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

An equivariant Poincaré duality for proper cocompact actions by matrix groups

Hao Guo Department of Mathematics, Texas A&M University [email protected]  and  Varghese Mathai School of Mathematical Sciences, University of Adelaide [email protected]
Abstract.

Let GG be a linear Lie group acting properly on a GG-spinc manifold MM with compact quotient. We give a short proof that Poincaré duality holds between GG-equivariant KK-theory of MM, defined using finite-dimensional GG-vector bundles, and GG-equivariant KK-homology of MM, defined through the geometric model of Baum and Douglas.

2020 Mathematics Subject Classification:
19K33, 19L47, 53C27

1. Introduction

Poincaré duality in KK-theory asserts that the KK-theory group of a closed spinc manifold is naturally isomorphic to its KK-homology group via cap product with the fundamental class in KK-homology. This class can be represented geometrically by the spinc-Dirac operator. More generally, if a compact Lie group acts on the manifold preserving the spinc structure, the analogous map implements Poincaré duality between the equivariant versions of KK-theory and KK-homology. In the case when the Lie group is non-compact but has finite component group, induction on KK-theory and KK-homology allow one to establish the analogous result [7] for proper actions. The observation underlying Poincaré duality in all of these cases is that there exist enough equivariant vector bundles with which to pair the fundamental class.

In contrast, Phillips [14] showed through a counter-example with a non-linear group that, for proper actions by a general Lie group GG on a manifold XX with compact quotient space, finite-dimensional vector bundles do not exhaust the GG-equivariant KK-theory of XX, and that it is necessary to introduce infinite-dimensional vector bundles into the description of KK-theory (see also [10]).

One case in which finite-dimensional bundles are sufficient is when the group GG is linear (see [13]), owing to the key fact that in this case every GG-equivariant vector bundle over XX is a direct summand of a GG-equivariantly trivial bundle, ie. one that is isomorphic to X×VX\times V for some finite-dimensional representation of GG on VV.

Motivated by this, we give a short proof of Poincaré duality in this setting. That is, we show that the natural map from GG-equivariant KK-theory to GG-equivariant KK-homology (which for us means Baum-Douglas’ geometric KK-homology [2]), given by pairing with the fundamental class, is an isomorphism for proper cocompact GG-spinc manifolds where GG is a matrix group.

Theorem 1.1.

Suppose a linear Lie group GG acts properly and cocompactly on a GG-equivariantly spinc manifold XX. Then there is a natural isomorphism

ψ:KG(X)KG(X),\psi\colon K_{G}^{*}(X)\to K^{G}_{*}(X), (1.1)

where the left and right-hand sides denote GG-equivariant KK-theory and geometric KK-homology respectively.

For compact Lie group actions, Theorem 1.1 is implied by the work of [5] on the isomorphism between the equivariant geometric and analytic models of KK-homology. Meanwhile, by the Peter-Weyl theorem, such groups form a subclass of linear Lie groups. Consequently, Theorem 1.1 provides another approach to some of the results in [5].

While Theorem 1.1 makes no reference to the analytic model of KK-homology [1, 9], we note that (1.1) still holds when the right-hand side is replaced by the analytic KK-homology group KKG(C0(X),)KK_{*}^{G}(C_{0}(X),\mathbb{C}). Indeed, as a special case of Emerson-Meyer’s general second duality result in [6, Section 6], (in particular, see the first display after (1.5) in [6]), we have

KKG(C0(X),)KG(TX).KK_{*}^{G}(C_{0}(X),\mathbb{C})\cong K_{G}^{*}(TX).

By the Thom isomorphism from [14, Theorem 8.11] (see also Theorem 2.6 below), we have

KG(TX)KG(X).K_{G}^{*}(TX)\cong K^{*}_{G}(X).

Putting this together gives KG(X)KKG(C0(X),)K_{G}^{*}(X)\cong KK_{*}^{G}(C_{0}(X),\mathbb{C}). Combined with Theorem 1.1, this gives:

Corollary 1.2.

Suppose a linear Lie group GG acts properly and cocompactly on a GG-equivariantly spinc manifold XX. Then we have an isomorphism between the GG-equivariant geometric and analytic KK-homology groups of XX:

KG(X)KKG(C0(X),).K^{G}_{*}(X)\cong KK_{*}^{G}(C_{0}(X),\mathbb{C}). (1.2)

In particular, the geometric KK-homology groups KGK_{*}^{G} are part of a GG-equivariant extraordinary homology theory.

The natural map realizing this isomorphism (1.2) is the Baum-Douglas map [2, 7], which can be described as follows. Let (M,E,f)(M,E,f) be a GG-equivariant geometric KK-cycle for KG(X)K_{*}^{G}(X) (see Definition 2.8 below). Then MM is a GG-equivariantly spinc manifold with Dirac operator DMD_{M} acting on sections of a spinor bundle SMS_{M}, EE is a GG-equivariant vector bundle over MM, and f:MXf\colon M\to X is a GG-equivariant continuous map. The Baum-Douglas map takes

[M,E,f]f[L2(SME),φ,DE(1+DE2)1/2],[M,E,f]\mapsto f_{*}\left[L^{2}(S_{M}\otimes E),\varphi,D_{E}(1+D_{E}^{2})^{-1/2}\right],

where φ\varphi is the *-representation of C0(M)C_{0}(M) on (L2(E))\mathcal{B}(L^{2}(E)) given by pointwise multiplication, and the right-hand side is the pushforward under ff of a class in KKG(C0(M),)KK_{*}^{G}(C_{0}(M),\mathbb{C}). Corollary 1.2 then implies:

Corollary 1.3.

Suppose a linear Lie group GG acts properly and cocompactly on a GG-equivariantly spinc manifold XX. Then for any eKKG(C0(X),)e\in KK_{*}^{G}(C_{0}(X),\mathbb{C}), there exist a GG-equivariantly cocompact spinc manifold MM and a GG-equivariant continuous map f:MXf\colon M\to X such that ee is the pushforward under ff of the class of a Dirac-type operator on MM in KKG(C0(M),)KK_{*}^{G}(C_{0}(M),\mathbb{C}). More precisely, there exists a vector bundle EME\to M such that

e=f[L2(SME),φ,DE(1+DE2)1/2].e=f_{*}\left[L^{2}(S_{M}\otimes E),\varphi,D_{E}(1+D_{E}^{2})^{-1/2}\right].

Acknowledgements. The authors would like to thank Peter Hochs and Hang Wang for their useful feedback on this paper.

Hao Guo was partially supported by funding from the Australian Research Council through the Discovery Project DP200100729, and partially by the National Science Foundation through the NSF DMS-2000082. Varghese Mathai was partially supported by funding from the Australian Research Council, through the Australian Laureate Fellowship FL170100020.

2. Preliminaries

We begin by recalling the definitions and facts we will need. Unless specified otherwise, GG will always denote a closed subgroup of GL(n,)GL(n,\mathbb{R}) for some nn. All vector bundles will be complex. For this section, let XX be a locally compact proper GG-space.

2.1. Equivariant KK-theory

In [13], Phillips showed that the GG-equivariant KK-theory of the space XX with GG-cocompact supports can be defined in a such a way that is directly analogous to non-equivariant, compactly supported KK-theory.

Definition 2.1 ([13, Definition 1.1]).

A GG-equivariant KK-cocycle for XX is a triple (E,F,t)(E,F,t) consisting of two finite-dimensional complex GG-vector bundles EE and FF over XX and a GG-equivariant bundle bundle map t:EFt:E\to F whose restriction to the complement of some GG-cocompact subset of XX is an isomorphism. Two KK-cocycles (E,F,t)(E,F,t) and (E,F,t)(E^{\prime},F^{\prime},t^{\prime}) are said to be equivalent if there exist finite-dimensional GG-vector bundles HH and HH^{\prime} and GG-equivariant isomorphisms

a:EHEH,b:FHFHa:E\oplus H\to E^{\prime}\oplus H,\qquad b:F\oplus H\to F^{\prime}\oplus H^{\prime}

such that bx1(txid)ax=txidb_{x}^{-1}(t^{\prime}_{x}\oplus\operatorname{id})a_{x}=t_{x}\oplus\operatorname{id} for all xx in the complement of a GG-cocompact subset of XX. The set of equivalence classes [E,F,t][E,F,t] of KK-cocycles forms a semigroup under the direct sum operation, and we define the group KG0(X)K^{0}_{G}(X) is the Grothendieck completion of the semigroup of finite-dimensional complex GG-vector bundles over XX.

Remark 2.2.

When it is clear from context, or when XX is GG-cocompact, we will omit the map tt from the cycle, and simply denote a class in KG0(X)K^{0}_{G}(X) by [E][F][E]-[F].

Remark 2.3.

For general locally compact groups, Definition 2.1 needs to be modified to include infinite-dimensional bundles [14, Chapter 3]. For GG linear, this is not necessary [13, Theorem 2.3].

Definition 2.4.

For each non-negative integer ii, let

KGi(X)=KG0(X×i),K^{i}_{G}(X)=K^{0}_{G}(X\times\mathbb{R}^{i}),

where GG acts trivially on i\mathbb{R}^{i}.

By [13, Lemma 2.2], KGiK^{i}_{G} satisfies Bott periodicity, so that we have a natural isomorphism KGi(X)KGi+2(X)K^{i}_{G}(X)\cong K^{i+2}_{G}(X) for each ii. We will use the notation

KG(X)=KG0(X)KG1(X).K^{*}_{G}(X)=K^{0}_{G}(X)\oplus K^{1}_{G}(X).

In addition, KGiK^{i}_{G} are contravariant functors from the category of proper GG-spaces and proper GG-equivariant maps to the category of abelian groups, and form an equivariant extraordinary cohomology theory with a continuity property. In particular, Bott periodicity implies that for any GG-invariant open subset UXU\subseteq X, there is a six-term exact sequence of abelian groups

KG0(U){K^{0}_{G}(U)}KG0(X){K^{0}_{G}(X)}KG0(X\U){K^{0}_{G}(X\backslash U)}KG1(X\U){K^{1}_{G}(X\backslash U)}KG1(X){K^{1}_{G}(X)}KG1(U).{K^{1}_{G}(U).}\scriptstyle{\partial}\scriptstyle{\partial} (2.1)

Here, the map KGi(U)KGi(X)K^{i}_{G}(U)\to K^{i}_{G}(X) is induced by the extension-by-zero homomorphism – see Remark 2.5 below, and the boundary maps \partial are defined as in equivariant KK-theory for compact group actions [15].

Remark 2.5 (Extension-by-zero).

Any inclusion of GG-invariant open subsets U1U2U_{1}\hookrightarrow U_{2} induces in the obvious way an extension-by-zero *-homomorphism C0(U1)C0(U2)C_{0}(U_{1})\to C_{0}(U_{2}). This extends to a *-homomorphism C0(U1)GC0(U2)GC_{0}(U_{1})\rtimes G\to C_{0}(U_{2})\rtimes G between crossed products. The induced map on operator KK-theory, together with the identification KGi(Uj)Ki(C0(Uj)G)K^{i}_{G}(U_{j})\cong K_{i}(C_{0}(U_{j})\rtimes G), gives a map

KGi(U1)KGi(U2).K_{G}^{i}(U_{1})\to K_{G}^{i}(U_{2}).

To prove Poincaré duality, we will make use of the following Thom isomorphism theorem for GG-spinc bundles:

Theorem 2.6 ([14, Theorem 8.11]).

Let EE be a finite-dimensional GG-equivariant spinc vector bundle over XX. Then there is a natural isomorphism

TG:KGi(X)KGi+dimE(E)T_{G}\colon K_{G}^{i}(X)\xrightarrow{\cong}K_{G}^{i+\dim E}(E)

for i=0,1i=0,1, where dimE\dim E is the real dimension of EE and i+dimEi+\dim E is taken mod 22.

2.2. The Gysin homomorphism

Theorem 2.6 can be used to give an explicit geometric description of the Gysin (pushforward) homomorphism in GG-equivariant KK-theory, which we will need later.

Let Y1Y_{1} and Y2Y_{2} be two GG-cocompact GG-spinc manifolds and f:Y1Y2f\colon Y_{1}\to Y_{2} a GG-equivariant continuous map. By cocompactness, both Y1Y_{1} and Y2Y_{2} contain only finitely many orbit types. Together with the fact that GG is a linear group, this implies, by [11, Theorem 4.4.3], that there exists a GG-equivariant embedding

jY1:Y12nj_{Y_{1}}\colon Y_{1}\to\mathbb{R}^{2n}

for some nn, where GG is considered as a subgroup of GL(2n,)GL(2n,\mathbb{R}). Let

iY2:Y2Y2×2ni_{Y_{2}}\colon Y_{2}\to Y_{2}\times\mathbb{R}^{2n}

denote the zero section, and define the GG-equivariant embedding

iY1:Y1\displaystyle i_{Y_{1}}\colon Y_{1} Y2×2n,\displaystyle\to Y_{2}\times\mathbb{R}^{2n},
y\displaystyle y (f(y),jY1(y)).\displaystyle\mapsto(f(y),j_{Y_{1}}(y)).

Let ν1\nu_{1} be the normal bundle of iY1i_{Y_{1}}, which we identify with a GG-invariant tubular neighbourhood U1U_{1} of its image. Note that it follows from the two-out-of-three lemma for GG-equivariant spinc-structures (see [12, Section 3.1] and [8, Remark 2.6]), together with the assumption that Y1Y_{1} and Y2Y_{2} are GG-spinc, that ν1\nu_{1} has a GG-spinc structure, and so Theorem 2.6 applies. Identifying the normal bundle with U1U_{1}, we have the Thom isomorphism

TG:KGi(Y1)\displaystyle T_{G}\colon K_{G}^{i}(Y_{1}) KGi+dimν1(U1),\displaystyle\xrightarrow{\cong}K_{G}^{i+\dim\nu_{1}}(U_{1}), (2.2)

for i=0,1i=0,1. The Gysin homomorphism

f!:KGi(Y1)KGi+dimY2dimY1(Y2)f_{!}\colon K^{i}_{G}(Y_{1})\to K^{i+\dim Y_{2}-\dim Y_{1}}_{G}(Y_{2}) (2.3)

associated to ff is then the composition

KGi(Y1){K^{i}_{G}(Y_{1})}KGi+dimY2+2ndimY1(U1){K_{G}^{i+\dim Y_{2}+2n-\dim Y_{1}}(U_{1})}KGi+dimY2+2ndimY1(Y2×2n){K_{G}^{i+\dim Y_{2}+2n-\dim Y_{1}}(Y_{2}\times\mathbb{R}^{2n})}KGi+dimY2dimY1(Y2),{K_{G}^{i+\dim Y_{2}-\dim Y_{1}}(Y_{2}),}TG\scriptstyle{T_{G}}λ\scriptstyle{\lambda}\scriptstyle{\cong}

where λ\lambda is induced by the extension-by-zero map associated to the inclusion of U1U_{1} into Y2×2nY_{2}\times\mathbb{R}^{2n} (see Remark 2.5 below), and the right horizontal isomorphism is due to Bott periodicity.

Remark 2.7.

It can be seen from the above that f!f_{!} depends only on the GG-homotopy class of ff and that the Gysin map is functorial under compositions.

2.3. Equivariant geometric KK-homology

We briefly review the equivariant version of Baum and Douglas’ geometric definition of KK-homology [2]; see [3], [4], [5], or [7] for more details. As before, XX is a locally compact proper GG-space.

Definition 2.8.

A GG-equivariant geometric KK-cycle for XX is a triple (M,E,f)(M,E,f), where

  • MM is a proper GG-cocompact manifold with a GG-equivariant spinc-structure;

  • EE is a smooth GG-equivariant Hermitian vector bundle over MM;

  • f:MXf\colon M\to X is a GG-equivariant continuous map.

For i=0i=0 or 11, the GG-equivariant geometric KK-homology group KiG(X)K_{i}^{G}(X) is the abelian group generated by geometric KK-cycles (M,E,f)(M,E,f) where dimM=i\dim M=i mod 22, subject to the equivalence relation generated by the following three elementary relations:

  1. (i)

    (Direct sum – disjoint union) For two GG-equivariant Hermitian vector bundles E1E_{1} and E2E_{2} over MM and a GG-equivariant continuous map f:MXf\colon M\to X,

    (MM,E1E2,ff)(M,E1E2,f);(M\sqcup M,E_{1}\sqcup E_{2},f\sqcup f)\sim(M,E_{1}\oplus E_{2},f);
  2. (ii)

    (Bordism) Suppose two cycles (M1,E1,f1)(M_{1},E_{1},f_{1}) and (M2,E2,f2)(M_{2},E_{2},f_{2}) are bordant, so that there exists a GG-cocompact proper GG-spinc manifold WW with boundary, a smooth GG-equivariant Hermitian vector bundle EWE\to W and a continuous GG-equivariant map f:WXf\colon W\to X such that (W,E|W,f|W)(\partial W,E|_{\partial W},f|_{\partial W}) is isomorphic to (M1,E1,f1)(M2,E2,f2),(M_{1},E_{1},f_{1})\sqcup(-M_{2},E_{2},f_{2}), where M2-M_{2} denotes M2M_{2} with the opposite GG-spinc structure. Then

    (M1,E1,f1)(M2,E2,f2);(M_{1},E_{1},f_{1})\sim(M_{2},E_{2},f_{2});
  3. (iii)

    (Vector bundle modification) Let VV be a GG-spinc vector bundle of real rank 2k2k over MM. Upon choosing a GG-invariant metric on VV, let M^\widehat{M} be the unit sphere bundle of (M×)V(M\times\mathbb{R})\oplus V, where the bundle M×M\times\mathbb{R} is equipped with the trivial GG-action. Let FF be the Bott bundle over M^\widehat{M}, which is fibrewise the non-trivial generator of K0(S2k)K^{0}(S^{2k}). (See [4, Section 3] for a more detailed description.) Then

    (M,E,f)(M^,Fπ(E),fπ),(M,E,f)\sim(\widehat{M},F\otimes\pi^{*}(E),f\circ\pi),

    where π:M^M\pi:\widehat{M}\to M is the canonical projection.

Addition in Kigeo,G(X)K_{i}^{\textnormal{geo},G}(X) is given by

[M1,E1,f1]+[M2,E2,f2]=[M1M2,E1E2,f1f2],[M_{1},E_{1},f_{1}]+[M_{2},E_{2},f_{2}]=[M_{1}\sqcup M_{2},E_{1}\sqcup E_{2},f_{1}\sqcup f_{2}],

the additive inverse of [M,E,f][M,E,f] is its opposite [M,E,f][-M,E,f], while the additive identity is given by the empty cycle where M=M=\emptyset.

Remark 2.9.

The above definition of classes [M,E,f][M,E,f] continues to make sense if we replace the bundle EE by a KK-theory class. Indeed, if

e=[E1][E2]=[E1][E2]KG0(M),e=[E_{1}]-[E_{2}]=[E_{1}^{\prime}]-[E_{2}^{\prime}]\in K^{0}_{G}(M),

then there exists a GG-vector bundle FF over XX such that

E1E2FE1E2F.E_{1}\oplus E_{2}^{\prime}\oplus F\cong E_{1}^{\prime}\oplus E_{2}\oplus F.

By Definition 2.8 (i), this means

[M,E1,f]+[M,E2,f]+[M,F,f]=[M,E1,f]+[M,E2,f]+[M,F,f].[M,E_{1},f]+[M,E_{2}^{\prime},f]+[M,F,f]=[M,E_{1}^{\prime},f]+[M,E_{2},f]+[M,F,f].

Adding the inverse of [M,F,f][M,F,f] to both sides and rearranging shows that the class

[M,e,f][M,E1,f][M,E2,id]\displaystyle[M,e,f]\coloneqq[M,E_{1},f]-[M,E_{2},\textnormal{id}] =[M,E1,id][M,E2,id]\displaystyle=[M,E_{1}^{\prime},\textnormal{id}]-[M,E_{2}^{\prime},\textnormal{id}]

is well-defined.

Finally, we can describe vector bundle modification using the Gysin homomorphism 2.3. To do this, let M^\widehat{M} be the manifold underlying the vector bundle modification of a cycle (M,E,f)(M,E,f) by a bundle VV, as in Definition 2.8 (iii). Then M^\widehat{M} is the unit sphere bundle of (M×)V(M\times\mathbb{R})\oplus V. We will refer to the GG-equivariant embedding

s:M\displaystyle s\colon M M^(M×)V,\displaystyle\to\widehat{M}\subseteq(M\times\mathbb{R})\oplus V,
m\displaystyle m (m,1,0)\displaystyle\mapsto(m,1,0)

as the north pole section.

Lemma 2.10.

Let (M,E,f)(M,E,f) be a geometric cycle for XX. Let (M^,Fπ(E),fπ)(\widehat{M},F\otimes\pi^{*}(E),f\circ\pi) be its modification by a GG-spinc vector bundle VV of even real rank, and let π:M^M\pi\colon\widehat{M}\to M be the projection. Let s:MM^s\colon M\to\widehat{M} be the north pole section. Then

(M^,Fπ(E),fπ)(M^,s![E],fπ).(\widehat{M},F\otimes\pi^{*}(E),f\circ\pi)\sim(\widehat{M},s![E],f\circ\pi).
Proof.

The proof we give is similar to that of [5, Lemma 3.5] concerning the case of compact Lie group actions; compare also the discussion following [4, Definition 6.9]. To begin, observe that the total space of VV can be identified GG-equivariantly with a GG-invariant tubular neighbourhood UU of the embedding s:MM^s\colon M\to\widehat{M}. The Gysin map s!s_{!} is then the composition

KG(M)TGKG(U)𝜆KG(M^),K_{G}^{*}(M)\xrightarrow{T_{G}}K_{G}^{*}(U)\xrightarrow{\lambda}K_{G}^{*}(\widehat{M}), (2.4)

where TGT_{G} is the Thom isomorphism in the form (2.2), while λ\lambda is the homomorphism induced by the extension-by-zero map C0(U)C0(M^)C_{0}(U)\to C_{0}(\widehat{M}). Note that TGT_{G} is essentially given via tensor product with a “Bott element”, and (2.4) admits the following geometric description. Let FF be the Bott bundle over M^\widehat{M}, and let F0F_{0} be the bundle defined by pulling back the restriction F|MF|_{M} along π\pi. The composition (2.4) is then given by pulling back a vector bundle over MM along π\pi and tensoring with the class [F][F0][F]-[F_{0}]. On the other hand, since M^\widehat{M} is the boundary of the unit sphere bundle of (M×)W(M\times\mathbb{R})\oplus W, and the bundle F0F_{0} is pulled back from MM, the cycle (M^,F0π(E),fπ)(\widehat{M},F_{0}\otimes\pi^{*}(E),f\circ\pi) is bordant to the empty cycle. Thus we have a bordism of cycles

(M^,Fπ(E),fπ)(M^,([F][F0])π(E),fπ),(\widehat{M},F\otimes\pi^{*}(E),f\circ\pi)\sim(\widehat{M},([F]-[F_{0}])\otimes\pi^{*}(E),f\circ\pi),

whence the right-hand side is equal to (M^,s![E],fπ)(\widehat{M},s![E],f\circ\pi) by the description of the composition (2.4) given above. ∎

3. Poincaré duality for geometric KK-homology

In this section we prove Theorem 1.1. For the rest of this section, let XX be a proper GG-spinc manifold with X/GX/G is compact.

We can define the following natural map between KG(X)K^{*}_{G}(X) and KG(X)K_{*}^{G}(X), which can be thought of as cap product with the fundamental KK-homology class on XX. For this, let S1S^{1} be the unit circle in \mathbb{C}, and define the map

c:X\displaystyle c\colon X X×S1\displaystyle\to X\times S^{1}
x\displaystyle x (x,1).\displaystyle\mapsto(x,1). (3.1)
Definition 3.1.

Define ϕ:KG(X)KG(X)\phi\colon K_{G}^{*}(X)\to K_{*}^{G}(X) by

ϕ:KGi(X)\displaystyle\phi\colon K^{i}_{G}(X) Ki+dimXG(X),\displaystyle\to K_{i+\dim X}^{G}(X),
x\displaystyle x {[X,e,id], if i=0,[X×S1,c!(e),pr1], if i=1,\displaystyle\mapsto\begin{cases}[X,e,\textnormal{id}],&\textnormal{ if }i=0,\\ [X\times S^{1},c_{!}(e),\textnormal{pr}_{1}],&\textnormal{ if }i=1,\end{cases}

where pr1:X×S1\textnormal{pr}_{1}\colon X\times S^{1} is the projection onto the first factor, and we have used the notation from Remark 2.9.

We now show that ϕ\phi is an isomorphism by defining explicitly a map ψ\psi that will turn out to be its inverse.

Definition 3.2.

Define the map ψ:KG(X)KG(X)\psi\colon K_{*}^{G}(X)\to K_{G}^{*}(X) by

ψ:Ki+dimXG(X)\displaystyle\psi\colon K_{i+\dim X}^{G}(X) KGi(X),\displaystyle\to K^{i}_{G}(X),
[M,E,f]\displaystyle[M,E,f] f![E],\displaystyle\mapsto f_{!}[E],

for i=0,1i=0,1, where f!f_{!} is the Gysin homomorphism from (2.3).

Remark 3.3.

Note that

f![E]f!(KG0(M))KGdimXdimM(X)=KGdimX(i+dimX)(X)=KGi(X),f_{!}[E]\in f_{!}(K_{G}^{0}(M))\subseteq K_{G}^{\dim X-\dim M}(X)=K^{\dim X-(i+\dim X)}_{G}(X)=K^{i}_{G}(X),

so the degrees make sense.

We first need to show that ψ\psi is well-defined. For this, we will use the following:

Lemma 3.4.

Let WW be a GG-cocompact, GG-spinc manifold-with-boundary, and let XX be a GG-cocompact GG-spinc manifold. Let h:WXh\colon W\to X be a GG-equivariant map, and let i:WWi\colon\partial W\hookrightarrow W be the natural inclusion. Then the composition

KG(W)iKG(W)(h|W)!KG(X)K^{*}_{G}(W)\xrightarrow{i^{*}}K^{*}_{G}(\partial W)\xrightarrow{(h|_{\partial W})_{!}}K^{*}_{G}(X)

is the zero map.

Proof.

We give the proof when dimW=dimX\dim\partial W=\dim X mod 22 and show that

KG0(W)iKG0(W)(h|W)!KG0(X)K^{0}_{G}(W)\xrightarrow{i^{*}}K^{0}_{G}(\partial W)\xrightarrow{(h|_{\partial W})_{!}}K^{0}_{G}(X) (3.2)

is the zero map; the proofs for the other cases are similar.

Let us consider the composition (3.2) upon applying the Thom isomorphism, Theorem 2.6, in the form of (2.2), and use the description of the Gysin map from (2.3).

Let jWj_{W} be a GG-equivariant embedding of WW into 2n\mathbb{R}^{2n} for some nn, where GG is realized as a subgroup of GL(2n,)GL(2n,\mathbb{R}); note that this is possible because GG is assumed to be linear. Let jWj_{\partial W} denote the restriction of jWj_{W} to W\partial W. Let iX:XX×2ni_{X}\colon X\to X\times\mathbb{R}^{2n} be the zero section. Define the embedding

iW:W\displaystyle i_{W}\colon W X×2n,\displaystyle\to X\times\mathbb{R}^{2n},
w\displaystyle w (h(w),jW(w)),\displaystyle\mapsto\left(h(w),j_{W}(w)\right),

and let iWi_{\partial W} be the restriction of iWi_{W} to W\partial W. Let νW\nu_{W} and νW\nu_{\partial W} denote the respective normal bundles of the embeddings iWi_{W} and iWi_{\partial W}. We may identify these normal bundles with GG-invariant tubular neighbourhoods UWU_{W} and UWU_{\partial W} in X×2nX\times\mathbb{R}^{2n}, noting that in general UWU_{W} has boundary. Since the normal bundle of W\partial W in WW is trivial and one-dimensional, there is a natural GG-equivariant identification

UWUW×(ε,ε)U_{\partial W}\cong\partial U_{W}\times(-\varepsilon,\varepsilon) (3.3)

for some ε>0\varepsilon>0. It follows from the two-out-of-three lemma for GG-equivariant spinc-structures (see [12, Section 3.1] and [8, Remark 2.6]), together with the fact that WW and XX are GG-spinc, that νW\nu_{W} and νW\nu_{\partial W} are GG-spinc vector bundles, and hence Theorem 2.6 applies. The resulting Thom isomorphisms for WW, W\partial W, and XX (in the notation of (2.2)) are shown as vertical arrows in the following commutative diagram:

KG0(W){K^{0}_{G}(W)}KG0(W){K^{0}_{G}(\partial W)}KG0(X){K^{0}_{G}(X)}KG1(UW){K^{1}_{G}(U_{W})}KG0(UW){K^{0}_{G}(U_{\partial W})}KG0(X×2n),{K^{0}_{G}(X\times\mathbb{R}^{2n}),}i\scriptstyle{i^{*}}TG\scriptstyle{T_{G}}(h|W)!\scriptstyle{(h|_{\partial W})_{!}}TG\scriptstyle{T_{G}}TG\scriptstyle{T_{G}}TGi\scriptstyle{T_{G}i^{*}}λ\scriptstyle{\lambda} (3.4)

where the map TGiT_{G}i^{*} is determined uniquely by commutativity, and the homomorphism λ\lambda is induced by the extension-by-zero map C0(UW)C0(X×2n)C_{0}(U_{\partial W})\to C_{0}(X\times\mathbb{R}^{2n}) as in Remark 2.5. It thus suffices to show that the composition λTGi\lambda\circ T_{G}i^{*} vanishes.

By [13, Lemma 2.2] or [14, Chapter 5], we have a six-term exact sequence

KG0(UW\UW){K^{0}_{G}(U_{W}\backslash\partial U_{W})}KG0(UW){K^{0}_{G}(U_{W})}KG0(UW){K^{0}_{G}(\partial U_{W})}KG1(UW){K^{1}_{G}(\partial U_{W})}KG1(UW){K^{1}_{G}(U_{W})}KG1(UW\UW),{K^{1}_{G}(U_{W}\backslash\partial U_{W}),}λ\scriptstyle{\lambda}j\scriptstyle{j^{*}}δ\scriptstyle{\delta}δ\scriptstyle{\delta}j\scriptstyle{j^{*}}λ\scriptstyle{\lambda} (3.5)

where jj^{*} is induced by the inclusion j:UWUWj\colon\partial U_{W}\hookrightarrow U_{W} and the maps λ\lambda are again induced by the extension-by-zero map C0(UW\UW)C0(UW)C_{0}(U_{W}\backslash\partial U_{W})\to C_{0}(U_{W}). The identification (3.3) gives a natural isomorphism KG1(UW)KG0(UW)K_{G}^{1}(\partial U_{W})\cong K_{G}^{0}(U_{\partial W}). Using this, the bottom row of (3.4) fits into the following commutative diagram:

KG1(UW){K^{1}_{G}(U_{W})}KG0(UW){K^{0}_{G}(U_{\partial W})}KG0(UX){K^{0}_{G}(U_{X})}KG1(UW){K^{1}_{G}(\partial U_{W})}KG0(UW\UW).{K^{0}_{G}(U_{W}\backslash U_{\partial W}).}TGi\scriptstyle{T_{G}i^{*}}j\scriptstyle{j^{*}}λ\scriptstyle{\lambda}δ\scriptstyle{\delta}\scriptstyle{\cong}λ\scriptstyle{\lambda} (3.6)

It follows from exactness of (3.5) that λTGi=0\lambda\circ T_{G}i^{*}=0, and hence (h|W)!i=0(h|_{\partial W})_{!}\circ i^{*}=0. ∎

Proposition 3.5.

The map ψ\psi is well-defined.

Proof.

That ψ\psi respects disjoint union/direct sum is clear, since for any element of the form [M,E1E2,f]KG(M)[M,E_{1}\oplus E_{2},f]\in K^{G}_{*}(M), we have

ψ[M,E1E2,f]=f![E1E2]=f![E1]+f![E2]KG(M).\psi[M,E_{1}\oplus E_{2},f]=f_{!}[E_{1}\oplus E_{2}]=f_{!}[E_{1}]+f_{!}[E_{2}]\in K^{*}_{G}(M).

Next, let [W,E,h][W,E,h] be an equivariant bordism between two elements [M1,E1,h1][M_{1},E_{1},h_{1}] and [M2,E2,h2][-M_{2},E_{2},h_{2}]. Then Lemma 3.4 applied to W=M1M2\partial W=M_{1}\sqcup-M_{2} implies that

(h1)![E1]=(h2)![E2],(h_{1})_{!}[E_{1}]=(h_{2})_{!}[E_{2}],

hence ψ\psi is well-defined with respect to the bordism relation. To see that ψ\psi is well-defined with respect to vector bundle modification, let (M^,Fπ(E),fπ)(\widehat{M},F\oplus\pi^{*}(E),f\circ\pi) be the modification of a cycle (M,E,f)(M,E,f) for XX by a bundle VV, as in Definition 2.8 (iii). By Lemma 2.10, we have

[M^,Fπ(E),fπ]=[M^,s![E],fπ].[\widehat{M},F\oplus\pi^{*}(E),f\circ\pi]=[\widehat{M},s_{!}[E],f\circ\pi].

Functoriality of the Gysin map with respect to composition, together with the fact that πs=id\pi\circ s=\textnormal{id}, now implies

ψ[M^,s![E],fπ]\displaystyle\psi[\widehat{M},s_{!}[E],f\circ\pi] =(fπ)!s![E]\displaystyle=(f\circ\pi)_{!}s_{!}[E]
=f!(πs)![E]\displaystyle=f_{!}\circ(\pi\circ s)_{!}[E]
=f![E]\displaystyle=f_{!}[E]
=ψ[M,E,f].\displaystyle=\psi[M,E,f].\qed
Proposition 3.6.

The map ϕ\phi is injective.

Proof.

For any eKGi(X)e\in K_{G}^{i}(X), we have

ψϕ(e)\displaystyle\psi\circ\phi(e) ={ψ[X,e,id]=id!(e),if i=0,ψ[X×S1,c!(e),pr1]=(pr1)!(id!(e))=(pr1id)!(e),if i=1,\displaystyle=\begin{cases}\psi[X,e,\textnormal{id}]=\textnormal{id}_{!}(e),&\textnormal{if $i=0$},\\ \psi[X\times S^{1},c_{!}(e),\textnormal{pr}_{1}]=(\textnormal{pr}_{1})_{!}(\textnormal{id}_{!}(e))=(pr_{1}\circ\textnormal{id})_{!}(e),&\textnormal{if $i=1$},\end{cases}

which are both equal to ee, where we have used functoriality of the Gysin map. Hence ψϕ=id\psi\circ\phi=\textnormal{id}, so ϕ\phi is injective. ∎

For surjectivity, we will use the Gysin homomorphism from subsection 2.2, together with the following result, which is a special case of [5, Theorem 4.1] but applied to linear instead of compact GG.

Lemma 3.7.

Let M,N,XM,N,X be three GG-cocompact GG-spinc manifolds, g:NXg\colon N\to X a GG-equivariant continuous map, and f:MNf\colon M\to N a GG-equivariant embedding with even codimension. Then for any GG-vector bundle EME\to M, we have

[M,E,gf]=[N,f![E],g]KG(X).[M,E,g\circ f]=[N,f![E],g]\in K_{*}^{G}(X).
Proof.

The proof of Theorem [5, Theorem 4.1], which was stated for compact Lie groups, goes through with no changes to our setting. ∎

Proposition 3.8.

The map ϕ\phi is surjective.

Proof.

Examining Definitions 3.1 and 3.2, one sees that ϕψ\phi\circ\psi is given, at the level of geometric cycles, by

[M,E,f]f![E]{[X,f![E],id], if dimM=dimXmod2,[X×S1,c!(f![E]),pr1], otherwise,[M,E,f]\mapsto f_{!}[E]\mapsto\begin{cases}[X,f_{!}[E],\textnormal{id}],&\textnormal{ if }\dim M=\dim X\mod{2},\\ [X\times S^{1},c_{!}(f_{!}[E]),\textnormal{pr}_{1}],&\textnormal{ otherwise,}\end{cases}

where the map cc was defined in (3). Let i:M2ni\colon M\to\mathbb{R}^{2n} be a GG-equivariant embedding for some nn, and let j:X2n×Xj\colon X\to\mathbb{R}^{2n}\times X be the zero section. Upon compactifying 2n\mathbb{R}^{2n}, ff factors as

S2n×X{S^{2n}\times X}M{M}X,{X,}pr2\scriptstyle{\textnormal{pr}_{2}}i×f\scriptstyle{i\times f}f\scriptstyle{f} (3.7)

where pr2\textnormal{pr}_{2} is the projection onto the second factor, and jj becomes an embedding XS2n×XX\to S^{2n}\times X.

Suppose first that dimM=dimXmod2\dim M=\dim X\mod{2}. Then it suffices to prove that any geometric cycle of the form (M,E,f)(M,E,f) is equivalent to (X,f![E],id)(X,f_{!}[E],\textnormal{id}). By Lemma 3.7 applied to the embedding i×fi\times f, we have

[M,E,f]=[S2n×X,(i×f)![E],pr2]KG(X).[M,E,f]=[S^{2n}\times X,(i\times f)_{!}[E],\textnormal{pr}_{2}]\in K_{*}^{G}(X). (3.8)

Meanwhile, Lemma 3.7 applied to jj, together with functoriality of the Gysin map, yields

[X,f![E],id]\displaystyle[X,f_{!}[E],\textnormal{id}] =[X,f![E],pr2j]\displaystyle=[X,f_{!}[E],\textnormal{pr}_{2}\circ j]
=[S2n×X,j!(f![E]),pr2]\displaystyle=[S^{2n}\times X,j_{!}(f_{!}[E]),\textnormal{pr}_{2}]
=[S2n×X,(jf)![E],pr2].\displaystyle=[S^{2n}\times X,(j\circ f)_{!}[E],\textnormal{pr}_{2}]. (3.9)

Finally, the maps i×fi\times f and jfj\circ f are GG-homotopic through

F:M×I\displaystyle F\colon M\times I 2n×XS2n×X,\displaystyle\to\mathbb{R}^{2n}\times X\hookrightarrow S^{2n}\times X,
(m,t)\displaystyle(m,t) ((1t)i(m),f(m)).\displaystyle\mapsto((1-t)i(m),f(m)).

Invariance of the Gysin map under GG-homotopy now implies that (3.8) and (3) are equal, and we conclude.

The case of dimMdimX(mod 2)\dim M\neq\dim X\ (\mathrm{mod}\ 2) proceeds analogously as follows. Here, we need to show that any geometric cycle of the form (M,E,f)(M,E,f) is equivalent to (X×S1,c!(f![E]),pr1)(X\times S^{1},c_{!}(f_{!}[E]),\textnormal{pr}_{1}), where cc was defined in (3). Similar to (3.8), we have

[M,E,f]=[S2n×X×S1,(i×(cf))![E],pr2]KG(X),[M,E,f]=[S^{2n}\times X\times S^{1},(i\times(c\circ f))_{!}[E],\textnormal{pr}_{2}]\in K_{*}^{G}(X), (3.10)

Define

j~:X×S1\displaystyle\tilde{j}\colon X\times S^{1} S2n×X×S1,\displaystyle\to S^{2n}\times X\times S^{1},
(x,s)\displaystyle(x,s) (j(x),s).\displaystyle\mapsto(j(x),s).

Then

[X×S1,c!(f![E]),pr1]\displaystyle[X\times S^{1},c_{!}(f_{!}[E]),\textnormal{pr}_{1}] =[X×S1,c!(f![E]),pr2j~!]\displaystyle=[X\times S^{1},c_{!}(f_{!}[E]),\textnormal{pr}_{2}\circ\tilde{j}_{!}]
=[S2n×X×S1,j~!(c!(f![E]),pr2]\displaystyle=[S^{2n}\times X\times S^{1},\tilde{j}_{!}(c_{!}(f_{!}[E]),\textnormal{pr}_{2}]
=[S2n×X,(j~cf)![E],pr2].\displaystyle=[S^{2n}\times X,(\tilde{j}\circ c\circ f)_{!}[E],\textnormal{pr}_{2}]. (3.11)

The maps i×(cf)i\times(c\circ f) and j~cf\tilde{j}\circ c\circ f are GG-homotopic through

F:M×I\displaystyle F\colon M\times I 2n×X×S1S2n×X×S1,\displaystyle\to\mathbb{R}^{2n}\times X\times S^{1}\hookrightarrow S^{2n}\times X\times S^{1},
(m,t)\displaystyle(m,t) ((1t)i(m),f(m),1),\displaystyle\mapsto((1-t)i(m),f(m),1),

and the claim follows from GG-homotopy invariance of the Gysin map. ∎

Propositions 3.8 and equation (3.6) together imply that ϕ:KGi(X)Ki+dimXG(X)\phi\colon K^{i}_{G}(X)\to K^{G}_{i+\dim X}(X) is an isomorphism for i=0,1i=0,1, which establishes Theorem 1.1.

References

  • [1] Paul Baum, Alain Connes, and Nigel Higson. Classifying space for proper actions and KK-theory of group CC^{\ast}-algebras. In CC^{\ast}-algebras: 1943–1993 (San Antonio, TX, 1993), volume 167 of Contemp. Math., pages 240–291. Amer. Math. Soc., Providence, RI, 1994.
  • [2] Paul Baum and Ronald G. Douglas. Index theory, bordism, and KK-homology. In Operator algebras and KK-theory (San Francisco, Calif., 1981), volume 10 of Contemp. Math., pages 1–31. Amer. Math. Soc., Providence, R.I., 1982.
  • [3] Paul Baum, Nigel Higson, and Thomas Schick. On the equivalence of geometric and analytic KK-homology. Pure Appl. Math. Q., 3(1, part 3):1–24, 2007.
  • [4] Paul Baum, Nigel Higson, and Thomas Schick. A geometric description of equivariant KK-homology for proper actions. In Quanta of maths, volume 11 of Clay Math. Proc., pages 1–22. Amer. Math. Soc., Providence, RI, 2010.
  • [5] Paul Baum, Hervé Oyono-Oyono, Thomas Schick, and Michael Walter. Equivariant geometric KK-homology for compact Lie group actions. Abh. Math. Semin. Univ. Hambg., 80(2):149–173, 2010.
  • [6] Heath Emerson and Ralf Meyer. Dualities in equivariant Kasparov theory. New York J. Math., 16:245–313, 2010.
  • [7] Hao Guo, Varghese Mathai, and Hang Wang. Positive scalar curvature and Poincaré duality for proper actions. J. Noncommut. Geom., 13(4):1381–1433, 2019.
  • [8] Peter Hochs and Varghese Mathai. Geometric quantization and families of inner products. Adv. Math., 282:362–426, 2015.
  • [9] G. G. Kasparov. Equivariant KKKK-theory and the Novikov conjecture. Invent. Math., 91(1):147–201, 1988.
  • [10] Wolfgang Lück and Bob Oliver. The completion theorem in KK-theory for proper actions of a discrete group. Topology, 40(3):585–616, 2001.
  • [11] Richard S. Palais. On the existence of slices for actions of non-compact Lie groups. Ann. of Math. (2), 73:295–323, 1961.
  • [12] M. G. Penington and R. J. Plymen. The Dirac operator and the principal series for complex semisimple Lie groups. J. Funct. Anal., 53(3):269–286, 1983.
  • [13] N. Christopher Phillips. Equivariant KK-theory for proper actions and CC^{*}-algebras. In Index theory of elliptic operators, foliations, and operator algebras (New Orleans, LA/Indianapolis, IN, 1986), volume 70 of Contemp. Math., pages 175–204. Amer. Math. Soc., Providence, RI, 1988.
  • [14] N. Christopher Phillips. Equivariant KK-theory for proper actions, volume 178 of Pitman Research Notes in Mathematics Series. Longman Scientific & Technical, Harlow; copublished in the United States with John Wiley & Sons, Inc., New York, 1989.
  • [15] Graeme Segal. Equivariant KK-theory. Inst. Hautes Études Sci. Publ. Math., (34):129–151, 1968.