This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The equivariant cohomology ring of a cohomogeneity-one action

Jeffrey D. Carlson, Oliver Goertsches, Chen He, and Augustin-Liviu Mare
Abstract

We compute the rational Borel equivariant cohomology ring of a cohomogeneity-one action of a compact Lie group.

1. Introduction

Cohomogeneity-one Lie group actions—that is, those whose orbit space is one-dimensional—form an intensively-studied class of examples which are a next natural object of study after homogeneous actions. In lieu of a necessarily incomplete attempt to summarize the vast geometric literature surrounding these actions, we content ourselves with a gesture toward the substantial bibliography to be found in the recent classificatory work of Galaz-García and Zarei [GGZ15].

Given the prominence of these actions, it is natural to wonder what can be said of their equivariant cohomology. Due to earlier work of two of the authors [GM14, GM17], it is known the rational Borel equivariant cohomology ring is Cohen–Macaulay, and structure theorems for this ring have been worked out in special cases [GM14, Cor. 4.2, Props. 5.1, 5.10] along with topological consequences for the manifold acted on. In this work, we describe the equivariant cohomology rings of a certain broad class of cohomogeneity-one actions (delineated precisely in the discussion after Theorem˜3.3), obtaining more explicit expressions in the case of actions on manifolds.

In the most interesting case, where the space XX acted on by a compact, connected Lie group GG is a manifold and the orbit space X/GX/G is a closed interval, Theorem˜3.2, due mostly to Mostert, implies XX can be written up to GG-equivariant homeomorphism as the double mapping cylinder G/K([1,1]×G/H)G/K+\smash{G/K^{-}\cup\big{(}[-1,1]\times G/H\big{)}\cup G/K^{+}} of a pair of quotient maps G/HG/K±G/H\rightrightarrows G/K^{\pm} for some closed subgroups HK±{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}H}\leq{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}K^{\pm}} of GG. By work of Galaz-García–Searle (Theorem˜3.3), the same holds if XX is an Alexandrov space. As such, the equivariant Mayer–Vietoris sequence is applicable to the cover of XX by the preimages of two subintervals of X/GX/G.111 The non-equivariant Mayer–Vietoris sequence of the same cover has also long been used to study such spaces [GrH87, Hoel10, EU11]. As the equivariant cohomology of the restricted actions is well-known, this approach would in general recover the additive structure up to an extension problem, but in our case, surprisingly, we are able to determine the entire ring structure. This is Section˜3.2. In Section˜5, we prove more explicit formulas depending on the numbers dimK±/H(mod2)\dim K^{\pm}/H\allowbreak\mkern 8.0mu({\operator@font mod}\mkern 6.0mu2) in the case XX is a manifold MM, so that K±/HK^{\pm}/H are homology spheres. In the following result and throughout, HΓHΓ(;Q)=H(BΓ;Q){\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}H^{*}_{\Gamma}}\coloneqq H^{*}_{\Gamma}(*\mspace{2.0mu};\mathbb Q)=H^{*}(B\Gamma;\mathbb Q) will denote the Borel Γ\Gamma-equivariant cohomology ring of a point with rational coefficients. In fact, all cohomology will take rational coefficients unless explicitly specified otherwise.

{restatable}

theoremHodd Let MM be the double mapping cylinder of the quotient maps G/H\ext@arrow20202020\arrowfill@----G/K±G/H\ext@arrow{20}{20}{20}{20}{\arrowfill@\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}{\mathrel{\mathrlap{\raisebox{2.5pt}{$\rightarrow$}}\mathrlap{\raisebox{-2.5pt}{$\rightarrow$}}\hphantom{\vbox{\hbox{$\rightarrow$}\hbox{$\rightarrow$}}}}}}{}{}G/K^{\pm} for closed subgroups H<K±H<K^{\pm} of a compact Lie group GG such that K±/HK^{\pm}/H are homology spheres.
 
(a) Assume K+/HK^{+}/H is odd-dimensional and K/HK^{-}/H even-dimensional, and the bundle BHBK+BH\longrightarrow BK^{+} is orientable. Then we have an HG{H^{*}_{G}}-algebra isomorphism

HGMHKeHH[e]<HH[e],{H^{*}_{G}}M\cong H^{*}_{K^{-}}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}eH^{*}_{H}[e]<H^{*}_{H}[e],

where HK+HH[e]H^{*}_{K^{+}}\cong H^{*}_{H}[e] for a certain class eHK+1+dimK+/He\in H_{K^{+}}^{1+\dim K^{+}/H}, the product HK×HH[e]HH[e]H^{*}_{K^{-}}\times H^{*}_{H}[e]\longrightarrow H^{*}_{H}[e] is determined by the injection HK HHH^{*}_{K^{-}}\mathrel{\leavevmode\hbox to23.05pt{\vbox to9.73pt{\pgfpicture\makeatletter\hbox{\thinspace\lower-4.86601pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{} {{}{}}{}{}{}{{{}{}}}{{}}{}{}{{{}{}}}{{}}{}{{ {\pgfsys@beginscope \pgfsys@setdash{}{0.0pt}\pgfsys@roundcap\pgfsys@roundjoin{} {}{}{} {}{}{} \pgfsys@moveto{-2.07999pt}{2.39998pt}\pgfsys@curveto{-1.69998pt}{0.95998pt}{-0.85318pt}{0.28pt}{0.0pt}{0.0pt}\pgfsys@curveto{-0.85318pt}{-0.28pt}{-1.69998pt}{-0.95998pt}{-2.07999pt}{-2.39998pt}\pgfsys@stroke\pgfsys@endscope}} }{}{}{{}}{{ {\pgfsys@beginscope{{}} \pgfsys@setdash{}{0.0pt}\pgfsys@roundcap\pgfsys@roundjoin{} {}{}{} {}{}{} \pgfsys@moveto{2.07999pt}{2.39998pt}\pgfsys@curveto{1.69998pt}{0.95998pt}{0.85318pt}{0.28pt}{0.0pt}{0.0pt}\pgfsys@curveto{0.85318pt}{-0.28pt}{1.69998pt}{-0.95998pt}{2.07999pt}{-2.39998pt}\pgfsys@stroke\pgfsys@endscope}} }{}{}{{}}\pgfsys@moveto{2.07999pt}{0.0pt}\pgfsys@lineto{21.98894pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{-1.0}{0.0}{0.0}{-1.0}{2.27998pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{22.18893pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{11.19446pt}{-1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle$}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{3.0pt}{1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle\ \ \ \ $}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}}}H^{*}_{H}, and the HG{H^{*}_{G}}-module structure is induced by the inclusions K±GK^{\pm}\xhookrightarrow{\ \ \ \ }G. If K+/HK^{+}/H is a sphere, then ee is the Euler class of the sphere bundle BHBK+BH\longrightarrow BK^{+}.
 
(b) Assume that both K±/HK^{\pm}/H are odd-dimensional and the bundles BHBK±BH\longrightarrow BK^{\pm} are both orientable. Then we have an HG{H^{*}_{G}}-algebra isomorphism

HGMHH[e,e+]/(ee+),H^{*}_{G}M\,\cong\,{\raisebox{1.99997pt}{$H^{*}_{H}[e_{-},e_{+}]\,$}}\ \!\!\big{/}\!\!\ {\raisebox{-1.99997pt}{$(e_{-}e_{+})$}},

where HK±HH[e±]H^{*}_{K^{\pm}}\cong H^{*}_{H}[e_{\pm}] for classes e±HK±1+dimK±/He_{\pm}\in H_{K^{\pm}}^{1+\dim K^{\pm}/H} and the HG{H^{*}_{G}}-module structure is induced by the inclusions K±GK^{\pm}\xhookrightarrow{\ \ \ \ }G. If K±/HK^{\pm}/H is a sphere, then e±e_{\pm} is the Euler class of the sphere bundle BHBK±BH\longrightarrow BK^{\pm}.

In the event both spheres are even-dimensional, the generators of the Weyl groups W(K±)W(K^{\pm}) with respect to a shared maximal torus generate a dihedral subgroup of the automorphisms of this torus, of order 2k2k. It is this kk that figures in the following result.

{restatable}

theoremHeven Let MM the double mapping cylinder of the quotient maps G/H\ext@arrow20202020\arrowfill@----G/K±G/H\ext@arrow{20}{20}{20}{20}{\arrowfill@\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}{\mathrel{\mathrlap{\raisebox{2.5pt}{$\rightarrow$}}\mathrlap{\raisebox{-2.5pt}{$\rightarrow$}}\hphantom{\vbox{\hbox{$\rightarrow$}\hbox{$\rightarrow$}}}}}}{}{}G/K^{\pm} for closed subgroups H<K±H<K^{\pm} of a compact Lie group GG such that K±/HS2n±K^{\pm}/H\approx S^{2n_{\pm}} are even-dimensional spheres, and the bundles BHBK±BH\longrightarrow BK^{\pm} are both orientable. Then the number k(n+n+)k(n_{-}+n_{+}) is even, and we have an HG{H^{*}_{G}}-algebra isomorphism

HGM(imρ+imρ)HSk(n+n+)+1,{H^{*}_{G}}M\cong(\operatorname{im}\rho_{+}^{*}\cap\operatorname{im}\rho_{-}^{*})\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}S^{k(n_{-}+n_{+})+1},

where the injections ρ±:HK±HH\rho_{\pm}^{*}\colon H^{*}_{K^{\pm}}\longrightarrow H^{*}_{H} are induced by the inclusions HK±H\xhookrightarrow{\ \ \ \ }K^{\pm} and the HG{H^{*}_{G}}-module structure is induced by K±GK^{\pm}\xhookrightarrow{\ \ \ \ }G.

Cohomogeneity-one actions whose orbit space is S1S^{1} arise as mapping tori of right translations rnr_{n} of homogeneous spaces G/KG/K by elements nNG(K)n\in N_{G}(K), and this case admits a parallel but much more easily-proved statement we discuss in Section˜3.1.

{restatable}

theoremcircle Let MM be the mapping torus of the right translation by nNG(K)n\in N_{G}(K) on the homogeneous space G/KG/K of a compact Lie group GG. Then one has HS1H^{*}S^{1}- and (HGHS1)({H^{*}_{G}}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}S^{1})-algebra isomorphisms

HM\displaystyle H^{*}M H(G/K)rnHS1,\displaystyle\cong H^{*}(G/K)^{\langle r_{n}^{*}\rangle}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}S^{1},
HGM\displaystyle{H^{*}_{G}}M H(BK)rnHS1\displaystyle\cong H^{*}(BK)^{\langle r_{n}^{*}\rangle}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}S^{1}

respectively, where the HS1H^{*}S^{1}-module structure is given by pullback from M/GM/G in both cases and the HG{H^{*}_{G}}-algebra structure is induced by the inclusion KGK\xhookrightarrow{\ \ \ \ }G.

The unexpectedly great utility of the Mayer–Vietoris sequence in our situation results from an additional structural feature of the sequence that seems not to be frequently noted, namely the fact that the connecting map preserves a module structure over the cohomology ring of the whole space. This result is proved in Section 2.

Acknowledgments. The authors would like to thank the referee for careful proofreading, for suggesting a reference, and for making an important correction to their statement of Theorem˜3.2. The first author would like to thank Omar Antolín Camarena for helpful conversations and the National Center for Theoretical Sciences (Taiwan) for its hospitality during a phase of this work.

2. The Mayer–Vietoris sequence

Let GG be a topological group, XX a GG-space, and A,BXA,B\subseteq X two GG-invariant subsets whose interiors cover XX. The rings HGA×HGB{H^{*}_{G}}A\times{H^{*}_{G}}B and HG(AB){H^{*}_{G}}(A\cap B) in the Mayer–Vietoris sequence inherit an HGX{H^{*}_{G}}X-module structure by restriction. It is clear the restriction maps between these rings are HGX{H^{*}_{G}}X-module homomorphisms and we claim the connecting map is as well. It is enough to prove the analogous result for singular cohomology, as the Mayer–Vietoris sequence in equivariant cohomology is just the Mayer–Vietoris sequence of the associated cover (AG,BG)(A_{G},B_{G}) of the homotopy orbit space XGX_{G}.

Proposition 2.1.

Let XX be a topological space, AA and BB a pair of subspaces whose interiors cover XX, and kk any commutative ring with unity. Then the connecting map δMV:H(AB;k)H(X;k)[1]{\delta_{\mathrm{MV}}}\colon H^{*}(A\cap B;k)\longrightarrow H^{*}(X;k)[1] in the Mayer–Vietoris sequence of (X;A,B)(X;A,B) is a homomorphism of H(X;k)H^{*}(X;k)-modules.

Proof.

The covering hypothesis makes the inclusion C(A)+C(B)C(X)C_{*}(A)+C_{*}(B)\xhookrightarrow{\ \ \ \ }C_{*}(X) of singular chain complexes a quasi-isomorphism, whose dual C(X;k) HomZ(C(A)+C(B),k)LC^{*}\mspace{-2.0mu}(X;k)\mathrel{\leavevmode\hbox to23.05pt{\vbox to9.73pt{\pgfpicture\makeatletter\hbox{\thinspace\lower-4.86601pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{} {{}{}}{}{}{}{{{}{}}}{{}}{}{}{{{}{}}}{{}}{}{}{}{}{}{{}}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@lineto{21.98894pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{19.70895pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{22.18893pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{11.19446pt}{-1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle$}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{3.0pt}{1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle\ \ \ \ $}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}}}\operatorname{Hom}Z\!\big{(}C_{*}(A)+C_{*}(B),k\big{)}\eqqcolon{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}L} is then again a quasi-isomorphism. The Mayer–Vietoris sequence in cohomology is the long exact sequence arising from the short exact sequence of cochain complexes

0L𝑖C(A;k)×C(B;k)𝑗C(AB;k)0,0\to L\overset{{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}i}}{\longrightarrow}C^{*}(A;k)\times C^{*}(B;k)\overset{{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}j}}{\longrightarrow}C^{*}(A\cap B;k)\to 0,

where i(c)=(c|A,c|B)i(c)=(c|_{A},c|_{B}) and j(cA,cB)=cA|ABcB|ABj(c_{A},c_{B})=c_{A}|_{A\cap B}-c_{B}|_{A\cap B}. To define the connecting map, given a homogeneous cocycle zZq(AB;k)z\in Z^{q}(A\cap B;k), one selects cochains cACq(A;k)c_{A}\in C^{q}(A;k) and cBCq(B;k)c_{B}\in C^{q}(B;k) such that j(cA,cB)=zj(c_{A},c_{B})=z, and then δMV[z]Hq+1(X;k)\delta_{\mathrm{MV}}[z]\in H^{q+1}(X;k) is the class represented by the unique element zz^{\prime} of Zq+1(L)Z^{q+1}(L) such that i(z)=(δcA,δcB)i(z^{\prime})=(\delta c_{A},\delta c_{B}). Now given a class in Hp(X;k)H^{p}(X;k), the restrictions of some representative xx to A,BA,B, and ABA\cap B factor through the representative x|LZp(L)x|_{L}\in Z^{p}(L). Observe that x|ABz=j(x|AcA,x|BcB)x|_{A\cap B}\mspace{-1.0mu}\smile\mspace{-1.0mu}z=j(x|_{A}\mspace{-1.0mu}\smile\mspace{-1.0mu}c_{A},x|_{B}\mspace{-1.0mu}\smile\mspace{-1.0mu}c_{B}). But since δx=0\delta x=0, we have (δ(x|AcA),δ(x|BcB))=(x|AδcA,x|BδcB)=i(x|Lz)\big{(}\delta(x|_{A}\mspace{-1.0mu}\smile\mspace{-1.0mu}c_{A}),\delta(x|_{B}\mspace{-1.0mu}\smile\mspace{-1.0mu}c_{B})\big{)}=({x|_{A}\mspace{-1.0mu}\smile\mspace{-1.0mu}\delta c_{A}},{x|_{B}\mspace{-1.0mu}\smile\mspace{-1.0mu}\delta c_{B}})=i(x|_{L}\mspace{-1.0mu}\smile\mspace{-1.0mu}z^{\prime}), so δMV([x][z])=[x]δMV[z]\delta_{\rm{MV}}\big{(}[x]\mspace{-1.0mu}\smile\mspace{-1.0mu}[z]\big{)}=[x]\mspace{-1.0mu}\smile\mspace{-1.0mu}\delta_{\rm{MV}}[z] as claimed. ∎

Remark 2.2.

This feature turns out not to be specific to Borel equivariant cohomology, but applies to multiplicative cohomology theories in general, and the first author proves this fact and an extension of Section˜3.2 to other cohomology theories in an accompanying paper [Car18], along with analogues of the other main results in equivariant K-theory.

3. Equivariant cohomology rings

To deploy Proposition 2.1 as promised, we need the structure theorem for cohomogeneity-one actions on manifolds.

Figure 3.1: Schematic of a double mapping cylinder
Refer to caption
Theorem 3.2 ([Most57a, Thm. 4][Most57b][GGZ15, Thm. A]).

Let GG be a compact, connected Lie group acting continuously with cohomogeneity one on a connected topological manifold MM without boundary. Then MM is, up to GG-equivariant homeomorphism, as follows.

  • If M/G(1,1)M/G\approx(-1,1), there is a closed subgroup KGK\leq G such that M(1,1)×G/KM\approx(-1,1)\times G/K.

  • If M/GS1M/G\approx S^{1}, there are a closed subgroup KGK\leq G and an element nNG(K)n\in N_{G}(K) such that MM is the mapping torus of the right translation of G/KG/K by nn. (The equivariant homeomorphism type of the resulting space depends only on the class of nn in π0NG(K)\pi_{0}N_{G}(K).)

  • If M/G[0,1)M/G\approx[0,1), there are closed subgroups H<KGH<K\leq G such that K/HK/H is either a sphere SnS^{n} or the Poincaré homology sphere P3P^{3} and MM is the open mapping cylinder G/K𝜋([0,1)×G/H)\smash{G/K\underset{\pi}{\cup}\big{(}[0,1)\times G/H\big{)}} of the projection π:G/HG/K\pi\colon G/H\longrightarrow G/K.

  • If M/G[1,1]M/G\approx[-1,1], there are closed subgroups H<K±GH<K^{\pm}\leq G such that each of K±/HK^{\pm}/H is either a sphere SnS^{n} or the Poincaré homology sphere P3P^{3} and MM is the double mapping cylinder

    G/Kπ([1,1]×G/H)π+G/K{G/K^{-}\underset{\pi^{-}}{\cup}\big{(}[-1,1]\times G/H\big{)}\underset{\pi^{+}}{\cup}G/K}

    of the projections π±:G/H\ext@arrow20202020\arrowfill@----G/K±\pi^{\pm}\colon G/H\ext@arrow{20}{20}{20}{20}{\arrowfill@\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}{\mathrel{\mathrlap{\raisebox{2.5pt}{$\rightarrow$}}\mathrlap{\raisebox{-2.5pt}{$\rightarrow$}}\hphantom{\vbox{\hbox{$\rightarrow$}\hbox{$\rightarrow$}}}}}}{}{}G/K^{\pm}.

Conversely, these constructions yield only cohomogeneity-one GG-actions on manifolds. In the cases where M/GM/G has boundary, MM admits a GG-invariant smooth structure if and only if no isotropy quotient K/HK/H or K±/HK^{\pm}/H is P3P^{3}.

Before proceeding, we note the noncompact cases are trivial for our purposes, since in these cases MM equivariantly deformation retracts to the cohomogeneity-zero case G/KG/K. There is a similar classification of cohomogeneity-one actions on closed Alexandrov spaces.

Theorem 3.3 ([GGS11, Thm. A]).

Let GG be a compact Lie group acting effectively and isometrically with cohomogeneity one on a closed (i.e., compact and without boundary) Alexandrov space XX. Then XX is, up to GG-equivariant homeomorphism, as follows.

  • If X/GS1X/G\approx S^{1}, there are a closed subgroup KGK\leq G and an element nNG(K)n\in N_{G}(K) such that XX is the mapping torus of the right translation of G/KG/K by nn (and hence, by Theorem˜3.2, a smooth manifold).

  • If X/G[1,1]X/G\approx[-1,1], there are closed subgroups H<K±GH<K^{\pm}\leq G such that K±/HK^{\pm}/H are positively-curved homogeneous spaces and XX is the double mapping cylinder of the projections G/H\ext@arrow20202020\arrowfill@----G/K±G/H\ext@arrow{20}{20}{20}{20}{\arrowfill@\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}{\mathrel{\mathrlap{\raisebox{2.5pt}{$\rightarrow$}}\mathrlap{\raisebox{-2.5pt}{$\rightarrow$}}\hphantom{\vbox{\hbox{$\rightarrow$}\hbox{$\rightarrow$}}}}}}{}{}G/K^{\pm}.

Conversely, these constructions yield only cohomogeneity-one GG-actions on Alexandrov spaces.

Thus our Section˜3.2 will apply more generally than just to manifolds. Because of these two classification results, it is reasonable to focus our attention in the rest of the paper on cohomogeneity-one actions of the following types:

  • the mapping torus of a right translation on a homogeneous space G/KG/K or

  • the double mapping cylinder of a span of projections G/KG/HG/K+G/K^{-}\leftarrow G/H\to G/K^{+}.

3.1.   Mapping tori

By Theorem˜3.2, if the orbit space of a cohomogeneity-one action is a circle, the space in question can be assumed to be a manifold MM, the mapping torus of the right-translation rn{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}r_{n}} of some element nNG(K){\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}n}\in N_{G}(K) on G/KG/K, and hence actually a smooth manifold MM.

Lemma 3.4.

Let YY be a topological space, φ\varphi a self-homeomorphism of YY such that some finite power φ\varphi^{\ell} is homotopic to idY\operatorname{id}_{Y}, and XX the mapping torus of φ\varphi. Then

HXH(Y)φHS1.H^{*}X\cong{H^{*}(Y)^{\langle\varphi^{*}\rangle}}\underset{}{\mathbin{\raisebox{0.85pt}{$\displaystyle\otimes$}}}{H^{*}S^{1}}.
Proof.

Note that XX admits an \ell-sheeted cyclic covering by the mapping torus of φ\varphi^{\ell}, which is homeomorphic to the mapping torus Y×S1Y\times S^{1} of the identity. The covering action is conjugate to a Z/\mathbb Z/\ell-action on Y×S1Y\times S^{1} under which 1+Z1+\ell\mathbb Z acts, up to homotopy, as (y,θ)(φ(y),θ+2π)(y,\theta)\longmapsto\big{(}\varphi(y),\theta+\frac{2\pi}{\ell}\big{)}, which, rotating the circle component, is in turn homotopic to (y,θ)(φ(y),θ)(y,\theta)\longmapsto\big{(}\varphi(y),\theta\big{)}. A standard lemma on the transfer map [Hat02, Prop. 3G.1] then gives

HXH(Y×S1)Z/H(Y)φHS1.H^{*}X\cong H^{*}(Y\times S^{1})^{\mathbb Z/\ell}\cong H^{*}(Y)^{\langle\varphi^{*}\rangle}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}S^{1}.\qed

We continue to denote by rn:G/KG/Kr_{n}\mspace{-1.5mu}\colon G/K\longrightarrow G/K right multiplication by an element nNG(K)n\in N_{G}(K). As NG(K)N_{G}(K) is compact, it has finitely many path-components, so some power nn^{\ell} lies in the path-component of the identity and hence the corresponding power rnr_{n}^{\ell} is homotopic to the identity. Applying Lemma˜3.4 to MM gives the first displayed isomorphism and applying it to MGM_{G} gives the second.

3.2.   Double mapping cylinders

Let GG be a compact Lie group and HK±GH\leq K^{\pm}\leq G any closed subgroups. Then the double mapping cylinder XX of π±:G/H\ext@arrow20202020\arrowfill@----G/K±\pi^{\pm}\colon G/H\ext@arrow{20}{20}{20}{20}{\arrowfill@\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}{\mathrel{\mathrlap{\raisebox{2.5pt}{$\rightarrow$}}\mathrlap{\raisebox{-2.5pt}{$\rightarrow$}}\hphantom{\vbox{\hbox{$\rightarrow$}\hbox{$\rightarrow$}}}}}}{}{}G/K^{\pm} admits the obvious invariant open cover by the respective inverse images UU^{-} and U+U^{+} of the subintervals [1,ε)[-1,\varepsilon) and (ε,1](-\varepsilon,1], for some small ε>0\varepsilon>0, of X/G[1,1]X/G\approx[-1,1] depicted in Figure˜3.1. Their intersection W=UU+W=U^{-}\cap U^{+} equivariantly deformation retracts to G/HG/H and U±U^{\pm} to G/K±G/K^{\pm} in such a way that the inclusions WU±W\xhookrightarrow{\ \ \ \ }U^{\pm} correspond to the projections π±\pi^{\pm}. Now we can apply Proposition˜2.1.

{restatable}

theoremmainH Let XX be the double mapping cylinder of the projections π±:G/H\ext@arrow20202020\arrowfill@----G/K±\pi^{\pm}\colon G/H\ext@arrow{20}{20}{20}{20}{\arrowfill@\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}\mathrel{\mathrlap{\raisebox{2.5pt}{$\relbar$}}\raisebox{-2.5pt}{$\relbar$}}{\mathrel{\mathrlap{\raisebox{2.5pt}{$\rightarrow$}}\mathrlap{\raisebox{-2.5pt}{$\rightarrow$}}\hphantom{\vbox{\hbox{$\rightarrow$}\hbox{$\rightarrow$}}}}}}{}{}G/K^{\pm}. Then one has a graded HG{H^{*}_{G}}-algebra and a graded HHH^{*}_{H}-module isomorphism, respectively:

HGevenXHK×HHHK+,HGoddX(HHimρ+imρ+)[1],H_{G}^{\mathrm{even}}X\,\cong\,\smash{{H^{*}_{K^{-}}}\underset{H^{*}_{H}}{\times}{H^{*}_{K^{+}}}},\qquad H_{G}^{\mathrm{odd}}X\,\cong\mspace{1.0mu}\Big{(}\frac{H^{*}_{H}}{\operatorname{im}\rho^{*}_{-}+\operatorname{im}\rho^{*}_{+}}\Big{)}[1],

where ρ±:HK±HH{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\rho^{*}_{\pm}}\colon H^{*}_{K^{\pm}}\longrightarrow H^{*}_{H} are induced by the inclusions HK±H\xhookrightarrow{\ \ \ \ }K^{\pm} and HK×HHHK+H^{*}_{K^{-}}\times_{H^{*}_{H}}H^{*}_{K^{+}} denotes the fiber product.222 That is, HK×HHHK+<HK×HK+H^{*}_{K^{-}}\times_{H^{*}_{H}}H^{*}_{K^{+}}<H^{*}_{K^{-}}\times H^{*}_{K^{+}} is the subring of pairs (x,x+)(x_{-},x_{+}) such that ρ(x)=ρ+(x+)\rho^{*}_{-}(x_{-})=\rho^{*}_{+}(x_{+}). The multiplication of odd-degree elements is zero, and the product HGevenX×HGoddXHGoddXH_{G}^{\mathrm{even}}X\times H_{G}^{\mathrm{odd}}X\longrightarrow H_{G}^{\mathrm{odd}}X descends from the multiplication of HHH^{*}_{H} in that

(x,x+)q¯=ρ(x)q¯.(x_{-},x_{+})\cdot\bar{q}=\overline{\rho^{*}_{-}(x_{-})\cdot q}.

for (x,x+)HK×HHHK+(x_{-},x_{+})\in\smash{{H^{*}_{K^{-}}}\underset{H^{*}_{H}}{\times}{H^{*}_{K^{+}}}} and q¯HGoddX\bar{q}\in H_{G}^{\mathrm{odd}}X the image of qHHq\in H^{*}_{H}.

Proof.

For any ΓG\Gamma\leq G we have homeomorphisms (G/Γ)G=EG×GG/ΓEG/Γ=BΓ(G/\Gamma)_{G}=EG\times_{G}G/\Gamma\approx EG/\Gamma=B\Gamma. As HK±\smash{H^{*}_{K^{\pm}}} and HH\smash{H^{*}_{H}} are concentrated in even degree, the Mayer–Vietoris sequence of the cover just discussed reduces to a four-term exact sequence

0HGevenXiHK×HK+HHHGoddXi0,0\to H_{G}^{\mathrm{even}}X\overset{{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}i^{*}}}{\longrightarrow}H^{*}_{K^{-}}\times H^{*}_{K^{+}}\longrightarrow H^{*}_{H}\longrightarrow H_{G}^{\mathrm{odd}}X\overset{i^{*}}{\to}0,

so that HGevenXH_{G}^{\mathrm{even}}X is the kernel and HGoddXH_{G}^{\mathrm{odd}}X the cokernel of the middle map (x,x+)ρ+(x+)ρ(x)(x_{-},x_{+})\longmapsto\rho_{+}^{*}(x_{+})-\rho_{-}^{*}(x_{-}). The multiplicative structure on HGevenH_{G}^{\mathrm{even}} follows from the fact i:HGXHK×HK+i^{*}\colon{H^{*}_{G}}X\longrightarrow H^{*}_{K^{-}}\times H^{*}_{K^{+}} is a ring map, the description of the product HGevenX×HGoddXHGoddXH_{G}^{\mathrm{even}}X\times H_{G}^{\mathrm{odd}}X\longrightarrow H_{G}^{\mathrm{odd}}X follows from Proposition˜2.1, and the fact the product HGoddX×HGoddXHGevenXH_{G}^{\mathrm{odd}}X\times H_{G}^{\mathrm{odd}}X\longrightarrow H_{G}^{\mathrm{even}}X is zero follows from the observation ii^{*} is injective on HGevenXH_{G}^{\mathrm{even}}X and yet i(xy)=i(x)i(y)=0i^{*}(xy)=i^{*}(x)i^{*}(y)=0 for any elements x,yHGoddXx,y\in H_{G}^{\mathrm{odd}}X. ∎

Example 3.5.

Let G=O(n)G=\mathrm{O}(n) with K=K±=O(3)K=K^{\pm}=\mathrm{O}(3) and H=O(2)H=\mathrm{O}(2) block-diagonal. We have HKQ[p1]HH{H^{*}_{K}}\cong\mathbb Q[p_{1}]\cong H^{*}_{H}, where p1p_{1} is the first Pontrjagin class of the tautological bundle over the infinite Grassmannian Gr(3,R)=BO(3)\operatorname{Gr}(3,\mathbb R^{\infty})=B\mathrm{O}(3), so HGXQ[p1]{H^{*}_{G}}X\cong\mathbb Q[p_{1}].

Example 3.6.

In the situation where G=K±G=K^{\pm}, the resulting double mapping cylinder is just the unreduced suspension S(G/H)S(G/H). One has

HGevenS(G/H)=HG,\displaystyle\phantom{\mbox{and}}\qquad H_{G}^{\mathrm{even}}S(G/H)={H^{*}_{G}}, HGoddS(G/H)=HH/im(HGHH)[1].\displaystyle\qquad H_{G}^{\mathrm{odd}}S(G/H)={\raisebox{1.99997pt}{$H^{*}_{H}\,$}}\ \!\!\big{/}\!\!\ {\raisebox{-1.99997pt}{$\operatorname{im}(\mspace{-1.0mu}{H^{*}_{G}}\to H^{*}_{H})$}}[1].

4. Maps of classifying spaces

In the event the cohomogeneity-one double mapping cylinder is a manifold MM, we will presently see that more precise descriptions of HGMH^{*}_{G}M can be obtained depending on the dimensions of the isotropy spheres K±/HK^{\pm}/H, subject to an orientation hypothesis on the left action of K±K^{\pm} in case these groups are disconnected, and these descriptions depend crucially on the structure of HHH^{*}_{H} as a module over HK±H^{*}_{K^{\pm}}.

Lemma 4.1.

Let H<KH<K be compact Lie groups such that K/HK/H is a homology sphere. The K/HK/H-bundle ρ:BHBK\smash{\rho\colon BH\longrightarrow BK} is orientable if and only if left multiplication on K/HK/H by any element of KK induces the identity in cohomology.

Proof.

Recall orientability of a fiber bundle is defined as triviality of the action of the fundamental group of the base on the cohomology of the fiber. To determine the action-up-to-homotopy of a class of π1(BK,e0K)\pi_{1}(BK,e_{0}K) on the fiber, lift a representative loop η\eta to a path η~\widetilde{\eta} in EKEK starting at e0e_{0} and ending at some e0k1e_{0}k_{1}. Then for any e0kHe_{0}kH in the fiber ρ1(e0K)\rho^{-1}(e_{0}K), the path η~kH\widetilde{\eta}kH lifts η\eta to BHBH, starting at e0kHe_{0}kH and ending at e0k1kHe_{0}k_{1}kH, which we may define to be ηe0kH\eta\cdot e_{0}kH. Under the identification ρ1(e0K)K/H\rho^{-1}(e_{0}K)\approx K/H given by e0kHkHe_{0}kH\leftrightarrow kH, this is just the action of π0K\pi_{0}K induced by the defining homogeneous action of KK on K/HK/H. The generator 1H0(K/H)1\in H^{0}(K/H) is invariant trivially, so the action is trivial in cohomology if and only if the fundamental class [K/H][K/H] is fixed by the π0K\pi_{0}K-action. ∎

Remark 4.2.

Particularly, this rules out the case K/HS0K/H\approx S^{0} going forward.

Proposition 4.3 (Cf. Goertsches–Mare [GM14, Prop. 3.1][GM17, Prop. 4.2]).

Let H<KH<K be compact Lie groups such that K/HK/H is a homology sphere of odd dimension nn and BHBKBH\longrightarrow BK is an orientable K/HK/H-bundle. Then ρ:HKHH\rho^{*}\colon{H^{*}_{K}}\longrightarrow H^{*}_{H} is a surjection and can be written

HH[e]e0HH,H^{*}_{H}[e]\xrightarrow{e\mapsto 0}H^{*}_{H},

where eHHn+1{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}e}\in H^{n+1}_{H} is the generalized Euler class of the bundle K/HBHBKK/H\to BH\to BK.

Proof.

Let K0K_{0} be the identity component of KK and write H1=HK0{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}H_{1}}=H\cap K_{0}, so that K0/H1K/HK_{0}/H_{1}\longrightarrow K/H is a homeomorphism. We claim HH1H^{*}_{H_{1}} is a polynomial ring. This follows if K/HK/H is a sphere of dimension at least 22 since then K0/H1K/HK_{0}/H_{1}\approx K/H is both simply-connected and covered by K0/H0K_{0}/H_{0}, forcing H0=H1H_{0}=H_{1}. If K/HK/H is homeomorphic to S1S^{1} or the Poincaré homology sphere P3P^{3}, then we still know HH1(HH0)H1/H0H^{*}_{H_{1}}\cong(H^{*}_{H_{0}})^{H_{1}/H_{0}}, but the action of H1/H0H_{1}/H_{0} on HH0H^{*}_{H_{0}} is trivial because it is known [GM14, Pf., Prop. 3.1][GM17, Pf., Prop. 4.2] that H0H_{0} is normal in K0K_{0} in these cases, so the action of H1/H0H_{1}/H_{0} is the restriction of an action of K0/H0K_{0}/H_{0}, which is homotopically trivial since K0/H0K_{0}/H_{0} is path-connected.333 To make this account self-contained, the proof of normality is thus. The transitive action of K0K_{0} on K0/H1K_{0}/H_{1} induces a map λ:K0HomeoK0/H1{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\lambda}\colon K_{0}\longrightarrow\operatorname{Homeo}K_{0}/H_{1} whose image, which acts effectively by definition, can only be S1S^{1} itself if K0/H1S1K_{0}/H_{1}\approx S^{1} and SO(3)\mathrm{SO}(3) if K0/H1P3K_{0}/H_{1}\approx P^{3} [Bre61, Thm. 1.1]. As kerλ\ker\lambda stabilizes all points, it is in particular contained in H1H_{1}. The stabilizer of the coset 1H1K0/H11H_{1}\in K_{0}/H_{1} under the effective action of imλ\operatorname{im}\lambda is λ(H1)H1/kerλ\lambda(H_{1})\cong H_{1}/\ker\lambda, which must be finite since imλ\operatorname{im}\lambda is of rank one, so kerλ\ker\lambda is of finite index in H1H_{1}; particularly, its identity component must be H0H_{0}. Since kerλ\ker\lambda is normal in K0K_{0} by definition, so also must be H0H_{0}.

We now consider the map of K/HK/H-bundles

EK/H1\textstyle{EK/H_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EK/H\textstyle{EK/H\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EK/K0\textstyle{EK/K_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EK/K.\textstyle{EK/K.} (4.1)

The left map is equivariant with respect to the right HH-action, inducing an effective right action of πH/H1π0K{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\pi}\coloneqq H/H_{1}\cong\pi_{0}K such that the right map is the quotient, and so we may identify [Hat02, Prop. 3G.1] the map HKHH{H^{*}_{K}}\longrightarrow H^{*}_{H} with the map of invariants (HK0)π(HH1)π(H^{*}_{K_{0}})^{\pi}\longrightarrow(H^{*}_{H_{1}})^{\pi}. Now let us consider a portion of the induced map of generalized Gysin sequences [MiT00, SS3.7]:

HK\textstyle{{H^{*}_{K}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e\scriptstyle{\mspace{-1.0mu}\smile\mspace{-1.0mu}e}HK+n+1\textstyle{H^{*+n+1}_{K}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ\scriptstyle{\rho^{*}}HH+n+1\textstyle{H^{*+n+1}_{H}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}HK0\textstyle{H^{*}_{K_{0}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e\scriptstyle{\mspace{-1.0mu}\smile\mspace{-1.0mu}e}HK0+n+1\textstyle{H^{*+n+1}_{K_{0}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}HH1+n+1.\textstyle{H^{*+n+1}_{H_{1}}.}

The commutativity of the diagram implies the identification HK(HK0)π{H^{*}_{K}}\cong(H^{*}_{K_{0}})^{\pi} takes the one generalized Euler class to the other, or in other words that the class in the lower sequence is π\pi-invariant. Since HK0H^{*}_{K_{0}} is a polynomial ring, multiplication by ee is injective, so the horizontal maps before and after are zero, so this is actually an inclusion of short exact sequences. As the image of multiplication by ee is precisely the principal ideal (e)(e), we obtain an π\pi-equivariant isomorphism HH1HK0/(e)H^{*}_{H_{1}}\cong H^{*}_{K_{0}}/(e). As HK0HH1H^{*}_{K_{0}}\longrightarrow H^{*}_{H_{1}} is a π\pi-equivariant surjection between graded polynomial rings over Q\mathbb Q on n+1n+1 and nn generators, respectively, whose kernel is generated by the π\pi-invariant element ee, Lemma˜4.6 applies to yield a π\pi-equivariant isomorphism HH1[e]HK0H^{*}_{H_{1}}[e]\overset{\sim}{\longrightarrow}H^{*}_{K_{0}}. This restricts to an isomorphism of π\pi-invariants HH[e]HK\smash{H^{*}_{H}[e]\overset{\sim}{\longrightarrow}{H^{*}_{K}}}. ∎

Remark 4.4.

When K/HK/H is a sphere, the generalized Euler class featuring in Proposition˜4.3 is well known to be the standard Euler class of a sphere bundle [MiT00, Thm. 5.17, pp. 145–6]. If K/HP3K/H\approx P^{3}, on the other hand, from the fact that the action KHomeoP3K\longrightarrow\operatorname{Homeo}P^{3} factors through an SO(3)\mathrm{SO}(3) subgroup [Bre61, Thm. 1.1], one can associate to BHBKBH\longrightarrow BK the principal SO(3)\mathrm{SO}(3)-bundle ξ:EK×KSO(3)BK\xi\colon EK\times_{K}\mathrm{SO}(3)\longrightarrow BK and show ee is 6060 times the first Pontrjagin class p1(ξ)p_{1}(\xi).

Remark 4.5.

The persistent orientability hypothesis is necessary; we will see in Remark˜5.3 that Section˜5.2 fails without this hypothesis. For now, consider the case of K=O(2)K=\mathrm{O}(2) and HH a subgroup of order 22 generated by an element hh of determinant 1-1. Then for zSO(2)z\in\mathrm{SO}(2) we have hzH=z1Hh\cdot zH=z^{-1}H. We have HK0=HSO(2)=Q[c1]H^{*}_{K_{0}}=H^{*}_{\mathrm{SO}(2)}=\mathbb Q[c_{1}] and c1=ec_{1}=e in the notation of the proof since HH0=H{1}=Q\smash{H^{*}_{H_{0}}=H^{*}_{\{1\}}=\mathbb Q}. The proof would go through if we had e=c1HKe=c_{1}\in{H^{*}_{K}}, but we do not; in fact HO(2)Q[c12]H^{*}_{\mathrm{O}(2)}\cong\mathbb Q[c_{1}^{2}], where c12c_{1}^{2} is represented by the first Pontrjagin class p1p_{1} of the tautological 22-plane bundle over the infinite Grassmannian G(2,R)=BO(2)G(2,\mathbb R^{\infty})=B\mathrm{O}(2) [Hat09, Thm. 3.16(a)]. Thus, although it is incidentally true in this case that HKHH[c12]{H^{*}_{K}}\cong H^{*}_{H}[c_{1}^{2}], the proof of Proposition˜4.3 cannot possibly go through.

Lemma 4.6.

Suppose AA is a graded polynomial ring in finitely many variables over Q\mathbb Q equipped with an action of a finite group π\pi fixing the field of constants Q\mathbb Q, and xx is a π\pi-invariant homogeneous element of AA such that B=A/(x)B=A/(x) is again a polynomial ring. Then there is a π\pi-invariant graded Q\mathbb Q-subalgebra A<AA^{\prime}<A such that A=A[x]A=A^{\prime}[x] and AAφB\smash{A^{\prime}\hookrightarrow A\mathrel{\leavevmode\hbox to11.93pt{\vbox to14.11pt{\pgfpicture\makeatletter\hbox{\thinspace\lower-4.86601pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{} {{}{}}{}{}{}{{{}{}}}{{}}{}{}{{{}{}}}{{}}{}{}{}{}{}{{}}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@lineto{10.85977pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{8.57979pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{11.05977pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{5.62988pt}{-1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle$}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{3.00002pt}{2.89409pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\varphi}$}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}}}B} is a ring isomorphism.

Proof.

We consider AA and BB as modules over the group ring Qπ\mathbb Q\pi. It is easy to see that any Qπ\mathbb Q\pi-module CC complementary to (x)(x) will be taken bijectively and Qπ\mathbb Q\pi-linearly to A/(x)A/(x) by φ\varphi, so we just need to show such a complement can be chosen to be a ring and generators of this ring can be chosen such that together with xx they form a set of Q\mathbb Q-algebra generators for AA. We write QB=B1/B1B1{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}Q}B=B^{\geq 1}/B^{\geq 1}\cdot B^{\geq 1} for the graded Q\mathbb Q-module of indecomposables, in this case a free module. As QBQB is a finite-dimensional π\pi-representation and the order of π\pi is invertible over Q\mathbb Q, we may break QBQB into irreducible representations of π\pi, which are cyclic Qπ\mathbb Q\pi-modules, each generated by an element b¯j\bar{b}_{j}. We may lift each of these to a homogeneous element bjB1b_{j}\in B^{\geq 1}. By construction, the union of the π\pi-orbits of the bjb_{j} forms a set of Q\mathbb Q-algebra generators for BB. Now let ajCa_{j}\in C be a φ\varphi-preimage of bjb_{j}. Then the Qπ\mathbb Q\pi-algebra AA^{\prime} generated by the aja_{j} is taken bijectively onto BB by φ\varphi, so it is a polynomial subalgebra and in fact another Qπ\mathbb Q\pi-linear complement to (x)(x). It is clear from the isomorphism ABA^{\prime}\xrightarrow{\sim}B that each element of AA can be represented uniquely as a polynomial in xx over AA^{\prime}. ∎

The case K/HK/H is even-dimensional is simpler.

Proposition 4.7 ([Bor53, Thm. 26.1(a)][Sam41, p. 1121]).

Let HKH\leq K be compact Lie groups of equal rank. Then HKHH{H^{*}_{K}}\longrightarrow H^{*}_{H} is injective. This applies particularly if K/HK/H is an even-dimensional sphere. In this case, if the bundle ρ:BHBK\rho\colon BH\longrightarrow BK is orientable and n=dimK/H2n=\dim K/H\geq 2, then HHH^{*}_{H} is a free HK{H^{*}_{K}}-module of rank two on 11 and a lift eHHn{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}e}\in H_{H}^{n} of the fundamental class of K/HK/H under the surjection H(BH)H(K/H)H^{*}(BH)\longrightarrow H^{*}(K/H).

Samelson showed the ranks are equal if K/HK/H is an even-dimensional sphere and the injectivity statement is due to Borel.

Proof.

The covering π0KBK0BK\pi_{0}K\to BK_{0}\to BK coming from the action of π0K=K/K0\pi_{0}K=K/K_{0} on BK0=EK/K0BK_{0}=EK/K_{0}, induces an isomorphism HK(HK0)π0K{H^{*}_{K}}\cong(H^{*}_{K_{0}})^{\pi_{0}K} by a standard transfer lemma [Hat02, Prop. 3G.1]. As HK0H^{*}_{K_{0}} is a polynomial ring by Borel’s theorem, the Serre spectral sequence of K/HBHBKK/H\to BH\to BK is concentrated in even degree and so collapses. By Lemma˜4.1, the coefficients are simple, so HHHKH(K/H)HK{1,e}H^{*}_{H}\cong{H^{*}_{K}}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}(K/H)\cong{H^{*}_{K}}\{1,e\} as an HK{H^{*}_{K}}-module for some ee represented by 1[K/H]1\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}{[K/H]} in the associated graded algebra. ∎

Remark 4.8.

Note that the basis {1,e}\{1,e\} is preserved under the map

HHHKH(K/H) HK0H(K0/H0)HH0H^{*}_{H}\cong{H^{*}_{K}}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}(K/H)\mathrel{\leavevmode\hbox to23.05pt{\vbox to9.73pt{\pgfpicture\makeatletter\hbox{\thinspace\lower-4.86601pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{} {{}{}}{}{}{}{{{}{}}}{{}}{}{}{{{}{}}}{{}}{}{}{}{}{{}}{}{}{}{{}}\pgfsys@moveto{2.07999pt}{0.0pt}\pgfsys@lineto{21.98894pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{-1.0}{0.0}{0.0}{-1.0}{2.27998pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{22.18893pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{11.19446pt}{-1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle$}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{3.0pt}{1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle\ \ \ \ $}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}}}H^{*}_{K_{0}}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}(K_{0}/H_{0})\cong H^{*}_{H_{0}}

induced by the map of K/HK/H-bundles (BH0BK0)(BHBK)(BH_{0}\to BK_{0})\longrightarrow(BH\to BK), so eHH0e\in H^{*}_{H_{0}} may be chosen π0H\pi_{0}H-invariant.

Remark 4.9.

The lift ee in Proposition˜4.7 can be chosen to be the pullback under BHBSO(n)BH\longrightarrow B\mathrm{SO}(n) of the universal Euler class eSO(n)HSO(n)ne_{\mathrm{SO}(n)}\in H^{n}_{\mathrm{SO}(n)}. To see this, note that by the classification of transitive Lie group actions on spheres [Bes87, Ex. 7.13], the action KHomeoK/HK\longrightarrow\operatorname{Homeo}K/H must factor through a subgroup isomorphic to SO(n+1)\mathrm{SO}(n+1), sending HH into an SO(n)\mathrm{SO}(n) subgroup and hence inducing a map of SnS^{n}-bundles from BHBKBH\to BK to BSO(n)BSO(n+1)B\mathrm{SO}(n)\to B\mathrm{SO}(n+1). Both Serre spectral sequences collapse at E2E_{2}, and the E2E_{2} map HSO(n)HSnHKH(K/H)H^{*}_{\mathrm{SO}(n)}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}S^{n}\longrightarrow{H^{*}_{K}}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}(K/H) sends 1[Sn]1[K/H]1\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}{[S^{n}]}\longmapsto 1\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}{[K/H]}. But ee represents 1[K/H]1\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}{[K/H]}, and since HSO(n)HSO(n+1){1,eSO(n)}H^{*}_{\mathrm{SO}(n)}\cong H^{*}_{\mathrm{SO}(n+1)}\{1,e_{\mathrm{SO}(n)}\} as an HSO(n+1)H^{*}_{\mathrm{SO}(n+1)}-module, eSO(n)e_{\mathrm{SO}(n)} represents 1[Sn]1\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}{[S^{n}]}.

5. Double mapping cylinders which are manifolds

In this section we are in the situation of Section˜3.2 and additionally the isotropy quotients K±/HK^{\pm}/H are homology spheres.

5.1.   The case when one of K±/HK^{\pm}/H is odd-dimensional

If K+/HK^{+}/H is odd-dimensional, then HK+HHH^{*}_{K^{+}}\longrightarrow H^{*}_{H} is surjective, so HGoddMH_{G}^{\mathrm{odd}}M vanishes by Section˜3.2 and HGM=HGevenM{H^{*}_{G}}M=H_{G}^{\mathrm{even}}M is easily described.

\Hodd

*

Proof.

(a)  Since the map HK+HH[e]HHH^{*}_{K^{+}}\xrightarrow{\sim}H^{*}_{H}[e]\to H^{*}_{H} is reduction modulo (e)(e) by Proposition˜4.3 and HKHHH^{*}_{K^{-}}\longrightarrow H^{*}_{H} is an injection by Proposition˜4.7, the fiber product is the subring of HK×HH[e]H^{*}_{K^{-}}\times H^{*}_{H}[e] consisting of the direct summands {(x,x)HK×HK}\big{\{}(x,x)\in H^{*}_{K^{-}}\times H^{*}_{K^{-}}\big{\}} and {0}×eHH[e]\{0\}\times e\cdot H^{*}_{H}[e]. We may identify the former with HK<HH<HH[e]H^{*}_{K^{-}}<H^{*}_{H}<H^{*}_{H}[e] and the latter with eHH[e]HH[e]eH^{*}_{H}[e]\lhd H^{*}_{H}[e], and the two interact multiplicatively via the rule xef(x,x)(0,ef)=(0,efx)efxx\cdot ef\leftrightarrow(x,x)\cdot(0,ef)=(0,efx)\leftrightarrow efx.

(b) Using Proposition˜4.3 to make identifications HK±HH[e±]H^{*}_{K^{\pm}}\cong H^{*}_{H}[e_{\pm}] such that HK±HHH^{*}_{K^{\pm}}\longrightarrow H^{*}_{H} is reduction modulo (e±)(e_{\pm}), we see the fiber product is the subring of HH[e]×HH[e+]H^{*}_{H}[e_{-}]\times H^{*}_{H}[e_{+}] comprising the three direct summands

{(x,x)HH×HH},eHH[e]×{0},{0}×e+HH[e+],\big{\{}(x,x)\in H^{*}_{H}\times H^{*}_{H}\big{\}},\qquad\qquad e_{-}H^{*}_{H}[e_{-}]\times\{0\},\qquad\qquad\{0\}\times e_{+}H^{*}_{H}[e_{+}],

on which multiplication is determined by the three rules

(x,x)(ef,0)=(efx,0),(x,x)(0,e+f+)=(0,e+f+x),(ef,0)(0,e+f+)=(0,0),(x,x)\cdot(e_{-}f_{-},0)=(e_{-}f_{-}x,0),\qquad(x,x)\cdot(0,e_{+}f_{+})=(0,e_{+}f_{+}x),\qquad(e_{-}f_{-},0)\cdot(0,e_{+}f_{+})=(0,0),

so the map to HH[e,e+]/(ee+)H^{*}_{H}[e_{-},e_{+}]/(e_{-}e_{+}) sending (x+ef,x+e+f+)x+ef+e+f+(x+e_{-}f_{-},x+e_{+}f_{+})\longmapsto x+e_{-}f_{-}+e_{+}f_{+} is a ring isomorphism. ∎

Remark 5.1.

The second and fourth author have shown [GM14, GM17] that for any cohomogeneity-one action of a compact, connected Lie group GG on a compact, connected topological manifold MM, the equivariant cohomology HGMH^{*}_{G}M is a Cohen–Macaulay module over HGH^{*}_{G}. In the special case when all the hypotheses of Section˜1 are fulfilled, this result can be recovered easily from that theorem. Concretely, in case (a), the equivariant cohomology is a direct sum of two Cohen–Macaulay modules over HGH^{*}_{G} and in case (b) it is an algebra over HGH^{*}_{G} finitely generated as an HG{H^{*}_{G}}-module and Cohen–Macaulay as a ring.

Remark 5.2.

Special cases of the actions investigated in this section are also considered by Choi and Kuroki [Kur11, CK11].

5.2.   The case when both of K±/HK^{\pm}/H are even-dimensional

In this subsection we assume that KK^{-}, K+K^{+}, and HH have all three the same rank, or equivalently that MM is a manifold with K±/HK^{\pm}/H even-dimensional. We start with the special case where K=K+K^{-}=K^{+}.

{restatable}

propositionspherefiber Assume KK+=K{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}K}\coloneqq K^{+}=K^{-}, that K/H=S2nK/H=S^{2n} is an even-dimensional sphere, and that the bundle BHBKBH\longrightarrow BK is orientable. Then we have an H(BG;Z)H^{*}(BG;\mathbb Z)-algebra isomorphism

HG(M;Z)H(BK;Z)H(S2n+1;Z),{H^{*}_{G}}(M;\mathbb Z)\cong H^{*}(BK;\mathbb Z)\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}(S^{2n+1};\mathbb Z),

where the H(BG;Z)H^{*}(BG;\mathbb Z)-algebra structure is induced by the inclusion KGK\xhookrightarrow{\ \ \ \ }G.

Proof.

In this case MGM_{G} is the homotopy pushout of BKBHBKBK\leftarrow BH\to BK, which we may write as the quotient of [1,1]×BH[-1,1]\times BH by the relation collapsing the ends {±1}×BH\{\pm 1\}\times BH to one copy of BKBK each. There is an obvious map

ξ:MG\displaystyle{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\xi}\colon M_{G}  BK,\displaystyle\mathrel{\leavevmode\hbox to23.05pt{\vbox to9.73pt{\pgfpicture\makeatletter\hbox{\thinspace\lower-4.86601pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{} {{}{}}{}{}{}{{{}{}}}{{}}{}{}{{{}{}}}{{}}{}{}{}{}{}{{}}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@lineto{21.98894pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{19.70895pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{22.18893pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{11.19446pt}{-1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle$}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{3.0pt}{1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle\ \ \ \ $}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}}}BK,
[t,eH]\displaystyle[t,eH] eK,t(1,1),\displaystyle\longmapsto eK,\quad t\in(-1,1),
[±1,eK]\displaystyle[\pm 1,eK] eK.\displaystyle\longmapsto eK.

The fiber of this map over eKBKeK\in BK is the unreduced suspension S(K/H)S2n+1S(K/H)\approx S^{2n+1}, so ξ\xi is a sphere bundle. We claim the Serre spectral sequence of ξ\xi has simple coefficients and collapses at E2E_{2}. Indeed, given eKBKeK\in BK and a loop η\eta representing a class in π1(BK,eK)\pi_{1}(BK,eK), if η~\widetilde{\eta} is a lift to EKEK starting at ee and ending at ekek, a lift of η\eta starting at (±1,eK)(\pm 1,eK) (respectively, (t,eH)(t,eH)) ends at (±1,ekK)=(±1,eK)(\pm 1,ekK)=(\pm 1,eK) (resp., (t,ekH)(t,ekH)). This action fixes the poles of the fiber S(K/H)S(K/H) and acts as left multiplication by kk on each latitude K/HK/H, fixing the orientation of the latitude by Lemma˜4.1, so the action preserves the orientation of the fiber S2n+1S^{2n+1} of ξ\xi as a whole, and thus, again by Lemma˜4.1, the bundle ξ\xi is orientable. As there are only two nonzero rows, to see the spectral sequence collapses at E2H(BK;H(S2n+1;Z))=H(BK;Z)H(S2n+1;Z)E_{2}\cong H^{*}\big{(}BK;H^{*}(S^{2n+1};\mathbb Z)\big{)}=H^{*}(BK;\mathbb Z)\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}(S^{2n+1};\mathbb Z), it is enough to show d2n+2[S2n+1]=0H2n+2(BK)d_{2n+2}[S^{2n+1}]=0\in H^{2n+2}(BK). But this differential must be zero because ξ\xi admits the section eK[1,eK]eK\longmapsto[-1,eK], showing ξ:H(BK;Z)HG(M;Z)\xi^{*}\colon H^{*}(BK;\mathbb Z)\longrightarrow{H^{*}_{G}}(M;\mathbb Z) is injective. The collapse shows HG(M;Z)H(S2n+1;Z){H^{*}_{G}}(M;\mathbb Z)\longrightarrow H^{*}(S^{2n+1};\mathbb Z) is surjective and HG(M;Z)E=H(BK;Z)H(S2n+1;Z){H^{*}_{G}}(M;\mathbb Z)\cong E_{\infty}=H^{*}(BK;\mathbb Z)\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}(S^{2n+1};\mathbb Z) as a module over H(BK;Z)H^{*}(BK;\mathbb Z). Thus 1HG0(M;Z)1\in H_{G}^{0}(M;\mathbb Z) and any preimage zHG2n+1(M;Z)z\in H^{2n+1}_{G}(M;\mathbb Z) of the fundamental class [S2n+1]H2n+1(S2n+1;Z)[S^{2n+1}]\in H^{2n+1}(S^{2n+1};\mathbb Z) generate HG(M;Z){H^{*}_{G}}(M;\mathbb Z) as an H(BK;Z)H^{*}(BK;\mathbb Z)-module.

By definition, this zz commutes with all elements of imξH(BK;Z)\operatorname{im}\xi^{*}\cong H^{*}(BK;\mathbb Z), and we claim it also squares to zero. Indeed, from Section˜3.2, it is in the image of the Mayer–Vietoris connecting map from H(BH;Z)H^{*}(BH;\mathbb Z), which factors as H(BH;Z)ΣH~+1(SBH;Z)H~+1(MG;Z)\smash{H^{*}(BH;\mathbb Z)\overset{\Sigma}{\to}\widetilde{H}^{*+1}(SBH;\mathbb Z)\to\widetilde{H}^{*+1}(M_{G};\mathbb Z)}, where Σ\Sigma is the suspension isomorphism and MGSBHM_{G}\to SBH is the map that collapses each end BKBK of the double mapping cylinder to a point [May99, SS19.1–3, pp. 146–7]. Since the cup product on H~(SBH)\widetilde{H}^{*}(SBH) is identically zero and z2z^{2} is the image of a square in H~(SBH)\widetilde{H}^{*}(SBH), we see that indeed z2=0z^{2}=0. It follows HG(M;Z)ΛH(BK;Z)[z]H(BK;Z)H(S2n+1;Z){H^{*}_{G}}(M;\mathbb Z)\cong\Lambda_{H^{*}(BK;\mathbb Z)}[z]\cong H^{*}(BK;\mathbb Z)\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}(S^{2n+1};\mathbb Z) as an H(BK;Z)H^{*}(BK;\mathbb Z)-algebra. ∎

Remark 5.3.

The orientability hypothesis in the proposition above is essential, as can be seen from the action of SO(3){\rm S}{\rm O}(3) on RP3#RP3\mathbb RP^{3}\,\#\,\mathbb RP^{3} described by Mostert [Most57a, Thm. 7]. The isotropies are given by K=K±=S(O(2)×O(1))O(2)K=K^{\pm}=\mathrm{S}\big{(}\mathrm{O}(2)\times\mathrm{O}(1)\mspace{-1.0mu}\big{)}\cong\mathrm{O}(2) and H=SO(2)×{1}H=\mathrm{SO}(2)\times\{1\}, so K/HS0K/H\approx S^{0} and the bundle is not orientable by Remark˜4.2. If the conclusion of Proposition 5.2 held in this instance, we would have (with Q\mathbb Q coefficients as usual)

HSO(3)(RP3#RP3)HO(2)HS1.H^{*}_{\mathrm{SO}(3)}(\mathbb R\mathrm{P}^{3}\,\#\,\mathbb R\mathrm{P}^{3})\cong H^{*}_{\mathrm{O}(2)}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}S^{1}. (5.1)

But the SO(3)\mathrm{SO}(3)-action at hand is equivariantly formal [GM14, Cor. 1.3], so HSO(3)(RP3#RP3)\smash{H^{*}_{\mathrm{SO}(3)}(\mathbb R\mathrm{P}^{3}\,\#\,\mathbb R\mathrm{P}^{3})} is isomorphic as an HSO(3)H^{*}_{\mathrm{SO}(3)}-module to H(RP3#RP3)HSO(3)H^{*}(\mathbb R\mathrm{P}^{3}\,\#\,\mathbb R\mathrm{P}^{3})\otimes H^{*}_{\mathrm{SO}(3)}, which unlike HO(2)HS1H^{*}_{\mathrm{O}(2)}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}S^{1} is zero in dimension 11.

Example 5.4.

Let us now use Section˜5.2 to compute the equivariant cohomology of the action arising from the inclusion diagram (G,K,K+,H)=(Sp(2),Sp(1)2,Sp(1)2,Sp(1)×U(1))(G,K^{-},K^{+},H)=\big{(}\mathrm{Sp}(2),\mathrm{Sp}(1)^{2},\mathrm{Sp}(1)^{2},\mathrm{Sp}(1)\times\mathrm{U}(1)\big{)}. For a nice treatment of this manifold we refer to Püttmann [Pütt09, Sect. 4.3]. From the Mayer–Vietoris sequence, one sees the manifold MM has the same integral cohomology as the direct product S3×S4S^{3}\times S^{4}. By Proposition 5.2,

HSp(2)(M;Z)H(BSp(1)2;Z)H(S3;Z).H^{*}_{\mathrm{Sp}(2)}(M;\mathbb Z)\cong H^{*}\big{(}B\mathrm{Sp}(1)^{2};\mathbb Z\big{)}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}(S^{3};\mathbb Z).

Equivariant formality of the Sp(2)\mathrm{Sp}(2)-action on MM was already known over Q\mathbb Q [GM14, Cor. 1.3], but the inclusion Sp(1)2Sp(2)\mathrm{Sp}(1)^{2}\xhookrightarrow{\ \ \ \ }\mathrm{Sp}(2) induces an H(BSp(2);Z)H^{*}\big{(}B\mathrm{Sp}(2);\mathbb Z\big{)}-module isomorphism H(BSp(1)2;Z)H(BSp(2);Z)H(S4;Z)H^{*}\big{(}B\mathrm{Sp}(1)^{2};\mathbb Z\big{)}\cong H^{*}\big{(}B\mathrm{Sp}(2);\mathbb Z\big{)}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}(S^{4};\mathbb Z), so the action is actually equivariantly formal over Z\mathbb Z.

We will now generalize the proposition above to the case when KK^{-} and K+K^{+} are not necessarily equal. Assume that K±/H=S2n±K^{\pm}/H=S^{2n_{\pm}} for n±1{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}n_{\pm}}\geq 1. Let SH{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}S}\leq H be a maximal torus and Ξ{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\Xi} the subgroup of AutS\operatorname{Aut}S generated by the Weyl groups W(K±)W(K^{\pm}) and W(H)W(H). The Weyl groups, and hence Ξ\Xi, are all contained in the image of the conjugation map NG(S)AutSN_{G}(S)\to\operatorname{Aut}S sending g(sgsg1)g\longmapsto(s\mapsto gsg^{-1}). Since image of this map is compact and AutSZdimS\operatorname{Aut}S\cong\mathbb Z^{\dim S} is discrete, Ξ\Xi is finite. Because K±/HK^{\pm}/H are even-dimensional spheres, by Proposition˜4.7, HH=(HS)WHH^{*}_{H}=({H^{*}_{S}})^{W_{H}} is of rank two over HK±=(HS)WK±H^{*}_{K^{\pm}}=({H^{*}_{S}})^{W_{K^{\pm}}}, so W(H)W(H) is an index-two subgroup of each of W(K±)W(K^{\pm}) and hence normal. It follows W(H)W(H) is also normal in Ξ\Xi, and we will be particularly interested in the quotient group Ξ/W(H)\Xi/W(H). Note that the involutions w±W(K±)/W(H)Z/2{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}w_{\pm}}\in W(K^{\pm})/W(H)\cong\mathbb Z/2 also generate Ξ/W(H)\Xi/W(H), so the latter must be a dihedral group. Because we will be considering functions on the Lie algebra 𝔰{\mathfrak{s}}, in the rest of this section, cohomology will take real or complex coefficients.

Definition 5.5.

For a compact, disconnected Lie group Γ\Gamma with maximal torus TT, we continue to define its Weyl group W(Γ)W(\Gamma) as NΓ(T)/ZΓ(T)N_{\Gamma}(T)/Z_{\Gamma}(T).444 We note that this is not the only definition used in the literature, and does not agree with the common definition invoking a Cartan subgroup as per Segal [Seg68, Def. 1.1][BtD85, SSIV.4].

It is easy to see every component of Γ\Gamma contains some element of NΓ(T)N_{\Gamma}(T), and such elements in the same component differ by an element of NΓ0(T)N_{\Gamma_{0}}(T), so there is a well defined action of π0Γ\pi_{0}\Gamma on the fixed point set R[𝔱]W(Γ0)\mathbb R[{\mathfrak{t}}]^{W(\Gamma_{0})} and one has

H(BΓ;R)H(BΓ0;R)π0Γ(R[𝔱]W(Γ0))=π0ΓR[𝔱]W(Γ)\smash{H^{*}(B\Gamma;\mathbb R)\cong H^{*}(B\Gamma_{0};\mathbb R)^{\pi_{0}\Gamma}\cong\big{(}\mathbb R[{\mathfrak{t}}]^{W(\Gamma_{0})}\big{)}\mspace{-2.0mu}{}^{\pi_{0}\Gamma}=\mathbb R[{\mathfrak{t}}]^{W(\Gamma)}}

as in the connected case.

Lemma 5.6.

The action of the dihedral group Ξ/W(H)\Xi/W(H) on H(BH;R)=R[𝔰]W(H)H^{*}(BH;\mathbb R)=\mathbb R[{\mathfrak{s}}]^{W(H)} is effective.

Proof.

We show any element cg:sgsg1{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}c_{g}}\mspace{-2.0mu}\colon s\mapsto gsg^{-1} of Ξ\Xi that fixes R[𝔰]W(H)\mathbb R[{\mathfrak{s}}]^{W(H)} pointwise is already in W(H)W(H). If Υ{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\Ups} is the subgroup of Ξ\Xi generated by W(H)W(H) and cgc_{g}, then clearly

R[𝔰]Υ=R[𝔰]W(H).\mathbb R[{\mathfrak{s}}]^{\Ups}=\mathbb R[{\mathfrak{s}}]^{W(H)}. (5.2)

By Molien’s theorem [Kan01, SS17-3], given any action of a finite group Γ\Gamma on a real vector space VV, the Poincaré series in tt of R[V]Γ\mathbb R[V]^{\Gamma} is a polynomial in the variable (1t)1(1-t)^{-1} with leading coefficient 1/|Γ|1/|\Gamma|. In our case, this shows |Υ|=|W(H)||\Ups|=|W(H)|, so cgc_{g} is in W(H)W(H) as claimed. ∎

Remark 5.7.

Assume that GG is connected and MM smooth and equipped with a GG-invariant Riemannian metric and a complete geodesic γ{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\gamma} in MM meeting each orbit orthogonally. The Weyl group of γ\gamma is defined [PT87, SS4][AA93, SS5] to be W(γ)NG(γ)/ZG(γ){\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}W(\gamma)}\coloneqq N_{G}(\gamma)/Z_{G}(\gamma), where NG(γ)<G{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}N_{G}(\gamma)}<G is the setwise stabilizer of γ\gamma and ZG(γ){\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}Z_{G}(\gamma)} the pointwise. The account of Alekseevsky–Alekseevsky [AA93, SS4,5] shows GG acts transitively on the set of such γ\gamma, so we may assume γ\gamma passes through (0,1H)(1,1)×G/HM(0,1H)\in(-1,1)\times G/H\subsetneq M; it can also be shown HH is the common stabilizer of all points of γ\gamma in (1,1)×G/H(-1,1)\times G/H, so W(γ)=NG(γ)/HW(\gamma)=N_{G}(\gamma)/H, and it follows γ\gamma passes through (±1,1K±){±1}×G/K±(\pm 1,1K^{\pm})\in\{\pm 1\}\times G/K^{\pm} as well. Further, there are unique involutions σ±NK±(H)/H{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\sigma_{\pm}}\in N_{K^{\pm}}(H)/H acting antipodally on the spheres K±/HK^{\pm}/H and generating W(γ)W(\gamma) as a dihedral subgroup of NG(H)/HN_{G}(H)/H.

One might hope from this account that the groups W(γ)W(\gamma) and Ξ/W(H)\Xi/W(H) are isomorphic, but they are typically not, as we will see in Example˜5.15. There is at least a homomorphism from the one to the other, which is an isomorphism if rkG=rkS\operatorname{rk}G=\operatorname{rk}S. To construct it, observe first that since all maximal tori in HH are conjugate, for any gNG(H)g\in N_{G}(H) there exists an hgH{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}h_{g}}\in H such that ghgNG(S)gh_{g}\in N_{G}(S), and this specification uniquely determines the left coset hgNH(S)h_{g}N_{H}(S), defining a homomorphism gghgNH(S)g\longmapsto gh_{g}N_{H}(S) from NG(H)(NG(S)NG(H))/NH(S)N_{G}(H)\longrightarrow\big{(}N_{G}(S)\cap N_{G}(H)\big{)}/N_{H}(S); to see multiplicativity, note that for given gH,gHNG(H)/HgH,g^{\prime}H\in N_{G}(H)/H, we may make the choice hgg=(g)1hgghgh_{gg^{\prime}}=(g^{\prime})^{-1}h_{g}g^{\prime}h_{g^{\prime}}. It is not hard to see the kernel is HH, so there is an induced monomorphism

ψ:NG(H)H\displaystyle{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\psi}\colon\frac{N_{G}(H)}{H}  NG(S)NG(H)NH(S),\displaystyle\mathrel{\leavevmode\hbox to23.05pt{\vbox to9.73pt{\pgfpicture\makeatletter\hbox{\thinspace\lower-4.86601pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{} {{}{}}{}{}{}{{{}{}}}{{}}{}{}{{{}{}}}{{}}{}{}{}{}{{}}{}{}{}{{}}\pgfsys@moveto{2.07999pt}{0.0pt}\pgfsys@lineto{21.98894pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{-1.0}{0.0}{0.0}{-1.0}{2.27998pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{22.18893pt}{0.0pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{11.19446pt}{-1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle$}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}}\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{3.0pt}{1.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{\hbox{$\scriptstyle\ \ \ \ $}}} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}}}\frac{N_{G}(S)\cap N_{G}(H)}{N_{H}(S)},
gH\displaystyle gH ghgNH(S).\displaystyle\longmapsto gh_{g}N_{H}(S).

Following with the map c:(NG(S)NG(H))/NH(S)NAutS(W(H))/W(H){\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}c}\colon\big{(}N_{G}(S)\cap N_{G}(H)\big{)}/N_{H}(S)\longrightarrow N_{\operatorname{Aut}S}\big{(}W(H)\big{)}/W(H) sending gg to conjugation of SS by gg, we obtain a map NG(H)/HNAutS(W(H))/W(H)N_{G}(H)/H\longrightarrow N_{\operatorname{Aut}S}\big{(}W(H)\big{)}/W(H). When restricted to NK±(H)/HN_{K^{\pm}}(H)/H, this cψc\circ\psi takes values in W(K±)/W(H)W(K^{\pm})/W(H), so there is a restricted map ψ¯:W(γ)Ξ/W(H){\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\overline{\psi}}\colon W(\gamma)\longrightarrow\Xi/W(H) as claimed. When rkG=rkS\operatorname{rk}G=\operatorname{rk}S, the map cc is injective, so ψ¯(σ±)=w±\overline{\psi}(\sigma_{\pm})=w_{\pm} and ψ¯\overline{\psi} is an isomorphism.

We henceforth write D2k{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}D_{2k}} for the dihedral group Ξ/W(H)\Xi/W(H), where k{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}k} is the order of w+ww_{+}w_{-}. Our remaining goal is the following.

\Heven

*

This will follow from an analysis of the action of D2k{D_{2k}} on HHH^{*}_{H}.

Lemma 5.8.

Let EE be a complex representation of D2k=w,w+w2,w+2,(w+w)kD_{2k}=\langle w_{-},w_{+}\mid w_{-}^{2},w_{+}^{2},(w_{+}w_{-})^{k}\rangle. Set rw+w{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}r}\coloneqq w_{+}w_{-} and sw+{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}s}\coloneqq w_{+} so that sr=wsr=w_{-}. Write ζ{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\zeta} for the root of unity e2πi/ke^{2\pi i/k} and E{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}E_{\ell}} for the ζ\zeta^{\ell}-eigenspace of rr.

  • The transformations ww_{-} and w+w_{+} agree on E0E_{0} and are opposite on Ek/2E_{k/2} if kk is even.

  • Both ww_{-} and w+w_{+} exchange each EE_{\ell} with EE_{-\ell}.

Proof.

Multiplying the relation id=rr1\operatorname{id}=rr^{-1} on the left by w+w_{+} yields the key equation

w+=wr1=ζwonE.\phantom{E_{\ell}}\quad w_{+}=w_{-}r^{-1}=\zeta^{-\ell}w_{-}\quad\mbox{on}\ E_{\ell}. (5.3)
  • The =0\ell=0 and =k/2\ell=k/2 cases of (5.3) show w=w+w_{-}=w_{+} on E0E_{0} and w=w+w_{-}=-w_{+} on Ek/2E_{k/2}.

  • Since rw=w+w2=w+rw_{-}=w_{+}w_{-}^{2}=w_{+}, (5.3) gives rwv=w+v=ζwvrw_{-}v=w_{+}v=\zeta^{-\ell}w_{-}v for vEv\in E_{\ell}. On the other hand, multiplying (5.3) on the left by rr yields rw+v=ζrwv=ζw+v.rw_{+}v=\zeta^{-\ell}rw_{-}v=\zeta^{-\ell}w_{+}v.

From now on, we specialize Lemma˜5.8 to the case E=HH=H(BH;C)E={\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}H^{*}_{H}}=H^{*}(BH;\mathbb C). We write also EmEHHm{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}E_{\ell}^{m}}\coloneqq E_{\ell}\cap H_{H}^{m}.

Lemma 5.9.

There exist p±HH2n±\smash{{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}p_{\pm}}\in H^{2n_{\pm}}_{H}} such that w±p±=p±w_{\pm}p_{\pm}=-p_{\pm} and HH=HK±p±HK±H^{*}_{H}=H^{*}_{K^{\pm}}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}p_{\pm}H^{*}_{K^{\pm}}.

Proof.

Note that since HK±=(HH)w±H^{*}_{K^{\pm}}=(H^{*}_{H})^{\langle w_{\pm}\rangle}, the 11-eigenspace of w±w_{\pm} is HK±H^{*}_{K^{\pm}}. Recall that Proposition˜4.7 gives a HK±H^{*}_{K^{\pm}}-basis {1,e±}\{1,e_{\pm}\} for HHH^{*}_{H}; now p±(e±w±e±)/2p_{\pm}\coloneqq(e_{\pm}-w_{\pm}e_{\pm})/2 is a 1-1-eigenvector for w±w_{\pm} and {1,p±}\{1,p_{\pm}\} is another HK±H^{*}_{K^{\pm}}-module basis for HHH^{*}_{H}. ∎

Let us now consider the rr-eigenspace decompositions p±=q±p_{\pm}=\sum{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}q^{\pm}_{\ell}} where q±Eq^{\pm}_{\ell}\in E_{\ell} for Z/k\ell\in\mathbb Z/k. Since the w±w_{\pm} interchange EE_{\ell} with EE_{-\ell} by Lemma˜5.8, one finds q±=w±q±q^{\pm}_{-\ell}=-w_{\pm}q^{\pm}_{\ell}, and specifically that w±q0±=q0±w_{\pm}q_{0}^{\pm}=-q_{0}^{\pm}, and if kk is even, w±qk/2±=qk/2±w_{\pm}q^{\pm}_{k/2}=-q^{\pm}_{k/2}. All told, the decomposition is

p±=q0±+0<<k/2(q±w±q±)+qk/2±,p_{\pm}=q^{\pm}_{0}+\sum_{0\mspace{1.0mu}<\mspace{1.0mu}\ell\mspace{1.0mu}<\mspace{1.0mu}k/2}(q^{\pm}_{\ell}-w_{\pm}q^{\pm}_{\ell})+q^{\pm}_{k/2}, (5.4)

where the last term is taken to be zero if kk is odd.

Lemma 5.10.

In (5.4) only one term is non-zero. Explicitly, the elements p±p_{\pm} each lie either in E0E_{0}, in Ek/2E_{k/2}, or in EEE_{\ell}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{-\ell} for 0<<k/20<\ell<k/2.

Proof.

Each of terms q0±q^{\pm}_{0}, qk/2±q^{\pm}_{k/2}, and q±w±q±q^{\pm}_{\ell}-w_{\pm}q^{\pm}_{\ell} in (5.4) lies in the 1-1-eigenspace of w±w_{\pm} on HH2n±H_{H}^{2n_{\pm}}, but by Lemma˜5.9 this is the one-dimensional Cp±\mathbb Cp_{\pm}, which meets at most one of E0E_{0}, Ek/2E_{k/2}, and EEE_{\ell}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{-\ell} nontrivially. ∎

Lemma 5.11.

Exactly one of the following cases obtains:

  1. (i)

    k=1k=1 and n+=nnn_{+}=n_{-}\eqqcolon{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}n}, and one can rescale p±p_{\pm} in such a way that p+=pE02np_{+}=p_{-}\in E_{0}^{2n}.

  2. (ii)

    k=2k=2 and p±E1p_{\pm}\in E_{1}.

  3. (iii)

    k2k\geq 2 and n+=nnn_{+}=n_{-}\eqqcolon{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}n}, and there is one jj relatively prime to kk such that 0<j<k/20<j<k/2 and up to rescaling,

    p±=qw±qp_{\pm}=q-w_{\pm}q

    for the same single element qEj2n{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}q}\in E_{j}^{2n}. Moreover, dimEj2n=dimEj2n=1\dim E_{j}^{2n}=\dim E_{-j}^{2n}=1.

Proof.

This is a case analysis following Lemma˜5.10.

(i) The case one of p±p_{\pm} lies in E0E_{0}

Suppose p+E0p_{+}\in E_{0} and, for a contradiction, suppose mm is minimal such that there exists a nonzero xEmx\in E_{\ell}^{m} for some \ell indivisible by kk. Then xw+xx-w_{+}x is a 1-1-eigenvector of w+w_{+}, hence divisible by p+p_{+} by Lemma˜5.9. We must have x=w+xx=w_{+}x for otherwise the component of (xw+x)/p+(x-w_{+}x)/p_{+} in Em2n+\smash{E_{\ell}^{m-2n_{+}}} would contradict minimality of mm. Thus xEk/2x\in E_{k/2} by Lemma˜5.8, so wx=w+x=xw_{-}x=-w_{+}x=-x by Lemma˜5.8 and pp_{-} divides xx by Lemma˜5.9. It follows, again lest we contradict minimality of mm, that x/pE0x/p_{-}\in E_{0} and so pEk/2p_{-}\in E_{k/2}. But then, by Lemma˜5.8 again, wp+=w+p+=p+w_{-}p_{+}=w_{+}p_{+}=-p_{+} so pp_{-} divides p+p_{+} by Lemma˜5.9 and we have

w+(p+p)=w+p+w+p=p+wp=p+p,w_{+}\Big{(}\frac{p_{+}}{p_{-}}\Big{)}=\frac{w_{+}p_{+}}{w_{+}p_{-}}=\frac{-p_{+}}{-w_{-}p_{-}}=-\frac{p_{+}}{p_{-}},

contradicting the fact p+p_{+} has minimal degree in the 1-1-eigenspace of w+w_{+}. Thus in fact HH=E0H^{*}_{H}=E_{0}. Since rr then acts trivially but Lemma˜5.6 states the action of D2kD_{2k} is effective, it follows k=1k=1. By Lemma˜5.8 the actions of ww_{-} and w+w_{+} agree on E0E_{0}, so the w±w_{\pm}-eigenspace decompositions of HHH^{*}_{H} in Lemma˜5.9 coincide and by rescaling we may assume p=p+p_{-}=p_{+}.

(ii) The case one of p±p_{\pm} lies in Ek/2E_{k/2}

If p+Ek/2p_{+}\in E_{k/2}, then in fact HH=E0Ek/2H^{*}_{H}=E_{0}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{k/2}, for we could otherwise produce from any nonzero homogeneous xEx\in E_{\ell}, where k/2k/2 does not divide \ell, an element (xw+x)/p+(x-w_{+}x)/p_{+} of smaller degree with a nonzero component in Ek/2E_{\ell-k/2}. It follows r2r^{2} acts as the identity on HHH^{*}_{H}, and since the action of D2kD_{2k} is effective by Lemma˜5.6, we have k=2k=2 and p+E1p_{+}\in E_{1}. As for pp_{-}, consulting Lemma˜5.10, it must lie in E0E_{0} or E1E_{1}, but if it lay in E0E_{0} we would be in the previous case.

(iii) The case p±p_{\pm} both lie in Ej±Ej±E_{j^{\pm}}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{-j^{\pm}} for some j±j^{\pm} with 0<j±<k/20<j^{\pm}<k/2

We first note that n+=nn_{+}=n_{-}, for if we had, say, n+>nn_{+}>n_{-}, then by Lemma˜5.9, w+w_{+} would act as the identity on HH2n\smash{H^{2n_{-}}_{H}}; but this would imply p=qjwqj\smash{p_{-}=q_{j^{-}}-w_{-}q_{j^{-}}} is equal to w+p=w+qjζjqj\smash{w_{+}p_{-}=w_{+}q_{j^{-}}-\zeta^{j^{-}}q_{j^{-}}}, which could only happen if ζj=1\zeta^{j^{-}}=-1 and jk/2(modk)j^{-}\equiv k/2\allowbreak\mkern 8.0mu({\operator@font mod}\mkern 6.0muk).

Next we claim j+=jj^{+}=j^{-}. As already mentioned, the 1-1-eigenspace of w+w_{+} in HH2nH^{2n}_{H} is Cp+\mathbb Cp_{+}, which lies in Ej+Ej+E_{j^{+}}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{-j^{+}} Since the ±1\pm 1-eigenspaces of w+w_{+} on E2nE2nE_{\ell}^{2n}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{-\ell}^{2n} are {x±w+x:xE2n}\{x\pm w_{+}x:x\in E_{\ell}^{2n}\} and hence both are of dimension dimE2n\dim E_{\ell}^{2n}, but the 1-1-eigenspace is trivial except for =j\ell=j^{-}, it follows all the other terms in 0<<k/2(E2nE2n)\bigoplus_{0<\ell<k/2}(E_{\ell}^{2n}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{-\ell}^{2n}) are zero, so j+=jjj^{+}=j^{-}\eqqcolon{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}j} and dimEj2n=dimEj2n=1\dim E_{j}^{2n}=\dim E_{-j}^{2n}=1. Thus we may rescale to take q=q+q_{-}=q_{+} as claimed.

To see that jj and kk are coprime, first note that HH=iEigcd(j,k)H^{*}_{H}=\sum_{i}E_{i\cdot\gcd(j,k)}, for given a putative nonzero homogeneous element xEx\in E_{\ell} such that gcd(j,k)\gcd(j,k) does not divide \ell, the component of (xw+x)/p+(x-w_{+}x)/p_{+} in EjE_{\ell-j} would be nonzero of smaller degree. Thus rk/gcd(j,k)D2kr^{k/\!\gcd(j,k)}\in D_{2k} acts trivially on HHH^{*}_{H}, and since the D2kD_{2k}-action is effective by Lemma˜5.6, it follows gcd(j,k)=1\gcd(j,k)=1. ∎

Lemma 5.12.

The elements p±p_{\pm} are prime elements of HHH^{*}_{H}.

Proof.

Recall from Remark˜4.8 that the basis element ee from Proposition˜4.7 may be taken π0H\pi_{0}H-invariant. Thus the same holds of the eigenvectors p±=(e±w±e±)/2p_{\pm}=(e_{\pm}-w_{\pm}e_{\pm})/2 defined in Lemma˜5.9. It is clear in HH0H^{*}_{H_{0}} that p±p_{\pm} are irreducible, for all elements of lesser degree lie in the 11-eigenspace HK0±\smash{H\mspace{-1.0mu}\begin{subarray}{c}*\phantom{k}\\ K_{0}^{\pm}\end{subarray}}. Since HH0H^{*}_{H_{0}} is a polynomial ring, the principal ideals (p±)(p_{\pm}) are prime. In fact, the ideals (p±)(p_{\pm}) are prime in HHH^{*}_{H} as well, for given x,yHHx,y\in H^{*}_{H} such that xy(p±)xy\in(p_{\pm}), we know one of the two, say xx, is divisible by p±p_{\pm} in HH0H^{*}_{H_{0}}. But then as x/p±x/p_{\pm} is π0H\pi_{0}H-invariant as well, xx is also divisible by p±p_{\pm} in HHH^{*}_{H}. ∎

Lemma 5.13.

Write V{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}V} for the joint 1-1-eigenspace of w±w_{\pm}. Then

VE0(HK+HK+)E0HHHK+HK+V\longrightarrow\frac{E_{0}}{(H^{*}_{K^{-}}+H^{*}_{K^{+}})\cap E_{0}}\longrightarrow\frac{H^{*}_{H}}{H^{*}_{K^{-}}+H^{*}_{K^{+}}}

are isomorphisms.

All unelaborated claims in the proof are clauses of Lemma˜5.8.

Proof.

Write E0=1k1E{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}E_{\neq 0}}\coloneqq\bigoplus_{\ell=1}^{k-1}E_{\ell} so that we have a direct sum decomposition HH=E0E0H^{*}_{H}=E_{0}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{\neq 0}. This decomposition is invariant under w±w_{\pm}, so HK±H^{*}_{K^{\pm}} inherit such decompositions and

HK+HK+=([HKE0]+[HK+E0])([HKE0]+[HK+E0]).H^{*}_{K^{-}}+H^{*}_{K^{+}}=\big{(}[H^{*}_{K^{-}}\cap E_{0}]+[H^{*}_{K^{+}}\cap E_{0}]\big{)}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}\big{(}[H^{*}_{K^{-}}\cap E_{\neq 0}]+[H^{*}_{K^{+}}\cap E_{\neq 0}]\big{)}.

Because w+|E0=w|E0w_{+}|_{E_{0}}=w_{-}|_{E_{0}}, the first direct summand is the common 11-eigenspace of w±w_{\pm} on E0E_{0}, whose complement is VV. We will be done if we can show the second summand is all of E0E_{\neq 0}.

  • If kk is even, r=w+wr=w_{+}w_{-} acts on Ek/2E_{k/2} as multiplication by ζk/2=1\zeta^{k/2}=-1, so w+|Ek/2=w|Ek/2w_{+}|_{E_{k/2}}=-w_{-}|_{E_{k/2}}. Thus Ek/2E_{k/2} decomposes as the sum of the 11-eigenspace HK+Ek/2H^{*}_{K^{+}}\cap E_{k/2} of w+w_{+} on Ek/2E_{k/2} and its 1-1-eigenspace, which is HKEk/2H^{*}_{K^{-}}\cap E_{k/2}.

  • If 0<<k/20<\ell<k/2, then since rr acts as ζ±1\zeta^{\ell}\neq\pm 1 on EE_{\ell}, we have w+vwvw_{+}v\neq w_{-}v for nonzero vEv\in E_{\ell}, so for nonzero u,vEu,v\in E_{\ell}, we cannot have u+w+u=v+wvu+w_{+}u=v+w_{-}v in EEE_{\ell}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{-\ell}. Thus the 11-eigenspaces (id+w±)E(\operatorname{id}+w_{\pm})E_{\ell} of w±w_{\pm} are disjoint, and since dimEm=dimEm\dim E_{\ell}^{m}=\dim E_{-\ell}^{m} for all m0m\geq 0, their sum is all of EEE_{\ell}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{-\ell}.∎

We are now finally in a position to prove Proposition 1.

Proof of Proposition 1.

We proceed through the trichotomy of Lemma˜5.11, in each case applying Section˜3.2.

(i) The case p=p+=pE02np=p_{+}=p_{-}\in E_{0}^{2n} and HH=E0H^{*}_{H}=E_{0}

In this case the actions of w±w_{\pm} coincide by Lemma˜5.8, so the images HK{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}{H^{*}_{K}}} of the injections HK±HHH^{*}_{K^{\pm}}\xhookrightarrow{\ \ \ \ }H^{*}_{H} agree. Thus

HGevenM\displaystyle H_{G}^{\mathrm{even}}M =HKHK+=HK;\displaystyle=H^{*}_{K^{-}}\cap H^{*}_{K^{+}}={H^{*}_{K}};
HGoddM\displaystyle H_{G}^{\mathrm{odd}}M =HKpHKHK[1]pHK[1]HK[2n+1].\displaystyle=\frac{{H^{*}_{K}}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}p{H^{*}_{K}}}{{H^{*}_{K}}}[1]\cong p{H^{*}_{K}}[1]\cong{H^{*}_{K}}[2n+1].

(ii) The case k=2k=2 and HH=E0E1H^{*}_{H}=E_{0}\mathbin{\raisebox{0.85pt}{$\displaystyle\oplus$}}E_{1} and p±E1p_{\pm}\in E_{1}.

We show V=p+p(HKHK+)V=p_{+}p_{-}\cdot(H^{*}_{K^{-}}\cap H^{*}_{K^{+}}) and apply Lemma˜5.13; since degp+p=2n+2n+\deg p_{+}p_{-}=2n_{-}+2n_{+}, the result will then follow from Section˜3.2. Suppose xx lies in VV. From Lemma˜5.9 we know xx is divisible by both pp_{-} and p+p_{+}. Note that wp±=ζk/2w±p±=p±w_{\mp}p_{\pm}=\zeta^{-k/2}w_{\pm}p_{\pm}=p_{\pm} by Lemma˜5.8. As ww_{-} sends x/p+x/p_{+} to wx/wp+=(x/p+)w_{-}x/w_{-}p_{+}=-(x/p_{+}), we see pp_{-} divides the latter as well. The quotient w±(x/p+p)=x/p+p=x/p+pw_{\pm}(x/p_{+}p_{-})=-x/{-p_{+}p}_{-}=x/p_{+}p_{-} is in the joint 11-eigenspace HKHK+H^{*}_{K^{-}}\cap H^{*}_{K^{+}}.

(iii) The case p±=qw±qp_{\pm}=q-w_{\pm}q for some qEj2nq\in E_{j}^{2n}.

We will find an element P{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}P} of degree 2nk2nk such that V=P(HKHK+)V=P\cdot(H^{*}_{K^{-}}\cap H^{*}_{K^{+}}) and apply Lemma˜5.13. Since w±w_{\pm} generate all wD2kw\in D_{2k}, we have wx=±xw\cdot x=\pm x for any xVx\in V. Particularly, xx is also divisible by the wp±wp_{\pm}, which we explicitly enumerate. Writing η=ζj{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}\eta}=\zeta^{j}, from Lemma˜5.8 we obtain relations p=qηw+qp_{-}=q-\eta w_{+}q and r(qηw+q)=η(qη2w+q)r(q-\eta^{\ell}w_{+}q)=\eta(q-\eta^{\ell-2}w_{+}q) which suffice to show that if we consider elements only up to nonzero complex multiples, the sets

{wp,wp+:wD2k}and{qηw+q:Z}={qηwq:Z}\{wp_{-},wp_{+}:w\in D_{2k}\}\qquad\mbox{and}\qquad\{q-\eta^{\ell}w_{+}q:\ell\in\mathbb Z\}=\{q-\eta^{\ell}w_{-}q:\ell\in\mathbb Z\}

are equal. Since jj and kk are relatively prime by Lemma˜5.11.(iii), the elements qηw+qq-\eta^{\ell}w_{+}q for 0<k0\leq\ell<k are all distinct elements, none a scalar multiple of any other since their EjE_{j}-components are equal, and prime by Lemma˜5.12. Since each divides xx, their product P{\color[rgb]{.255,.41,.884}\definecolor[named]{pgfstrokecolor}{rgb}{.255,.41,.884}P} also divides xx. Note that

P==0k1(qηw±q)=qkw±qk.P=\prod_{\ell=0}^{k-1}(q-\eta^{\ell}w_{\pm}q)=q^{k}-w_{\pm}q^{k}.

But the right-hand side is a 1-1-eigenvector of w±w_{\pm}, so w±xP=xP=xP\displaystyle\smash{w_{\pm}\frac{x}{P}=\frac{-x}{-P}=\frac{x}{P}} lies in HKHK+H^{*}_{K^{-}}\cap H^{*}_{K^{+}}. ∎

We now illustrate Section˜1 with some examples.

Example 5.14.

There is a cohomogeneity one action of G=SU(3)G=\mathrm{SU}(3) on the sphere S7S^{7} with isotropy groups K=S(U(2)×U(1))K^{-}={\rm S}\big{(}\mathrm{U}(2)\times\mathrm{U}(1)\big{)}, K+=S(U(1)×U(2))K^{+}={\rm S}\big{(}\mathrm{U}(1)\times\mathrm{U}(2)\big{)}, and H=S(U(1)×U(1)×U(1))H={\rm S}\big{(}\mathrm{U}(1)\times\mathrm{U}(1)\times\mathrm{U}(1)\big{)} [GWZ08, Table E]. Observe that rkG=rkH\operatorname{rk}G=\operatorname{rk}H and we have n=n+=1n_{-}=n_{+}=1. Using Remark˜5.7 and the table in Grove et al. [GWZ08], one deduces that k=3k=3. Here HHH^{*}_{H} is just the quotient ring Q[t1,t2,t3]/(t1+t2+t3)\mathbb Q[t_{1},t_{2},t_{3}]/(t_{1}+t_{2}+t_{3}), which admits a natural Σ3\Sigma_{3}-action permuting the generators tjt_{j}. Within HHH^{*}_{H} we have HK=(HH)(1 2)H^{*}_{K^{-}}=(H^{*}_{H})^{\langle(1\ 2)\rangle} and HK+=(HH)(2 3)H^{*}_{K^{+}}=(H^{*}_{H})^{\langle(2\ 3)\rangle}, whose intersection is (HH)Σ3HSU(3)(H^{*}_{H})^{\Sigma_{3}}\cong H^{*}_{\mathrm{SU}(3)}. Section˜1 implies

HSU(3)S7=HSU(3)HS7.H^{*}_{\mathrm{SU}(3)}S^{7}=H^{*}_{\mathrm{SU}(3)}\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}S^{7}.

This is in fact a known result, being equivalent to the equivariant formality of the action [GM14, Cor. 1.3] since there is only one possible graded HSU(3)H^{*}_{\mathrm{SU}(3)}-algebra structure for a free graded HSU(3)H^{*}_{\mathrm{SU}(3)}-module on generators of degrees 0 and 77.

Similar calculations can be made for any of the last five examples in Grove et al. [GWZ08, Table E]. Note that they are all equivariantly formal actions. This is not the case in the next example.

Example 5.15.

The left action of SU(3)\mathrm{SU}(3) on itself given by (A,B)ABA(A,B)\longmapsto ABA^{\top} has cohomogeneity one [Pütt09, Example 5.5]. One can see that relative to the canonical metric on SU(3)\mathrm{SU}(3), there is a transversal geodesic segment joining the two singular orbits, along which the isotropies are K+=SO(3)K^{+}=\mathrm{SO}(3), K=SU(2)×{1}K^{-}=\mathrm{SU}(2)\times\{1\}, and H=SO(2)×{1}H=\mathrm{SO}(2)\times\{1\}. The Weyl group W(γ)W(\gamma) turns out to be D4D_{4}, but the induced symmetry group Ξ/W(H)\Xi/W(H) is just AutSO(2)=D2\operatorname{Aut}\mathrm{SO}(2)=D_{2}, so we are in the case k=1k=1. If we write xx for a generator of HSO(2)=Q[x]H^{*}_{\mathrm{SO}(2)}=\mathbb Q[x], then the canonical maps send HSO(3)H^{*}_{\mathrm{SO}(3)} and HSU(2)H^{*}_{\mathrm{SU}(2)} both isomorphically to Q[x2]\mathbb Q[x^{2}]. Since n=n+=1n_{-}=n_{+}=1, one concludes

HSU(3)SU(3)Q[x2]HS3.H^{*}_{\mathrm{SU}(3)}\mathrm{SU}(3)\cong\mathbb Q[x^{2}]\operatorname*{\mathchoice{\raisebox{0.85pt}{$\displaystyle\otimes$}}{\raisebox{0.85pt}{$\otimes$}}{\raisebox{0.7pt}{$\scriptstyle\otimes$}}{\raisebox{0.2pt}{$\scriptscriptstyle\otimes$}}}H^{*}S^{3}.
Remark 5.16.

Again, as in Remark˜5.1, one can deduce directly that for K±K^{\pm} and HH as in Section˜1, HGMH^{*}_{G}M is Cohen–Macaulay as an HG{H^{*}_{G}}-module. This time, one notices that the equivariant cohomology is the direct sum of two copies of HK+HK=(HH)D2kH^{*}_{K^{+}}\cap H^{*}_{K^{-}}=(H^{*}_{H})^{D_{2k}}, the latter being a HGH^{*}_{G}-algebra which is finitely generated as an HG{H^{*}_{G}}-module and Cohen–Macaulay as a ring.

Remark 5.17.

When rkG\operatorname{rk}G is also equal to rkH\operatorname{rk}H, the GG-action is equivariantly formal. In this case MM is odd-dimensional and the situation is particularly nice because the restricted action of a maximal torus of GG is of GKM type in a sense recently defined by the third author [He16, Example 4.17].

References

  • [AA93] Andrey V. Alekseevsky and Dmitry V. Alekseevsky. Riemannian GG-manifold with one-dimensional orbit space. Ann. Global Anal. Geom., 11(3):197–211, 1993. http://link.springer.com/article/10.1007/BF00773366.
  • [Bes87] Arthur Besse. Einstein manifolds, volume 10 of Ergebnisse der Mathematik und ihrer Grenzgebiete, 3. Folge. Springer, 1987.
  • [Bor53] Armand Borel. Sur la cohomologie des espaces fibrés principaux et des espaces homogènes de groupes de Lie compacts. Ann. of Math. (2), 57(1):115–207, 1953. http://jstor.org/stable/1969728.
  • [Bre61] Glen E. Bredon. On homogeneous cohomology spheres. Ann. of Math., pages 556–565, 1961. http://jstor.org/stable/1970317.
  • [BtD85] Theodor Bröcker and Tammo tom Dieck. Representations of compact Lie groups, volume 98 of Grad. Texts in Math. Springer, 1985.
  • [Car18] Jeffrey D. Carlson. The equivariant K-theory of a cohomogeneity-one action. 2018. arXiv:1805.00502.
  • [CK11] Suyoung Choi and Shintarô Kuroki. Topological classification of torus manifolds which have codimension one extended actions. Algebr. Geom. Top., 11(5):2655–2679, 2011. http://projecteuclid.org/euclid.agt/1513715302, arXiv:0906.1335.
  • [EU11] Christine M. Escher and Shari K. Ultman. Topological structure of candidates for positive curvature. Topology Appl., 158(1):38–51, 2011. doi:10.1016/j.topol.2010.10.001.
  • [GGS11] Fernando Galaz-García and Catherine Searle. Cohomogeneity one Alexandrov spaces. Transform. Groups, 16(1):91–107, 2011. arXiv:0910.5207.
  • [GGZ15] Fernando Galaz-García and Masoumeh Zarei. Cohomogeneity one topological manifolds revisited. Math. Z., 2015. URL: http://link.springer.com/article/10.1007/s00209-017-1915-y, arXiv:1503.09068.
  • [GM14] Oliver Goertsches and Augustin-Liviu Mare. Equivariant cohomology of cohomogeneity one actions. Topology Appl., 167:36–52, 2014. arXiv:1110.6318.
  • [GM17] Oliver Goertsches and Augustin-Liviu Mare. Equivariant cohomology of cohomogeneity-one actions: The topological case. Topology Appl., 2017. arXiv:1609.07316.
  • [GrH87] Karsten Grove and Stephen Halperin. Dupin hypersurfaces, group actions and the double mapping cylinder. J. Differential Geom., 26(3):429–459, 1987. doi:10.4310/jdg/1214441486.
  • [GWZ08] Karsten Grove, Burkhard Wilking, and Wolfgang Ziller. Positively curved cohomogeneity one manifolds and 3-Sasakian geometry. J. Differential Geom., 78:33–111, 2008. http://projecteuclid.org/euclid.jdg/1197320603.
  • [Hat02] Allen Hatcher. Algebraic topology. Cambridge Univ. Press, 2002. http://math.cornell.edu/˜hatcher/AT/ATpage.html.
  • [Hat09] Allen Hatcher. Vector bundles and K-theory. 2009 manuscript. http://math.cornell.edu/˜hatcher/VBKT/VBpage.html.
  • [He16] Chen He. Localization of certain odd-dimensional manifolds with torus actions. 2016. arXiv:1608.04392.
  • [Hoel10] Corey A. Hoelscher. On the homology of low-dimensional cohomogeneity one manifolds. Transform. Groups, 15(1):115–133, 2010.
  • [Kan01] Richard Kane. Reflection groups and invariant theory, volume 5 of C. M. S. Books in Mathematics. Springer, 2001.
  • [Kur11] Shintarô Kuroki. Classification of torus manifolds with codimension one extended actions. Transf. Groups, 16(2):481–536, Jun 2011. doi:10.1007/s00031-011-9136-7.
  • [May99] J. Peter May. A concise course in algebraic topology. University of Chicago Press, 1999.
  • [MiT00] Mamoru Mimura and Hiroshi Toda. Topology of Lie groups, I and II, volume 91 of Transl. Math. Monogr. Amer. Math. Soc., Providence, RI, 2000.
  • [Most57a] Paul S. Mostert. On a compact Lie group acting on a manifold. Ann. of Math., 65(3):447–455, 1957. http://jstor.org/stable/1970056.
  • [Most57b] Paul S. Mostert. Errata: On a compact Lie group acting on a manifold. Ann. of Math., 66(3):589, 1957. http://jstor.org/stable/1969911.
  • [PT87] Richard Palais and Chuu-Lian Terng. A general theory of canonical forms. Trans. Amer. Math. Soc., 300:771–789, 1987. URL: http://ams.org/journals/tran/1987-300-02/S0002-9947-1987-0876478-4/.
  • [Pütt09] Thomas Püttmann. Cohomogeneity one manifolds and selfmaps of nontrivial degree. Transform. Groups, 14:225–247, 2009. URL: http://link.springer.com/article/10.1007/s00031-008-9037-6, arXiv:0710.3770.
  • [Sam41] Hans Samelson. Beiträge zur Topologie der Gruppen-Mannigfaltigkeiten. Ann. of Math., 42(1):1091–1137, Jan 1941. http://jstor.org/stable/1970463.
  • [Seg68] Graeme Segal. The representation-ring of a compact Lie group. Publ. Math. Inst. Hautes Études Sci., 34:113–128, 1968. http://numdam.org/item/PMIHES_1968__34__113_0.

Department of Mathematics, University of Toronto
jcarlson@math.toronto.edu

Fachbereich Mathematik und Informatik, Philipps-Universität Marburg
goertsch@mathematik.uni-marburg.de

Yau Mathematical Sciences Center, Tsinghua University
che@math.tsinghua.edu.cn

Department of Mathematics and Statistics, University of Regina
mareal@math.uregina.ca