Positive scalar curvature and an equivariant Callias-type index theorem for proper actions
Abstract.
For a proper action by a locally compact group on a manifold with a -equivariant -structure, we obtain obstructions to the existence of complete -invariant Riemannian metrics with uniformly positive scalar curvature. We focus on the case where is noncompact. The obstructions follow from a Callias-type index theorem, and relate to positive scalar curvature near hypersurfaces in . We also deduce some other applications of this index theorem. If 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 -invariant Riemannian metrics with uniformly positive scalar curvature, under an equivariant bounded geometry assumption.
1. Introduction
Let be a locally compact group, acting properly on a manifold . Suppose that has a -equivariant -structure. The results in this paper are about the following question.
Question 1.1.
When does admit a complete, -invariant Riemannian metric with uniformly positive scalar curvature?
We are mainly interested in the case where is noncompact.
The literature on the non-equivariant case of Question 1.1 is too vast to summarise, but important work where is noncompact was done by Gromov and Lawson [13]. A more refined perspective on the non-equivariant case is to consider a manifold , and let be its universal cover, and is fundamental group. This allows one to construct obstructions to metrics of positive scalar curvature in terms of -equivariant index theory on , refining index-theoretic obstructions on . If is compact, this is the origin of the even more refined Rosenberg index [36, 35, 37], in -theory of the real maximal group -algebra of .
More generally, if is a discrete group not necessarily acting freely on , then is an orbifold, whence Question 1.1 becomes the question of whether admits an orbifold metric of positive scalar curvature.
We consider the case where is not necessarily discrete, and does not necessarily act freely. Results on this case of Question 1.1 in the case where is an almost-connected Lie group and is compact were obtained in [18, 34]. In this paper, we focus on the case where is noncompact.
Results on positive scalar curvature
We first obtain obstructions to -invariant Riemannian metrics with positive scalar curvature, both in the -theory of the maximal or reduced group -algebra of , and in terms of numerical topological invariants generalising the -genus. If is a connected Lie group, then these obstructions vanish under a mild assumption, as shown in [21]. In that case, we construct -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 be a -invariant, cocompact hypersurface with trivial normal bundle, that partitions into two open sets. If admits a complete, -invariant Riemannian metric with nonnegative scalar curvature, and positive scalar curvature in a neighbourhood of , then
for a -Dirac operator on .
See Theorem 2.1. In the case where is the universal cover of a manifold and is its fundamental group, this becomes Theorem A in [9].
Theorem 1.2 implies topological obstructions to -invariant Riemannian metrics with positive scalar curvature. Let , and let be the fixed point set of . Let be its normal bundle in , and let be the curvature of the Levi–Civita connection restricted to . The -localised -genus of is
for a cutoff function on . If acts freely, then if , and is the -genus of . In general, for example in the orbifold case, may be nonzero for different .
Corollary 1.3.
Consider the setting of Theorem 1.2. Suppose that either
-
•
is any locally compact group and ;
-
•
is discrete and finitely generated and is any element; or
-
•
is a connected semisimple Lie group and is a semisimple element.
Then is independent of the choice of -invariant Riemannian metric, and .
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 is connected, and not every stabiliser of its action on is maximal compact, then . This implies that as well. In view of Theorem 1.2 and Corollary 1.3, this makes it a natural question if admits a -invariant Riemannian metric with positive scalar curvature if 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 is a connected Lie group, and let be a maximal compact subgroup. Abels’ slice theorem [1] implies that there is a diffeomorphism , for a -invariant submanifold . Consider the infinitesimal action map
mapping to . The action by on is said to have no shrinking orbits with respect to a -invariant Riemannian metric on , if the pointwise operator norm of as a map from to a tangent space is uniformly positive outside a neighbourhood of the fixed point set . We say that has -bounded geometry if it has bounded geometry and no shrinking orbits.
Theorem 1.4.
Suppose that is non-abelian, and that acts effectively on with compact fixed point set. If there exists a -invariant Riemannian metric on for which has -bounded geometry, then the manifold admits a -invariant metric with uniformly positive scalar curvature.
In the compact case, the Atiyah–Hirzebruch vanishing theorem [3] implies that the obstruction to Riemannian metrics of positive scalar curvature vanishes if acts nontrivially on . 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 is for now, and letting be any locally compact group, we consider a -equivariant Dirac-type operator on a -equivariant vector bundle . A Callias-type operator is of the form , for a -equivariant endomorphism of such that is uniformly positive outside a cocompact set. Then this operator has an index , constructed in [15]. (See Theorem 4.2 in [17] for a realisation of this index in terms of coarse geometry.)
Theorem 1.5 (-Callias-type index theorem).
We have
for a Dirac operator on a -invariant, cocompact hypersurface .
See Theorem 3.4. Versions of this result where 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 , then under conditions, there is a subalgebra such that , and there is a well-defined trace on given by
where is the centraliser of . In various settings, including the three cases in Corollary 1.3, there are index formulas for the number , see [24, 41, 40]. These index formulas imply that, in the setting of Theorem 1.2,
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 to .
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 -Dirac operators (Corollary 7.3); on induction of Callias-type indices from compact groups to noncompact groups (Corollary 7.4); and on the -version [31] of the quantisation commutes with reduction problem [14, 28, 30, 39] for -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 7.1–7.5.
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 is mentioned, the inner product on that space will be denoted by , and the corresponding norm by . Spaces of continuous sections are denoted by ; spaces of smooth sections by . Subscripts denote compact supports.
If is a group, and is a subgroup acting on a set , then we write for the quotient of by the -action given by
for , and . If is a manifold, is Lie group and is a closed subgroup, then this action is proper and free, and 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, will be be a locally group, acting properly on a manifold . We do not assume that is compact and are in fact interested mainly in the case where it is not. The group 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 on a manifold , a -equivariant elliptic differential operator on has an equivariant index defined by the analytic assembly map [4]. Here is the maximal or reduced group -algebra of , and the index takes values in its even -theory if is odd with respect to a grading, and in odd -theory otherwise.
Our most general obstruction result is the following.
Theorem 2.1.
Let be a complete Riemannian -manifold, on which a locally compact group acts properly and isometrically. Let be a -invariant, cocompact hypersurface with trivial normal bundle, such that for disjoint open subsets . If the scalar curvature on is nonnegative, and positive in a neighbourhood of , then the -Dirac operator on , acting on sections of the restriction of the spinor bundle on to , satisfies
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.
Theorem 2.1 implies a set of topological obstructions to -invariant positive scalar curvature metrics on . Let be any Riemannian manifold on which acts properly, isometrically and cocompactly. Let , and let be its fixed point set. (Properness of the action implies that if is not contained in a compact subset of .) Let be the normal bundle to in . The connected components of are submanifolds of of possibly different dimensions, so the rank of may jump between these components. In what follows, we implicitly apply all constructions to the connected components of and add the results together.
By a cutoff function we will mean a smooth function such that has compact intersection with each -orbit, and for each we have
We will also use cutoff functions for other group actions, which are defined analogously.
Let be the curvature of the Levi–Civita connection restricted to . Let be the -class of . Let be the centraliser of . Let be a cutoff function for the action by on .
Definition 2.3.
The -localised -genus of is
Corollary 2.4.
Suppose that either
-
•
is any locally compact group and ;
-
•
is discrete and finitely generated and is any element;
-
•
is a connected semisimple Lie group and is a semisimple element.
Let be a manifold on which acts properly and that admits a -equivariant -structure. Let be a -invariant, cocompact hypersurface such that for disjoint open subsets . The localised -genus is independent of the choice of a Riemannian metric. If admits a complete, -invariant Riemannian metric whose scalar curvature is nonnegative, and positive in a neighbourhood of , then .
Theorem 2 in [21] is a generalisation of Atiyah and Hirzebruch’s vanishing theorem [3] to actions by noncompact groups. It states that if is a connected Lie group, and not all stabilisers of the action by on are maximal compact, then . 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 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 has infinitely many connected components. (The vanishing result generalises directly to the case where has finitely many connected components.) As noted in Remark 2.2, Theorem 2.1 implies Theorem A in [9], in the case where is the fundamental group of and is its universal cover. More generally, if is discrete, then is an orbifold. Then Theorem 2.1 and Corollary 2.4 lead to obstructions to orbifold metrics on with nonnegative scalar curvature, and positive scalar curvature near the sub-orbifold . If acts freely, then is zero if (and ), but in the orbifold case the localised -genera for nontrivial elements are additional obstructions to positive scalar curvature.
2.2. Existence for connected Lie groups
In this subsection, we suppose that is a connected Lie group. As pointed out at the end of the previous subsection, for such groups the obstructions to -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 is connected, Abels’ global slice theorem [1] implies we have a diffeomorphism , for a -invariant submanifold . Our existence result, Theorem 2.9, supposes that such a slice has -bounded geometry, a notion introduced in Definition 2.8.
Suppose a compact, connected Lie group acts isometrically on a complete Riemannian manifold . Let denote a bi-invariant Riemannian metric on . For each we have a linear map
(2.1) |
defined by , for . Define a pointwise norm function
(2.2) |
where denotes the linear operator norm with respect to and .
Definition 2.5.
We say that the action of on has no shrinking orbits if, for any neighbourhood of the fixed point set , there exists a constant such that for all we have
where the norm function is taken with respect to the Riemannian metric .
We remark that the condition of the action having no shrinking orbits is independent of .
Example 2.6.
Suppose that , on which acts in the natural way. Let be positive and rotation-invariant, and consider the Riemannian metric on equal to times the Euclidean metric. Then for all and ,
So where is the Euclidean norm of . Hence the action has no shrinking orbits if and only if the function has a positive lower bound outside a neighbourhood of .
Now let us define the notion of -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 there exists such that , where is the Riemann curvature tensor.
Definition 2.8.
The action of on is said to have -bounded geometry if it has no shrinking orbits and has bounded geometry.
Our main existence result, proved in Subsection 6.2, is the following.
Theorem 2.9.
Let be a connected Lie group, and a maximal compact subgroup with non-abelian identity component. Let be a manifold admitting an effective action by with compact fixed point set. If there exists a Riemannian metric on such that the -action has -bounded geometry, then the manifold admits a -invariant metric with uniformly positive scalar curvature.
This result may be viewed as a strengthening of the vanishing of the obstructions to -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 -genus as in Atiyah and Hirzebruch’s vanishing theorem [3] in the compact case. (See the diagram on page 233 of [27].)
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 -Callias-type index
From now on, will be a complete Riemannian manifold on which acts properly and isometrically. Let denote a -graded, -equivariant Clifford module over , and an odd-graded Dirac operator on , associated to a -invariant Clifford connection on via the Clifford action by on .
Let be an odd, -equivariant, fibrewise Hermitian vector bundle endomorphism of .
Definition 3.1.
The endomorphism is admissible for if
-
•
the operator on is a vector bundle endomorphism; and
-
•
there are a cocompact subset and a constant such that we have the pointwise estimate
(3.1) on .
In this setting the operator is called a -Callias-type operator.
In the rest of the paper, we will use the following notation. Let denote the modular function on . Let be either the reduced or maximal group -algebra of . Let act on sections of the bundle by
for a section, and .
Equip the space with a right -action defined by
(3.2) |
and a -valued inner product defined by
(3.3) |
for , , , and . Let be the Hilbert -module completion of with respect to this structure.
The definition of the equivariant index of -Callias-type operators is based on the following result, Theorem 4.19 in [15].
Theorem 3.2.
There is a continuous -invariant cocompactly supported function on such that
(3.4) |
is a well-defined, adjointable operator on , such that is a Kasparov -cycle. Its class in is independent of the function chosen.
For details about the definition of the operator , we refer to Definition 4.11 in [15].
Definition 3.3.
3.2. Hypersurfaces and the index theorem
In the setting of the previous subsection, we now suppose that for an ungraded, -equivariant Clifford module over , where the first copy of is the even part of , and the second copy is the odd part. Suppose that
(3.5) |
for a Dirac operator on , and that
(3.6) |
for a Hermitian endomorphism of . (The conditions on then become conditions on .)
Let be as in Definition 3.1. Let be a -invariant, cocompact subset containing in its interior, such that is a smooth submanifold of . Let be the closure of the complement of , so that and . In this and similar settings, we write
By (3.1), the restriction of to is fibrewise invertible. Let and 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 to pointing into , times , defines -invariant gradings on .
Let be the Clifford connection on used to define . By restriction and projection, it defines connections on . The Clifford action by on preserves by the first condition in Definition 3.1; see also Remark 1.2 in [2]. Hence the connections define Dirac operators on . These operators are odd-graded. Because is cocompact, has an equivariant index
defined by the analytic assembly map [4].
Theorem 3.4 (-Callias-type index theorem).
We have
(3.7) |
Versions of this result where 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 -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 -modules from [15]. Let , , and be as in Subsection 3.1.
Definition 4.1.
For each nonnegative integer , define to be the pre-Hilbert -module whose underlying vector space is , equipped with the right -action defined by (3.2), and -valued inner product defined by
where and is as in (3.3). Here we set equal to the identity operator. Denote by the vector space completion of with respect to the norm induced by , and extend naturally the -action to a -action, and to a -valued inner product on , to give it the structure of a Hilbert -module. We call the -th -Sobolev module with respect to .
The module defined above Theorem 3.2 equals . The following version of the Rellich lemma holds for Sobolev modules (Theorem 3.12 in [15]).
Theorem 4.2.
Let be a continuous -invariant cocompactly supported function on . Then multiplication by defines an element of whenever .
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 and . 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, -invariant, uniformly bounded bundle endomorphism of defines an element of .
Proof.
Let be a smooth, -invariant, uniformly bounded bundle endomorphism of . Since is uniformly bounded, it defines a bounded operator on . Let denote its operator norm, let be a cutoff function on , and let be the pointwise adjoint of . Since the operator is positive on , it has a positive square root that one observes is -invariant. For a fixed , the function
has compact support in , by properness of the -action. Thus the map defined by has compact support in .
It follows that for any unitary representation of on a Hilbert space and ,
is a well-defined vector in . By computations similar to those in the proof of Proposition 5.4 in [15], one sees that equals
Thus, for any unitary representation of ,
is a positive operator, where we let act on by
It follows that the element
is positive in . Hence extends to an operator on all of . Similarly, defines an operator on all of that one checks is the adjoint of . ∎
Proposition 4.4.
Suppose that there are a -invariant, cocompact subset and a -invariant, cocompact hypersurface such that there is a -equivariant isometry , and a -equivariant vector bundle isomorphism . Let be a -equivariant vector bundle endomorphism of . Suppose that, on , and are constant in the factor of . Then defines an element of .
The proof uses the next lemma. To state it, let be the completion of with respect to the inner product
Lemma 4.5.
Let and be as in Proposition 4.4. Let be a bounded, positive operator on such that
-
•
preserves the subspace ;
-
•
for any , the function given by has compact support in .
Then
is a positive element of .
Proof.
Let be the positive square root of in . Since has compact support, and , the map defined by has compact support in . As in the proof of Proposition 4.3, one finds that for any unitary representation of on a Hilbert space and ,
is a well-defined vector in , and that
Similarly to the proof of Proposition 4.3, we deduce that is a positive element of . ∎
Proof of Proposition 4.4.
Because of the forms of and , there is a canonical (up to equivalence) first Sobolev norm on sections of that is -invariant, and invariant under the relevant class of translations in the factor of . Because is an order zero differential operator constant on the factor , it defines a bounded operator with respect to . Due to the form of , the norm on is equivalent to , and so defines a bounded operator on .
Let be the natural projection. Let be a cutoff function on such that
for a cutoff function on . Let and denote the respective adjoints of and in . Then the operator
is bounded and self-adjoint on with norm at most , where the norms are taken in .
Let be a smooth, nonnegative function on that is identically on the support of , and such that
for a compactly supported function on . Consider the endomorphism of . For the same reasons as for , it defines a bounded operator on . Fix . Because is a positive bounded operator on , we may apply Lemma 4.5 with to conclude that
(4.1) |
is positive in . Define by
By construction of , the quantities
are bounded as functions of (see also Remark 4.6 below). A direct calculation shows that
for some constant . Together with positivity of (4.1), this implies that
so that extends to an operator on all of . Similarly, the -adjoint defines an operator on that one checks is the adjoint of . ∎
Remark 4.6.
As can be seen from the proof, the conclusion of Proposition 4.4 holds more broadly for any on which the functions and on 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 , then .
Proof.
In this setting, the operator in (3.4) is invertible. This implies that the -cycle is operator homotopic to the degenerate cycle . ∎
Proof.
In this setting, is bounded, and the cycle is operator homotopic to .
In general, let be a -algebra, let be a Hilbert -module, and set . and an adjointable operator on such that is a Kasparov -cycle, and is of the form
(4.2) |
for a (necessarily self-adjoint) . Then . Because 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 be an odd, -equivariant Dirac-type operator on a -graded Clifford module , and let be an odd, smooth -equivariant, uniformly bounded Hermitian vector bundle endomorphism of . Fix . For , consider the operator .
Proposition 4.9 (Homotopy invariance).
Suppose that
-
(1)
for , the endomorphism defines an adjointable operator on the Sobolev module of Definition 4.1;
-
(2)
there is a nonnegative, -invariant, cocompactly supported function such that for all , the operator is invertible, with inverse in .
Then is independent of .
Proof.
Corollary 4.10.
If is a -Callias-type operator on , and is a -equivariant, odd vector bundle endomorphism of that equals zero outside a cocompact set, then .
Proof.
We set and apply Proposition 4.9. Since is cocompactly supported, the first condition in Proposition 4.9 holds by Proposition 3.5 in [15]. For the same reason, is a Callias-type potential for all , so by Theorem 5.6 in [15], the second condition in Proposition 4.9 holds for , where a priori the function may depend on . But since is zero outside a cocompact set, we can choose independent of . The claim then follows from Proposition 4.9. ∎
Remark 4.11.
4.4. A relative index theorem
We will use an analogue of Bunke’s relative index theorem, Theorem 1.2 in [7]. For , let , , and , respectively, be as , , and were before. Suppose that there are a -invariant hypersurface and a -invariant tubular neighbourhood of , and that there is a -equivariant isometry such that
-
•
;
-
•
;
-
•
, where is the Clifford connection used to define ; and
-
•
corresponds to via .
Suppose that for closed, -invariant subsets . We identify and via and write for this manifold when we do not want to distinguish between and . Using the map , form
For , let , and be obtained from the corresponding data on and by cutting and gluing along via .
Theorem 4.13.
In the above situation,
Proof.
This proof is an adaptation of the proof of Theorem 1.14 in [7], with some results from [15] added. For , let and be as and above and in Theorem 3.2, for the data indexed by . Using superscripts to denote opposite gradings, we write and . We will show that
(4.3) |
which is equivalent to the theorem.
For , let be real-valued functions such that
-
•
and ;
-
•
and ; and
-
•
.
We view pointwise multiplication by these functions as operators
(4.4) |
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 , we form the operator
on . Then is an odd, self-adjoint, adjointable operator on such that . As such, it generates a Clifford algebra .
We claim that is a compact operator. This is based on the Rellich lemma for Hilbert -modules, Theorem 4.2. Let be one of the functions or , viewed as an operator from to as in (4.4). Let be as in Theorem 3.2, for the operator . For and , the operator on is invertible by Lemma 4.6 in [15], and we denote its inverse by .
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) |
equals
(4.6) |
Theorem 4.2, together with Lemma 4.6(a) in [15], implies that for all cocompactly supported continuous functions on , the compositions , and are compact operators on if , and adjointable operators if . 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 for constants . So the integral converges in the operator norm on , and we conclude that (4.5) is a compact operator on . This implies that is a compact operator.
Corollary 4.14.
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 can be identified with the corresponding data near .
5. Proof of the -Callias-type index theorem
The first and most important step in the proof of Theorem 3.4 is Proposition 5.1, which states that equals the index of a -Callias-type operator on the manifold , which we will call the cylinder on . 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].

In this section, we consider the setting of Subsection 3.2. In particular, and are assumed to be of the forms (3.5) and (3.6).
5.1. An index on the cylinder
Let be the pullbacks of the bundles of defined in Subsection 3.2 along the projection . They are Clifford modules, with Clifford actions
for and , where is the Clifford action by on (which preserves as pointed out in Subsection 3.2), and is the normal vector field to in the direction of . Let be the Dirac operator on defined by this Clifford action, and the pullback to of the restriction to of the Clifford connection used to define .
Let be an odd function such that for all . We also denote its pullback to by . Then pointwise multiplication by is an admissible endomorphism for . Whenever a Dirac operator with a subscript 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
on . We also consider the admissible endomorphism
of .
Proposition 5.1.
We have
(5.1) |
5.2. Attaching a half-cylinder
Let be the pullback of . We choose small enough so that .
Because the sets are cocompact in Corollary 4.14, we initially compare the left-hand side of (5.1) to an index on a manifold where only is replaced by a half-cylinder . To be more precise, is invariant under changes in the Riemannian metric on cocompact sets because the Kasparov -cycles corresponding to two -invariant Riemannian metrics differing only a cocompact set are homotopic by convexity of the space of -invariant Riemannian metrics. We choose a metric such that there is a neighbourhood of that is isometric to (see Figure 2).

By Corollary 4.10, the index of does not change if we change in a cocompact set. So we may assume that is constant in the direction normal to inside ; i.e. for all and , , for an endomorphism of . We further choose such that a set as in Definition 3.1 is contained in .
Let be the pullback of to a connection on . We choose the Clifford connection to define so that on , it equals the restriction of to .
For this structure near , we can form the Riemannian manifold (see Figure 3), and define the Clifford module such that it equals on and on . Let be the Clifford connection on corresponding to on and to on . Let be the resulting Dirac operator.

We define an endomorphism of that is equal to on and to the pullback of on . Recall that by removing the subscript from we refer to the construction (3.5). Similarly, when we remove the subscript from , we will be referring to the endomorphism defined by as in (3.6). Then Corollary 4.14 immediately implies that
(5.2) |
The connection , and therefore the corresponding Dirac operator, does not preserve the decomposition . With respect to this decomposition, that Dirac operator has the form
(5.3) |
for vector bundle homomorphisms and . (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 equals this operator on . Let be the pullback of to a connection on . Consider a Clifford connection on that is equal to the direct sum of and on and to on . Then the corresponding Dirac operator is equal to
on and to on .
Lemma 5.2.
There exists a such that is admissible for . For such ,
(5.4) |
Proof.
Existence of 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
This follows from Proposition 4.9 by setting , , with and as in (5.3), and . Note that although is not a Callias-type potential here, Proposition 4.9 still applies, since defines an element of by Proposition 4.3 and an element of by Proposition 4.4. (The conditions of Proposition 4.4 hold because of the forms of , and .) A more straightforward application of Proposition 4.9 yields . ∎
Let be an odd function such that
(The property that 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 in Subsection 5.1, but Proposition 5.1 for general 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 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 .)
Let be the grading operator on that equals on . Let be its pullback to . Let be the endomorphism of equal to on and equal to zero on the rest of .
Lemma 5.3.
We have
Proof.
This can be proved via a linear homotopy between and . 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 be the manifold with reversed orientation. Form the manifold
See Figure 4. (The notation is motivated by the fact that with reversed orientation is naturally equal to , which can be identified with via a shift over a distance .)

Let be equal to the vector bundle , but with the opposite Clifford action (where acts as ). Let be the Clifford module that is equal to on and to on . From the Clifford connections on and a Clifford connection on , construct a Clifford connection on by
Using this connection, we obtain the Dirac operator on . Then equals on .
The function equals on . Thus on this interval, both and are equal to . (Here we use matrix notation with respect to the decomposition .) So we can define the endomorphism of by setting it equal to on and equal to
(5.5) |
on , where .
Lemma 5.4.
We have
Proof.
Proof of Proposition 5.1..
Consider the cylinder as in Figure 1. The data and coincide in a neighbourhood of . By cutting along and gluing, we obtain the corresponding data and . See Figure 5. To be explicit,
(5.6) |
on .

The connection on obtained from cutting and gluing the connections and is the direct sum connection . So the corresponding Dirac operator equals
By the explicit form (5.6) of , the operator on is the direct sum of the operators on and on . So
Lemma 4.7 then implies that the second term on the right-hand side is zero, whence
5.4. Proof of Theorem 3.4
Let be the Hilbert -module constructed from as in Subsection 3.1. We write for brevity.
Lemma 5.5.
For all the operator on 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 can only be bounded below by a function that is -proper, in the sense that the inverse image of a compact set is cocompact instead of compact. One chooses the bump functions in the proof of Theorem 2.40 in [11] to be -invariant and cocompactly supported. Theorem 4.2, and Lemma 4.6(a) in [15], imply that is a compact operator on . And as in the proof of Theorem 2.40 in [11], one shows that converges to in the operator norm on . ∎
We will use an analogue of Theorem 6.6 in [9].
Lemma 5.6.
The operator
(5.7) |
lies in .
Proof.
Proposition 5.7.
In the setting of Proposition 5.1,
Proof.
Let be the Hilbert -module constructed from as in Subsection 3.1. Then . So Lemma 5.6 implies that is represented by the unbounded Kasparov cycle
(5.9) |
The Callias-type operator on
is equal to
(5.10) |
where is the grading operator on , equal to times Clifford multiplication by the unit normal vector field on pointing into (so that ). Let be the even and odd-graded parts of , and let be the restriction of to even and odd-graded sections of , respectively. With respect to the decomposition
(5.11) |
the operator (5.10) equals
(5.12) |
The kernel of in is one-dimensional, and spanned by the function
By the properties of , , whereas . It follows that is invertible on the appropriate domain, while is zero on and invertible on .
Consider the submodules
of (5.11). These are preserved by the operator . (For , this is immediate from (5.12); for , this follows from the facts that is symmetric and preserves .) We find that the cycle (5.9) decomposes as
(5.13) |
The operator 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 -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 . ∎
Remark 5.8.
A similar argument in the case where 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 is odd-dimensional. Let denote scalar curvature. Let be a cocompact subset of such that , on , and the distance from to is positive. Let be a function such that for all and for all . Consider the operator as in (3.5), where is the -Dirac operator on , and the admissible endomorphism as in (3.6), where is pointwise multiplication by . Then the set in Subsection 3.2 can be chosen to be cocompact as required, and so that , where . In this setting,
(6.1) |
where is the spinor bundle. (This is consistent with Corollary 7.3.)
For any , Lichnerowicz’ formula implies that
On , the function on the right-hand side equals . Since is -invariant, and positive on the cocompact set , it has a positive lower bound on that set. Further, is -invariant and cocompactly supported, hence bounded. So we can choose small enough so that on . It follows that has a positive lower bound. This implies that is invertible, so . The claim then follows by Theorem 3.4 and (6.1).
If is even-dimensional, then the claim follows by applying the result in the odd-dimensional case to the manifold . ∎
Proof of Corollary 2.4.
This follows from Theorem 2.1 and index formulas for traces defined by orbital integrals applied to . These index formulas are:
-
•
Theorem 6.10 in [41] if is any locally compact group and ;
-
•
Theorem 6.1 in [40] if is discrete and finitely generated and is any element;
-
•
Proposition 4.11 in [24] if is a connected semisimple Lie group and is a semisimple element.
These results imply that in the setting of Corollary 2.4,
for a trace . As is independent of the choice of Riemannian metric, so is . Finally, vanishing of implies vanishing of for all 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 , on a compact manifold .
Let be a compact Lie group acting on . Consider the principal -bundle defined by the map
Take a -invariant Riemannian metric on . Let be a bi-invariant Riemannian metric on . Let denote the lift of to the orthogonal complement to in with respect to the product metric on .
For each , let be the lift of the metric on to . Then
is a Riemannian metric on the total space . One can check that, for each , is invariant under the left -action on defined by
Thus descends, via the projection onto the second factor,
to a -invariant metric on . Further, one sees that is a Riemannian submersion with respect to the metrics and on the total space and base respectively.
The following proposition shows that, under the conditions stated, for all sufficiently small , 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 is non-compact but has -bounded geometry; see Definition 2.8.
Proposition 6.1.
Let be a manifold with an action by a non-abelian, compact, connected Lie group . Fix a bi-invariant metric on . If is a -invariant Riemannian metric on with -bounded geometry, then for any neighbourhood of the fixed point set , there exists such that for all , the metric constructed above has uniform positive scalar curvature on .
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 , there is an orthogonal splitting , where is the Lie subalgebra of the isotropy subgroup of . The map from (2.1) restricts to an injection on . Denote the orthogonal complement of in by . Then
For each , choose an orthonormal basis of with respect to such that for all ,
for some continuous, positive functions . For each , define a function by
By the calculations in the proofs of Propositions 3.6 and 3.7 in [27] and the assumption that has bounded geometry, for any neighbourhood of , the scalar curvature of at any is bounded below by
(6.2) |
as , where the term is independent of . Since the -action has no shrinking orbits with respect to , there exists such that for each and . In particular, for , the expression (6.2) is bounded below by
(6.3) |
Now, without loss of generality we may assume that or , as any compact, connected, non-abelian Lie group has such a subgroup. Since has no subgroups of codimension , we have at each . For all and , is 4 times the sectional curvature of the plane spanned by and with respect to the metric , and this is constant in , and positive for or . Thus for any neighbourhood of , there exists such that for all , the expression (6.3), and hence also (6.2), is uniformly positive outside . It follows that for all such , the scalar curvature of is uniformly positive outside . ∎
Theorem 6.2.
Let be a manifold that admits an effective action by a compact, connected, non-abelian Lie group , such that the fixed point set is compact. If there exists a -invariant Riemannian metric on such that the -action has -bounded geometry, then admits a -invariant metric with uniformly positive scalar curvature.
Proof.
Since is compact and the action is effective, by Section 4 of [27] there exists , a -invariant neighbourhood of with compact closure, and a -invariant Riemannian metric on such that each metric , constructed from as in subsection 6.2, has positive scalar curvature on for .
Fix a bi-invariant metric on . Let be a -invariant metric on for which the -action has -bounded geometry. Let be a smooth, -invariant partition of unity on such that on and on , where is a relatively compact neighbourhood of containing the closure of . Then
is a -invariant Riemannian metric on . Applying the prescription in Subsection 6.2 to , we obtain a family of -invariant metrics on . We claim that for sufficiently small , has uniformly positive scalar curvature on .
To see this, let be the norm function (2.2) associated to the metric . Since and coincide on , and has -bounded geometry, there exists such that for all . One sees that has -bounded geometry, and so by Proposition 6.1, there exists such that for all , has uniformly positive scalar curvature on . It follows that for all , has uniformly positive scalar curvature on . ∎
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 is noncompact, and is not known to satisfy Baum–Connes surjectivity, then it is a priori unclear if 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 -Callias-type operators as defined above:
Corollary 7.1.
The Callias-index of 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 be an odd-dimensional Riemannian manifold with boundary , on which acts properly and isometrically, preserving , such that is compact. Suppose that a neighbourhood of is -equivariantly isometric to , for some . Let be a -equivariant Clifford module, and consider the Clifford module , graded by times Clifford multiplication by the inward-pointing unit normal. Suppose that . Let be the Dirac operator on associated to a -invariant Clifford connection for the Clifford action by on . Then .
Proof.
Form the manifold by attaching the cylinder to along . Extend and the Clifford action to in the natural way. We write for the extension of to . The connection pulls back to a Clifford connection on . Because the -invariant Clifford connections form an affine space, we can extend this pulled-back connection to all of using a -invariant Clifford connection on and a partition of unity. Let be the associated Dirac operator.
7.3. -Dirac operators
Corollary 7.3.
Consider the setting of Theorem 3.4, and assume that is odd-dimensional and is a -Dirac operator. Then there is a union of connected components of and a -structure on with spinor bundle such that
(7.1) |
where is a -Dirac operator on the spinor bundle . If is connected, then .
Proof.
Since is an irreducible Clifford module, and is invariant under the Clifford action of , over each connected component of the bundle is either zero or . Since is odd-dimensional, is the spinor bundle of the -structure of that it inherits from . So (7.1) follows.
There is a converse to Corollary 7.3 in the following sense. Let be any union of connected components; there is a finite number of such subsets of since is compact. We can define an admissible endomorphism such that (7.1) holds, by taking to be multiplication by a -invariant function on that equals on and on , and is constant or outside a cocompact set. Thus given any hypersurface bounding a cocompact set, and any set of connected components of , we have an index
(7.2) |
with and related as in (3.6), independent of the choice of with the property that is positive definite on and negative definite on .
7.4. Induction
Suppose that is an almost connected, reductive Lie group, and let be maximal compact. In [18, 19, 24] some results were proved relating -equivariant indices to -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 , for a -invariant submanifold , and . Let be a Cartan decomposition. Then . We assume that the -invariant Riemannian metric on is induced by a -invariant Riemannian metric on and an -invariant inner product on via this identification.
We assume for simplicity that the adjoint representation lifts to the double cover of . (This is true for a double cover of .) Then the standard representation of may be viewed as a representation of . We assume that for a Clifford module . The Clifford action by on equals , for a Clifford action by on , and the Clifford action of on . We choose the -invariant connection so that , for Clifford connections on and on
Since is -equivariant, it is determined by its restriction to , which is a -equivariant endomorphism of . We assume that , for a -equivariant endomorphism of . (What follows remains true if for an -invariant endomorphism of , but this requires a small extra argument that we omit here.)
Consider the Dirac operator on . Form from as in (3.5) and from as in (3.6). Let be the representation ring of and
(7.3) |
be the Dirac induction map [4].
Corollary 7.4.
The operator is a -equivariant Callias-type operator, and
Proof.
Theorem 3.4 implies that . Write , so that . Then is a compact manifold. Define analogously to . The induction result for cocompact actions, Theorem, 4.5 in [19], Theorem 5.3 in [24] or Theorem 46 in [18], implies that
Another application of Theorem 3.4, now with replaced by , or Theorem 1.5 in [2] with a compact group action added, shows that . ∎
7.5. Callias quantisation commutes with reduction
Theorem 3.11 in [16] is a quantisation commutes with reduction result for the equivariant index of -Callias-type operators. This result applies to reduction at the trivial representation of ; i.e. to an index defined in terms of -invariant sections of . Using Theorem 3.4, we can generalise this result to reduction at more general representations, or more precisely, at arbitrary generators of . 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 is odd-dimensional, and that is the spinor bundle for a -equivariant -structure. Let be the -Dirac operator on , defined by the Clifford connection corresponding to a connection on the determinant line bundle .
The -moment map associated to is the map such that for all ,
where denotes the Lie derivative with respect to , and is the vector field induced by . The reduced space at an element is defined as , where is the stabiliser of . This reduced space is noncompact in general, and may not be smooth. But the reduced space is compact. It is not always a smooth manifold, but if it is, and , then we have an identification , with and as in Subsection 7.4, including -structures. See Propositions 3.13 and 3.14 in [22]. Let be as in Corollary 7.3. Then we similarly have in the smooth case, including -structures.
There is a nontrivial way to define a -quantisation , even when is not smooth, described in detail in Section 5.1 of [31]. Motivated by the identification in the smooth case, we define for .
Let be a maximal torus, and fix a positive root system for . Let have highest weight . (The first inclusion is defined by , the second by the Cartan decomposition.) Following [31, 32], we call an element an ancestor of if the coadjoint orbit is admissible in the sense of [32], and its -equivariant -quantisation is . There exists a finite set of ancestors representing all different such coadjoint orbits.
Let be the reduced group -algebra of and the Dirac induction map (7.3). By the Connes–Kasparov conjecture, proved in [10, 26, 42], the abelian group is free, with generators , where runs over .
Recall the definition of the Callias index of -Dirac operators (7.2).
Corollary 7.5 (Callias quantisation commutes with reduction).
We have
(7.4) |
Proof.
Remark 7.6.
In cases where is smooth and is a hypersurface in , which is a transversality condition between and , one can use Theorem 1.5 in [2] (more precisely, its special case for -Dirac operators, which is the non-equivariant case of Corollary 7.3) to express the -quantisation as the index of a Callias-type operator on .
References
- [1] Herbert Abels. Parallelizability of proper actions, global -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 -theory of group -algebras. In -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 -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 -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 -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, -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. -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 -Dirac operator. Adv. Math., 134(2):240–277, 1998.
- [29] Alexandr S. Miščenko and Anatoliǐ T. Fomenko. The index of elliptic operators over -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. -algebras, positive scalar curvature, and the Novikov conjecture. Inst. Hautes Études Sci. Publ. Math., (58):197–212 (1984), 1983.
- [36] Jonathan Rosenberg. -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. -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 -Lefschetz fixed-point formula for orbifolds. J. Differential Geom., 102(2):285–349, 2016.
- [41] Hang Wang. -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.