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

Positive scalar curvature and an equivariant Callias-type index theorem for proper actions

Hao Guo Texas A&M University [email protected] Peter Hochs University of Adelaide [email protected]  and  Varghese Mathai University of Adelaide [email protected]
Abstract.

For a proper action by a locally compact group GG on a manifold MM with a GG-equivariant Spin\operatorname{Spin}-structure, we obtain obstructions to the existence of complete GG-invariant Riemannian metrics with uniformly positive scalar curvature. We focus on the case where M/GM/G is noncompact. The obstructions follow from a Callias-type index theorem, and relate to positive scalar curvature near hypersurfaces in MM. We also deduce some other applications of this index theorem. If GG is a connected Lie group, then the obstructions to positive scalar curvature vanish under a mild assumption on the action. In that case, we generalise a construction by Lawson and Yau to obtain complete GG-invariant Riemannian metrics with uniformly positive scalar curvature, under an equivariant bounded geometry assumption.

1. Introduction

Let GG be a locally compact group, acting properly on a manifold MM. Suppose that MM has a GG-equivariant Spin\operatorname{Spin}-structure. The results in this paper are about the following question.

Question 1.1.

When does MM admit a complete, GG-invariant Riemannian metric with uniformly positive scalar curvature?

We are mainly interested in the case where M/GM/G is noncompact.

The literature on the non-equivariant case of Question 1.1 is too vast to summarise, but important work where MM is noncompact was done by Gromov and Lawson [13]. A more refined perspective on the non-equivariant case is to consider a manifold XX, and let MM be its universal cover, and GG is fundamental group. This allows one to construct obstructions to metrics of positive scalar curvature in terms of GG-equivariant index theory on MM, refining index-theoretic obstructions on XX. If XX is compact, this is the origin of the even more refined Rosenberg index [36, 35, 37], in KOKO-theory of the real maximal group CC^{*}-algebra of GG.

More generally, if GG is a discrete group not necessarily acting freely on MM, then XM/GX\coloneqq M/G is an orbifold, whence Question 1.1 becomes the question of whether XX admits an orbifold metric of positive scalar curvature.

We consider the case where GG is not necessarily discrete, and does not necessarily act freely. Results on this case of Question 1.1 in the case where GG is an almost-connected Lie group and M/GM/G is compact were obtained in [18, 34]. In this paper, we focus on the case where M/GM/G is noncompact.

Results on positive scalar curvature

We first obtain obstructions to GG-invariant Riemannian metrics with positive scalar curvature, both in the KK-theory of the maximal or reduced group CC^{*}-algebra of GG, and in terms of numerical topological invariants generalising the A^\hat{A}-genus. If GG is a connected Lie group, then these obstructions vanish under a mild assumption, as shown in [21]. In that case, we construct GG-invariant Riemannian metrics with uniformly positive scalar curvature, under an equivariant bounded geometry assumption.

Our most general obstruction result is the following.

Theorem 1.2.

Let HMH\subset M be a GG-invariant, cocompact hypersurface with trivial normal bundle, that partitions MM into two open sets. If MM admits a complete, GG-invariant Riemannian metric with nonnegative scalar curvature, and positive scalar curvature in a neighbourhood of HH, then

indexG(DH)=0K(C(G)),\operatorname{index}_{G}(D^{H})=0\quad\in K_{*}(C^{*}(G)),

for a Spin\operatorname{Spin}-Dirac operator DHD^{H} on HH.

See Theorem 2.1. In the case where MM is the universal cover of a manifold XX and GG is its fundamental group, this becomes Theorem A in [9].

Theorem 1.2 implies topological obstructions to GG-invariant Riemannian metrics with positive scalar curvature. Let gGg\in G, and let HgHH^{g}\subset H be the fixed point set of gg. Let 𝒩Hg\mathcal{N}\to H^{g} be its normal bundle in HH, and let R𝒩R^{\mathcal{N}} be the curvature of the Levi–Civita connection restricted to 𝒩\mathcal{N}. The gg-localised A^\hat{A}-genus of HH is

A^g(H)HgcgA^(Hg)det(1geR𝒩/2πi)1/2,\hat{A}_{g}(H)\coloneqq\int_{H^{g}}c^{g}\frac{\hat{A}(H^{g})}{\det(1-ge^{-R^{\mathcal{N}}/2\pi i})^{1/2}},

for a cutoff function cgc^{g} on XgX^{g}. If GG acts freely, then A^g(H)=0\hat{A}_{g}(H)=0 if geg\neq e, and A^e(H)\hat{A}_{e}(H) is the A^\hat{A}-genus of H/GH/G. In general, for example in the orbifold case, A^g(H)\hat{A}_{g}(H) may be nonzero for different gg.

Corollary 1.3.

Consider the setting of Theorem 1.2. Suppose that either

  • GG is any locally compact group and g=eg=e;

  • GG is discrete and finitely generated and gg is any element; or

  • GG is a connected semisimple Lie group and gg is a semisimple element.

Then A^g(H)\hat{A}_{g}(H) is independent of the choice of GG-invariant Riemannian metric, and A^g(H)=0\hat{A}_{g}(H)=0.

See Theorem 2.4.

Theorem 2 in [21] is a generalisation of Atiyah and Hirzebruch’s vanishing result [3] to the noncompact case. It states that, if GG is connected, and not every stabiliser of its action on HH is maximal compact, then indexG(DH)=0\operatorname{index}_{G}(D^{H})=0. This implies that A^g(H)=0\hat{A}_{g}(H)=0 as well. In view of Theorem 1.2 and Corollary 1.3, this makes it a natural question if MM admits a GG-invariant Riemannian metric with positive scalar curvature if GG is a connected Lie group. The answer, given in the present paper, turns out to be yes under a certain equivariant bounded geometry assumption.

Suppose that GG is a connected Lie group, and let K<GK<G be a maximal compact subgroup. Abels’ slice theorem [1] implies that there is a diffeomorphism MG×KNM\cong G\times_{K}N, for a KK-invariant submanifold NMN\subset M. Consider the infinitesimal action map

φ:N×𝔨TN\varphi\colon N\times\mathfrak{k}\to TN

mapping (y,X)N×𝔨(y,X)\in N\times\mathfrak{k} to ddt|t=0exp(tX)y\left.\frac{d}{dt}\right|_{t=0}\exp(tX)y. The action by KK on NN is said to have no shrinking orbits with respect to a KK-invariant Riemannian metric on NN, if the pointwise operator norm of φ\varphi as a map from 𝔨\mathfrak{k} to a tangent space is uniformly positive outside a neighbourhood of the fixed point set NKN^{K}. We say that NN has KK-bounded geometry if it has bounded geometry and no shrinking orbits.

Theorem 1.4.

Suppose that KK is non-abelian, and that KK acts effectively on NN with compact fixed point set. If there exists a KK-invariant Riemannian metric on NN for which NN has KK-bounded geometry, then the manifold G×KNG\times_{K}N admits a GG-invariant metric with uniformly positive scalar curvature.

In the compact case, the Atiyah–Hirzebruch vanishing theorem [3] implies that the obstruction A^(N)\hat{A}(N) to Riemannian metrics of positive scalar curvature vanishes if KK acts nontrivially on NN. Then Lawson and Yau [27] constructed such metrics, under mild conditions. In a similar way, Theorem 1.4 complements the vanishing result in [21].

A Callias-type index theorem

Two effective sources of index-theoretic obstructions to metrics of positive scalar curvature on noncompact manifolds are coarse index theory and Callias-type index theory. For some results involving coarse index theory, see for example [38] and the literature on the coarse Novikov conjecture, in particular [12] for the equivariant setting we are interested in here. We will use Callias-type index theory.

Not assuming that MM is Spin\operatorname{Spin} for now, and letting GG be any locally compact group, we consider a GG-equivariant Dirac-type operator DD on a GG-equivariant vector bundle SMS\to M. A Callias-type operator is of the form D+ΦD+\Phi, for a GG-equivariant endomorphism Φ\Phi of SS such that D+ΦD+\Phi is uniformly positive outside a cocompact set. Then this operator has an index indexG(D+Φ)K(C(G))\operatorname{index}_{G}(D+\Phi)\in K_{*}(C^{*}(G)), constructed in [15]. (See Theorem 4.2 in [17] for a realisation of this index in terms of coarse geometry.)

Theorem 1.5 (GG-Callias-type index theorem).

We have

indexG(D+Φ)=indexG(DN),\operatorname{index}_{G}(D+\Phi)=\operatorname{index}_{G}(D_{N}),

for a Dirac operator DND_{N} on a GG-invariant, cocompact hypersurface NMN\subset M.

See Theorem 3.4. Versions of this result where GG is trivial were proved in [2, 5, 7, 8, 25]. Versions for operators on bundles of modules over operator algebras were proved in [6, 9]. Parts of our proof of Theorem 1.5 are based on a similar strategy as the proof of the index theorem in [9].

We deduce Theorem 2.1 from Theorem 1.5. This approach is an equivariant generalisation of the obstructions to metrics of positive scalar obtained in [2, 7, 9].

If gGg\in G, then under conditions, there is a subalgebra A(G)C(G)A(G)\subset C^{*}(G) such that K(A(G))=K(C(G))K_{*}(A(G))=K_{*}(C^{*}(G)), and there is a well-defined trace τg\tau_{g} on A(G)A(G) given by

τg(f)=G/Zf(hgh1)d(hZ),\tau_{g}(f)=\int_{G/Z}f(hgh^{-1})d(hZ),

where ZZ is the centraliser of gg. In various settings, including the three cases in Corollary 1.3, there are index formulas for the number τg(indexG(D))\tau_{g}(\operatorname{index}_{G}(D)), see [24, 41, 40]. These index formulas imply that, in the setting of Theorem 1.2,

τg(indexG(DH))=A^g(H).\tau_{g}(\operatorname{index}_{G}(D^{H}))=\hat{A}_{g}(H).

Hence Theorem 1.2 implies Corollary 1.3.

Theorem 1.4 is proved via a generalisation of Lawson and Yau’s arguments [27], together with a result from [17] that allows one to induce up metrics of positive scalar curvature from NN to M=G×KNM=G\times_{K}N.

Apart from using Theorem 1.5 to prove Theorem 2.1, we obtain some further applications, on the image (Corollary 7.1) and cobordism invariance (Corollary 7.2) of the analytic assembly map; on the Callias-type index of Spinc\operatorname{Spin}^{c}-Dirac operators (Corollary 7.3); on induction of Callias-type indices from compact groups to noncompact groups (Corollary 7.4); and on the Spinc\operatorname{Spin}^{c}-version [31] of the quantisation commutes with reduction problem [14, 28, 30, 39] for Spinc\operatorname{Spin}^{c}-Callias type operators (Corollary 7.5).

Outline of this paper

We state our obstruction and existence results in Section 2: Theorem 2.1, Corollary 2.4 and Theorem 2.9. In Section 3, we state the equivariant Callias-type index theorem, Theorem 3.4. Theorem 3.4 is proved in Sections 4 and 5. We then deduce Theorem 2.1 and Corollary 2.4 in Subsection 6.1. Theorem 2.9 is proved in Subsection 6.2. In Section 7, we obtain some further applications of the Callias-type index theorem, Corollaries 7.17.5.

Acknowledgements

HG was supported in part by funding from the National Science Foundation under grant no. 1564398. PH thanks Guoliang Yu and Texas A&M University for their hospitality during a research visit. VM was supported by funding from the Australian Research Council, through the Australian Laureate Fellowship FL170100020.

Notation and conventions

All manifolds, vector bundles, group actions and other maps between manifolds are implicitly assumed to be smooth. If a Hilbert space HH is mentioned, the inner product on that space will be denoted by (-,-)H(\relbar,\relbar)_{H}, and the corresponding norm by H\|\cdot\|_{H}. Spaces of continuous sections are denoted by Γ\Gamma; spaces of smooth sections by Γ\Gamma^{\infty}. Subscripts cc denote compact supports.

If GG is a group, and HH is a subgroup acting on a set SS, then we write G×HSG\times_{H}S for the quotient of G×SG\times S by the HH-action given by

h(g,s)=(gh1,hs),h\cdot(g,s)=(gh^{-1},hs),

for hGh\in G, gGg\in G and sSs\in S. If SS is a manifold, GG is Lie group and HH is a closed subgroup, then this action is proper and free, and G×HSG\times_{H}S is a manifold.

A continuous group action, and also the space acted on, is said to be cocompact if the quotient space is compact.

2. Results on positive scalar curvature

In all parts of this paper except Subsections 2.2 and 6.2, which concern the existence results, GG will be be a locally group, acting properly on a manifold MM. We do not assume that M/GM/G is compact and are in fact interested mainly in the case where it is not. The group GG may have infinitely many connected components, and for may for example be an infinite discrete group.

2.1. Obstructions

For a proper, cocompact action by GG on a manifold NN, a GG-equivariant elliptic differential operator DD on NN has an equivariant index indexG(D)K(C(G))\operatorname{index}_{G}(D)\in K_{*}(C^{*}(G)) defined by the analytic assembly map [4]. Here C(G)C^{*}(G) is the maximal or reduced group CC^{*}-algebra of GG, and the index takes values in its even KK-theory if DD is odd with respect to a grading, and in odd KK-theory otherwise.

Our most general obstruction result is the following.

Theorem 2.1.

Let MM be a complete Riemannian Spin\operatorname{Spin}-manifold, on which a locally compact group GG acts properly and isometrically. Let HMH\subset M be a GG-invariant, cocompact hypersurface with trivial normal bundle, such that MH=XYM\setminus H=X\cup Y for disjoint open subsets X,YMX,Y\subset M. If the scalar curvature on MM is nonnegative, and positive in a neighbourhood of HH, then the Spin\operatorname{Spin}-Dirac operator DHD^{H} on HH, acting on sections of the restriction of the spinor bundle on MM to HH, satisfies

indexG(DH)=0K(C(G)).\operatorname{index}_{G}(D^{H})=0\quad\in K_{*}(C^{*}(G)).

We will deduce this result from an equivariant index theorem for Callias-type operators, Theorem 3.4, which may be of independent interest and has some other applications as well.

Remark 2.2.

If MM is the universal cover of a manifold XX, and GG is the fundamental group of XX acting on MM in the usual way, then Theorem 2.1 reduces to Theorem A in [9], by the Miščenko–Fomenko realisation of the analytic assembly map in that case [29].

Theorem 2.1 implies a set of topological obstructions to GG-invariant positive scalar curvature metrics on MM. Let XX be any Riemannian manifold on which GG acts properly, isometrically and cocompactly. Let gGg\in G, and let XgXX^{g}\subset X be its fixed point set. (Properness of the action implies that Xg=X^{g}=\emptyset if gg is not contained in a compact subset of GG.) Let 𝒩Xg\mathcal{N}\to X^{g} be the normal bundle to XgX^{g} in XX. The connected components of XgX^{g} are submanifolds of XX of possibly different dimensions, so the rank of 𝒩\mathcal{N} may jump between these components. In what follows, we implicitly apply all constructions to the connected components of XgX^{g} and add the results together.

By a cutoff function we will mean a smooth function c:M[0,1]{c}\colon M\rightarrow[0,1] such that supp(c)\operatorname{supp}({c}) has compact intersection with each GG-orbit, and for each xMx\in M we have

Gc(g1x)𝑑g=1.\int_{G}{c}(g^{-1}x)\,dg=1.

We will also use cutoff functions for other group actions, which are defined analogously.

Let R𝒩R^{\mathcal{N}} be the curvature of the Levi–Civita connection restricted to 𝒩\mathcal{N}. Let A^(Xg)\hat{A}(X^{g}) be the A^\hat{A}-class of XgX^{g}. Let ZG(g)<GZ_{G}(g)<G be the centraliser of gg. Let cgc^{g} be a cutoff function for the action by ZG(g)Z_{G}(g) on XgX^{g}.

Definition 2.3.

The gg-localised A^\hat{A}-genus of XX is

A^g(X)XgcgA^(Xg)det(1geR𝒩/2πi)1/2.\hat{A}_{g}(X)\coloneqq\int_{X^{g}}c^{g}\frac{\hat{A}(X^{g})}{\det(1-ge^{-R^{\mathcal{N}}/2\pi i})^{1/2}}.

If g=eg=e, then

A^e(X)=XceA^(X)\hat{A}_{e}(X)=\int_{X}c^{e}\hat{A}(X)

is the L2L^{2}-A^\hat{A} genus of XX used in [41]. If GG is discrete and acts properly and freely on XX, then A^e(X)=A^(X/G)\hat{A}_{e}(X)=\hat{A}(X/G).

Corollary 2.4.

Suppose that either

  • GG is any locally compact group and g=eg=e;

  • GG is discrete and finitely generated and gg is any element;

  • GG is a connected semisimple Lie group and gg is a semisimple element.

Let MM be a manifold on which GG acts properly and that admits a GG-equivariant Spin\operatorname{Spin}-structure. Let HMH\subset M be a GG-invariant, cocompact hypersurface such that MH=XYM\setminus H=X\cup Y for disjoint open subsets X,YMX,Y\subset M. The localised A^\hat{A}-genus A^g(H)\hat{A}_{g}(H) is independent of the choice of a Riemannian metric. If MM admits a complete, GG-invariant Riemannian metric whose scalar curvature is nonnegative, and positive in a neighbourhood of HH, then A^g(H)=0\hat{A}_{g}(H)=0.

Theorem 2 in [21] is a generalisation of Atiyah and Hirzebruch’s vanishing theorem [3] to actions by noncompact groups. It states that if GG is a connected Lie group, and not all stabilisers of the action by GG on HH are maximal compact, then indexG(DH)=0\operatorname{index}_{G}(D^{H})=0. So in this setting, the obstructions in Theorem 2.1 and Corollary 2.4 vanish. This makes it a natural question whether Riemannian metrics as in Theorem 2.1 exist if GG is a connected Lie group. A partial affirmative answer to that question is given in Subsection 2.2.

This also means that the natural place to look for examples and applications where Theorem 2.1 and Corollary 2.4 yield nontrivial obstructions is the setting where GG has infinitely many connected components. (The vanishing result generalises directly to the case where GG has finitely many connected components.) As noted in Remark 2.2, Theorem 2.1 implies Theorem A in [9], in the case where GG is the fundamental group of M/GM/G and MM is its universal cover. More generally, if GG is discrete, then M/GM/G is an orbifold. Then Theorem 2.1 and Corollary 2.4 lead to obstructions to orbifold metrics on M/GM/G with nonnegative scalar curvature, and positive scalar curvature near the sub-orbifold N/GN/G. If GG acts freely, then A^g(H)\hat{A}_{g}(H) is zero if geg\not=e (and A^e(H)=A^(H/G)\hat{A}_{e}(H)=\hat{A}(H/G)), but in the orbifold case the localised A^\hat{A}-genera for nontrivial elements gg are additional obstructions to positive scalar curvature.

2.2. Existence for connected Lie groups

In this subsection, we suppose that GG is a connected Lie group. As pointed out at the end of the previous subsection, for such groups the obstructions to GG-invariant Riemannian metrics with positive scalar curvature in Theorem 2.1 and Corollary 2.4 vanish under a mild assumption on the action, so it is natural to investigate existence of such metrics.

If GG is connected, Abels’ global slice theorem [1] implies we have a diffeomorphism MG×KNM\cong G\times_{K}N, for a KK-invariant submanifold NMN\subset M. Our existence result, Theorem 2.9, supposes that such a slice NN has KK-bounded geometry, a notion introduced in Definition 2.8.

Suppose a compact, connected Lie group KK acts isometrically on a complete Riemannian manifold (N,gN)(N,g_{N}). Let bb denote a bi-invariant Riemannian metric on KK. For each yNy\in N we have a linear map

(2.1) φy:𝔨TyN\varphi_{y}\colon\mathfrak{k}\rightarrow T_{y}N

defined by φy(X):=ddt|t=0exp(tX)y\varphi_{y}(X):=\left.\frac{d}{dt}\right|_{t=0}\exp(tX)y, for X𝔨X\in\mathfrak{k}. Define a pointwise norm function

(2.2) φ:N,yφy,\|\varphi\|\colon N\rightarrow\mathbb{R},\qquad y\mapsto\|\varphi_{y}\|,

where \|\,\cdot\| denotes the linear operator norm with respect to bb and gNg_{N}.

Definition 2.5.

We say that the action of KK on NN has no shrinking orbits if, for any neighbourhood UU of the fixed point set NKN^{K}, there exists a constant CU>0C_{U}>0 such that for all yNUy\in N\setminus U we have

φ(y)CU,\|{\varphi}(y)\|\geq C_{U},

where the norm function is taken with respect to the Riemannian metric gg.

We remark that the condition of the action having no shrinking orbits is independent of bb.

Example 2.6.

Suppose that N=2N=\mathbb{R}^{2}, on which K=SO(2)K=\operatorname{SO}(2) acts in the natural way. Let ψC(2)\psi\in C^{\infty}(\mathbb{R}^{2}) be positive and rotation-invariant, and consider the Riemannian metric on 2\mathbb{R}^{2} equal to ψ2\psi^{2} times the Euclidean metric. Then for all y2y\in\mathbb{R}^{2} and X𝔰𝔬(2)X\in\mathbb{R}\cong\mathfrak{so}(2),

φy(X)=X(0110)y.\varphi_{y}(X)=X\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}y.

So φ(y)=ψ(y)y,\|\varphi\|(y)=\psi(y)\|y\|, where y\|y\| is the Euclidean norm of yy. Hence the action has no shrinking orbits if and only if the function yψ(y)yy\mapsto\psi(y)\|y\| has a positive lower bound outside a neighbourhood of (2)SO(2)={0}(\mathbb{R}^{2})^{\operatorname{SO}(2)}=\{0\}.

Now let us define the notion of KK-bounded geometry, which is a strengthening of the standard notion of bounded geometry.

Definition 2.7.

A Riemannian manifold has bounded geometry if

  • its injectivity radius is positive;

  • for each l0l\geq 0 there exists Cl>0C_{l}>0 such that lRCl\|{\nabla^{l}R}\|_{\infty}\leq C_{l}, where RR is the Riemann curvature tensor.

Definition 2.8.

The action of KK on NN is said to have KK-bounded geometry if it has no shrinking orbits and NN has bounded geometry.

Our main existence result, proved in Subsection 6.2, is the following.

Theorem 2.9.

Let GG be a connected Lie group, and K<GK<G a maximal compact subgroup with non-abelian identity component. Let NN be a manifold admitting an effective action by KK with compact fixed point set. If there exists a Riemannian metric on NN such that the KK-action has KK-bounded geometry, then the manifold G×KNG\times_{K}N admits a GG-invariant metric with uniformly positive scalar curvature.

This result may be viewed as a strengthening of the vanishing of the obstructions to GG-invariant metrics of positive scalar curvature in Theorem 2.1 and Corollary 2.4 in the case of connected Lie groups, by the result in [21], in the same way that Lawson and Yau’s [27] construction of metrics of positive scalar curvature strengthens the vanishing of the A^\hat{A}-genus as in Atiyah and Hirzebruch’s vanishing theorem [3] in the compact case. (See the diagram on page 233 of [27].)

Remark 2.10.

By Abels’ slice theorem [1], every manifold with a proper action by a connected Lie group GG is of the form G×KNG\times_{K}N in Theorem 2.9. The condition that NKN^{K} is compact is equivalent to the condition that the points in G×KNG\times_{K}N whose stabilisers in GG are maximal compact form a cocompact set.

3. An index theorem

We will deduce Theorem 2.1, and hence Corollary 2.4, from an equivariant index theorem for Callias-type operators, Theorem 3.4. This is based on equivariant index theory for such operators with respect to proper actions, developed in [15]. The proof of the index theorem involves several arguments analogous to those in [9].

3.1. The GG-Callias-type index

From now on, MM will be a complete Riemannian manifold on which GG acts properly and isometrically. Let SMS\to M denote a /2\mathbb{Z}/2-graded, GG-equivariant Clifford module over MM, and DD an odd-graded Dirac operator on Γ(S)\Gamma^{\infty}(S), associated to a GG-invariant Clifford connection on SS via the Clifford action by TMTM on SS.

Let Φ\Phi be an odd, GG-equivariant, fibrewise Hermitian vector bundle endomorphism of SS.

Definition 3.1.

The endomorphism Φ\Phi is admissible for DD if

  • the operator DΦ+ΦDD\Phi+\Phi D on Γ(S)\Gamma^{\infty}(S) is a vector bundle endomorphism; and

  • there are a cocompact subset ZMZ\subset M and a constant C>0C>0 such that we have the pointwise estimate

    (3.1) Φ2DΦ+ΦD+C\Phi^{2}\geq\|D\Phi+\Phi D\|+C

    on MZM\setminus Z.

In this setting the operator D+ΦD+\Phi is called a GG-Callias-type operator.

In the rest of the paper, we will use the following notation. Let μ\mu denote the modular function on GG. Let C(G)C^{*}(G) be either the reduced or maximal group CC^{*}-algebra of GG. Let GG act on sections of the bundle SS by

(gs)(x)g(s(g1x)),(gs)(x)\coloneqq g(s(g^{-1}x)),

for ss a section, gGg\in G and xMx\in M.

Equip the space Γc(S)\Gamma_{c}(S) with a right Cc(G)C_{c}(G)-action defined by

(3.2) (s1b)(x)G(gs)(x)b(g1)μ(g)1/2𝑑g(s_{1}\cdot b)(x)\coloneqq\int_{G}(gs)(x)\cdot b(g^{-1})\mu(g)^{-1/2}\,dg

and a Cc(G)C_{c}(G)-valued inner product defined by

(3.3) (s1,s2)C(G)(g)μ(g)1/2(s1,gs2)L2(S),(s_{1},s_{2})_{C^{*}(G)}(g)\coloneqq\mu(g)^{-1/2}(s_{1},gs_{2})_{L^{2}(S)},

for s1,s2Γc(S)s_{1},s_{2}\in\Gamma^{\infty}_{c}(S), bCc(G)b\in C_{c}(G), gGg\in G, and xMx\in M. Let \mathcal{E} be the Hilbert C(G)C^{*}(G)-module completion of Γc(S)\Gamma^{\infty}_{c}(S) with respect to this structure.

The definition of the equivariant index of GG-Callias-type operators is based on the following result, Theorem 4.19 in [15].

Theorem 3.2.

There is a continuous GG-invariant cocompactly supported function ff on MM such that

(3.4) F(D+Φ)((D+Φ)2+f)1/2,F\coloneqq(D+\Phi)\bigl{(}(D+\Phi)^{2}+f\bigr{)}^{-1/2},

is a well-defined, adjointable operator on \mathcal{E}, such that (,F)(\mathcal{E},F) is a Kasparov (,C(G))(\mathbb{C},C^{*}(G))-cycle. Its class in KK(,C(G))K\!K(\mathbb{C},C^{*}(G)) is independent of the function ff chosen.

For details about the definition of the operator FF, we refer to Definition 4.11 in [15].

Definition 3.3.

The GG-index of the GG-Callias-type operator D+ΦD+\Phi is the class

indexG(D+Φ)[,F]K0(C(G))=KK(,C(G))\operatorname{index}_{G}(D+\Phi)\coloneqq[\mathcal{E},F]\in K_{0}(C^{*}(G))=K\!K(\mathbb{C},C^{*}(G))

as in Theorem 3.2.

3.2. Hypersurfaces and the index theorem

In the setting of the previous subsection, we now suppose that S=S0S0S=S_{0}\oplus S_{0} for an ungraded, GG-equivariant Clifford module S0S_{0} over MM, where the first copy of S0S_{0} is the even part of SS, and the second copy is the odd part. Suppose that

(3.5) D=(0D0D00)D=\begin{pmatrix}0&D_{0}\\ D_{0}&0\end{pmatrix}

for a Dirac operator D0D_{0} on S0S_{0}, and that

(3.6) Φ=(0iΦ0iΦ00)\Phi=\begin{pmatrix}0&i\Phi_{0}\\ -i\Phi_{0}&0\end{pmatrix}

for a Hermitian endomorphism Φ0\Phi_{0} of S0S_{0}. (The conditions on DΦ+ΦDD\Phi+\Phi D then become conditions on [D0,Φ0][D_{0},\Phi_{0}].)

Let ZZ be as in Definition 3.1. Let MMM_{-}\subset M be a GG-invariant, cocompact subset containing ZZ in its interior, such that NMN\coloneqq\partial M_{-} is a smooth submanifold of MM. Let M+M_{+} be the closure of the complement of MM_{-}, so that N=MM+N=M_{-}\cap M_{+} and M=MM+M=M_{-}\cup M_{+}. In this and similar settings, we write

M=MNM+.M=M_{-}\cup_{N}M_{+}.

By (3.1), the restriction of Φ0\Phi_{0} to NN is fibrewise invertible. Let S+NNS^{N}_{+}\to N and SNNS^{N}_{-}\to N be its positive and negative eigenbundles. (These are vector bundles, even though eigenbundles for single eigenvalues may not be.) Clifford multiplication by the unit normal vector field n^\hat{n} to NN pointing into M+M_{+}, times i-i, defines GG-invariant gradings on S±NS^{N}_{\pm}.

Let S0\nabla^{S_{0}} be the Clifford connection on S0S_{0} used to define D0D_{0}. By restriction and projection, it defines connections S±N\nabla^{S^{N}_{\pm}} on S±NS^{N}_{\pm}. The Clifford action by TM|NTM|_{N} on S0|NS_{0}|_{N} preserves S±NS^{N}_{\pm} by the first condition in Definition 3.1; see also Remark 1.2 in [2]. Hence the connections S±N\nabla^{S^{N}_{\pm}} define Dirac operators DS±ND^{S^{N}_{\pm}} on Γ(S±N)\Gamma^{\infty}(S^{N}_{\pm}). These operators are odd-graded. Because NN is cocompact, DS+ND^{S^{N}_{+}} has an equivariant index

indexG(DS+N)K0(C(G))\operatorname{index}_{G}(D^{S^{N}_{+}})\in K_{0}(C^{*}(G))

defined by the analytic assembly map [4].

Theorem 3.4 (GG-Callias-type index theorem).

We have

(3.7) indexG(D+Φ)=indexG(DS+N)K0(C(G)).\operatorname{index}_{G}(D+\Phi)=\operatorname{index}_{G}(D^{S^{N}_{+}})\quad\in K_{0}(C^{*}(G)).

Versions of this result where GG is trivial were proved in [2, 5, 7, 8, 25]. Versions for operators on bundles of modules over operator algebras are proved in [6, 9].

There are various index theorems for the the image of the right hand side of (3.7) under traces [24, 41, 40] or pairings with higher cyclic cocycles [23, 33, 34]. Via these results, Theorem 3.4 yields topological expressions for the corresponding images of the left hand side of (3.7). The results in [24, 41, 40] will be used to deduce Corollary 2.4 from Theorem 2.1.

4. Properties of the GG-Callias-type index

To prove Theorem 3.4, we will make use of the properties of the index of Definition 3.3 that we describe below.

4.1. Sobolev modules

We start by recalling the definition of Sobolev Hilbert C(G)C^{*}(G)-modules from [15]. Let MM, GG, SS and DD be as in Subsection 3.1.

Definition 4.1.

For each nonnegative integer jj, define Γc,j(S)\Gamma_{c}^{\infty,j}(S) to be the pre-Hilbert Cc(G)C_{c}(G)-module whose underlying vector space is Γc(S)\Gamma_{c}^{\infty}(S), equipped with the right Cc(G)C_{c}(G)-action defined by (3.2), and Cc(G)C_{c}(G)-valued inner product defined by

e1,e2j=k=0j(Dke1,Dke2)C(G),\langle e_{1},e_{2}\rangle_{\mathcal{E}^{j}}=\sum_{k=0}^{j}(D^{k}e_{1},D^{k}e_{2})_{C^{*}(G)},

where e1,e2Γc(S)e_{1},e_{2}\in\Gamma_{c}^{\infty}(S) and (-,-)C(G)(\relbar,\relbar)_{C^{*}(G)} is as in (3.3). Here we set D0D^{0} equal to the identity operator. Denote by j(S)=j\mathcal{E}^{j}(S)=\mathcal{E}^{j} the vector space completion of Γc,j(S)\Gamma_{c}^{\infty,j}(S) with respect to the norm induced by -,-j\langle\relbar,\relbar\rangle_{\mathcal{E}^{j}}, and extend naturally the Cc(G)C_{c}^{\infty}(G)-action to a C(G)C^{*}(G)-action, and -,-j\langle\relbar,\relbar\rangle_{\mathcal{E}^{j}} to a C(G)C^{*}(G)-valued inner product on j\mathcal{E}^{j}, to give it the structure of a Hilbert C(G)C^{*}(G)-module. We call j\mathcal{E}^{j} the jj-th GG-Sobolev module with respect to DD.

The module \mathcal{E} defined above Theorem 3.2 equals 0\mathcal{E}^{0}. The following version of the Rellich lemma holds for Sobolev modules (Theorem 3.12 in [15]).

Theorem 4.2.

Let ff be a continuous GG-invariant cocompactly supported function on MM. Then multiplication by ff defines an element of 𝒦(s,t)\mathcal{K}(\mathcal{E}^{s},\mathcal{E}^{t}) whenever s>ts>t.

We will state and prove a homotopy invariance result, Proposition 4.9, for the index in Definition 3.3, that will be of use later. A hypothesis in this result is that a certain vector bundle endomorphism defines adjointable operators on the Sobolev modules 0\mathcal{E}^{0} and 1\mathcal{E}^{1}. In order to check this condition in some geometric situations relevant to us, we will need Propositions 4.3 and 4.4 below.

Proposition 4.3.

A smooth, GG-invariant, uniformly bounded bundle endomorphism of SS defines an element of (0)\mathcal{L}(\mathcal{E}^{0}).

Proof.

Let Ψ\Psi be a smooth, GG-invariant, uniformly bounded bundle endomorphism of SS. Since Ψ\Psi is uniformly bounded, it defines a bounded operator on L2(S)L^{2}(S). Let Ψ\|\Psi\| denote its operator norm, let c{c} be a cutoff function on MM, and let Ψ\Psi^{*} be the pointwise adjoint of Ψ\Psi. Since the operator Ψ1ΨΨΨ2\Psi_{1}\coloneqq\Psi^{*}\Psi-\|\Psi\|^{2} is positive on L2(S)L^{2}(S), it has a positive square root QQ that one observes is GG-invariant. For a fixed eΓc(S)e\in\Gamma^{\infty}_{c}(S), the function

g(cΨ1(ge),ge)L2(S)=(cQ(ge),cQ(ge))L2(S)g\mapsto({c}\Psi_{1}(ge),ge)_{L^{2}(S)}=(\sqrt{{c}}Q(ge),\sqrt{{c}}Q(ge))_{L^{2}(S)}

has compact support in GG, by properness of the GG-action. Thus the map GL2(S)G\to L^{2}(S) defined by gcQ(ge)g\mapsto\sqrt{{c}}Q(ge) has compact support in GG.

It follows that for any unitary representation π\pi of GG on a Hilbert space HH and hHh\in H,

vGμ(g)1/2cQ(ge)π(g)h𝑑gv\coloneqq\int_{G}\mu(g)^{-1/2}\sqrt{{c}}Q(ge)\otimes\pi(g)h\,dg

is a well-defined vector in L2(S)HL^{2}(S)\otimes H. By computations similar to those in the proof of Proposition 5.4 in [15], one sees that vL2(S)H\|v\|_{L^{2}(S)\otimes H} equals

GGgcΨ1g1e,e0(g)𝑑g(π(g)h,h)H𝑑g.\int_{G}\int_{G}\left\langle gc\Psi_{1}g^{-1}e,e\right\rangle_{\mathcal{E}^{0}}(g^{\prime})\,dg\cdot(\pi(g^{\prime})h,h)_{H}\,dg^{\prime}.

Thus, for any unitary representation π\pi of GG,

π(GgcΨ1g1e,e0𝑑g)=π(Ψ1e,e0)\pi\left(\int_{G}\left\langle gc\Psi_{1}g^{-1}e,e\right\rangle_{\mathcal{E}^{0}}\,dg\right)=\pi(\langle\Psi_{1}e,e\rangle_{\mathcal{E}^{0}})

is a positive operator, where we let fCc(G)f\in C_{c}(G) act on HH by

π(f)hGf(g)π(g)h𝑑g.\pi(f)h\coloneqq\int_{G}f(g)\pi(g)h\,dg.

It follows that the element

Ψ1e,e0=(ΨΨΨ2)e,e0=Ψe,Ψe0Ψ2e,e0.\langle\Psi_{1}e,e\rangle_{\mathcal{E}^{0}}=\left\langle\left(\Psi^{*}\Psi-\|\Psi\|^{2}\right)e,e\right\rangle_{\mathcal{E}^{0}}=\langle\Psi e,\Psi e\rangle_{\mathcal{E}^{0}}-\|\Psi\|^{2}\langle e,e\rangle_{\mathcal{E}^{0}}.

is positive in C(G)C^{*}(G). Hence Ψ\Psi extends to an operator on all of 0\mathcal{E}^{0}. Similarly, Ψ\Psi^{*} defines an operator on all of 0\mathcal{E}^{0} that one checks is the adjoint of Ψ\Psi. ∎

Proposition 4.4.

Suppose that there are a GG-invariant, cocompact subset KMK\subset M and a GG-invariant, cocompact hypersurface NMN\subset M such that there is a GG-equivariant isometry MKN×(0,)M\setminus K\cong N\times(0,\infty), and a GG-equivariant vector bundle isomorphism S|MKS|N×(0,)S|_{M\setminus K}\cong S|_{N}\times(0,\infty). Let Ψ\Psi be a GG-equivariant vector bundle endomorphism of SS. Suppose that, on MKM\setminus K, Ψ\Psi and DD are constant in the factor (0,)(0,\infty) of MKN×(0,)M\setminus K\cong N\times(0,\infty). Then Ψ\Psi defines an element of (1)\mathcal{L}(\mathcal{E}^{1}).

The proof uses the next lemma. To state it, let H1(S)H^{1}(S) be the completion of Γc(S)\Gamma_{c}^{\infty}(S) with respect to the inner product

(-,-)H1(S)=(-,-)L2(S)+(D-,D-)L2(S).(\relbar,\relbar)_{H^{1}(S)}=(\relbar,\relbar)_{L^{2}(S)}+(D\relbar,D\relbar)_{L^{2}(S)}.
Lemma 4.5.

Let MM and SS be as in Proposition 4.4. Let Θ\Theta be a bounded, positive operator on H1(S)H^{1}(S) such that

  • Θ\Theta preserves the subspace Γc(S)\Gamma_{c}^{\infty}(S);

  • for any eΓc(S)e\in\Gamma_{c}^{\infty}(S), the function a:Ga\colon G\rightarrow\mathbb{\mathbb{R}} given by g(Θ(ge),ge)H1(S)g\mapsto(\Theta(ge),ge)_{H^{1}(S)} has compact support in GG.

Then

GgΘg1e,e1𝑑g\int_{G}\left\langle g\Theta g^{-1}e,e\right\rangle_{\mathcal{E}^{1}}\,dg

is a positive element of C(G)C^{*}(G).

Proof.

Let QQ be the positive square root of Θ\Theta in (H1(S)){\mathcal{B}}(H^{1}(S)). Since aa has compact support, and (Θ(ge),ge)H1(S)=(Q(ge),Q(ge))H1(S)(\Theta(ge),ge)_{H^{1}(S)}=(Q(ge),Q(ge))_{H^{1}(S)}, the map GH1(S)G\to H^{1}(S) defined by gQ(ge)g\mapsto Q(ge) has compact support in GG. As in the proof of Proposition 4.3, one finds that for any unitary representation π\pi of GG on a Hilbert space HH and hHh\in H,

vGμ(g)1/2Q(ge)π(g)h𝑑gv\coloneqq\int_{G}\mu(g)^{-1/2}Q(ge)\otimes\pi(g)h\,dg

is a well-defined vector in H1(S)HH^{1}(S)\otimes H, and that

GGgQ2g1e,e1(g)𝑑g(π(g)h,h)H𝑑g=vH1(S)H0.\int_{G}\int_{G}\left\langle gQ^{2}g^{-1}e,e\right\rangle_{\mathcal{E}^{1}}(g^{\prime})\,dg\cdot(\pi(g^{\prime})h,h)_{H}\,dg^{\prime}=\|v\|_{H^{1}(S)\otimes H}\geq 0.

Similarly to the proof of Proposition 4.3, we deduce that GgΘg1e,e1𝑑g\int_{G}\langle g\Theta g^{-1}e,e\rangle_{\mathcal{E}^{1}}\,dg is a positive element of C(G)C^{*}(G). ∎

Proof of Proposition 4.4.

Because of the forms of MM and SS, there is a canonical (up to equivalence) first Sobolev norm 1\|\cdot\|_{1} on sections of SS that is GG-invariant, and invariant under the relevant class of translations in the factor (0,)(0,\infty) of N×(0,)N\times(0,\infty). Because Ψ\Psi is an order zero differential operator constant on the factor (0,)(0,\infty), it defines a bounded operator with respect to 1\|\cdot\|_{1}. Due to the form of DD, the norm on H1(S)H^{1}(S) is equivalent to 1\|\cdot\|_{1}, and so Ψ\Psi defines a bounded operator on H1(S)H^{1}(S).

Let πN:MKN×(0,)N\pi_{N}\colon M\setminus K\cong N\times(0,\infty)\rightarrow N be the natural projection. Let c{c} be a cutoff function on MM such that

c|MK=πNcN{c}|_{M\setminus K}=\pi_{N}^{*}{c}_{N}

for a cutoff function cN{c}_{N} on NN. Let Ψ\Psi^{*} and c{c}^{*} denote the respective adjoints of Ψ\Psi and c{c} in (H1(S)){\mathcal{B}}(H^{1}(S)). Then the operator

Ψ1cΨΨ+ΨΨc2\Psi_{1}\coloneqq\frac{{c}\Psi^{*}\Psi+\Psi^{*}\Psi{c}^{*}}{2}

is bounded and self-adjoint on H1(S)H^{1}(S) with norm at most Ψ2c\|\Psi\|^{2}\|{c}\|, where the norms are taken in (H1(S)){\mathcal{B}}(H^{1}(S)).

Let c{c}^{\prime} be a smooth, nonnegative function on MM that is identically 11 on the support of cc, and such that

c|MK=πNcN{c}^{\prime}|_{M\setminus K}=\pi_{N}^{*}{c}^{\prime}_{N}

for a compactly supported function cN{c}^{\prime}_{N} on NN. Consider the endomorphism Ψ2(c)cΨ2cΨ1\Psi_{2}\coloneqq({c}^{\prime})^{*}{c}^{\prime}\|\Psi\|^{2}\|{c}\|-\Psi_{1} of SS. For the same reasons as for Ψ\Psi, it defines a bounded operator on H1(S)H^{1}(S). Fix eΓc(S)e\in\Gamma_{c}^{\infty}(S). Because Ψ\Psi is a positive bounded operator on H1(S)H^{1}(S), we may apply Lemma 4.5 with Θ=Ψ2\Theta=\Psi_{2} to conclude that

(4.1) G(gΨ2g1)e,e1𝑑g=G(g((c)c)g1Ψ2c)e,e1𝑑gΨΨe,e1\int_{G}\left\langle\left(g\Psi_{2}g^{-1}\right)e,e\right\rangle_{\mathcal{E}^{1}}\,dg=\int_{G}\left\langle\left(g(({c}^{\prime})^{*}{c}^{\prime})g^{-1}\|\Psi\|^{2}\|{c}\|\right)e,e\right\rangle_{\mathcal{E}^{1}}\,dg-\langle\Psi^{*}\Psi e,e\rangle_{\mathcal{E}^{1}}

is positive in C(G)C^{*}(G). Define b:M×Gb\colon M\times G\rightarrow\mathbb{R} by

b(x,g)bx(g)c(g1x).b(x,g)\coloneqq b_{x}(g)\coloneqq{c}^{\prime}(g^{-1}x).

By construction of c{c}^{\prime}, the quantities

C1(x)Gc(g1x)2𝑑g,C2(x)dc2vol(suppG(bx))C_{1}(x)\coloneqq\int_{G}{c}^{\prime}(g^{-1}x)^{2}\,dg,\qquad C_{2}(x)\coloneqq\|d{c}^{\prime}\|^{2}_{\infty}\cdot\operatorname{vol}\left(\operatorname{supp}_{G}(b_{x})\right)

are bounded as functions of xMx\in M (see also Remark 4.6 below). A direct calculation shows that

G(g((c)c)g1Ψ2c)e,e1𝑑gC(G)\displaystyle\left\|\int_{G}\left\langle\left(g(({c}^{\prime})^{*}{c}^{\prime})g^{-1}\|\Psi\|^{2}\|{c}\|\right)e,e\right\rangle_{\mathcal{E}^{1}}\,dg\right\|_{C^{*}(G)} =Ψ2cGcg1e,cg1e1𝑑gC(G)\displaystyle=\|\Psi\|^{2}\|{c}\|\left\|\int_{G}\left\langle{c}^{\prime}g^{-1}e,{c}^{\prime}g^{-1}e\right\rangle_{\mathcal{E}^{1}}dg\right\|_{C^{*}(G)}
Ψ2c(C1e12+C2e02)\displaystyle\leq\|\Psi\|^{2}\|{c}\|\left(\|C_{1}\|_{\infty}\|e\|^{2}_{\mathcal{E}^{1}}+\|C_{2}\|_{\infty}\|e\|^{2}_{\mathcal{E}^{0}}\right)
C3Ψ2e12,\displaystyle\leq C_{3}\|\Psi\|^{2}\|e\|^{2}_{\mathcal{E}^{1}},

for some constant C3C_{3}. Together with positivity of (4.1), this implies that

Ψe12=e,ΨΨe1C(G)C3Ψ2e12,\|\Psi e\|_{\mathcal{E}^{1}}^{2}=\|\langle e,\Psi^{*}\Psi e\rangle_{\mathcal{E}^{1}}\|_{C^{*}(G)}\leq C_{3}\|\Psi\|^{2}\|e\|^{2}_{\mathcal{E}^{1}},

so that Ψ\Psi extends to an operator on all of 1\mathcal{E}^{1}. Similarly, the H1(S)H^{1}(S)-adjoint Ψ\Psi^{*} defines an operator on 1\mathcal{E}^{1} that one checks is the adjoint of Ψ\Psi. ∎

Remark 4.6.

As can be seen from the proof, the conclusion of Proposition 4.4 holds more broadly for any MM on which the functions C1C_{1} and C2C_{2} on MM are bounded.

4.2. Vanishing

Two cases where the index of Definition 3.3 vanishes are straightforward to prove, but we state them here because they will be used in various places.

Lemma 4.7.

If (3.1) holds on all of MM, then indexG(D+Φ)=0\operatorname{index}_{G}(D+\Phi)=0.

Proof.

In this setting, the operator FF in (3.4) is invertible. This implies that the KKK\!K-cycle (,F)(\mathcal{E},F) is operator homotopic to the degenerate cycle (,F(FF)1/2)(\mathcal{E},F(F^{*}F)^{-1/2}). ∎

Lemma 4.8.

If M/GM/G is compact, and DD and Φ\Phi are of the forms (3.5) and (3.6), then indexG(D)=0\operatorname{index}_{G}(D)=0.

Proof.

In this setting, Φ\Phi is bounded, and the cycle (,F)(\mathcal{E},F) is operator homotopic to (,DD2+1)\Bigl{(}\mathcal{E},\frac{D}{\sqrt{D^{2}+1}}\Bigr{)}.

In general, let AA be a CC^{*}-algebra, let 0\mathcal{E}_{0} be a Hilbert AA-module, and set 00\mathcal{E}\coloneqq\mathcal{E}_{0}\oplus\mathcal{E}_{0}. and an adjointable operator FF on \mathcal{E} such that (,F)(\mathcal{E},F) is a Kasparov (,A)(\mathbb{C},A)-cycle, and FF is of the form

(4.2) F=(0F0F00),F=\begin{pmatrix}0&F_{0}\\ F_{0}&0\end{pmatrix},

for a (necessarily self-adjoint) F0(0)F_{0}\in\mathcal{L}(\mathcal{E}_{0}). Then [,F]=0K0(A)[\mathcal{E},F]=0\in K_{0}(A). Because DD2+1\frac{D}{\sqrt{D^{2}+1}} is of the form (4.2), the claim follows. ∎

4.3. Homotopy invariance

The index of Definition 3.3 has a homotopy invariance property analogous to Proposition 4.1 in [9]. This homotopy invariance applies in a more general setting than Callias-type operators.

Let PP be an odd, GG-equivariant Dirac-type operator on a /2\mathbb{Z}/2-graded Clifford module 𝒮\mathcal{S}, and let Ψ\Psi be an odd, smooth GG-equivariant, uniformly bounded Hermitian vector bundle endomorphism of 𝒮\mathcal{S}. Fix t0<t1t_{0}<t_{1}\in\mathbb{R}. For t[t0,t1]t\in[t_{0},t_{1}], consider the operator PtP+tΨP_{t}\coloneqq P+t\Psi.

Proposition 4.9 (Homotopy invariance).

Suppose that

  1. (1)

    for j=0,1j=0,1, the endomorphism Ψ\Psi defines an adjointable operator on the Sobolev module j\mathcal{E}^{j} of Definition 4.1;

  2. (2)

    there is a nonnegative, GG-invariant, cocompactly supported function fC(M)f\in C^{\infty}(M) such that for all t[t0,t1]t\in[t_{0},t_{1}], the operator Pt2+f:20P_{t}^{2}+f\colon\mathcal{E}^{2}\to\mathcal{E}^{0} is invertible, with inverse in (0,2)\mathcal{L}(\mathcal{E}^{0},\mathcal{E}^{2}).

Then indexG(Pt)K0(C(G))\operatorname{index}_{G}(P_{t})\in K_{0}(C^{*}(G)) is independent of t[t0,t1]t\in[t_{0},t_{1}].

Proof.

The proof proceeds identically to the proof of Proposition 4.1 in [9], with the exceptions that Theorem 4.2 should be substituted for Lemma 4.2 in [9], and Lemmas 4.4 and 4.6(a) in [15] should be substituted for Lemmas 1.4 and 1.5 in [7], respectively. ∎

Corollary 4.10.

If D+ΦD+\Phi is a GG-Callias-type operator on SMS\to M, and Ψ\Psi is a GG-equivariant, odd vector bundle endomorphism of SS that equals zero outside a cocompact set, then indexG(D+Φ)=indexG(D+Φ+Ψ)\operatorname{index}_{G}(D+\Phi)=\operatorname{index}_{G}(D+\Phi+\Psi).

Proof.

We set P=D+ΦP=D+\Phi and apply Proposition 4.9. Since Ψ\Psi is cocompactly supported, the first condition in Proposition 4.9 holds by Proposition 3.5 in [15]. For the same reason, Φ+tΨ\Phi+t\Psi is a Callias-type potential for all tt\in\mathbb{R}, so by Theorem 5.6 in [15], the second condition in Proposition 4.9 holds for t[0,1]t\in[0,1], where a priori the function ff may depend on tt. But since Ψ\Psi is zero outside a cocompact set, we can choose ff independent of tt. The claim then follows from Proposition 4.9. ∎

Remark 4.11.

In Proposition 4.9, it is not assumed that Ψ\Psi is a Callias-type potential in the sense of Definition 3.1. We will use Proposition 4.9 in this greater generality in the proof of Lemma 5.2.

Remark 4.12.

Corollary 4.10 can be used to give an alternative proof of Lemma 4.7: this corollary implies that in the setting of that lemma,

indexG(D+Φ)=indexG(DΦ)=indexG((D+Φ))=indexG(D+Φ).\operatorname{index}_{G}(D+\Phi)=\operatorname{index}_{G}(D-\Phi)=\operatorname{index}_{G}\bigl{(}(D+\Phi)^{*}\bigr{)}=-\operatorname{index}_{G}(D+\Phi).

See also Corollary 4.9 in [9].

4.4. A relative index theorem

We will use an analogue of Bunke’s relative index theorem, Theorem 1.2 in [7]. For j=1,2j=1,2, let MjM_{j}, SjS_{j}, DjD_{j} and Φj\Phi_{j}, respectively, be as MM, SS, DD and Φ\Phi were before. Suppose that there are a GG-invariant hypersurface NjMjN_{j}\subset M_{j} and a GG-invariant tubular neighbourhood UjU_{j} of NjN_{j}, and that there is a GG-equivariant isometry ψ:U1U2\psi\colon U_{1}\to U_{2} such that

  • ψ(N1)=N2\psi(N_{1})=N_{2};

  • ψ(S2|U2)S1|U1\psi^{*}(S_{2}|_{U_{2}})\cong S_{1}|_{U_{1}};

  • ψ(S2|U2)=S1|U1\psi^{*}(\nabla^{S_{2}}|_{U_{2}})=\nabla^{S_{1}}|_{U_{1}}, where Sj\nabla^{S_{j}} is the Clifford connection used to define DjD_{j}; and

  • Φ1|U1\Phi_{1}|_{U_{1}} corresponds to Φ2|U2\Phi_{2}|_{U_{2}} via ψ\psi.

Suppose that Mj=XjNjYjM_{j}=X_{j}\cup_{N_{j}}Y_{j} for closed, GG-invariant subsets Xj,YjMjX_{j},Y_{j}\subset M_{j}. We identify N1N_{1} and N2N_{2} via ψ\psi and write NN for this manifold when we do not want to distinguish between N1N_{1} and N2N_{2}. Using the map φ\varphi, form

M3X1NY2;M4X2NY1.M_{3}\coloneqq X_{1}\cup_{N}Y_{2};\qquad M_{4}\coloneqq X_{2}\cup_{N}Y_{1}.

For j=3,4j=3,4, let SjS_{j}, DjD_{j} and Φj\Phi_{j} be obtained from the corresponding data on M1M_{1} and M2M_{2} by cutting and gluing along U1U2U_{1}\cong U_{2} via ψ\psi.

Theorem 4.13.

In the above situation,

indexG(D1+Φ1)+indexG(D2+Φ2)=indexG(D3+Φ3)+indexG(D4+Φ4).\operatorname{index}_{G}(D_{1}+\Phi_{1})+\operatorname{index}_{G}(D_{2}+\Phi_{2})=\operatorname{index}_{G}(D_{3}+\Phi_{3})+\operatorname{index}_{G}(D_{4}+\Phi_{4}).
Proof.

This proof is an adaptation of the proof of Theorem 1.14 in [7], with some results from [15] added. For j=1,2,3,4j=1,2,3,4, let j\mathcal{E}_{j} and FjF_{j} be as \mathcal{E} and FF above and in Theorem 3.2, for the data indexed by jj. Using superscripts op\operatorname{op} to denote opposite gradings, we write 123op4op\mathcal{E}\coloneqq\mathcal{E}_{1}\oplus\mathcal{E}_{2}\oplus\mathcal{E}_{3}^{\operatorname{op}}\oplus\mathcal{E}_{4}^{\operatorname{op}} and FF1F2F3F4F\coloneqq F_{1}\oplus F_{2}\oplus F_{3}\oplus F_{4}. We will show that

(4.3) [,F]=0KK0(,C(G)),[\mathcal{E},F]=0\quad\in K\!K_{0}(\mathbb{C},C^{*}(G)),

which is equivalent to the theorem.

For j=1,2j=1,2, let χXj,χYjC(Mj)\chi_{X_{j}},\chi_{Y_{j}}\in C^{\infty}(M_{j}) be real-valued functions such that

  • supp(χXj)XjUj\operatorname{supp}(\chi_{X_{j}})\subset X_{j}\cup U_{j} and supp(χYj)YjUj\operatorname{supp}(\chi_{Y_{j}})\subset Y_{j}\cup U_{j};

  • ψ(χX2|U2)=χX1|U1\psi^{*}(\chi_{X_{2}}|_{U_{2}})=\chi_{X_{1}}|_{U_{1}} and ψ(χY2|U2)=χY1|U1\psi^{*}(\chi_{Y_{2}}|_{U_{2}})=\chi_{Y_{1}}|_{U_{1}}; and

  • χXj2+χYj2=1\chi_{X_{j}}^{2}+\chi_{Y_{j}}^{2}=1.

We view pointwise multiplication by these functions as operators

(4.4) χX1:13;χX2:14;χY1:23;χY2:24.\begin{split}\chi_{X_{1}}\colon&\mathcal{E}_{1}\to\mathcal{E}_{3};\\ \chi_{X_{2}}\colon&\mathcal{E}_{1}\to\mathcal{E}_{4};\end{split}\qquad\begin{split}\chi_{Y_{1}}\colon&\mathcal{E}_{2}\to\mathcal{E}_{3};\\ \chi_{Y_{2}}\colon&\mathcal{E}_{2}\to\mathcal{E}_{4}.\end{split}

The adjoints of these operators map in the opposite directions, and are also given by pointwise multiplication by the respective functions. Using these multiplication operators, and the grading operator γ\gamma, we form the operator

Xγ(00χX1χX200χY1χY2χX1χY100χX2χY200)X\coloneqq\gamma\begin{pmatrix}0&0&-\chi_{X_{1}}^{*}&\chi_{X_{2}}^{*}\\ 0&0&-\chi_{Y_{1}}^{*}&\chi_{Y_{2}}^{*}\\ \chi_{X_{1}}&\chi_{Y_{1}}&0&0\\ \chi_{X_{2}}&-\chi_{Y_{2}}&0&0\end{pmatrix}

on \mathcal{E}. Then XX is an odd, self-adjoint, adjointable operator on \mathcal{E} such that X2=1X^{2}=1. As such, it generates a Clifford algebra Cl\operatorname{Cl}.

We claim that XF+FXXF+FX is a compact operator. This is based on the Rellich lemma for Hilbert C(G)C^{*}(G)-modules, Theorem 4.2. Let χ\chi be one of the functions χXj\chi_{X_{j}} or χYj\chi_{Y_{j}}, viewed as an operator from k\mathcal{E}_{k} to l\mathcal{E}_{l} as in (4.4). Let fjC(M)f_{j}\in C^{\infty}(M) be as in Theorem 3.2, for the operator Dj+ΦjD_{j}+\Phi_{j}. For j=1,2,3,4j=1,2,3,4 and λ\lambda\in\mathbb{R}, the operator (Dj2+fj2+λ2)(D_{j}^{2}+f_{j}^{2}+\lambda^{2}) on j\mathcal{E}_{j} is invertible by Lemma 4.6 in [15], and we denote its inverse by Rj(λ)R_{j}(\lambda).

Then, as in the proof of Theorem 1.14 in [7], and using Proposition 4.12 in [15], we find that the operator

(4.5) χFlFkχ:lk\chi^{*}\circ F_{l}-F_{k}\circ\chi^{*}\colon\mathcal{E}_{l}\to\mathcal{E}_{k}

equals

(4.6) 2π(grad(χ)Rl(λ)+Dk2Rk(λ)grad(χ)Rl(λ)+..DkRk(λ)grad(fk)Rk(λ)grad(χ)Rl(λ)+DkRk(λ)grad(χ)DlRl(λ))dλ.\frac{2}{\pi}\int_{\mathbb{R}}\Bigl{(}-\operatorname{grad}(\chi)R_{l}(\lambda)+D_{k}^{2}R_{k}(\lambda)\operatorname{grad}(\chi)R_{l}(\lambda)+\Bigr{.}\\ \Bigl{.}D_{k}R_{k}(\lambda)\operatorname{grad}(f_{k})R_{k}(\lambda)\operatorname{grad}(\chi)R_{l}(\lambda)+D_{k}R_{k}(\lambda)\operatorname{grad}(\chi)D_{l}R_{l}(\lambda)\Bigr{)}\,d\lambda.

Theorem 4.2, together with Lemma 4.6(a) in [15], implies that for all cocompactly supported continuous functions φ\varphi on MjM_{j}, the compositions φDjnRj(λ)\varphi D_{j}^{n}R_{j}(\lambda), DjnφRj(λ)D_{j}^{n}\varphi R_{j}(\lambda) and DjnRj(λ)φD_{j}^{n}R_{j}(\lambda)\varphi are compact operators on j\mathcal{E}_{j} if n=0,1n=0,1, and adjointable operators if n=1n=1. So all the terms in the integrand in (4.6) are compact operators. By Lemmas 4.6 and 4.8 in [15], the norm of the integrand in (4.6) is bounded by a(b+λ2)1a(b+\lambda^{2})^{-1} for constants a,b>0a,b>0. So the integral converges in the operator norm on ()\mathcal{L}(\mathcal{E}), and we conclude that (4.5) is a compact operator on \mathcal{E}. This implies that XF+FXXF+FX is a compact operator.

Because XX generates Cl\operatorname{Cl} and XX anticommutes with FF modulo compacts, the pair (,F)(\mathcal{E},F) is a Kasparov (Cl,C(G))(\operatorname{Cl},C^{*}(G))-cycle. Its class in KK(Cl,C(G))K\!K(\operatorname{Cl},C^{*}(G)) is mapped to the left-hand aside of (4.3) by the pullback along the inclusion map Cl\mathbb{C}\hookrightarrow\operatorname{Cl}. That map is zero by Lemma 1.15 in [7], so (4.3) follows. ∎

Theorem 4.13 implies the following version of Proposition 5.9 in [9].

Corollary 4.14.

In the setting of Theorem 4.13, suppose that for j=1,2j=1,2, the set XjX_{j} is cocompact, and contains a set ZjZ_{j} for Φj\Phi_{j} as in Definition 3.1. Then

indexG(D1+Φ1)=indexG(D2+Φ2).\operatorname{index}_{G}(D_{1}+\Phi_{1})=\operatorname{index}_{G}(D_{2}+\Phi_{2}).
Proof.

This fact can be deduced from Theorem 4.13 in exactly the same way Proposition 5.9 in [9] is deduced from Theorem 5.7 in [9]. Compared to that proof in [9], references to Corollaries 3.4 and 4.9 and Theorem 5.7 in that paper should be replaced by references to Lemmas 4.7 and 4.8 and Theorem 4.13, respectively, in the present paper. ∎

The crucial assumption in Corollary 4.14 is that all data near N1N_{1} can be identified with the corresponding data near N2N_{2}.

5. Proof of the GG-Callias-type index theorem

The first and most important step in the proof of Theorem 3.4 is Proposition 5.1, which states that indexG(D+Φ)\operatorname{index}_{G}(D+\Phi) equals the index of a GG-Callias-type operator on the manifold N×N\times\mathbb{R}, which we will call the cylinder on NN. See Figure 1. Such an approach is taken in proofs of various other index theorems for Callias-type operators; see for example [2, 6, 7, 9].

Refer to caption
Figure 1. The cylinder N×N\times\mathbb{R}

In this section, we consider the setting of Subsection 3.2. In particular, DD and Φ\Phi are assumed to be of the forms (3.5) and (3.6).

5.1. An index on the cylinder

Let S±N×N×S^{N\times\mathbb{R}}_{\pm}\to N\times\mathbb{R} be the pullbacks of the bundles of S±NNS^{N}_{\pm}\to N defined in Subsection 3.2 along the projection N×NN\times\mathbb{R}\to N. They are Clifford modules, with Clifford actions

c^(v,t)=c(v+tn^),\hat{c}(v,t)=c(v+t\hat{n}),

for vTNv\in TN and tt\in\mathbb{R}, where cc is the Clifford action by TMTM on SS (which preserves S+NS^{N}_{+} as pointed out in Subsection 3.2), and n^\hat{n} is the normal vector field to NN in the direction of M+M_{+}. Let D0S+N×RD^{S^{N\times R}_{+}}_{0} be the Dirac operator on Γ(S+N×)\Gamma^{\infty}(S^{N\times\mathbb{R}}_{+}) defined by this Clifford action, and the pullback to N×N\times\mathbb{R} of the restriction to NN of the Clifford connection S+N\nabla^{S^{N}_{+}} used to define D0S+ND_{0}^{S^{N}_{+}}.

Let χC()\chi\in C^{\infty}(\mathbb{R}) be an odd function such that χ(t)=t\chi(t)=t for all t2t\geq 2. We also denote its pullback to N×N\times\mathbb{R} by χ\chi. Then pointwise multiplication by χ\chi is an admissible endomorphism for D0S+N×D^{S^{N\times\mathbb{R}}_{+}}_{0}. Whenever a Dirac operator with a subscript 0 is given, we will remove that subscript to denote the corresponding Dirac operator on two copies of the Clifford module in question, as in (3.5). In the current setting, this gives us the Dirac operator

DS+N×=(0D0S+N×D0S+N×0)D^{S^{N\times\mathbb{R}}_{+}}=\begin{pmatrix}0&D^{S^{N\times\mathbb{R}}_{+}}_{0}\\ D^{S^{N\times\mathbb{R}}_{+}}_{0}&0\end{pmatrix}

on Γ(S+N×S+N×)\Gamma^{\infty}(S^{N\times\mathbb{R}}_{+}\oplus S^{N\times\mathbb{R}}_{+}). We also consider the admissible endomorphism

χN×=(0iχiχ0)\chi^{N\times\mathbb{R}}=\begin{pmatrix}0&i\chi\\ -i\chi&0\end{pmatrix}

of S+N×S+N×S^{N\times\mathbb{R}}_{+}\oplus S^{N\times\mathbb{R}}_{+}.

Proposition 5.1.

We have

(5.1) indexG(D+Φ)=indexG(DS+N×+χN×).\operatorname{index}_{G}(D+\Phi)=\operatorname{index}_{G}(D^{S^{N\times\mathbb{R}}_{+}}+\chi^{N\times\mathbb{R}}).

The proof of Proposition 5.1 that we give below is an analogue of the proof of Theorem 5.4 in [9]. We give this proof in Subsections 5.2 and 5.3, referring to [9] for details in some places, and using results from [15] and from Section 4.

5.2. Attaching a half-cylinder

Let S0N×N×S_{0}^{N\times\mathbb{R}}\to N\times\mathbb{R} be the pullback of S0|NNS_{0}|_{N}\to N. We choose UU small enough so that S0|US0N×|US_{0}|_{U}\cong S^{N\times\mathbb{R}}_{0}|_{U}.

Because the sets XjX_{j} are cocompact in Corollary 4.14, we initially compare the left-hand side of (5.1) to an index on a manifold where only M+M_{+} is replaced by a half-cylinder N×[1,)N\times[1,\infty). To be more precise, indexG(D+Φ)\operatorname{index}_{G}(D+\Phi) is invariant under changes in the Riemannian metric on cocompact sets because the Kasparov (,C(G))(\mathbb{C},C^{*}(G))-cycles corresponding to two GG-invariant Riemannian metrics differing only a cocompact set are homotopic by convexity of the space of GG-invariant Riemannian metrics. We choose a metric such that there is a neighbourhood UU of NN that is isometric to N×(1/4,7/4)N\times(1/4,7/4) (see Figure 2).

Refer to caption
Figure 2. The manifold MM

By Corollary 4.10, the index of D+ΦD+\Phi does not change if we change Φ0\Phi_{0} in a cocompact set. So we may assume that Φ0\Phi_{0} is constant in the direction normal to NN inside UU; i.e. for all nNn\in N and t(1/4,7/4)t\in(1/4,7/4), Φ0(n,t)=Φ0N(n)\Phi_{0}(n,t)=\Phi^{N}_{0}(n), for an endomorphism Φ0N\Phi^{N}_{0} of S0|NS_{0}|_{N}. We further choose UU such that a set ZZ as in Definition 3.1 is contained in MUM_{-}\setminus U.

Let S0N×\nabla^{S_{0}^{N\times\mathbb{R}}} be the pullback of S0|N\nabla^{S_{0}}|_{N} to a connection on S0N×S_{0}^{N\times\mathbb{R}}. We choose the Clifford connection S0\nabla^{S_{0}} to define D0D_{0} so that on UU, it equals the restriction of S0N×\nabla^{S_{0}^{N\times\mathbb{R}}} to N×(1/4,7/4)N\times(1/4,7/4).

For this structure near NN, we can form the Riemannian manifold MCMN[1,)M_{C}\coloneqq M_{-}\cup_{N}[1,\infty) (see Figure 3), and define the Clifford module S0CMCS^{C}_{0}\to M_{C} such that it equals S0S_{0} on MM_{-} and S0N×S^{N\times\mathbb{R}}_{0} on N×(1/4,)N\times(1/4,\infty). Let S0C\nabla^{S_{0}^{C}} be the Clifford connection on S0CS^{C}_{0} corresponding to S0\nabla^{S_{0}} on MM_{-} and to S0N×\nabla^{S^{N\times\mathbb{R}}_{0}} on N×(1/4,)N\times(1/4,\infty). Let D0CD^{C}_{0} be the resulting Dirac operator.

Refer to caption
Figure 3. The manifold MCM_{C}

We define an endomorphism Φ0C\Phi^{C}_{0} of S0CS^{C}_{0} that is equal to Φ0\Phi_{0} on MM_{-} and to the pullback of Φ0N\Phi^{N}_{0} on N×(1/4,)N\times(1/4,\infty). Recall that by removing the subscript 0 from D0CD^{C}_{0} we refer to the construction (3.5). Similarly, when we remove the subscript 0 from Φ0C\Phi^{C}_{0}, we will be referring to the endomorphism ΦC\Phi^{C} defined by Φ0C\Phi^{C}_{0} as in (3.6). Then Corollary 4.14 immediately implies that

(5.2) indexG(D+Φ)=indexG(DC+ΦC).\operatorname{index}_{G}(D+\Phi)=\operatorname{index}_{G}(D^{C}+\Phi^{C}).

The connection S0N×\nabla^{S^{N\times\mathbb{R}}_{0}}, and therefore the corresponding Dirac operator, does not preserve the decomposition S0N×=S+N×SN×S^{N\times\mathbb{R}}_{0}=S^{N\times\mathbb{R}}_{+}\oplus S^{N\times\mathbb{R}}_{-}. With respect to this decomposition, that Dirac operator has the form

(5.3) (D0S+N×ABD0SN×,)\begin{pmatrix}D^{S^{N\times\mathbb{R}}_{+}}_{0}&A\\ B&D^{S^{N\times\mathbb{R}}_{-}}_{0},\end{pmatrix}

for vector bundle homomorphisms A:SN×S+N×A\colon S^{N\times\mathbb{R}}_{-}\to S^{N\times\mathbb{R}}_{+} and B:S+N×SN×B\colon S^{N\times\mathbb{R}}_{+}\to S^{N\times\mathbb{R}}_{-}. (See Section 5.16 in [6], or use the fact that the difference of two connections is an endomorphism-valued one-form.) The Dirac operator D0CD^{C}_{0} equals this operator on N×(1/4,)N\times(1/4,\infty). Let S±N×\nabla^{S_{\pm}^{N\times\mathbb{R}}} be the pullback of S±N\nabla^{S_{\pm}^{N}} to a connection on S±N×S_{\pm}^{N\times\mathbb{R}}. Consider a Clifford connection S0C\nabla^{S_{0}^{C}} on S0CS^{C}_{0} that is equal to the direct sum of S+N×\nabla^{S_{+}^{N\times\mathbb{R}}} and SN×\nabla^{S_{-}^{N\times\mathbb{R}}} on N×(1/2,)N\times(1/2,\infty) and to S0\nabla^{S_{0}} on MUM_{-}\setminus U. Then the corresponding Dirac operator D~0C\tilde{D}^{C}_{0} is equal to

(D0S+N×00D0SN×)\begin{pmatrix}D^{S^{N\times\mathbb{R}}_{+}}_{0}&0\\ 0&D^{S^{N\times\mathbb{R}}_{-}}_{0}\end{pmatrix}

on N×(1/2,)N\times(1/2,\infty) and to D0D_{0} on MUM_{-}\setminus U.

Lemma 5.2.

There exists a λ1\lambda\geq 1 such that λΦC\lambda\Phi^{C} is admissible for D~0C\tilde{D}^{C}_{0}. For such λ\lambda,

(5.4) indexG(DC+ΦC)=indexG(D~C+λΦC).\operatorname{index}_{G}(D^{C}+\Phi^{C})=\operatorname{index}_{G}(\tilde{D}^{C}+\lambda\Phi^{C}).
Proof.

Existence of λ\lambda with the desired property can be established as in the proof of Lemma 5.13 in [9]. The equality (5.4) can be proved via a linear homotopy; again see the proof of Lemma 5.13 in [9] for details, where references to Proposition 4.1 in [9] should be replaced by references to Proposition 4.9 in the current paper, and one uses Propositions 4.3 and 4.4 to check that the first condition in Proposition 4.9 is satisfied.

To be explicit, an application of Proposition 4.9 shows that

indexG(DC+λΦC)=indexG(D~C+λΦC).\operatorname{index}_{G}(D^{C}+\lambda\Phi^{C})=\operatorname{index}_{G}(\tilde{D}^{C}+\lambda\Phi^{C}).

This follows from Proposition 4.9 by setting P=DC+λΦCP=D^{C}+\lambda\Phi^{C}, Ψ=(0AB0,)\Psi=\begin{pmatrix}0&A\\ B&0,\end{pmatrix}, with AA and BB as in (5.3), t0=1t_{0}=-1 and t1=0t_{1}=0. Note that although Ψ\Psi is not a Callias-type potential here, Proposition 4.9 still applies, since Ψ\Psi defines an element of (0)\mathcal{L}(\mathcal{E}^{0}) by Proposition 4.3 and an element of (1)\mathcal{L}(\mathcal{E}^{1}) by Proposition 4.4. (The conditions of Proposition 4.4 hold because of the forms of DCD^{C}, ΦC\Phi^{C} and Ψ\Psi.) A more straightforward application of Proposition 4.9 yields indexG(DC+ΦC)=indexG(DC+λΦC)\operatorname{index}_{G}(D^{C}+\Phi^{C})=\operatorname{index}_{G}(D^{C}+\lambda\Phi^{C}). ∎

Let χC()\chi\in C^{\infty}(\mathbb{R}) be an odd function such that

χ(t)={0if 1/4t1/4;1if 3/4t3/2;tif t2.\chi(t)=\left\{\begin{array}[]{ll}0&\text{if $-1/4\leq t\leq 1/4$};\\ 1&\text{if $3/4\leq t\leq 3/2$};\\ t&\text{if $t\geq 2$}.\end{array}\right.

(The property that χ\chi is unbounded in a way that makes it proper is only used in the proof of Lemma 5.6; see Lemma 5.5.) Such functions form a subset of the set of functions χ\chi in Subsection 5.1, but Proposition 5.1 for general χ\chi follows from the case for this class of functions, because the index on the right-hand side of (5.1) does not change if we modify χ\chi in a cocompact set. (And at any rate, to prove Theorem 3.4, we only need Proposition 5.1 to hold for one such function χ\chi.)

Let γN\gamma^{N} be the grading operator on S0|NS_{0}|_{N} that equals ±1\pm 1 on S±NS^{N}_{\pm}. Let γN×\gamma^{N\times\mathbb{R}} be its pullback to S0N×S^{N\times\mathbb{R}}_{0}. Let Φ0χ\Phi_{0}^{\chi} be the endomorphism of S0CS_{0}^{C} equal to χγN×\chi\gamma^{N\times\mathbb{R}} on N×(1/4,)N\times(1/4,\infty) and equal to zero on the rest of MCM_{C}.

Lemma 5.3.

We have

indexG(D~C+λΦC)=indexG(D~C+Φχ).\operatorname{index}_{G}(\tilde{D}^{C}+\lambda\Phi^{C})=\operatorname{index}_{G}(\tilde{D}^{C}+\Phi^{\chi}).
Proof.

This can be proved via a linear homotopy between λΦC\lambda\Phi^{C} and Φχ\Phi^{\chi}. The details are precisely as in the proof of Lemma 5.15 in [9], with references to Propositions 4.1 and 5.9 in [9] replaced by references to Proposition 4.9 (combined with Propositions 4.3 and 4.4) and Corollary 4.14, respectively, in the present paper. ∎

5.3. Proof of Proposition 5.1

Let MM_{-}^{-} be the manifold MM_{-} with reversed orientation. Form the manifold

MC(N×(,1])NM.M_{C}^{-}\coloneqq(N\times(-\infty,1])\cup_{N}M_{-}^{-}.

See Figure 4. (The notation is motivated by the fact that MCM_{C} with reversed orientation is naturally equal to (N×(,1])NM(N\times(-\infty,-1])\cup_{N}M_{-}^{-}, which can be identified with MCM_{C}^{-} via a shift over a distance 22.)

Refer to caption
Figure 4. The manifold MCM_{C}^{-}

Let S0MS_{0}^{-}\to M^{-}_{-} be equal to the vector bundle S0|MS_{0}|_{M_{-}}, but with the opposite Clifford action (where vTMv\in TM^{-}_{-} acts as c(v)c(-v)). Let S0C,MCS_{0}^{C,-}\to M_{C}^{-} be the Clifford module that is equal to S0N×S^{N\times\mathbb{R}}_{0} on N×(,5/4]N\times(-\infty,5/4] and to S0S_{0}^{-} on MM^{-}_{-}. From the Clifford connections S±N×\nabla^{S^{N\times\mathbb{R}}_{\pm}} on S±N×S^{N\times\mathbb{R}}_{\pm} and a Clifford connection S0\nabla^{S_{0}^{-}} on S0S_{0}^{-}, construct a Clifford connection S0C,\nabla^{S_{0}^{C,-}} on S0C,S_{0}^{C,-} by

S0C,{S+N×SN×on N×(,5/4);S0on M.\nabla^{S_{0}^{C,-}}\coloneqq\left\{\begin{array}[]{ll}\nabla^{S^{N\times\mathbb{R}}_{+}}\oplus\nabla^{S^{N\times\mathbb{R}}_{-}}&\text{on $N\times(-\infty,5/4)$};\\ \nabla^{S_{0}^{-}}&\text{on $M_{-}^{-}$}.\end{array}\right.

Using this connection, we obtain the Dirac operator D0C,D^{C,-}_{0} on Γ(S0C,)\Gamma^{\infty}(S_{0}^{C,-}). Then D0C,D^{C,-}_{0} equals D~0C\tilde{D}^{C}_{0} on N×(1/2,5/4)N\times(1/2,5/4).

The function χ\chi equals 11 on (3/4,5/4)(3/4,5/4). Thus on this interval, both Φ0χ\Phi_{0}^{\chi} and (χ001)\begin{pmatrix}\chi&0\\ 0&-1\end{pmatrix} are equal to γN×\gamma^{N\times\mathbb{R}}. (Here we use 2×22\times 2 matrix notation with respect to the decomposition S0N×=S+N×SN×S_{0}^{N\times\mathbb{R}}=S_{+}^{N\times\mathbb{R}}\oplus S_{-}^{N\times\mathbb{R}}.) So we can define the endomorphism Φ0C,\Phi^{C,-}_{0} of S0C,S_{0}^{C,-} by setting it equal to Φ0χ\Phi^{\chi}_{0} on (N×(3/4,1])NM(N\times(3/4,1])\cup_{N}M_{-}^{-} and equal to

(5.5) (χ001)\begin{pmatrix}\chi&0\\ 0&-1\end{pmatrix}

on N×(,5/4)N\times(-\infty,5/4), where S0C,=S0N×=S+N×SN×S_{0}^{C,-}=S_{0}^{N\times\mathbb{R}}=S_{+}^{N\times\mathbb{R}}\oplus S_{-}^{N\times\mathbb{R}}.

Lemma 5.4.

We have

indexG(DC,+ΦC,)=0.\operatorname{index}_{G}(D^{C,-}+\Phi^{C,-})=0.
Proof.

Because χ=1\chi=-1 on (,3/4)(-\infty,-3/4), we can define the endomorphism Φ~0C,\tilde{\Phi}^{C,-}_{0} of S0C,S_{0}^{C,-} by setting it equal to Φ0C,\Phi^{C,-}_{0} on N×(,3/4)N\times(-\infty,-3/4) (where it equals (5.5)) and equal to 1-1 on (N×(1,1])NM(N\times(-1,1])\cup_{N}M_{-}^{-}. For that endomorphism, the estimate (3.1) holds on all of MM. Therefore by Lemma 4.7,

indexG(DC,+Φ~C,)=0.\operatorname{index}_{G}(D^{C,-}+\tilde{\Phi}^{C,-})=0.

The claim now follows from Corollary 4.10. ∎

Proof of Proposition 5.1..

Consider the cylinder N×N\times\mathbb{R} as in Figure 1. The data (MC,S0C,Φ0χ)(M_{C},S^{C}_{0},\Phi^{\chi}_{0}) and (MC,S0C,,Φ0C,)(M_{C}^{-},S_{0}^{C,-},\Phi_{0}^{C,-}) coincide in a neighbourhood of N×{1}N\times\{1\}. By cutting along N×{1}N\times\{1\} and gluing, we obtain the corresponding data (N×,S0N×,Φ0N×)(N\times\mathbb{R},S_{0}^{N\times\mathbb{R}},\Phi^{N\times\mathbb{R}}_{0}) and (MNM,S0MNM,Φ0MNM)(M_{-}\cup_{N}M_{-}^{-},S_{0}^{M_{-}\cup_{N}M_{-}^{-}},\Phi_{0}^{M_{-}\cup_{N}M_{-}^{-}}). See Figure 5. To be explicit,

(5.6) Φ0N×=(χ001)\Phi_{0}^{N\times\mathbb{R}}=\begin{pmatrix}\chi&0\\ 0&-1\end{pmatrix}

on S+N×SN×S^{N\times\mathbb{R}}_{+}\oplus S^{N\times\mathbb{R}}_{-}.

Refer to caption
Figure 5. The manifold MNMM_{-}\cup_{N}M_{-}^{-}

Theorem 4.13 implies that

indexG(D~C+Φχ)+indexG(DC,+ΦC,)=indexG(DN×+ΦN×)+indexG(DMNM+ΦMNM).\operatorname{index}_{G}(\tilde{D}^{C}+\Phi^{\chi})+\operatorname{index}_{G}(D^{C,-}+\Phi^{C,-})\\ =\operatorname{index}_{G}(D^{N\times\mathbb{R}}+\Phi^{N\times\mathbb{R}})+\operatorname{index}_{G}(D^{M_{-}\cup_{N}M_{-}^{-}}+\Phi^{M_{-}\cup_{N}M_{-}^{-}}).

By Lemmas 4.8 and 5.4, this implies that

indexG(D~C+Φχ)=indexG(DN×+ΦN×).\operatorname{index}_{G}(\tilde{D}^{C}+\Phi^{\chi})=\operatorname{index}_{G}(D^{N\times\mathbb{R}}+\Phi^{N\times\mathbb{R}}).

The connection ~S0N×\tilde{\nabla}^{S_{0}^{N\times\mathbb{R}}} on S0N×S_{0}^{N\times\mathbb{R}} obtained from cutting and gluing the connections S0C\nabla^{S_{0}^{C}} and S0C,\nabla^{S_{0}^{C,-}} is the direct sum connection S+N×S+N×\nabla^{S_{+}^{N\times\mathbb{R}}}\oplus\nabla^{S_{+}^{N\times\mathbb{R}}}. So the corresponding Dirac operator D0S0N×D_{0}^{S_{0}^{N\times\mathbb{R}}} equals

D0S0N×=(D0S+N×00D0S+N×).D_{0}^{S_{0}^{N\times\mathbb{R}}}=\begin{pmatrix}D_{0}^{S_{+}^{N\times\mathbb{R}}}&0\\ 0&D_{0}^{S_{+}^{N\times\mathbb{R}}}\end{pmatrix}.

By the explicit form (5.6) of Φ0N×\Phi^{N\times\mathbb{R}}_{0}, the operator D0S0N×±iΦ0N×D_{0}^{S_{0}^{N\times\mathbb{R}}}\pm i\Phi^{N\times\mathbb{R}}_{0} on Γ(S0N×)\Gamma^{\infty}(S^{N\times\mathbb{R}}_{0}) is the direct sum of the operators D0S+N×±iχD_{0}^{S_{+}^{N\times\mathbb{R}}}\pm i\chi on Γ(S+N×)\Gamma^{\infty}(S^{N\times\mathbb{R}}_{+}) and D0SN×iD_{0}^{S_{-}^{N\times\mathbb{R}}}\mp i on Γ(SN×)\Gamma^{\infty}(S^{N\times\mathbb{R}}_{-}). So

indexG(DN×+ΦN×)=indexG(DS+N×+χN×)+indexG(DSN×+(0ii0)).\operatorname{index}_{G}(D^{N\times\mathbb{R}}+\Phi^{N\times\mathbb{R}})=\operatorname{index}_{G}(D^{S^{N\times\mathbb{R}}_{+}}+\chi^{N\times\mathbb{R}})+\operatorname{index}_{G}\left(D^{S^{N\times\mathbb{R}}_{-}}+\begin{pmatrix}0&-i\\ i&0\end{pmatrix}\right).

Lemma 4.7 then implies that the second term on the right-hand side is zero, whence

indexG(D~C+Φχ)=indexG(DS+N×+χN×).\operatorname{index}_{G}(\tilde{D}^{C}+\Phi^{\chi})=\operatorname{index}_{G}(D^{S^{N\times\mathbb{R}}_{+}}+\chi^{N\times\mathbb{R}}).

The claim now follows from (5.2) in conjunction with Lemmas 5.2 and 5.3. As pointed out above Lemma 5.3, the case for the class of functions χ\chi we have used implies the case of the more general functions χ\chi allowed in Proposition 5.1. ∎

5.4. Proof of Theorem 3.4

Let N×\mathcal{E}_{N\times\mathbb{R}} be the Hilbert C(G)C^{*}(G)-module constructed from Γc(S+N×)\Gamma_{c}(S_{+}^{N\times\mathbb{R}}) as in Subsection 3.1. We write DχDS+N×+χN×D_{\chi}\coloneqq D^{S^{N\times\mathbb{R}}_{+}}+\chi^{N\times\mathbb{R}} for brevity.

Lemma 5.5.

For all a>0a>0 the operator (Dχ2+a)1(D_{\chi}^{2}+a)^{-1} on N×\mathcal{E}_{N\times\mathbb{R}} is compact.

Proof.

This is (a special case of) an analogue of Theorem 2.40 in [11]. The proof proceeds in the same way, with the difference that the operator Dχ2D_{\chi}^{2} can only be bounded below by a function hh that is GG-proper, in the sense that the inverse image of a compact set is cocompact instead of compact. One chooses the bump functions gng_{n} in the proof of Theorem 2.40 in [11] to be GG-invariant and cocompactly supported. Theorem 4.2, and Lemma 4.6(a) in [15], imply that gn(Dχ2+a)1g_{n}(D_{\chi}^{2}+a)^{-1} is a compact operator on N×\mathcal{E}_{N\times\mathbb{R}}. And as in the proof of Theorem 2.40 in [11], one shows that gn(Dχ2+a)1g_{n}(D_{\chi}^{2}+a)^{-1} converges to (Dχ2+a)1(D_{\chi}^{2}+a)^{-1} in the operator norm on (N×)\mathcal{L}(\mathcal{E}_{N\times\mathbb{R}}). ∎

We will use an analogue of Theorem 6.6 in [9].

Lemma 5.6.

The operator

(5.7) Dχ(Dχ2+f)1/2Dχ(Dχ2+1)1/2D_{\chi}\bigl{(}D_{\chi}^{2}+f\bigr{)}^{-1/2}-D_{\chi}\bigl{(}D_{\chi}^{2}+1\bigr{)}^{-1/2}

lies in 𝒦(N×){\mathcal{K}}(\mathcal{E}_{N\times\mathbb{R}}).

Proof.

By Proposition 4.12 in [15], and as in (6.6) in [9], the operator (5.7) equals

(5.8) 2π0Dχ(Dχ2+f+λ2)1(f1)(Dχ2+1+λ2)1𝑑λ.\frac{2}{\pi}\int_{0}^{\infty}D_{\chi}(D_{\chi}^{2}+f+\lambda^{2})^{-1}(f-1)(D_{\chi}^{2}+1+\lambda^{2})^{-1}\,d\lambda.

The operator (Dχ2+1+λ2)1(D_{\chi}^{2}+1+\lambda^{2})^{-1} is compact by Lemma 5.5. Lemmas 4.6 and 4.8 in [15] imply that the integrand in (5.8) is bounded by a(b+λ2)1a(b+\lambda^{2})^{-1} for certain a,b>0a,b>0, so the integral converges in operator norm. It therefore defines a compact operator. ∎

Theorem 3.4 follows from Proposition 5.1 and the following fact.

Proposition 5.7.

In the setting of Proposition 5.1,

indexG(Dχ)=indexG(DS+N).\operatorname{index}_{G}(D_{\chi})=\operatorname{index}_{G}(D^{S^{N}_{+}}).
Proof.

Let N\mathcal{E}_{N} be the Hilbert C(G)C^{*}(G)-module constructed from Γc(S+N)\Gamma_{c}(S_{+}^{N}) as in Subsection 3.1. Then N×=NL2()\mathcal{E}_{N\times\mathbb{R}}=\mathcal{E}_{N}\otimes L^{2}(\mathbb{R}). So Lemma 5.6 implies that indexG(Dχ)\operatorname{index}_{G}(D_{\chi}) is represented by the unbounded Kasparov cycle

(5.9) (NL2()2,Dχ).\left(\mathcal{E}_{N}\otimes L^{2}(\mathbb{R})\otimes\mathbb{C}^{2},D_{\chi}\right).

The Callias-type operator DχD_{\chi} on

Γ(S+N×S+N×)=Γ(S+N)C()2\Gamma^{\infty}(S^{N\times\mathbb{R}}_{+}\oplus S^{N\times\mathbb{R}}_{+})=\Gamma^{\infty}(S^{N}_{+})\otimes C^{\infty}(\mathbb{R})\otimes\mathbb{C}^{2}

is equal to

(5.10) (0DS+NDS+N0)1C()+γS+N(0iddtiddt0)+1Γc(S+N)(0iχiχ0),\begin{pmatrix}0&D^{S^{N}_{+}}\\ D^{S^{N}_{+}}&0\end{pmatrix}\otimes 1_{C^{\infty}(\mathbb{R})}+\gamma_{S^{N}_{+}}\otimes\begin{pmatrix}0&i\frac{d}{dt}\\ i\frac{d}{dt}&0\end{pmatrix}+1_{\Gamma^{\infty}_{c}(S^{N}_{+})}\otimes\begin{pmatrix}0&-i\chi\\ i\chi&0\end{pmatrix},

where γS+N\gamma_{S^{N}_{+}} is the grading operator on S+NS^{N}_{+}, equal to i-i times Clifford multiplication by the unit normal vector field n^\hat{n} on NN pointing into M+M_{+} (so that γS+Niddt=c(n^)ddt\gamma_{S^{N}_{+}}\otimes i\frac{d}{dt}=c(\hat{n})\otimes\frac{d}{dt}). Let N±\mathcal{E}_{N}^{\pm} be the even and odd-graded parts of N\mathcal{E}_{N}, and let D±S+ND^{S^{N}_{+}}_{\pm} be the restriction of DS+ND^{S^{N}_{+}} to even and odd-graded sections of S+NS^{N}_{+}, respectively. With respect to the decomposition

(5.11) NL2()2=(N+L2())(NL2())(N+L2())(NL2()),\mathcal{E}_{N}\otimes L^{2}(\mathbb{R})\otimes\mathbb{C}^{2}=(\mathcal{E}_{N}^{+}\otimes L^{2}(\mathbb{R}))\oplus(\mathcal{E}_{N}^{-}\otimes L^{2}(\mathbb{R}))\oplus(\mathcal{E}_{N}^{+}\otimes L^{2}(\mathbb{R}))\oplus(\mathcal{E}_{N}^{-}\otimes L^{2}(\mathbb{R})),

the operator (5.10) equals

(5.12) Dχ=(001N+(iddtiχ)DS+N1L2()00D+S+N1L2()1N(iddtiχ)1N+(iddt+iχ)DS+N1L2()00D+S+N1L2()1N(iddt+iχ)00)D_{\chi}=\begin{pmatrix}0&0&1_{\mathcal{E}_{N}^{+}}\otimes(i\frac{d}{dt}-i\chi)&D^{S^{N}_{+}}_{-}\otimes 1_{L^{2}(\mathbb{R})}\\ 0&0&D^{S^{N}_{+}}_{+}\otimes 1_{L^{2}(\mathbb{R})}&1_{\mathcal{E}_{N}^{-}}\otimes(-i\frac{d}{dt}-i\chi)&\\ 1_{\mathcal{E}_{N}^{+}}\otimes(i\frac{d}{dt}+i\chi)&D^{S^{N}_{+}}_{-}\otimes 1_{L^{2}(\mathbb{R})}&0&0\\ D^{S^{N}_{+}}_{+}\otimes 1_{L^{2}(\mathbb{R})}&1_{\mathcal{E}_{N}^{-}}\otimes(-i\frac{d}{dt}+i\chi)&0&0\end{pmatrix}

The kernel of iddt±iχi\frac{d}{dt}\pm i\chi in C()C^{\infty}(\mathbb{R}) is one-dimensional, and spanned by the function

f±(t)=e0tχ(u)𝑑u.f_{\pm}(t)=e^{\mp\int_{0}^{t}\chi(u)\,du}.

By the properties of χ\chi, f+L2()f_{+}\in L^{2}(\mathbb{R}), whereas fL2()f_{-}\not\in L^{2}(\mathbb{R}). It follows that iddtiχi\frac{d}{dt}-i\chi is invertible on the appropriate domain, while iddt+iχi\frac{d}{dt}+i\chi is zero on f+\mathbb{C}f_{+} and invertible on f+f_{+}^{\perp}.

Consider the submodules

1(N+f+)00(Nf+);21=(N+f+)(NL2())(N+L2())(Nf+)\begin{split}\mathcal{E}_{1}&\coloneqq(\mathcal{E}_{N}^{+}\otimes\mathbb{C}f_{+})\oplus 0\oplus 0\oplus(\mathcal{E}_{N}^{-}\otimes\mathbb{C}f_{+});\\ \mathcal{E}_{2}\coloneqq\mathcal{E}_{1}^{\perp}&=(\mathcal{E}_{N}^{+}\otimes f_{+}^{\perp})\oplus(\mathcal{E}_{N}^{-}\otimes L^{2}(\mathbb{R}))\oplus(\mathcal{E}_{N}^{+}\otimes L^{2}(\mathbb{R}))\oplus(\mathcal{E}_{N}^{-}\otimes f_{+}^{\perp})\end{split}

of (5.11). These are preserved by the operator DχD_{\chi}. (For 1\mathcal{E}_{1}, this is immediate from (5.12); for 2\mathcal{E}_{2}, this follows from the facts that DχD_{\chi} is symmetric and preserves 1\mathcal{E}_{1}.) We find that the cycle (5.9) decomposes as

(5.13) (1,(000DS+N1f+00000000D+S+N1f+000))(2,Dχ|2).\left(\mathcal{E}_{1},\begin{pmatrix}0&0&0&D^{S^{N}_{+}}_{-}\otimes 1_{\mathbb{C}f_{+}}\\ 0&0&0&0\\ 0&0&0&0\\ D^{S^{N}_{+}}_{+}\otimes 1_{\mathbb{C}f_{+}}&0&0&0\end{pmatrix}\right)\oplus(\mathcal{E}_{2},D_{\chi}|_{\mathcal{E}_{2}}).

The operator Dχ|2D_{\chi}|_{\mathcal{E}_{2}} is essentially self-adjoint on the initial domain of compactly supported smooth sections by Proposition 5.5 in [15], and its square has a positive lower bound in the Hilbert C(G)C^{*}(G)-module sense. Thus its self-adjoint closure is invertible, so that the second term in (5.13) is homotopic to a degenerate cycle. The first term represents indexG(DS+N)\operatorname{index}_{G}(D^{S^{N}_{+}}). ∎

Remark 5.8.

A similar argument in the case where GG is trivial is hinted at below Lemma 4.1 in [25].

6. Proofs of results on positive scalar curvature

6.1. Obstruction results

We now deduce Theorem 2.1 from Theorem 3.4, and Corollary 2.4 from Theorem 2.1 and index theorems in [24, 40, 41].

Proof of Theorem 2.1.

This proof is an adaptation of the proof of Theorem 2.1 in [2].

First suppose that MM is odd-dimensional. Let κ\kappa denote scalar curvature. Let KX¯K\subset\overline{X} be a cocompact subset of MM such that NKN\subset K, κ>0\kappa>0 on KK, and the distance from XKX\setminus K to YY is positive. Let χC(M)G\chi\in C^{\infty}(M)^{G} be a function such that χ(x)=1\chi(x)=1 for all xYx\in Y and χ(x)=1\chi(x)=-1 for all xXKx\in X\setminus K. Consider the operator DD as in (3.5), where D0D_{0} is the Spin\operatorname{Spin}-Dirac operator on MM, and the admissible endomorphism Φ\Phi as in (3.6), where Φ0\Phi_{0} is pointwise multiplication by χ\chi. Then the set MM_{-} in Subsection 3.2 can be chosen to be cocompact as required, and so that N=NHN=N^{-}\cup H, where f|N=1f|_{N^{-}}=-1. In this setting,

(6.1) S+N=S0|H,S^{N}_{+}=S_{0}|_{H},

where S0MS_{0}\to M is the spinor bundle. (This is consistent with Corollary 7.3.)

For any λ\lambda\in\mathbb{R}, Lichnerowicz’ formula implies that

(D0±iλχ)2=D02±iλc(dχ)+λ2χ2κ/4λdχ+λ2χ2.(D_{0}\pm i\lambda\chi)^{2}=D_{0}^{2}\pm i\lambda c(d\chi)+\lambda^{2}\chi^{2}\geq\kappa/4-\lambda\|d\chi\|+\lambda^{2}\chi^{2}.

On MKM\setminus K, the function on the right-hand side equals κ/4+λ2λ2\kappa/4+\lambda^{2}\geq\lambda^{2}. Since κ\kappa is GG-invariant, and positive on the cocompact set KK, it has a positive lower bound on that set. Further, dχd\chi is GG-invariant and cocompactly supported, hence bounded. So we can choose λ>0\lambda>0 small enough so that κ/4λdχ>0\kappa/4-\lambda\|d\chi\|>0 on KK. It follows that (D0±iλχ)2(D_{0}\pm i\lambda\chi)^{2} has a positive lower bound. This implies that D+ΦD+\Phi is invertible, so indexG(D+Φ)=0\operatorname{index}_{G}(D+\Phi)=0. The claim then follows by Theorem 3.4 and (6.1).

If MM is even-dimensional, then the claim follows by applying the result in the odd-dimensional case to the manifold M×S1M\times S^{1}. ∎

Proof of Corollary 2.4.

This follows from Theorem 2.1 and index formulas for traces defined by orbital integrals applied to indexG(DH)\operatorname{index}_{G}(D^{H}). These index formulas are:

  • Theorem 6.10 in [41] if GG is any locally compact group and g=eg=e;

  • Theorem 6.1 in [40] if GG is discrete and finitely generated and gg is any element;

  • Proposition 4.11 in [24] if GG is a connected semisimple Lie group and gg is a semisimple element.

These results imply that in the setting of Corollary 2.4,

τg(indexG(DH))=A^g(H),\tau_{g}(\operatorname{index}_{G}(D^{H}))=\hat{A}_{g}(H),

for a trace τg\tau_{g}. As indexG(DH)\operatorname{index}_{G}(D^{H}) is independent of the choice of Riemannian metric, so is A^g(H)\hat{A}_{g}(H). Finally, vanishing of indexG(DH)\operatorname{index}_{G}(D^{H}) implies vanishing of A^g(H)\hat{A}_{g}(H) for all gg as above. ∎

6.2. Existence result

In the remainder of this section, we prove Theorem 2.9 by generalising a construction by Lawson and Yau [27]. We first prove in this subsection an extension of Theorem 3.8 in [27], namely Proposition 6.1.

To prepare, let us recall the steps in the construction of Lawson and Yau’s positive scalar curvature metrics [27], which we denote by g~t\tilde{g}_{t}, on a compact manifold NN.

Let KK be a compact Lie group acting on NN. Consider the principal KK-bundle defined by the map

p:K×NN,(k,y)k1y.p\colon K\times N\rightarrow N,\qquad(k,y)\mapsto k^{-1}y.

Take a KK-invariant Riemannian metric gNg_{N} on NN. Let bb be a bi-invariant Riemannian metric on KK. Let g^\hat{g} denote the lift of gNg_{N} to the orthogonal complement to ker(Tp)\ker(Tp) in T(K×N)T(K\times N) with respect to the product metric bgNb\oplus g_{N} on K×NK\times N.

For each t>0t>0, let b^t2\hat{b}_{t^{2}} be the lift of the metric t2bt^{2}b on KK to TK×NT(K×N)TK\times N\subset T(K\times N). Then

gtg^b^t2g_{t}\coloneqq\hat{g}\oplus\hat{b}_{t^{2}}

is a Riemannian metric on the total space K×NK\times N. One can check that, for each tt, gtg_{t} is invariant under the left KK-action on K×NK\times N defined by

l(k,y)=(kl1,y).l\cdot(k,y)=(kl^{-1},y).

Thus gtg_{t} descends, via the projection onto the second factor,

π:K×NN,(k,y)y,\pi\colon K\times N\rightarrow N,\qquad(k,y)\mapsto y,

to a KK-invariant metric g~t\tilde{g}_{t} on NN. Further, one sees that π\pi is a Riemannian submersion with respect to the metrics gtg_{t} and g~t\tilde{g}_{t} on the total space and base respectively.

The following proposition shows that, under the conditions stated, for all sufficiently small tt, g~t\tilde{g}_{t} has positive scalar curvature outside a neighbourhood of the fixed point set. It is an adaptation of the proof of Theorem 3.8 in [27] to the more general setting when NN is non-compact but has KK-bounded geometry; see Definition 2.8.

Proposition 6.1.

Let NN be a manifold with an action by a non-abelian, compact, connected Lie group KK. Fix a bi-invariant metric on KK. If gNg_{N} is a KK-invariant Riemannian metric on NN with KK-bounded geometry, then for any neighbourhood UU of the fixed point set NKN^{K}, there exists tU>0t_{U}>0 such that for all ttUt\leq t_{U}, the metric g~t\tilde{g}_{t} constructed above has uniform positive scalar curvature on NUN\setminus U.

Proof.

We will follow the steps in the proof of Theorem 3.8 in [27] and show where the assumptions of bounded geometry and no shrinking orbits (Definition 2.5) are needed to obtain the conclusion.

For each yNNKy\in N\setminus N^{K}, there is an orthogonal splitting 𝔨=𝔨y𝔭y\mathfrak{k}=\mathfrak{k}_{y}\oplus\mathfrak{p}_{y}, where 𝔨y\mathfrak{k}_{y} is the Lie subalgebra of the isotropy subgroup KyK_{y} of yy. The map φy\varphi_{y} from (2.1) restricts to an injection on 𝔭y\mathfrak{p}_{y}. Denote the orthogonal complement of φy(𝔭y)\varphi_{y}(\mathfrak{p}_{y}) in TyNT_{y}N by VyV_{y}. Then

T(e,y)(K×N)𝔨y𝔭yφy(𝔭y)Vy.T_{(e,y)}(K\times N)\cong\mathfrak{k}_{y}\oplus\mathfrak{p}_{y}\oplus\varphi_{y}(\mathfrak{p}_{y})\oplus V_{y}.

For each yNNKy\in N\setminus N^{K}, choose an orthonormal basis {e1(y),,ely(y)}\{e_{1}(y),\ldots,e_{l_{y}}(y)\} of 𝔭y\mathfrak{p}_{y} with respect to bb such that for all j,k=1,,lyj,k=1,\ldots,l_{y},

g(φy(ej(y)),φy(ek(y)))=σj2(y)δjkg\bigl{(}\varphi_{y}(e_{j}(y)),\varphi_{y}(e_{k}(y))\bigr{)}=\sigma^{2}_{j}(y)\delta_{jk}

for some continuous, positive functions σj\sigma_{j}. For each j=1,,lyj=1,\ldots,l_{y}, define a function λj:NNK(0,)\lambda_{j}\colon N\setminus N^{K}\to(0,\infty) by

λj(y)σj(y)(1+σj(y)2).\lambda_{j}(y)\coloneqq\sigma_{j}(y)(1+\sigma_{j}(y)^{2}).

By the calculations in the proofs of Propositions 3.6 and 3.7 in [27] and the assumption that gNg_{N} has bounded geometry, for any neighbourhood UU of NKN^{K}, the scalar curvature of g~t\tilde{g}_{t} at any yNUy\in N\setminus U is bounded below by

(6.2) j,k=1ly1t2λj(y)2λk(y)2(t2+λj(y)2)(t2+λk(y)2)[ej(y),ek(y)]b2+O(1)\sum_{j,k=1}^{l_{y}}\frac{1}{t^{2}}\frac{\lambda_{j}(y)^{2}\lambda_{k}(y)^{2}}{(t^{2}+\lambda_{j}(y)^{2})(t^{2}+\lambda_{k}(y)^{2})}\|[e_{j}(y),e_{k}(y)]\|_{b}^{2}+O(1)

as t0t\rightarrow 0, where the O(1)O(1) term is independent of yy. Since the KK-action has no shrinking orbits with respect to gNg_{N}, there exists cU>0c_{U}>0 such that λj(y)>cU\lambda_{j}(y)>c_{U} for each yNUy\in N\setminus U and j=1,,lyj=1,\ldots,l_{y}. In particular, for tcUt\leq c_{U}, the expression (6.2) is bounded below by

(6.3) j,k=1ly14t2[ej(y),ek(y)]b2+O(1).\sum_{j,k=1}^{l_{y}}\frac{1}{4t^{2}}\|[e_{j}(y),e_{k}(y)]\|_{b}^{2}+O(1).

Now, without loss of generality we may assume that K=SU(2)K=\operatorname{SU}(2) or K=SO(3)K=\operatorname{SO}(3), as any compact, connected, non-abelian Lie group has such a subgroup. Since KK has no subgroups of codimension 11, we have ly=dim𝔭y2l_{y}=\dim\mathfrak{p}_{y}\geq 2 at each yNNKy\in N\setminus N^{K}. For all jj and kk, [ej(y),ek(y)]b2\|[e_{j}(y),e_{k}(y)]\|_{b}^{2} is 4 times the sectional curvature of the plane spanned by ej(y)e_{j}(y) and ek(y)e_{k}(y) with respect to the metric bb, and this is constant in yy, and positive for K=SU(2)K=\operatorname{SU}(2) or K=SO(3)K=\operatorname{SO}(3). Thus for any neighbourhood UU of NKN^{K}, there exists tU>0t_{U}>0 such that for all ttUt\leq t_{U}, the expression (6.3), and hence also (6.2), is uniformly positive outside UU. It follows that for all such tt, the scalar curvature of g~t\tilde{g}_{t} is uniformly positive outside UU. ∎

We now deduce Theorem 2.9 from the following noncompact generalisation of the main result in [27].

Theorem 6.2.

Let NN be a manifold that admits an effective action by a compact, connected, non-abelian Lie group KK, such that the fixed point set NKN^{K} is compact. If there exists a KK-invariant Riemannian metric on NN such that the KK-action has KK-bounded geometry, then NN admits a KK-invariant metric with uniformly positive scalar curvature.

Proof.

Since NKN^{K} is compact and the action is effective, by Section 4 of [27] there exists t0>0t_{0}>0, a KK-invariant neighbourhood UU of KK with compact closure, and a KK-invariant Riemannian metric gg^{\prime} on NN such that each metric g~t\tilde{g}_{t}^{\prime}, constructed from gg^{\prime} as in subsection 6.2, has positive scalar curvature on UU for 0<t<t00<t<t_{0}.

Fix a bi-invariant metric bb on KK. Let g′′g^{\prime\prime} be a KK-invariant metric on NN for which the KK-action has KK-bounded geometry. Let {f1,f2}\{f_{1},f_{2}\} be a smooth, KK-invariant partition of unity on NN such that f11f_{1}\equiv 1 on UU and f10f_{1}\equiv 0 on NUN\setminus U^{\prime}, where UU^{\prime} is a relatively compact neighbourhood of NKN^{K} containing the closure of UU. Then

gNf1g+f2g′′g_{N}\coloneqq f_{1}g^{\prime}+f_{2}g^{\prime\prime}

is a KK-invariant Riemannian metric on NN. Applying the prescription in Subsection 6.2 to gNg_{N}, we obtain a family {g~t}t>0\{\tilde{g}_{t}\}_{t>0} of KK-invariant metrics on NN. We claim that for sufficiently small tt, g~t\tilde{g}_{t} has uniformly positive scalar curvature on NN.

To see this, let φ\|\varphi\| be the norm function (2.2) associated to the metric gNg_{N}. Since g′′g^{\prime\prime} and gNg_{N} coincide on NUN\setminus U^{\prime}, and g′′g^{\prime\prime} has KK-bounded geometry, there exists CU>0C_{U^{\prime}}>0 such that φ(y)CU\|\varphi\|(y)\geq C_{U^{\prime}} for all yNUy\in N\setminus U^{\prime}. One sees that gNg_{N} has KK-bounded geometry, and so by Proposition 6.1, there exists t1=(t1)U>0t_{1}=(t_{1})_{U}>0 such that for all tt1t\leq t_{1}, g~t\tilde{g}_{t} has uniformly positive scalar curvature on NUN\setminus U. It follows that for all tmin{t0,t1}t\leq\min\{t_{0},t_{1}\}, g~t\tilde{g}_{t} has uniformly positive scalar curvature on NN. ∎

Proof of Theorem 2.9..

In the setting of Theorem 2.9, Theorem 6.2 implies that NN admits a KK-invariant metric with uniformly positive scalar curvature. By Theorem 4.6 in [17] (see also Theorem 58 in [18]), this metric induces a GG-invariant Riemannian metric on G×KNG\times_{K}N of uniformly positive scalar curvature. ∎

7. Further applications of the Callias-type index theorem

We used Theorem 3.4 to prove Theorem 2.1 in Subsection 6.1. We give some other applications of Theorem 3.4 here.

7.1. The image of the assembly map

If M/GM/G is noncompact, and GG is not known to satisfy Baum–Connes surjectivity, then it is a priori unclear if indexG(D+Φ)\operatorname{index}_{G}(D+\Phi) lies in the image of the Baum–Connes assembly map [4]; see the question raised on page 3 of [15]. Theorem 3.4 implies that this is in fact the case for GG-Callias-type operators as defined above:

Corollary 7.1.

The Callias-index of D+ΦD+\Phi lies in the image of the Baum–Connes assembly map.

7.2. Cobordism invariance of the assembly map

Theorem 3.4 leads to a perspective on cobordism invariance of the analytic assembly map.

Corollary 7.2.

Let XX be an odd-dimensional Riemannian manifold with boundary NN, on which GG acts properly and isometrically, preserving NN, such that M/GM/G is compact. Suppose that a neighbourhood UU of NN is GG-equivariantly isometric to N×[0,ε)N\times[0,\varepsilon), for some ε>0\varepsilon>0. Let S0XXS^{X}_{0}\to X be a GG-equivariant Clifford module, and consider the Clifford module S0X|NNS^{X}_{0}|_{N}\to N, graded by ii times Clifford multiplication by the inward-pointing unit normal. Suppose that S0X|US0X|N×[0,ε)S^{X}_{0}|_{U}\cong S^{X}_{0}|_{N}\times[0,\varepsilon). Let DND^{N} be the Dirac operator on S0X|NS^{X}_{0}|_{N} associated to a GG-invariant Clifford connection N\nabla^{N} for the Clifford action by TNTN on S0X|NS^{X}_{0}|_{N}. Then indexG(DN)=0\operatorname{index}_{G}(D^{N})=0.

Proof.

Form the manifold MM by attaching the cylinder N×(0,)N\times(0,\infty) to XX along UU. Extend S0XS^{X}_{0} and the Clifford action to MM in the natural way. We write S0S_{0} for the extension of S0XS^{X}_{0} to MM. The connection N\nabla^{N} pulls back to a Clifford connection on S0|N×(0,)S_{0}|_{N\times(0,\infty)}. Because the GG-invariant Clifford connections form an affine space, we can extend this pulled-back connection to all of MM using a GG-invariant Clifford connection on XNX\setminus N and a partition of unity. Let D0D_{0} be the associated Dirac operator.

Let Φ0\Phi_{0} be the identity endomorphism on S0S_{0}. Then Φ\Phi, as in (3.6) is admissible for DD as in (3.5), and (3.1) holds on all of MM. By Lemma 4.7, this implies that indexG(D+Φ)=0\operatorname{index}_{G}(D+\Phi)=0. In this case, S+N=S0X|NS^{N}_{+}=S^{X}_{0}|_{N}, so by Theorem 3.4, indexG(DN)=0\operatorname{index}_{G}(D^{N})=0. ∎

7.3. Spinc\operatorname{Spin}^{c}-Dirac operators

Corollary 7.3.

Consider the setting of Theorem 3.4, and assume that MM is odd-dimensional and D0D_{0} is a Spinc\operatorname{Spin}^{c}-Dirac operator. Then there is a union N+N^{+} of connected components of NN and a Spinc\operatorname{Spin}^{c}-structure on N+N^{+} with spinor bundle S0|N+S_{0}|_{N^{+}} such that

(7.1) indexG(D+Φ)=indexG(DS0|N+)K0(C(G)),\operatorname{index}_{G}(D+\Phi)=\operatorname{index}_{G}(D^{S_{0}|_{N^{+}}})\quad\in K_{0}(C^{*}(G)),

where DS0|N+D^{S_{0}|_{N^{+}}} is a Spinc\operatorname{Spin}^{c}-Dirac operator on the spinor bundle S0|N+N+S_{0}|_{N^{+}}\to N^{+}. If NN is connected, then indexG(D+Φ)=0\operatorname{index}_{G}(D+\Phi)=0.

Proof.

Since S0S_{0} is an irreducible Clifford module, and S+NS0|NS^{N}_{+}\subset S_{0}|_{N} is invariant under the Clifford action of TM|NTM|_{N}, over each connected component XX of NN the bundle S+N|XS^{N}_{+}|_{X} is either zero or S0|XS_{0}|_{X}. Since MM is odd-dimensional, S0|XS_{0}|_{X} is the spinor bundle of the Spinc\operatorname{Spin}^{c}-structure of XX that it inherits from MM. So (7.1) follows.

If NN is connected, then we either have N+=N^{+}=\emptyset, in which case indexG(D+Φ)=0\operatorname{index}_{G}(D+\Phi)=0 because S+NS^{N}_{+} is the zero bundle, or N+=NN^{+}=N, in which case (7.1) and Lemma 4.7 imply that

indexG(D+Φ)=indexG(D+(0ii0))=0.\operatorname{index}_{G}(D+\Phi)=\operatorname{index}_{G}\left(D+\begin{pmatrix}0&i\\ -i&0\end{pmatrix}\right)=0.

There is a converse to Corollary 7.3 in the following sense. Let N+NN^{+}\subset N be any union of connected components; there is a finite number of such subsets of NN since N/GN/G is compact. We can define an admissible endomorphism Φ0\Phi_{0} such that (7.1) holds, by taking Φ0\Phi_{0} to be multiplication by a GG-invariant function on MM that equals 11 on N+N^{+} and 1-1 on NN+N\setminus N^{+}, and is constant 11 or 1-1 outside a cocompact set. Thus given any hypersurface NN bounding a cocompact set, and any set N+N^{+} of connected components of NN, we have an index

(7.2) indexGN+(D):=indexG(D+Φ),\operatorname{index}_{G}^{N^{+}}(D):=\operatorname{index}_{G}(D+\Phi),

with Φ\Phi and Φ0\Phi_{0} related as in (3.6), independent of the choice of Φ0\Phi_{0} with the property that Φ0\Phi_{0} is positive definite on N+N^{+} and negative definite on NN+N\setminus N^{+}.

Versions of the index (7.2) are sometimes used in applications of Callias-type index theorems to obstructions to positive scalar curvature for Spin\operatorname{Spin}-manifolds, see [2, 9] and the proof of Theorem 2.1 in Subsection 6.1.

7.4. Induction

Suppose that GG is an almost connected, reductive Lie group, and let K<GK<G be maximal compact. In [18, 19, 24] some results were proved relating GG-equivariant indices to KK-equivariant indices via Dirac induction. Such results allow one to deduce results in equivariant index theory for actions by noncompact groups from corresponding results for compact groups. This was applied to obtain results in geometric quantisation [19, 20, 22] and geometry of group actions [18, 21]. Corollary 7.4 below is a version of this idea for the index of Definition 3.3.

To state this corollary, we consider the setting of Subsection 3.2. Using Abels’ slice theorem, we write M=G×KYM=G\times_{K}Y, for a KK-invariant submanifold YMY\subset M, and S0=G×KS0|YS_{0}=G\times_{K}S_{0}|_{Y}. Let 𝔤=𝔨𝔭\mathfrak{g}=\mathfrak{k}\oplus\mathfrak{p} be a Cartan decomposition. Then TMG×K(TY𝔭)TM\cong G\times_{K}(TY\oplus\mathfrak{p}). We assume that the GG-invariant Riemannian metric on MM is induced by a KK-invariant Riemannian metric on YY and an Ad(K)\operatorname{Ad}(K)-invariant inner product on 𝔭\mathfrak{p} via this identification.

We assume for simplicity that the adjoint representation Ad:KSO(𝔭)\operatorname{Ad}\colon K\to\operatorname{SO}(\mathfrak{p}) lifts to the double cover Spin(𝔭)\operatorname{Spin}(\mathfrak{p}) of SO(𝔭)\operatorname{SO}(\mathfrak{p}). (This is true for a double cover of GG.) Then the standard Spin\operatorname{Spin} representation S𝔭S_{\mathfrak{p}} of Spin(𝔭)\operatorname{Spin}(\mathfrak{p}) may be viewed as a representation of KK. We assume that S0|Y=S0YS𝔭S_{0}|_{Y}=S_{0}^{Y}\otimes S_{\mathfrak{p}} for a Clifford module S0YYS_{0}^{Y}\to Y. The Clifford action by TM|YTM|_{Y} on S0|YS_{0}|_{Y} equals cY1S𝔭+1S0Yc𝔭c_{Y}\otimes 1_{S_{\mathfrak{p}}}+1_{S_{0}^{Y}}\otimes c_{\mathfrak{p}}, for a Clifford action cYc_{Y} by TYTY on S0YS_{0}^{Y}, and the Clifford action c𝔭c_{\mathfrak{p}} of 𝔭\mathfrak{p} on S𝔭S_{\mathfrak{p}}. We choose the GG-invariant connection S0\nabla^{S_{0}} so that S0|Y=S0Y1S𝔭+1S0YS𝔭\nabla^{S_{0}}|_{Y}=\nabla^{S_{0}^{Y}}\otimes 1_{S_{\mathfrak{p}}}+1_{S_{0}^{Y}}\otimes\nabla^{S_{\mathfrak{p}}}, for Clifford connections S0Y\nabla^{S_{0}^{Y}} on S0YS_{0}^{Y} and S𝔭\nabla^{S_{\mathfrak{p}}} on Y×S𝔭Y\times S_{\mathfrak{p}}

Since Φ0\Phi_{0} is GG-equivariant, it is determined by its restriction to YY, which is a KK-equivariant endomorphism of S0|YS_{0}|_{Y}. We assume that Φ0|Y=Φ0Y1𝔭\Phi_{0}|_{Y}=\Phi_{0}^{Y}\otimes 1_{\mathfrak{p}}, for a KK-equivariant endomorphism Φ0Y\Phi_{0}^{Y} of S0YS_{0}^{Y}. (What follows remains true if Φ0=Φ0Y1𝔭+1Φ0𝔭\Phi_{0}=\Phi_{0}^{Y}\otimes 1_{\mathfrak{p}}+1\otimes\Phi_{0}^{\mathfrak{p}} for an Ad(K)\operatorname{Ad}(K)-invariant endomorphism of S𝔭S_{\mathfrak{p}}, but this requires a small extra argument that we omit here.)

Consider the Dirac operator D0Y:=cYS0YD_{0}^{Y}:=c_{Y}\circ\nabla^{S_{0}^{Y}} on Γ(S0Y)\Gamma^{\infty}(S_{0}^{Y}). Form DYD^{Y} from D0YD_{0}^{Y} as in (3.5) and ΦY\Phi^{Y} from Φ0Y\Phi_{0}^{Y} as in (3.6). Let R(K)R(K) be the representation ring of KK and

(7.3) DIndKG:R(K)K(C(G))\operatorname{D-Ind}_{K}^{G}\colon R(K)\to K_{*}(C^{*}(G))

be the Dirac induction map [4].

Corollary 7.4.

The operator DY+ΦYD^{Y}+\Phi^{Y} is a KK-equivariant Callias-type operator, and

indexG(D+Φ)=DIndKG(indexK(DY+ΦY)).\operatorname{index}_{G}(D+\Phi)=\operatorname{D-Ind}_{K}^{G}(\operatorname{index}_{K}(D^{Y}+\Phi^{Y})).
Proof.

Theorem 3.4 implies that indexG(D+Φ)=indexG(DS+N)\operatorname{index}_{G}(D+\Phi)=\operatorname{index}_{G}(D^{S^{N}_{+}}). Write YN:=YNY^{N}:=Y\cap N, so that N=G×KYNN=G\times_{K}Y^{N}. Then YNY^{N} is a compact manifold. Define DS+YND^{S^{Y^{N}}_{+}} analogously to DYD^{Y}. The induction result for cocompact actions, Theorem, 4.5 in [19], Theorem 5.3 in [24] or Theorem 46 in [18], implies that

indexG(DS+N)=DIndKG(indexK(DS+YN)).\operatorname{index}_{G}(D^{S^{N}_{+}})=\operatorname{D-Ind}_{K}^{G}(\operatorname{index}_{K}(D^{S^{Y^{N}}_{+}})).

Another application of Theorem 3.4, now with GG replaced by KK, or Theorem 1.5 in [2] with a compact group action added, shows that indexK(DS+YN)=indexK(DY+ΦY)\operatorname{index}_{K}(D^{S^{Y^{N}}_{+}})=\operatorname{index}_{K}(D^{Y}+\Phi^{Y}). ∎

7.5. Callias quantisation commutes with reduction

Theorem 3.11 in [16] is a quantisation commutes with reduction result for the equivariant index of Spinc\operatorname{Spin}^{c}-Callias-type operators. This result applies to reduction at the trivial representation of GG; i.e. to an index defined in terms of GG-invariant sections of SS. Using Theorem 3.4, we can generalise this result to reduction at more general representations, or more precisely, at arbitrary generators of K0(Cr(G))K_{0}(C^{*}_{r}(G)). Furthermore, this result is ‘exact’ rather than asymptotic as Theorem 3.11 in [16], in the sense that one does not need to consider high powers of a line bundle.

In the setting of Subsection 3.2, we now assume that MM is odd-dimensional, and that S0S_{0} is the spinor bundle for a GG-equivariant Spinc\operatorname{Spin}^{c}-structure. Let D0D_{0} be the Spinc\operatorname{Spin}^{c}-Dirac operator on Γ(S0)\Gamma(S_{0}), defined by the Clifford connection corresponding to a connection L\nabla^{L} on the determinant line bundle LL.

The Spinc\operatorname{Spin}^{c}-moment map associated to L\nabla^{L} is the map μ:M𝔤\mu\colon M\to\mathfrak{g}^{*} such that for all X𝔤X\in\mathfrak{g},

2πiμ,X=XXML,End(L)=C(M,),2\pi i\langle\mu,X\rangle=\mathcal{L}_{X}-\nabla^{L}_{X^{M}},\quad\in\operatorname{End}(L)=C^{\infty}(M,\mathbb{C}),

where X\mathcal{L}_{X} denotes the Lie derivative with respect to XX, and XMX^{M} is the vector field induced by XX. The reduced space at an element ξ𝔤\xi\in\mathfrak{g}^{*} is defined as Mξ:=μ1(ξ)/GξM_{\xi}:=\mu^{-1}(\xi)/G_{\xi}, where GξG_{\xi} is the stabiliser of ξ\xi. This reduced space is noncompact in general, and may not be smooth. But the reduced space Nξ:=(μ1(ξ)N)/GξN_{\xi}:=(\mu^{-1}(\xi)\cap N)/G_{\xi} is compact. It is not always a smooth manifold, but if it is, and ξ𝔨\xi\in\mathfrak{k}^{*}, then we have an identification NξYξN:=(μ1(ξ)YN)/KξN_{\xi}\cong Y^{N}_{\xi}:=(\mu^{-1}(\xi)\cap Y\cap N)/K_{\xi}, with YY and YNY^{N} as in Subsection 7.4, including Spinc\operatorname{Spin}^{c}-structures. See Propositions 3.13 and 3.14 in [22]. Let N+NN^{+}\subset N be as in Corollary 7.3. Then we similarly have Nξ+YξN+:=(μ1(ξ)YN+)/KξN^{+}_{\xi}\cong Y^{N^{+}}_{\xi}:=(\mu^{-1}(\xi)\cap Y\cap N^{+})/K_{\xi} in the smooth case, including Spinc\operatorname{Spin}^{c}-structures.

There is a nontrivial way to define a Spinc\operatorname{Spin}^{c}-quantisation QSpinc(YξN+)Q^{\operatorname{Spin}^{c}}(Y^{N^{+}}_{\xi})\in\mathbb{Z}, even when YξN+Y^{N^{+}}_{\xi} is not smooth, described in detail in Section 5.1 of [31]. Motivated by the identification Nξ+YξN+N^{+}_{\xi}\cong Y^{N^{+}}_{\xi} in the smooth case, we define QSpinc(Nξ+):=QSpinc(YξN+)Q^{\operatorname{Spin}^{c}}(N^{+}_{\xi}):=Q^{\operatorname{Spin}^{c}}(Y^{N^{+}}_{\xi}) for ξ𝔨\xi\in\mathfrak{k}^{*}.

Let T<KT<K be a maximal torus, and fix a positive root system for (K,T)(K,T). Let VK^V\in\hat{K} have highest weight λi𝔱𝔨𝔤\lambda\in i\mathfrak{t}^{*}\hookrightarrow\mathfrak{k}^{*}\hookrightarrow\mathfrak{g}^{*}. (The first inclusion is defined by 𝔱(𝔨)Ad(T)\mathfrak{t}^{*}\cong(\mathfrak{k}^{*})^{\operatorname{Ad}^{*}(T)}, the second by the Cartan decomposition.) Following [31, 32], we call an element ξ𝔨\xi\in\mathfrak{k}^{*} an ancestor of VV if the coadjoint orbit Ad(K)ξ\operatorname{Ad}^{*}(K)\xi is admissible in the sense of [32], and its KK-equivariant Spinc\operatorname{Spin}^{c}-quantisation is VV. There exists a finite set A(V)A(V) of ancestors representing all different such coadjoint orbits.

Let Cr(G)C^{*}_{r}(G) be the reduced group CC^{*}-algebra of GG and DIndKG\operatorname{D-Ind}_{K}^{G} the Dirac induction map (7.3). By the Connes–Kasparov conjecture, proved in [10, 26, 42], the abelian group K(Cr(G))K_{*}(C^{*}_{r}(G)) is free, with generators DIndKG[V]\operatorname{D-Ind}_{K}^{G}[V], where VV runs over K^\hat{K}.

Recall the definition of the Callias index of Spinc\operatorname{Spin}^{c}-Dirac operators (7.2).

Corollary 7.5 (Callias quantisation commutes with reduction).

We have

(7.4) indexGN+(D)=VK^ξA(V)QSpinc(Nξ+)DIndKG[V]K(Cr(G)).\operatorname{index}_{G}^{N^{+}}(D)=\bigoplus_{V\in\hat{K}}\sum_{\xi\in A(V)}Q^{\operatorname{Spin}^{c}}(N^{+}_{\xi})\operatorname{D-Ind}_{K}^{G}[V]\quad\in K_{*}(C^{*}_{r}(G)).
Proof.

By Corollary 7.3, indexGN+(D)=indexG(DS0|N+)\operatorname{index}_{G}^{N^{+}}(D)=\operatorname{index}_{G}(D^{S_{0}|_{N^{+}}}), where now DS0|N+D^{S_{0}|_{N^{+}}} is a Spinc\operatorname{Spin}^{c}-Dirac operator on N+N^{+}. Theorem 4.6 in [22] implies that indexG(DS0|N+)\operatorname{index}_{G}(D^{S_{0}|_{N^{+}}}) equals

VK^ξA(V)QSpinc(Nξ+)DIndKG[V].\bigoplus_{V\in\hat{K}}\sum_{\xi\in A(V)}Q^{\operatorname{Spin}^{c}}(N^{+}_{\xi})\operatorname{D-Ind}_{K}^{G}[V].

Remark 7.6.

In cases where MξM_{\xi} is smooth and NξN_{\xi} is a hypersurface in MξM_{\xi}, which is a transversality condition between NN and μ1(ξ)\mu^{-1}(\xi), one can use Theorem 1.5 in [2] (more precisely, its special case for Spinc\operatorname{Spin}^{c}-Dirac operators, which is the non-equivariant case of Corollary 7.3) to express the Spinc\operatorname{Spin}^{c}-quantisation QSpinc(Nξ+)Q^{\operatorname{Spin}^{c}}(N^{+}_{\xi}) as the index of a Callias-type operator on MξM_{\xi}.

References

  • [1] Herbert Abels. Parallelizability of proper actions, global KK-slices and maximal compact subgroups. Math. Ann., 212:1–19, 1974/75.
  • [2] Nicolae Anghel. On the index of Callias-type operators. Geom. Funct. Anal., 3(5):431–438, 1993.
  • [3] Michael Atiyah and Friedrich Hirzebruch. Spin-manifolds and group actions. In Essays on Topology and Related Topics (Mémoires dédiés à Georges de Rham), pages 18–28. Springer, New York, 1970.
  • [4] 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. American Mathematical Society, Providence, RI, 1994.
  • [5] Raoul Bott and Robert Seeley. Some remarks on the paper of Callias: “Axial anomalies and index theorems on open spaces” [Comm. Math. Phys. 62 (1978), no. 3, 213–234; MR 80h:58045a]. Comm. Math. Phys., 62(3):235–245, 1978.
  • [6] Maxim Braverman and Simone Cecchini. Callias-type operators in von Neumann algebras. J. Geom. Anal., 28(1):546–586, 2018.
  • [7] Ulrich Bunke. A KK-theoretic relative index theorem and Callias-type Dirac operators. Math. Ann., 303(2):241–279, 1995.
  • [8] Constantine Callias. Axial anomalies and index theorems on open spaces. Comm. Math. Phys., 62(3):213–234, 1978.
  • [9] Simone Cecchini. Callias-type operators in C{C}^{*}-algebras and positive scalar curvature on noncompact manifolds. ArXiv:1611.01800, 2016.
  • [10] Jérôme Chabert, Siegfried Echterhoff, and Ryszard Nest. The Connes-Kasparov conjecture for almost connected groups and for linear pp-adic groups. Publ. Math. Inst. Hautes Études Sci., (97):239–278, 2003.
  • [11] Johannes Ebert. Elliptic regularity for Dirac operators on families of noncompact manifolds. ArXiv:1608.01699, 2018.
  • [12] Benyin Fu, Xianjin Wang, and Guoliang Yu. The equivariant coarse novikov conjecture and coarse embedding. ArXiv:1909.00529, 2019.
  • [13] Mikhael Gromov and H. Blaine Lawson, Jr. Positive scalar curvature and the Dirac operator on complete Riemannian manifolds. Inst. Hautes Études Sci. Publ. Math., (58):83–196 (1984), 1983.
  • [14] Victor Guillemin and Schlomo Sternberg. Geometric quantization and multiplicities of group representations. Invent. Math., 67(3):515–538, 1982.
  • [15] Hao Guo. Index of equivariant Callias-type operators and invariant metrics of positive scalar curvature. ArXiv:1803.05558, 2018.
  • [16] Hao Guo, Peter Hochs, and Varghese Mathai. Coarse equivariant callias index theory and quantisation. in preparation.
  • [17] Hao Guo, Peter Hochs, and Varghese Mathai. Equivariant Callias index theory via coarse geometry. ArXiv:1902.07391, 2019.
  • [18] Hao Guo, Varghese Mathai, and Hang Wang. Positive scalar curvature and Poincaré duality for proper actions. J. Noncommut. Geometry, 2019. DOI:10.4171/JNCG/321, ArXiv:1609.01404.
  • [19] Peter Hochs. Quantisation commutes with reduction at discrete series representations of semisimple groups. Adv. Math., 222(3):862–919, 2009.
  • [20] Peter Hochs. Quantisation of presymplectic manifolds, KK-theory and group representations. Proc. Amer. Math. Soc., 143(6):2675–2692, 2015.
  • [21] Peter Hochs and Varghese Mathai. Spin-structures and proper group actions. Adv. Math., 292:1–10, 2016.
  • [22] Peter Hochs and Varghese Mathai. Quantising proper actions on Spinc-manifolds. Asian J. Math., 21(4):631–685, 2017.
  • [23] Peter Hochs, Yanli Song, and Xiang Tang. An index theorem for higher orbital integrals. in preparation, 2020.
  • [24] Peter Hochs and Hang Wang. A fixed point formula and Harish-Chandra’s character formula. Proc. London Math. Soc., 00(3):1–32, 2017.
  • [25] Dan Kucerovsky. A short proof of an index theorem. Proc. Amer. Math. Soc., 129(12):3729–3736, 2001.
  • [26] Vincent Lafforgue. K{K}-théorie bivariante pour les algèbres de Banach et conjecture de Baum-Connes. Invent. Math., 149(1):1–95, 2002.
  • [27] H. Blaine Lawson, Jr. and Shing Tung Yau. Scalar curvature, non-abelian group actions, and the degree of symmetry of exotic spheres. Comment. Math. Helv., 49:232–244, 1974.
  • [28] Eckhard Meinrenken. Symplectic surgery and the Spinc{\rm Spin}^{c}-Dirac operator. Adv. Math., 134(2):240–277, 1998.
  • [29] Alexandr S. Miščenko and Anatoliǐ T. Fomenko. The index of elliptic operators over CC^{\ast}-algebras. Izv. Akad. Nauk SSSR Ser. Mat., 43(4):831–859, 967, 1979.
  • [30] Paul-Émile Paradan. Localization of the Riemann-Roch character. J. Funct. Anal., 187(2):442–509, 2001.
  • [31] Paul-Émile Paradan and Michèle Vergne. Equivariant Dirac operators and differentiable geometric invariant theory. Acta Math., 218(1):137–199, 2017.
  • [32] Paul-Émile Paradan and Michèle Vergne. Admissible coadjoint orbits for compact Lie groups. Transform. Groups, 23(3):875–892, 2018.
  • [33] Markus Pflaum, Hessel Posthuma, and Xiang Tang. The transverse index theorem for proper cocompact actions of Lie groupoids. J. Differential Geom., 99(3):443–472, 2015.
  • [34] Paolo Piazza and Hessel Posthuma. Higher genera for proper actions of Lie groups. ArXiv:1801.06676, 2018.
  • [35] Jonathan Rosenberg. CC^{\ast}-algebras, positive scalar curvature, and the Novikov conjecture. Inst. Hautes Études Sci. Publ. Math., (58):197–212 (1984), 1983.
  • [36] Jonathan Rosenberg. CC^{\ast}-algebras, positive scalar curvature and the Novikov conjecture. II. In Geometric methods in operator algebras (Kyoto, 1983), volume 123 of Pitman Res. Notes Math. Ser., pages 341–374. Longman Sci. Tech., Harlow, 1986.
  • [37] Jonathan Rosenberg. CC^{\ast}-algebras, positive scalar curvature, and the Novikov conjecture. III. Topology, 25(3):319–336, 1986.
  • [38] Thomas Schick and Mostafa Esfahani Zadeh. Large scale index of multi-partitioned manifolds. J. Noncommut. Geom., 12(2):439–456, 2018.
  • [39] Youliang Tian and Weiping Zhang. An analytic proof of the geometric quantization conjecture of Guillemin-Sternberg. Invent. Math., 132(2):229–259, 1998.
  • [40] Bai-Ling Wang and Hang Wang. Localized index and L2L^{2}-Lefschetz fixed-point formula for orbifolds. J. Differential Geom., 102(2):285–349, 2016.
  • [41] Hang Wang. L2L^{2}-index formula for proper cocompact group actions. J. Noncommut. Geom., 8(2):393–432, 2014.
  • [42] Antony Wassermann. Une démonstration de la conjecture de Connes-Kasparov pour les groupes de Lie linéaires connexes réductifs. C. R. Acad. Sci. Paris Sér. I Math., 304(18):559–562, 1987.