg^{-1}H, Z)^K\\
&\cong \prod_{g\in S} Z^{(gKg^{-1})\cap H} \cong \prod_{g\in S} Z^{(g^{-1}Kg)\cap H}.
\end{align*}
This identification sends $f\mapsto\set{f(g)}$; the last equivalence is by using $S^{-1}$ as the set of
representatives instead of $S$.
In particular, the space-level Wirthmüller isomorphism may be written
\begin{equation}
\label{fixedwirth}
\bigvee_{\substack{g\in S\\g^{-1}Kg\subseteq H}} Z^{g^{-1}Kg}\stackrel\cong\longrightarrow \prod_{g\in S}
Z^{(g^{-1}Kg)\cap H}.
\end{equation}
This is nice-looking, but the indexing sets are slightly different. To overcome this, we'll
factor~\eqref{fixedwirth} as
\[\xymatrix{
\bigvee_{\substack{g\in S\\g^{-1}Kg\subseteq H}} Z^{g^{-1}Kg} \ar[r]^-{\vp_1}
& \prod_{\substack{g\in S\\g^{-1}Kg\subseteq H}} Z^{(g^{-1}Kg)\cap H}\ar[r]^-{\vp_2}
& \prod_{g\in S} Z^{(g^{-1}Kg)\cap H},
}\]
where $\vp_1$ is the natural map from the sum to the product and $\vp_2$ is inclusion. Then, we will estimate the
connectivity of $\vp_1$ and $\vp_2$ separately in the case when $Z = \Sigma^{m\rho}X$. Namely, we'd like to show
that they're both $(m[G:K] + O(1))$-connected.
First, what are the fixed points in $(S^{m\rho})^{g^{-1}Kg}$? This is a sphere whose dimension has been shrunk by
$K$, so smashing with it produces something $m[G:K]$-connected. Then, the usual argument about turning sums into
products says that $\vp_1$ is about $2m[G:K]$-connected.
For $\vp_2$, what's the connectivity of the missing factors in the domain? In this case, the connectivity is about
$m[G:K]$.
Thus, the connectivity of $\theta_{m,K}$ is $m([G,K] + O(1))$,\footnote{\TODO: I want to run through the
connectivity carefully and ensure I made no typos.} so the spectrum-level map $\overline\theta$ is a
$\pi_*^K$-isomorphism. More precisely, the cofiber of $\theta\colon G\wedge_H X\to F_H(G,X)$ has trivial $\pi_*^K$,
and there's a little bit to do here to check this.
\end{proof}
\begin{rem}
Let's see how this connects to the transfer. Let $V$ be a $G$-representation such that there's an embedding
$G/H\inj V$. Then, we obtain an $H$-map $G\wedge_H S^V\to S^V$ that takes points in $G\setminus H$ to the
basepoint. The adjoint of this map is the $G$-map $G\wedge_H S^V\to F_H(G,S^V)$.
On the other hand, we have a Pontrjagin-Thom map $S^V\to G\wedge_H S^V$ induced as follows: $G/H\inj V$ induces a
map $G\wedge_H D(V)\to V$, and therefore a map $S^V\to G\wedge_H D(V)/(G\wedge_H S(V))\cong G\wedge_H S^V$. The
observation is that the composition $S^V\to G\wedge_H S^V\to S^V$ is the identity (you can and should think about
this: it's possible to write down an explicit homotopy inverse). This motivates the slogan that ``the Wirthmüller
isomorphism is the inverse of the transfer map,'' which we'll say more about later.\index{Pontrjagin-Thom
construction}\index{Wirthmüller isomorphism!as an inverse of the transfer map}
\end{rem}
You can use \cref{wirthcor} to explicitly write down the transfer map: given subgroups $K\subseteq H\subseteq G$,
we have the ``right-way'' map $G/H\to G/K$, and therefore a map $\sus(G/K_+)\cong D(G/K_+)\to
D(G/H_+)\cong\sus(G/H_+)$. By the Wirthmüller isomorphism, this is the same as a map $F(G/K_+,S)\to F(G/H_+,S)$,
which is exactly the transfer map!\index{transfer map!from the Wirthmüller isomorphism}
This is a little circular; if you look carefully into the proof of \cref{Wirthiso}, the transfer map is already
there. But the point is more philosophical: the existence of transfers is the same thing as the Wirthmüller
isomorphism.
\begin{rem}
Transfer maps can be studied in more generality, e.g.\ in the context of equivariant vector bundles on homogeneous
spaces. Rothenberg wrote some stuff, but~\cite{LMS} is probably the best source.
\end{rem}