Higher localised -genera for proper actions and applications
Abstract.
For a finitely generated discrete group acting properly on a spin manifold , we formulate new topological obstructions to -invariant metrics of positive scalar curvature on that take into account the cohomology of the classifying space for proper actions.
In the cocompact case, this leads to a natural generalisation of Gromov-Lawson’s notion of higher -genera to the setting of proper actions by groups with torsion. It is conjectured that these invariants obstruct the existence of -invariant positive scalar curvature on . For classes arising from the subring of generated by elements of degree at most , we are able to prove this, under suitable assumptions, using index-theoretic methods for projectively invariant Dirac operators and a twisted -Lefschetz fixed-point theorem involving a weighted trace on conjugacy classes. The latter generalises a result of Wang-Wang [24] to the projective setting. In the special case of free actions and the trivial conjugacy class, this reduces to a theorem of Mathai [17], which provided a partial answer to a conjecture of Gromov-Lawson on higher -genera.
If is non-cocompact, we obtain obstructions to being a partitioning hypersurface inside a non-cocompact -manifold with non-negative scalar curvature that is positive in a neighbourhood of the hypersurface. Finally, we define a quantitative version of the twisted higher index, as first introduced in [12], and use it to prove a parameterised vanishing theorem in terms of the lower bound of the total curvature term in the square of the twisted Dirac operator.
Key words and phrases:
Higher index, positive scalar curvature, fixed point theorem, coarse geometry, twisted Roe algebra, quantitative index2010 Mathematics Subject Classification:
46L80, 58B34, 53C201. Introduction
In this paper we develop the theory of higher indices of projectively equivariant Dirac operators with respect to proper actions of discrete groups, and relate this to numerical invariants that are computable in terms of characteristic classes. In the spin setting, these results can be applied to give new obstructions to metrics of positive scalar curvature that are invariant under the proper action of a discrete group. This generalises the obstructions provided by the higher -genera of Gromov and Lawson [9] to the setting of proper actions.
On the operator-algebraic side, where the higher index resides, we consider appropriately weighted traces on twisted group -algebras. These traces generalise the traces associated to conjugacy classes in the setting of ordinary group -algebras by taking into account a -valued group -cocycle. On the geometric side, there is a corresponding algebra of projectively invariant operators of an appropriate trace class, in which the heat operator associated to a projectively invariant Dirac operator lies. The traces on these two sides are related through a weighted -Lefschetz fixed-point theorem that generalises a result of Wang-Wang [24, Theorem 6.1].
The special case where the group is torsion free and was considered in [6, 17], and led to a proof that the higher -genera of arising from the subring of generated by elements of degree at most are obstructions to positive scalar curvature on . This gave a partial answer to a conjecture of Gromov-Lawson [9, Conjecture]; see also [22, Conjecture 2.1].
Now let us turn to the general case where an arbitrary discrete group , possibly with torsion, acts properly on a manifold . Denote by the classifying space for proper actions [2], and let
be the classifying map for the action of on . For any integer , let . Let
(1.1) |
be the -invariant lift of a differential form on the orbifold belonging to the class in the orbifold de Rham cohomology group . (For background on orbifold de Rham cohomology, see for example [1, chapter 2].)
If the quotient is compact, then for each , the centraliser of acts properly and cocompactly on the fixed-point submanifold . Let be a cut-off function for this action, i.e. is non-negative and satisfies
for any . We define the higher localised -genus of with respect to to be
(1.2) |
where is the curvature of the Levi–Civita connection restricted to the normal bundle of in with respect to a -invariant Riemannian metric on .
We are led to the following conjecture:
Conjecture 1.1.
If a proper, cocompact, connected -spin manifold admits a -invariant Riemannian metric of positive scalar curvature, then all of the higher localised -genera of vanish. That is, if is any closed -form constructed as above, then for all , we have
Remark 1.2.
We give evidence for the validity of this conjecture in Theorem 1.5 and Corollary 1.6 by proving it for all arising from the cohomology ring generated by , under the assumption that is connected and is a regular element for the relevant group multiplier on (see subsection 2.2). In view of Remark 4.11, it seems likely that, in the special case of such and regular , Conjecture 1.1 is implied by the Baum-Connes conjecture, without any growth conditions on .
Remark 1.3.
The following is an immediate corollary to Conjecture 1.1:
Corollary 1.4.
Suppose is a proper, cocompact -spin manifold. Then there is no -invariant metric of positive scalar curvature on .
Indeed, suppose such a Riemannian metric existed. Letting be the associated volume form and , the quantity (1.2) is , where is a cut-off function for the -action on . This is positive, which contradicts Conjecture 1.1.
We give an index-theoretic approach to a special case of Conjecture 1.1 as follows. From the data above, we construct a twisted Dirac operator that is invariant under a projective representation of . We show that the integral (1.2) arises naturally as certain weighted traces of the associated heat operator, and relate this to the higher index of the twisted Dirac operator via a weighted trace map on the twisted group algebra. We prove:
Theorem 1.5.
Let be a finitely generated group, and let be a connected, -equivariantly spin manifold such that is compact and . Let be as in (2.1), and let be the associated projectively invariant Dirac operator on from Definition 2.4. Let and be the multipliers constructed in subsection 2.2, and suppose that is an -regular element, in the sense of (3.4), whose conjugacy class has polynomial growth.
-
(i)
Suppose is even-dimensional. Then
where be the -invariant higher index of from Definition 3.32, is the homomorphism (4.8), are the connected components of that intersect the support of a cut-off function for the action of on , is as in (2.2), is an arbitrary point in , and is the curvature of the Levi-Civita connection restricted to the normal bundle of in and with respect to a -invariant Riemannian metric on .
-
(ii)
If admits a -invariant Riemannian metric of positive scalar curvature, then for each integer , we have
(1.3) where the notation is as explained in part (i). If is connected, then
for each .
Corollary 1.6.
Next, we prove two obstruction results when is non-compact, again via index theory of projectively invariant Dirac operators. First, we show that the higher localised -genera are obstructions to -invariant Riemannian metrics with positive scalar curvature in a neighbourhood of . More precisely, we prove:
Theorem 1.7.
Let be a connected spin manifold on which a discrete group acts properly, preserving the spin structure, and suppose . Let be a connected -cocompact hypersurface in with trivial normal bundle. Let the multipliers and be as in subsection 5.2, and suppose that is an -regular element, in the sense of (3.4), whose conjugacy class has polynomial growth.
Suppose admits a complete -invariant Riemannian metric whose scalar curvature is non-negative everywhere on and positive in a neighbourhood of . Let be as above in (1.2). Then if is connected,
for each integer . In particular,
where is a -regular element, are the connected components of the fixed-point set that intersect the support of a cut-off function for the -action on , and is the normal bundle of in .
The proof of this theorem uses a Callias-type index theorem for projectively invariant Dirac operators, which is discussed in section 5.
Remark 1.8.
Second, we show that the projectively invariant higher index is compatible with the framework of quantitative -theory [21], which is a refinement of operator -theory in the context of geometric -algebras. As shown in subsection 6.1, this allows us to formulate a quantitative notion of projectively invariant higher index, generalising that defined in [12]. With respect to this index, we obtain a parameterised vanishing theorem where the vanishing propagation depends on the lower bound of the twisting curvature of the projective Dirac operator as well as the scalar curvature. More precisely, we prove:
Theorem 1.9.
Fix and . There exists a constant such that the following holds. Let be a smooth spin Riemannian manifold equipped with a proper isometric action by a discrete group , and suppose that . Let be the projection, the scalar curvature on , and the spinor bundle.
Let be as in (2.1), and let be the associated projectively invariant Dirac operator on from Definition 2.4, acting on sections of the bundle . Let denote the scalar curvature on , and let denote Clifford multiplication. If the estimate
holds for some positive constant , then the quantitative -equivariant higher index of on at scale vanishes for all :
where the algebra is as in Definition 3.27, and is defined as in subsection 6.2.
The paper is organised as follows. In section 2, we formulate the preliminary definitions and properties of the relevant operator algebras and projectively invariant operators. In section 4, we prove Theorem 1.5 and Corollary 1.6. In section 5, we develop some Callias-type index theory in the projective setting and use it to prove Theorem 1.7. In section 6, we formulate the quantitative twisted higher index and prove Theorem 1.9.
2. Preliminaries
We first fix some notation and recall the necessary operator-algebraic and geometric terminology we will need.
2.1. Notation
For a Riemannian manifold, we write , , , and to denote the -algebras of complex-valued functions on that are, respectively: bounded Borel, bounded continuous, continuous and vanishing at infinity, and continuous with compact support. If is a Borel subset, we write for the associated characteristic function.
For any -algebra , denote its unitization by . If is a Hilbert module over , let and denote the -algebras of bounded adjointable and compact operators on respectively.
For an element of a group , we use to denote the centraliser of in .
2.2. Multipliers and projective representations
Let be a discrete group.
Definition 2.1.
A multiplier on is a map satisfying
-
(i)
;
-
(ii)
,
for all , where is the identity element in .
In other words, a multiplier on is an element of , i.e. a -valued group -cocycle, satisfying the additional normalisation condition (ii). This condition is slightly stronger than the normalisation requirement in [17], namely , and we have adopted it to simplify some calculations. Nevertheless, every multiplier in the sense of [17] is cohomologous to a multiplier in our sense. Observe that given a multiplier , its pointwise complex conjugate is also a multiplier.
We will be concerned specifically with multipliers that arise from proper -actions on manifolds in the following way. Let be a smooth, connected manifold equipped with a proper action by a discrete group , preserving the spin structure. Suppose is compact and that . Let be the classifying space for proper -actions [2], and the classifying map for .
Let be a -cocycle on , and a closed -form on representing the de Rham cohomology class on the orbifold . Since lifts trivially to , the lift of to is exact, so there exists a one-form (not necessarily -invariant) such that
(2.1) |
Since is -invariant, for all . Thus is a closed -form on , and hence exact by assumption. It follows that there exists a family
of smooth functions on such that
(2.2) |
This implies that for any ,
(2.3) |
By way of normalisation, we may assume that there exists some such that
(2.4) |
for each . This, together with (2.3), implies that and that the formula
(2.5) |
defines an element of . For each , we have an associated -valued -cocycle
(2.6) |
which is a multiplier the sense of Definition 2.1. When it is clear from context, we will use the following abbreviations:
It can be shown that for a given class , different choices of , , and all lead to cohomologous . Nevertheless, it is useful to make specific choices, as the numerical obstructions we compute in Theorems 1.5 and 1.7 are expressed in terms of .
Remark 2.2.
Restricting (2.2) to the fixed-point submanifold , one sees that
hence on any connected component , the function is constant.
Definition 2.3.
Let be a -equivariant -vector bundle. For each and , define the unitary operators , , and on by:
-
•
;
-
•
;
-
•
,
for and . We refer to as a projective action on .
Note that for any and , we have
Thus for each , the map given by defines a projective representation of in the sense of subsection 2.3 below. An operator on that commutes with is said to be -invariant or simply projectively invariant if no confusion arises.
Now suppose is -equivariantly spin, equipped with a -invariant Riemannian metric. Let be the spinor bundle and the spin-Dirac operator. We can obtain a -invariant Dirac operator as follows. Let be a -equivariantly trivial line bundle. For each , consider the Hermitian connection
on . Equip with the obvious -grading. Then we have:
Definition 2.4.
For each , we will refer to the operator
(2.7) |
as the -invariant Dirac operator, or simply a projectively invariant Dirac operator if no confusion arises. We write .
2.3. Twisted group -algebras
Given a multiplier on a discrete group , a unitary -representation, or projective representation, of is a map
for some Hilbert space , such that
for all .
The twisted group algebra is the associative -algebra over with basis , where the multiplication and -operation are given on basis elements by
and extended linearly and conjugate-linearly respectively. It follows from condition (ii) in Definition 2.1 that the identity element of is and that .
The reduced twisted group -algebra is constructed as follows. Consider with its usual basis , and define a projective representation
by the following action on basis elements:
Then extends naturally to an injective -representation on , called the left regular representation of . The reduced twisted group -algebra is the completion of with respect to the induced norm. When convenient, we will simply abbreviate to .
When is the trivial multiplier, and are the ordinary group algebra and reduced group -algebra respectively.
Remark 2.5.
More generally, twisted group algebras can be formed using an arbitrary group -cocycle instead of a multipler. In this case, we would have . We will use this extra flexibility in Proposition 3.9 below.
3. Weighted traces and projective indices
We now define the traces associated to conjugacy classes, in an appropriate sense, on the algebraic part of the twisted group algebra that we will use in this paper.
When is the trival multiplier, we recover from the group algebra . In this case, for any conjugacy class , the map
(3.1) |
is a trace. When is non-trivial, this formula ceases in general to define a trace on . However, for conjugacy classes of certain -regular elements, one can define a trace via a weighted sum determined by .
Definition 3.1.
Let be a multiplier on . An element is -regular if
for all .
The following equivalent formulation is useful:
Lemma 3.2.
An element is -regular if and only if for any .
Proof.
Note that
where we have used that , while applying Definition 2.1 shows that
The two expressions on the right are equal if and only if is -regular. ∎
The property of being -regular is invariant under conjugation in , hence we may speak of -regular conjugacy classes. The next lemma implies that in order to consider any traces at all on , it is necessary to deal with -regular elements (see also [20, Lemma 1.2]):
Lemma 3.3.
If is not -regular, then for any trace map , .
Proof.
If is not -regular, then by Lemma 3.2 there exists such that for . Now
Since is a trace, and hence . ∎
To define traces for -regular conjugacy classes, we will use the following weighting function. Define a function by
(3.2) |
To prove that is well-defined, as well as for later calculations, we will make use of the following lemma.
Lemma 3.4.
For any multiplier on and , we have
Proof.
Note that
Multiplying and simplifying, this equals
The claim follows by equating coefficients of . ∎
Proposition 3.5.
The function in (3.2) is well-defined.
Proof.
First note that if for some , then
Since and is -regular, this equals
By Lemma 3.4, this equals
Equating coefficients of shows that is well-defined. ∎
Remark 3.6.
Observe that .
Remark 3.7.
When is finite, it is known that the set of -regular conjugacy classes in is in bijection with the set of distinct irreducible -representations.
Definition 3.8.
Suppose that is -regular for some multiplier on . Define the -weighted -trace by
We will show that is a trace in two steps. Define
where is a -coboundary. Then is a -cocycle cohomologous to . Let be the twisted group algebra defined using (see Remark 2.5), with basis . We first prove:
Proposition 3.9.
The map defined by
is a trace.
Proof.
By definition of the multiplication in , we have for any that
Using the definition of , this is equal to
Upon cancelling and applying Remark 3.6, together with the definition of , this simplifies to . It follows that
(3.3) |
for any . By extension, this holds if is replaced by any . Indeed, if , then
Applying Lemma 3.4 and (3.3), this is equal to
To see that is a trace, it suffices to show that for any , we have . We may assume that , so that for some . Then clearly . On the other hand,
By the preceding discussion, this is equal to
hence . ∎
We now use Proposition 3.9 to deduce:
Proposition 3.10.
is a trace.
Proof.
Using that is cohomologous to via the coboundary , one verifies that the map
is an isomorphism of -algebras. We have a commutative diagram
from which it follows that is a trace. ∎
For real-valued group cocycles, the notion of regularity also applies: for any , we say that is -regular if
(3.4) |
for all . For such , let us define a function by
(3.5) |
As with the function in (3.2), one checks that is well-defined.
Corollary 3.11.
Proof.
3.1. Traces on operators
We now define weighted traces on projectively invariant operators of an appropriate class. Let , , , and be as in subsection 2.2.
Definition 3.12.
A continuous function is called a cut-off function for the -action on if for any we have
Remark 3.13.
For any proper action, a cut-off function always exists. If the action is cocompact, the cut-off function can be chosen to be compactly supported.
We have the following useful lemma:
Lemma 3.14.
Suppose is compact, and let be any cut-off function.
-
(i)
Let be a smooth function on and its lift to . Then
-
(ii)
Let be a continuous function on that is invariant under the diagonal action of . If are integrable on , then
Proof.
Following [24, section 3], let us define the following special cut-off function. First write as
for finite subgroups of and -invariant, relatively compact subsets . Then admits the open cover , along with a subordinate partition of unity . Write for the -invariant lift of to . For each , define the function by
Then extends by zero to an -invariant function on all of . Define a smooth function by
(3.7) |
By [24, Lemma 3.9], is a compactly supported, smooth cut-off function on whenever the -action is cocompact.
Definition 3.15.
Let be a multiplier on , and let be a -regular conjugacy class.
-
•
A bounded -invariant operator on is said to be of -trace class if for all ,
-
(i)
the operator is of trace class for any ;
-
(ii)
the sum
(3.8) converges absolutely.
-
(i)
-
•
If is of -trace class, we define its -weighted -trace to be
for some such that is a cut-off function on .
Remark 3.16.
Lemma 3.17.
If is a -invariant -trace class operator, then
where is a cut-off function as in Definition 3.15, and denotes the Schwartz kernel of .
Proof.
Let such that . Note that for any and , we have
Since , this implies that
Taking the trace gives
Multiplying by and taking a sum finishes the proof. ∎
Our task now is to show that satisfies the tracial property and that it is independent of the choices of , , and made in Definition 3.15. This will be carried out in Proposition 3.20, after we record some preparatory observations in the form of Lemmas 3.18 and 3.19.
Lemma 3.18.
Suppose . Then for any ,
Proof.
Observe that is equal to
On the other hand, it is also equal to , which, can be written as
By Lemma 3.4, we have . We conclude by equating coefficients. ∎
Lemma 3.19.
Let be a -regular conjugacy class for some multiplier . Then for any and , we have
Proof.
We have
The first and third equalities follow from Lemma 3.4; the second follows from the fact that . Thus it suffices to show that
(3.9) |
For this, let . Then by Lemma 3.18,
It follows from the definition of that the left-hand side of (3.9) equals
which equals by repeated applications of Lemma 3.4. ∎
With these preparations, we are now ready to prove that is tracial and well-defined independently of the choice of functions used in Definition 3.15.
Proposition 3.20.
Proof.
We begin with . Let be a -trace class operator. By the formula in Lemma 3.17, it is clear that for any cut-off function , is independent of the choice of and such that . Define the function by
so that
where such that for some cut-off function . We will show that is -invariant, whence by Lemma 3.14, is independent of
By -invariance in Definition 3.28, we see that for any ,
(3.10) |
We claim that the -summand of equals the -summand of . To prove this, it suffices to show that
(3.11) |
Applying identity (2.3) with , , and , one sees that the function
is constant on . Letting , and using the definition of in terms of given by (2.6) together with (2.4), we see that
We then have
The first equality follows from the cocycle condition. The second follows from Lemma 3.19 and the third is again by the cocycle condition. The final equality follows from (2.6).
Again by (2.3), this equals . From this (3.11) follows. Summing over elements of then shows that is -invariant, which completes the proof of (i).
For (ii), note that by Lemma 3.17,
while
(3.12) |
where we have used the change of variable and the fact that is a cut-off function for any .
Now define by
so that
We claim that
-
(1)
is -equivariant for the diagonal action on ;
-
(2)
To prove claim (1), fix some . By the change of variable , we have
Since and are -invariant, Lemma 3.29 implies that equals
We claim that the -summand of equals the -summand of , that is:
(3.13) |
but this is precisely what we have already established in (3.11). Now summing over elements of proves claim (1).
Now for claim (2), let us denote the -summand of by . Then by (3.1), it suffices to show that for each , we have
(3.14) |
To see this, note that by Lemma 3.29, we have
Using the definition of , this means that
Thus to establish (3.14), it remains to show that
By the identity (2.3), the left-hand side equals , while the right-hand side equals . Thus equality follows from the normalisation condition . This completes the proof of claim (2).
Finally, since is invariant under the diagonal -action, we may apply Lemma 3.14 (ii) to , and use claim (2), to obtain
To summarise the notation, we now have:
-
•
the trace on ,
-
•
the trace on ;
-
•
the traces and on operators on .
Recall that for any trace map on operators on a Hilbert space , the associated supertrace applied to an operator
(3.15) |
on the -graded space is
The supercommutator of homogeneous elements is
This extends linearly to arbitrarily elements of . It follows that the supertrace vanishes on supercommutators.
Definition 3.21.
Denote by , , and the supertraces associated to , , and respectively.
Then Definition 3.15 generalises easily to:
Definition 3.22.
Let be a -graded operator on written in the form (3.15). If and are of -trace class, then we define the -weighted -supertrace of to be
(3.16) |
Remark 3.23.
Proposition 3.20 implies:
Corollary 3.24.
If and are -graded -invariant operators on such that the diagonal entries of and are of -trace class, then
3.2. Twisted Roe algebras
To formulate the higher index of projectively invariant operators, we will work with certain geometric -algebras. The analogous construction in the non-twisted case is well-known; see for instance [26]. Since the discussion applies to for any , let us fix .
Definition 3.26.
Let be an operator on .
- •
-
•
The support of , denoted , is the complement of all for which there exist such that , , and
-
•
The propagation of is the extended real number
where denotes the Riemannian distance on ;
-
•
is locally compact if and for all .
Definition 3.27.
The -equivariant algebraic Roe algebra of , denoted by , is the -subalgebra of consisting of -equivariant locally compact operators with finite propagation.
The -equivariant Roe algebra of , denoted by , is the completion of in .
Definition 3.28.
Consider the vector bundle . Let denote the convolution algebra of smooth sections of such that
-
(i)
is -invariant, in the sense that
for all and ;
-
(ii)
has finite propagation, in the sense that there exists an such that whenever .
An element acts on a section by
(3.17) |
Lemma 3.29.
Suppose a -invariant operator on with finite propagation has a smooth Schwartz kernel . Then .
Proof.
The finite-propagation property for follows from the fact that has finite propagation. The assumption , together with the fact that , implies that
Conversely, an element of defines an element of via the action (3.17).
Let denote the multiplier algebra of the -equivariant Roe algebra, and write . We have a short exact sequence of -algebras
(3.18) |
3.3. The twisted higher index
We discuss the index map first in a more general context. Recall that associated to any short exact sequence of -algebras
is a cyclic exact sequence in -theory:
where the connecting maps and are defined as follows.
Definition 3.30.
-
(i)
: let be an invertible matrix with entires in representing a class in . Write
Then lifts to an invertible matrix with entries in . Then
is an idempotent, and we define
(3.19) -
(ii)
: let be an idempotent matrix with entries in representing a class in . Let be a lift of to a matrix algebra over . Then we define
(3.20)
Now let be a Riemannian spin manifold on which acts properly and isometrically, respecting the spin structure. Let be a multiplier on . Let be a trivial line bundle, and let be the twisted Dirac operator as in (2.7) acting on smooth sections of . Pick any normalizing function , i.e. a continuous, odd function such that
and form the bounded self-adjoint operator on . When is even, is naturally a direct sum , and and are odd-graded:
Proposition 3.31.
The class of in is invertible and independent of the choice of , and is an idempotent modulo .
Proof.
Since , it suffices to show that for any , we have . In fact, functions in the Schwartz algebra with compactly supported Fourier transform form a dense subset of , we may assume that is such a function. In that case, the operator is given by a smooth -invariant Schwartz kernel, and hence an element of (see Definition 3.28). It follows that . Finally, since the difference of any two normalizing functions lies in , the class of does not depend on the choice of normalising function . ∎
Definition 3.32.
For , let be the connecting maps from Definition 3.30. The -invariant higher index of on is the element
From this, we can obtain an index in the -theory of as follows. Define
(3.21) |
where is the -representation on . Then extends to an isometry
The following is easily proved:
Lemma 3.33.
The map is equivariant with respect to the projective representation on and , where is the left-regular -representation on , as in subsection 2.3.
At the level of operators, the map induces two inclusion maps
defined by first identifying operators on with operators on via conjugation by and then extending them by the zero operator and identity operator, respectively, on the orthogonal complement of . We will write , .
These maps restrict to inclusions
For each , the map extends in the obvious way to maps between matrix algebras over and its unitisation, preserving idempotence and invertibility when and respectively. We get induced maps
This gives an index
(3.22) |
where (mod ).
Remark 3.34.
When , (3.22) recovers the usual -equivariant higher index of Dirac operators on cocompact manifolds.
4. Projective PSC obstructions on cocompact manifolds
In this section we prove Theorem 1.5. Thus let , , , , and be given as in the statement of the theorem.
We begin by describing how the trace from Definition 3.8 can be extended to a trace on a smooth dense subalgebra of . Our construction is based on [8, section 6], [27, subsection 2.2], and [16, section 3], but some extra arguments are needed to make the adaptation to the proper action case, having to do with the form of the embedding in (3.3).
Let and be as in Definition 2.4, with the underlying manifold being cocompact with respect to the action of a finitely generated group . We will work with the following choice of orthonormal basis of . Let be the cut-off function for the -action on used in the embedding from (3.3). Since the support of is compact, there exists an orthonormal basis of , for some relatively compact neighbourhood of consisting of eigenfunctions of . Asymptotically, the eigenvalues satisfy for some constant . Identifying with via this basis, one checks that a smoothing operator supported in lies in the algebra
of matrices with rapidly decreasing entries [7, chapter 3], which are dense inside . Complete to an orthonormal basis of , and embed inside itself via the map . Since this map preserves , the previous identification now extends to an identification of with the new copy of in such a way that for any (see Definition 3.28), the operator is an element of . Let
(4.1) |
Fix a set of generators of , and let denote the associated word length function, where is the word metric. Let be the unbounded self-adjoint operator on given by , where the notation is as explained in subsection 2.3. Let be the unbounded self-adjoint operator on given by
(4.2) |
for , where is the eigenvalue of , and the set is as in (4.1).
Consider also the unbounded operators on and on . Note that is a closed derivation with domain , which consists of all elements such that maps to itself, and the operator , defined initially on , extends to a bounded operator on . Define
and
(4.3) |
Note that is a left ideal as well as a Freéchet subalgebra of , with respect to the Fréchet topology given by the family of norms
for , where the norm on the right-hand side is the operator norm.
Lemma 4.1.
Any can be written as a strongly convergent sum
where for each . If for each and
for all , then .
Lemma 4.2.
The -algebra is dense in and stable under the holomorphic functional calculus.
Proof.
First note that if , then
(4.5) |
which is bounded since both and are bounded. It follows that is dense in .
Now recall that a subalgebra of a Banach algebra is called spectral invariant in if the invertible elements of the unitisation are precisely those which are invertible in . In the case that is a Frechet subalgebra of of , is spectral invariant if and only if is stable under the holomorphic functional calculus in , by [23]. Further, it is a purely algebraic fact that if is a left ideal in , then is itself a spectral invariant subalgebra of .
By [15, Theorem 1.2], is dense in and closed under the holomorphic functional calculus, and hence a spectral invariant subalgebra of . Since is a left ideal in as well as a Fréchet subalgebra, it follows from the above discussion that is holomorphically closed. ∎
Let denote the operator trace on , and denote by the amplified trace on . Recall that a conjugacy class is said to have polynomial growth if there exist constants and such that the number of elements such that is at most .
Lemma 4.3.
Let be a -regular element with respect to a multiplier . If the conjugacy class has polynomial growth, then extends to a continuous trace on .
Proof.
This follows from an adaptation of the proof of [27, Lemma 2.7]. Indeed, let be an arbitrary element of , where can be written as
(4.6) |
Define the function by
(4.7) |
It follows from Definition 3.8 that agrees with on . Now since , the right-hand side of (4.7) converges absolutely by the same estimates as in the proof of [27, Lemma 2.7], hence extends to a continuous trace on . ∎
We will continue to write for the extended trace on from the proof of Lemma 4.3. It follows from and Lemmas 4.2 and 4.3 that we have an induced map
(4.8) |
We are now ready to establish the connection between and the trace from section 3.
Proposition 4.4.
Let be a -invariant operator on and a -regular element with respect to . If is of -trace class and , then
Proof.
For any , denote by
the canonical projection. Let be as given, and write as a matrix , where
(4.9) |
as in (4.6). Letting be the element of given by the matrix , one observes that
(4.10) |
Now by (4.7), together with the fact that converges absolutely on elements of , we have
Thus by Definition 3.15 it suffices to show that for each ,
(4.11) |
To this end, let be the orthonormal basis of from the start of this section. By definition,
where is as in (3.3). Hence by (4.10), we have
(4.12) |
for each and . The map can be written explicitly as follows: given and ,
Applying this to and using that gives
Thus (4.12) equals . By (3.3), , hence
(4.13) |
It follows that for any ,
which establishes (4.11). ∎
Theorem 4.5.
Proof.
For (i), note that by (4.10) and (4.13), we can write
Under the identification of with given at the start of this section, the operator is an element of . Hence by Lemma 4.1, it suffices to prove that
(4.14) |
for any , where the norm is the operator norm on .
Note that by construction, the operator is local in the sense that for any , we have , while if , then acts on as . Since the operator is supported only on the subset , and for some operators of order zero, it follows from the Gaussian decay of the heat kernel that there exist constants , , and such that for all with , we have
Since is finitely generated, there exist constants such that the number of group elements with is at most . Setting
and combining the above observations shows that
which is finite. This establishes (4.14).
For (ii), note that we have a commutative diagram
where the map is induced by the operator trace on . It follows from [13, Exercise 12.7.3] that the element
can be represented explicitly by the following difference of idempotents constructed from the heat operator:
(4.15) |
By part (i), the diagonal entries are in the unitisation of , and a similar argument shows that this is also true of the off-diagonal entries. Thus
where we have used Proposition 4.4 for the last equality.
We now turn our attention to Theorem 1.5. To begin, we have:
Proposition 4.6.
The proof of this proposition uses the Bochner-Lichnerowicz formula.
Lemma 4.7 (Bochner-Lichnerowicz).
Let be a spin Riemannian manifold, and let be the spinor bundle with connection . Let be a Hermitian connection on a Hermitian vector bundle . Then
where is the Dirac operator associated to the connection on , is the scalar curvature on , and denotes Clifford multiplication by the curvature of .
Proof of Proposition 4.6.
Let be the Hermitian connection on the trivial bundle defined by , where is any one-form satisfying . Applying Lemma 4.7 with gives
where . Thus for all , we have
by our assumption. Then can be defined by choosing the normalising function equal to the sign function, and a routine computation then shows that the index representative (6.4) is exactly zero. The final statement follows by applying the homomorphism to , as in (3.22). ∎
Corollary 4.8.
Proof.
The curvature of the connection is . Thus by Proposition 4.6, it suffices to show that
(4.16) |
for sufficiently small. For , let be the pointwise eigenvalues of , which we view as a skew-symmetric endomorphism of using the Riemannian metric. Since the action of on is cocompact, is uniformly bounded below by some , while there exists a constant such that
for each . Since is -invariant, the right-hand side is again uniformly bounded over . It follows that (4.16) holds for sufficiently small. ∎
Proposition 4.9.
Let be a -regular conjugacy class, with as in (2.6). Then the operators and are of -trace class.
Proof.
Fix and . Let be the Schwartz kernel of . By standard estimates for the heat kernel on cocompact manifolds (see for example [24, Corollary 3.5]), there exists a constant , depending on , such that
for all , where the norm is taken in . Then
for some constant , by Fubini’s theorem. Hence is of -trace class. ∎
Next, let be a subset such that
-
(i)
for all , there exists some such that ;
-
(ii)
if and , then .
Let be a cut-off function for the action of on . One checks that the function defined by
(4.17) |
is a cut-off function for the action of on .
Proposition 4.10.
Proof.
It suffices to establish this identity for . The case of general follows by replacing and by and respectively, similar to the proof of Corollary 3.11. To proceed, note that by (2.3) applied with , , and , the function
is constant on , hence
(4.19) |
for some . Letting and using that gives
Now by the cocycle identity and Lemma 3.4,
Using this and (4.21), the claimed equality (4.18) becomes
(4.20) |
Applying (2.3) with , , and now shows that the function
is constant on , hence
(4.21) |
for some . Letting shows that
This, together with the computation
We can now finish the proof of Theorem 1.5.
Proof of Theorem 1.5.
For (i), first note that
It follows from Corollary 3.24 that
(4.22) |
hence the function is constant in . Let be the Schwartz kernel of the operator . By standard heat kernel estimates, in the limit the integral
localises to arbitrarily small neighbourhoods of the fixed-point submanifold ; see for example [14, Lemma 4.10]. Pick a sufficiently small tubular neighbourhood and identify it with the normal bundle of in . Since the curvature of the twisting bundle is -invariant, the standard asymptotic expansion of (see [3, Theorem 6.11]) applies. The same argument as in the compact case (see [3, Theorem 6.16]) then shows that the above integral equals
(4.23) |
where we have used that is a cut-off function for the action of on .
By Remark 2.2, is constant on each connected component of . Further, the support of is compact and thus intersects only finitely many of these connected components, . For each , pick a point , and let be the restriction of to . By Theorem 4.5 (ii), together with the above discussion, we have
For (ii), observe that we can, by a straightforward suspension argument, reduce to the case where is even-dimensional. If admits a -invariant metric of positive scalar curvature, then vanishes for all for some , by Proposition 4.8. Since the map
is a polynomial in , it must vanish identically on , hence
(4.24) |
for each , where we may also assume that each is even-dimensional.
Remark 4.11.
It should be possible to derive part (ii) of Theorem 1.5 independently of any growth assumptions on by using methods similar to the proof of [19, Theorem 3.4]; see also [27, Remark A.2]. In particular, this would imply that if the Baum-Connes conjecture holds for , then the map from (4.8) can be defined via the -weighted -supertrace from (3.16).
Proof of Corollary 1.6.
By a suspension argument, we may assume without loss of generality that both and are even-dimensional. Further, since the vanishing property is preserved under sums of forms, we may to restrict our attention to the case where
for some , where each is the lift of a differential form on representing a class in . For each , let . Then applying to argument from the proof of Theorem 1.5 to instead of , with the twisted Dirac operator defined accordingly, shows that if admits a -invariant metric of positive scalar curvature, then there exists a such that
whenever for all . Since the left-hand side is a polynomial in the variables , it vanishes identically on . In particular, the coefficient of is zero, and this is equal to
This concludes the proof. ∎
4.1. Special cases
We discuss two special cases of Theorem 1.5 that have already appeared elsewhere in the literature. These correspond to the extreme cases when either the conjugacy class or the multiplier is trivial.
4.1.1. Free action and
In this case, the canonical trace
(4.25) |
extends continuously to a trace and induces a linear map
Since the action of on is free, we may work with the classifying space instead of , as done in [17]. Let be the classifying map for , and let . Let be a differential form on the quotient manifold such that , and let be the -invariant lift of to .
Since , Theorem 1.5 (i) reduces to the twisted -index theorem in of Mathai [17, Theorem 3.6] for the case of the spin-Dirac operator, namely that
where is a cut-off function for the -action on . Theorem 1.5 (ii) recovers the result if the admits a metric of positive scalar curvature, then for each non-negative integer ,
This is [17, Corollary 1].
4.1.2. Trivial multiplier
In this case, the projectively invariant Dirac operator is simply the -invariant Dirac operator acting on sections of the spinor bundle . The element reduces to the usual -equivariant higher index for cocompact actions [2]. The trace is then the unweighted trace from (3).
Theorem 1.5 (i) then states that if has polynomial growth, then
where we have implicitly taken to be zero. This recovers the formula [24, Theorem 6.1] in the case of the spin-Dirac operator.
5. A neighbourhood PSC obstruction
In this section we prove Theorem 1.7. We will make use of a Callias-type index theorem for projectively invariant operators.
5.1. Projectively invariant Callias-type operators
Let us begin with a general definition and discussion of projectively invariant Callias-type operators and their higher indices.
Let be a complete Riemannian manifold on which a discrete gorup acts properly and isometrically, and suppose that . Let
be the classifying map of . Let , and let in the de Rham cohomology of the orbifold . Let be the -invariant lift of to . As in subsection 2.2, we obtain a one-form , a family of functions on , and a family of multipliers on parameterised by .
Remark 5.1.
The cocycles and from (2.5) and (2.6) can be defined equivalently in terms of the above data restricted to any submanifold preserved by the action of . Indeed, by working with the de Rham differential on instead of , (2.3) implies that the family
satisfies
It follows that and can be defined equivalently as
Since the discussion in rest of this subsection applies uniformly to for any with only minor and obvious adjustments, let us now fix and write .
Let be a -graded Clifford bundle over such that is equipped with the projective action from Definition 2.3. Let be an odd-graded Dirac operator acting on smooth sections of that commutes with .
Definition 5.2.
An odd-graded, -equivariant fibrewise Hermitian bundle endomorphism of is admissible for if is an endomorphism of such that there exists a cocompact subset and a constant such that the pointwise estimate
(5.1) |
holds over . In this setting, is called a -invariant Callias-type operator.
Remark 5.3.
A bundle endomorphism on commutes with the projective action from Definition 2.3 if and only if it commutes with the unitary action of the group.
A key property of the operator is that it has an index in . The indices of such operators can be defined either via Roe algebras, similar to what was done in subsection 3.3, or using Hilbert -modules. We will take the latter approach in order to frame the discussion in parallel with that in [11] for the untwisted case.
To this end, given elements and sections , the formulas
(5.2) |
define a pre-Hilbert -module structure on .
Definition 5.4.
Let be the Hilbert -module completion of with respect to (5.1).
The admissibility condition (5.1) implies that the operator is projectively Fredholm in the following sense:
Proposition 5.5.
There exists a cocompactly supported, -invariant continuous function on such that
(5.3) |
such that the pair is a cycle in . The class is independent of the choice of .
Proof.
The proof is analogous to that of [10, Theorem 4.19]. Instead of the Hilbert module -module used there, we work with . ∎
Definition 5.6.
The -index of is the class
5.2. Localisation of projective Callias-type indices
One of the key properties of Callias-type operators in the equivariant setting is that the their indices can be calculated by localising to a cocompact subset of the manifold [11, Theorem 3.4]. In the projective setting, a similar result holds. For this, we will assume that the -invariant Dirac operator from Definition 5.2 takes the form of a Dirac operator twisted by a line bundle, as in Definition 2.4.
Let be an ungraded -equivariant Clifford bundle over equipped with a -invariant Hermitian connection . Define the Hermitian connection
(5.4) |
on a -equivariantly trivial Hermitian line bundle , and form the connection on the bundle
(5.5) |
In the notation of subsection 5.1, we will take
where the first copy of is given the even grading, and the second copy the odd grading. Let be the Dirac operator on associated to , and define
(5.6) |
on . Let be -invariant a Hermitian endomorphism of such that holds outside a cocompact subset for some . Let
(5.7) |
Then is a -invariant Callias-type operator acting on sections of , in the sense of Definition 5.2.
Let be a -invariant, cocompact subset containing in its interior, such that is a (not necessary connected) smooth submanifold of . Let be the closure of the complement of , so that and . We will use the notation
Let denote the restriction of to , equipped with the restricted connection . By (5.1), the restriction of to is fibrewise invertible. Let
be the positive and negative eigenbundles of . Clifford multiplication by times the unit normal vector field to pointing into defines -invariant gradings on both and . Define the connections
(5.8) |
on , where are the orthogonal projections. Along with the Clifford action of on , these connections give rise to two Dirac operators
(5.9) |
both odd-graded, acting on sections of and respectively.
For each group element , define unitary operators , and on by:
-
•
;
-
•
;
-
•
,
where and . Note that (2.1) and (2.2) continue to hold if we restrict , , and to and work with the de Rham differential on instead of . It follows that the operator is equivariant with respect the projective action . Since acts on cocompactly, has a -invariant higher index
by Definition 3.32 and (3.22). Equivalently, this index can be formulated as in (5.3), using a bounded transform on a Hilbert module. As mentioned previously, we will adopt the latter in order to follow more closely the exposition of [11].
With these preparations, we have the following:
Theorem 5.7 (-Callias-type index theorem).
(5.10) |
This is the projective analogue of the equivariant Callias-type index theorem [11, Theorem 3.4]. The proof is analogous to that in the untwisted setting, once we make the following modifications:
Given these similarities, we will for the most part only sketch the proof of Theorem 5.7, and invite the reader who is interested in a more detailed discussion to [11, section 5]. Nevertheless, let us give a detailed example of how one of the key technical tools used in the proof of [11, Theorem 3.4] can be adapted to the projective setting, namely the relative index theorem for Callias-type operators [11, Theorem 4.13]. The twisted analogue of that theorem is as follows.
For , let , , and be as , , , and were in Definition 5.2. Suppose there exist -invariant, cocompact hypersurfaces , -invariant tubular neighbourhoods , and a -equivariant isometry such that
-
•
;
-
•
;
-
•
, where is the Clifford connection used to define ;
-
•
corresponds to via .
Suppose that for closed, -invariant subsets . We identify and via and simply write for this manifold. Construct
For , let , and be obtained from the corresponding data on and by cutting and gluing along via . For , form the Hilbert -modules as in Definition 5.4.
Proposition 5.8.
In the above situation,
Proof.
(Compare the proof of [11, Theorem 4.13].) Define
where a superscript indicates reversal of the -grading on the given module. Similar to (5.3), define
for , and
For , let be real-valued functions such that:
-
(i)
and ;
-
(ii)
and ; and
-
(iii)
.
We view pointwise multiplication by these functions as operators
(5.11) |
Define the operator
where is the grading operator on . Then is an odd, self-adjoint operator on . Further, using properties (ii) and (iii), one verifies directly that Let denote the Clifford algebra generated by . It follows from a discussion analogous to [11, section 4] that
Since generates and anticommutes with modulo , the pair is a Kasparov -cycle. Its class is mapped to by the homomorphism induced by the pullback along the inclusion . By [4, Lemma 1.15], that homomorphism is zero. Hence
which is equivalent to the theorem. ∎
Proof of Theorem 5.7.
The first and most important step is to use Proposition 5.8 to reduce the computation of to the index of a -invariant Callias-type operator on the cylinder . This is done by following the same geometric steps as in [11, subsection 5.3], only applied to the bundle instead of the bundle used there. Where [11, Theorem 4.13] was used, we now apply Proposition 5.8. In [11], a homotopy invariance property of equivariant Callias-type operators [11, Proposition 4.9] was proved, together with the fact that the index of a Callias-type operator does not change if one modifies the potential on a cocompact subset [11, Corollary 4.10]. The proofs of both of these properties carry over to the projective setting after making the modifications (1) – (4) from above.
More precisely, the above discussion reduces the computation of to the -index of the operator (5.12) below. To define this operator, let be as in (5.5), and denote by its restriction to . Let
be the pullbacks of along the canonical projection . Then are Clifford bundles over , with Clifford action
where , , is the normal vector field to pointing into , and is Clifford multiplication on . Let and be as in (5.8) and (5.9). By pulling back along and composing with , we obtain Dirac operators acting on sections of . In particular, operator is equivariant with respect to the pull-back of the projective action to . Let be an odd function such that for all . Let be its pullback along the projection . For our Callias-type operator, we will take two copies of the bundle . Define
acting on smooth sections of . Then the endomorphism
is admissible for in the sense of Definition 5.2, and
(5.12) |
is a -invariant Callias-type operator on . By the discussion in the first paragraph of this proof, we have
(5.13) |
It then suffices to prove that
(5.14) |
which is the projective analogue of [11, Proposition 5.7]. For this, note that the operator can be written explicitly as
where is a grading on defined as times Clifford multiplication by the unit normal vector field on pointing into . The equality (5.14) then follows from the fact that the kernel of in is one-dimensional. Combining (5.13) and (5.14) concludes the proof. ∎
5.3. Proof of Theorem 1.7
Proof of Theorem 1.7.
This proof is similar to, but more subtle than, that of [11, Theorem 2.1], as it involves an additional scaling argument along with the use of an appropriate partition of unity. Hence we will give the full details.
First note that by a suspension argument, we only need to consider odd-dimensional . In this case, let be the spinor bundle over , let be the spin-Dirac operator, and let for a -trivial line bundle . For , let be the Hermitian connection on defined by the one-form . In the notation of subsection 5.2, take and be the Dirac operator associated to the connection . Let
act on sections of . We now construct a potential that is admissible for , for all , in the sense of Definition 5.2.
Let be as in the statement of the theorem. Then for disjoint open subsets and . Pick a cocompact subset of such that and on , and the distance from to is positive. Pick a -invariant function such that equals on and on . Let be the endomorphism of given by pointwise multiplication by . Now define and according to (5.6) and (5.7) respectively. Define the endomorphism
One finds that . Together with the fact that , this implies that
By construction, the estimate holds pointwise on , hence is admissible for , for all , in the sense of Definition 5.2.
Next, in the notation of subsection 5.1, take . Then is a disjoint union for a cocompact subset such that . In this case,
(5.15) |
By Theorem 5.7 and the proof of Theorem 1.5 (see also Remark 5.9), it suffices to show that for all sufficiently small . For convenience, let us write for . By a homotopy argument, for any positive , so it suffices to show that
for some and all sufficiently small . Note that the endomorphism is still admissible for , and (5.15) continues to hold. Let be an arbitrary cocompact neighbourhood of . By construction,
(5.16) |
on . On the set , we can obtain an estimate as follows. Letting be the connection used to define , we have
(5.17) |
Let by cocompactness of . Then
-
•
on : there exist such for all and , the endomorphism (5.3) is bounded below by ;
-
•
on : since and , the endomorphism (5.3) is bounded below by By cocompactness of , there exist such that for all and .
Combining this with (5.16), we see that the estimate
holds on both and for all and . To combine these these estimates, let smooth functions such that
-
•
is a partition of unity on ;
-
•
and .
For any , we may take to be sufficiently large so that . Since is a perturbation of by an endomorphism,
(5.18) |
for , hence . For any , one computes that
where the norms and inner products are taken in and we have used that . It follows that
Hence so that
Taking small enough, and hence large enough, we see that is strictly positive. Thus is invertible for and , whence . ∎
6. A quantitative obstruction in the non-cocompact setting
When is non-compact, we can use quantitative -theory to give obstructions to the existence of -invariant metrics of positive scalar curvature on . This uses the fact that the twisted Roe algebra is naturally filtered by propagation, making it an example of a geometric -algebra. We now review these concepts.
6.1. Geometric -algebras and quantitative K-theory
Definition 6.1.
A unital -algebra is geometric if it admits a filtration satisfying the following properties:
-
(i)
if ;
-
(ii)
;
-
(iii)
is dense in .
If is non-unital, then its unitization , viewed as as as a vector space, is a geometric -algebra with filtration In addition, for each , the matrix algebra is a geometric -algebra with filtration
Definition 6.2 ([5, Definition 2.15]).
Let be a geometric -algebra. For , , and ,
-
•
an element is called an -quasiidempotent if
-
•
if is unital, an element is called an -quasiinvertible if , , and there exists with
The pair is called an -quasiinverse pair.
The quantitative -groups and are defined by collecting together all quasiidempotents and quasiinvertibles over all matrix algebras, quotienting by an equivalence relation, and taking the Gröthendieck completion.
Definition 6.3 ([5, subsection 3.1]).
Let be a unital geometric -algebra. Let , , and .
-
(i)
Denote by the set of -quasiidempotents in . For each positive integer , let
We have inclusions given by Set
Define an equivalence relation on by if and are -homotopic in . Denote the equivalence class of an element by . Define addition on by
With this operation, is an abelian monoid with identity . Let denote its Grothendieck completion.
-
(ii)
Denote by the set of -quasiinvertibles in . For each positive integer , let
We have inclusions given by Set
Define an equivalence relation on by if and are -homotopic in . Denote the equivalence class of an element by . Define addition on by
With this operation, is an abelian group .
Remark 6.4.
If is a non-unital geometric -algebra, then we have a canonical -homomorphism . Using contractivity of , we have homomorphisms
where or . Define .
The following result on quasiidempotents and quasiinvertibles is useful.
Lemma 6.5 ([12, Lemma 3.4]).
Let be a geometric -algebra. If is an -idempotent in , and satisfies
then is a quasiidempotent that is -homotopic to . In particular, if
then the class of is zero in .
Suppose that is unital and is an -quasiinverse pair in . If satisfies
then is a quasiinvertible that is -homotopic to . In particular, if
then the class of is zero in .
There is a homomorphism of abelian groups
(6.1) |
preserving the -grading, where (resp. ) denotes the direct sum of the quantitative (resp. operator) and -groups.
6.2. The quantitative higher index
Fix and . For each , let the algebras and be as in Definition 3.27, with and the projective representation replaced by and , as in Definition 2.3.
For each , define the subspace
(6.2) |
of . Then with respect to the filtration
is a geometric -algebra in the sense of Definition 6.1. As in [12], this structure allows us define a refinement of the higher index that takes values in the quantitative -groups of . The construction is similar to that in [12, subsection 3.2.2], so we will be brief.
The -invariant higher index given in Definition 3.32 can be represented explicitly as follows. Let be a normalising function. If is even, define the idempotent
(6.3) |
where the notation means the -th entry of the matrix . Then is represented by the difference of idempotents
(6.4) |
For odd, can be represented by the unitary
(6.5) |
Even-dimensional
Choose a normalizing function such that
(6.6) |
Let as in (6.4). Then the -quantitative maximal higher index of is the class
Odd-dimensional
For each integer define polynomials
(6.7) |
One finds that as , the difference converges uniformly to for in the interval . Let be the smallest number such that
(6.8) |
for all . Pick a normalizing function satisfying
(6.9) |
Then the operator
has propagation at most and spectrum contained in . Then is an -quasiinvertible in the unitisation of . It was shown in [12, subsection 3.2.2] that
The -quantitative higher index of is the class
Remark 6.6.
First, although in the above constructions we needed to make a choice of , the quantitative higher index obtained is independent of this choice.
Second, the -higher index of relates to its quantitative refinement by , where is the homomorphism from (6.1).
6.3. A quantitative obstruction
We now prove Theorem 1.9. This uses the construction of the twisted higher index from subsection 2.3 in terms of twisted Roe algebras, which are geometric -algebras in the sense of [21]. The result we obtain generalises [12, Theorem 1.1].
Proof.
The technique of the proof is as in [12, section 4]. The differences are that we now work with the reduced rather than the maximal version of the twisted Roe algebra, and that bounds on used in that paper are now replaced by bounds on the endomorphism . By Lemma 4.7, we have
Suppose that holds as an estimate on operators on . Let be a normalizing function whose distributional Fourier transform is supported on some finite interval for . For each , let be the normalizing function defined by
(6.10) |
. Let be the index representative defined using .
If is even-dimensional, let
Let and a function be such that
(6.11) |
whenever for all such that , where the norm of is taken in . Note that for , (6.11) also implies that if . By (6.10), we have
(6.12) |
whenever , while
It follows that is an -quasiidempotent in -matrices over the unitisation of with norm strictly less than . By Lemma 6.5,
Letting , we obtain . Note that for any , can also be represented by . The homomorphism
induced by the inclusion
takes to , which therefore vanishes.
If is odd-dimensional, let and the polynomial be as (6.2). Let be a normalizing function satisfying (6.9), and let . Let be such that
whenever holds for all such that or, equivalently, whenever
(6.13) |
for all . Meanwhile,
Thus is an -quasiinvertible in -matrices over the unitisation of satisfying
(6.14) |
By Lemma 6.5,
Letting , we obtain . For any , the element can also be represented by . The homomorphism
induced by the inclusion
takes to , which therefore vanishes. ∎
References
- [1] Alejandro Adem, Johann Leida, and Yongbin Ruan. Orbifolds and stringy topology, volume 171 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2007.
- [2] 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. Amer. Math. Soc., Providence, RI, 1994.
- [3] Nicole Berline, Ezra Getzler, and Michèle Vergne. Heat kernels and Dirac operators. Grundlehren Text Editions. Springer-Verlag, Berlin, 2004. Corrected reprint of the 1992 original.
- [4] Ulrich Bunke. A -theoretic relative index theorem and Callias-type Dirac operators. Math. Ann., 303(2):241–279, 1995.
- [5] Yeong Chyuan Chung. Quantitative -theory for Banach algebras. J. Funct. Anal., 274(1):278–340, 2018.
- [6] A. Connes, M. Gromov, and H. Moscovici. Group cohomology with Lipschitz control and higher signatures. Geom. Funct. Anal., 3(1):1–78, 1993.
- [7] Alain Connes. Noncommutative geometry. Academic Press, Inc., San Diego, CA, 1994.
- [8] Alain Connes and Henri Moscovici. Cyclic cohomology, the Novikov conjecture and hyperbolic groups. Topology, 29(3):345–388, 1990.
- [9] 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.
- [10] Hao Guo. Index of equivariant Callias-type operators and invariant metrics of positive scalar curvature. J. Geom. Anal., 31(1):1–34, 2021.
- [11] Hao Guo, Peter Hochs, and Varghese Mathai. Positive scalar curvature and an equivariant Callias-type index theorem for proper actions. Ann. K-theory, 6(2):319–356, 2021.
- [12] Hao Guo, Zhizhang Xie, and Guoliang Yu. Quantitative -theory, positive scalar curvature, and band width. In Perspectives in Scalar Curvature (Eds. M. Gromov and H. B. Lawson). World Sci., to appear.
- [13] Nigel Higson and John Roe. Analytic -homology. Oxford Mathematical Monographs. Oxford University Press, Oxford, 2000. Oxford Science Publications.
- [14] Peter Hochs and Hang Wang. A fixed point formula and Harish-Chandra’s character formula. Proc. Lond. Math. Soc. (3), 116(1):1–32, 2018.
- [15] Ronghui Ji. Smooth dense subalgebras of reduced group -algebras, Schwartz cohomology of groups, and cyclic cohomology. J. Funct. Anal., 107(1):1–33, 1992.
- [16] Y. Kordyukov, V. Mathai, and M. Shubin. Equivalence of spectral projections in semiclassical limit and a vanishing theorem for higher traces in -theory. J. Reine Angew. Math., 581:193–236, 2005.
- [17] Varghese Mathai. -theory of twisted group -algebras and positive scalar curvature. In Tel Aviv Topology Conference: Rothenberg Festschrift (1998), volume 231 of Contemp. Math., pages 203–225. Amer. Math. Soc., Providence, RI, 1999.
- [18] Varghese Mathai. Heat kernels and the range of the trace on completions of twisted group algebras. In The ubiquitous heat kernel, volume 398 of Contemp. Math., pages 321–345. Amer. Math. Soc., Providence, RI, 2006. With an appendix by Indira Chatterji.
- [19] H. Moscovici and F.-B. Wu. Localization of topological Pontryagin classes via finite propagation speed. Geom. Funct. Anal., 4(1):52–92, 1994.
- [20] J. M. Osterburg and D. S. Passman. Trace methods in twisted group algebras. II. Proc. Amer. Math. Soc., 130(12):3495–3506, 2002.
- [21] Hervé Oyono-Oyono and Guoliang Yu. On quantitative operator -theory. Ann. Inst. Fourier (Grenoble), 65(2):605–674, 2015.
- [22] Jonathan Rosenberg. -algebras, positive scalar curvature, and the Novikov conjecture. III. Topology, 25(3):319–336, 1986.
- [23] Larry B. Schweitzer. A short proof that is local if is local and Fréchet. Internat. J. Math., 3(4):581–589, 1992.
- [24] Bai-Ling Wang and Hang Wang. Localized index and -Lefschetz fixed-point formula for orbifolds. J. Differential Geom., 102(2):285–349, 2016.
- [25] Hang Wang. -index formula for proper cocompact group actions. J. Noncommut. Geom., 8(2):393–432, 2014.
- [26] Rufus Willett and Guoliang Yu. Higher Index Theory. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2020.
- [27] Zhizhang Xie and Guoliang Yu. Delocalized Eta Invariants, Algebraicity, and K-Theory of Group C*-Algebras. Int. Math. Res. Not. IMRN, 08 2019. rnz170.