Functoriality for Higher Rho Invariants of Elliptic Operators
Abstract.
Let be a closed spin manifold with positive scalar curvature and the Dirac operator on . Let and be two Galois covers of such that is a quotient of . Then the quotient map from to naturally induces maps between the geometric -algebras associated to the two manifolds. We prove, by a finite-propagation argument, that the maximal higher rho invariants of the lifts of to and behave functorially with respect to the above quotient map. This can be applied to the computation of higher rho invariants, along with other related invariants.
2010 Mathematics Subject Classification:
46L80, 58B34, 53C201. Introduction
An elliptic differential operator on a closed manifold has a Fredholm index. When such an operator is lifted to a covering space, one can, by taking into account the group of symmetries, define a far-reaching generalization of the Fredholm index, called the higher index [1, 2, 6, 13, 26]. The higher index serves as an obstruction to the existence of invariant metrics of positive scalar curvature. In the case that such a metric exists, so that the higher index of the lifted operator vanishes, a secondary invariant called the higher rho invariant [19, 12] can be defined. The higher rho invariant is an obstruction to the inverse of the operator being local [4]. For some recent applications of the higher index and higher rho invariant to problems in geometry and topology, we refer the reader to [5, 22, 23, 25, 27, 28, 29].
The main purpose of this paper is to prove that the higher rho invariant behaves functorially under appropriate maps between covering spaces.
More precisely, suppose and are two Galois covers of a closed spin manifold with deck transformation groups and , where is a normal subgroup of . The natural projection induces a family of maps between various geometric -algebras associated to and , called folding maps, which will be reviewed in section 2. In particular, there is a folding map
between maximal equivariant Roe algebras on and with the property that it preserves small propagation of operators. Consequently, induces a map
at the level of obstruction algebras. The main result of this paper is that the -theoretic map induced by relates the maximal higher rho invariants of Dirac operators on and :
Theorem 1.1.
Let be a closed, spin Riemannian manifold with positive scalar curvature. Let be the Dirac operator on . Let and be Galois covers of with deck transformation groups and respectively, for a normal subgroup of . Let and be the lifts of the Dirac operator on to and respectively. Then
where is the map on -theory induced by and is the maximal higher rho invariant of .
We also generalize this result to the non-cocompact setting using the same method of proof – see section 7 of this paper.
Theorem 1.1 is useful for computing higher rho and related invariants. It has been applied in the recent work of Wang, Xie, and Yu [24] to compute the delocalized eta invariant on a covering space of a closed manifold by showing that, under suitable geometric conditions, it can be approximated by delocalized eta invariants on finite-sheeted covers, which are more computable. We mention two of their results.
Let be a closed spin manifold equipped with a positive scalar curvature metric. First, the authors proved that given a finitely generated discrete group and a sequence of finite-index normal subgroups of that distinguishes, in a suitable sense ([25, Definition 2.3]), a given non-trivial conjugacy class in , the sequence of associated delocalized eta invariants on the -Galois covers of stabilizes, under the assumption that the maximal Baum-Connes assembly map for is a rational isomorphism. We refer to [25, Theorem 1.1] for more details.
Second, the authors proved that if is the Dirac operator associated to the -Galois cover of , then if the spectral gap of at zero is sufficiently large, the delocalized eta invariant of is equal to the limit of those of Dirac operators associated to the -Galois covers of [25, Theorem 1.4]. In particular, this is true if the group has subexponential growth [25, Corollary 1.5].
Finally, we mention that in a new preprint [16] by Liu, Tang, Xie, Yao, and Yu, the authors apply Theorem 1.1 to study rigidity of relative eta invariants. These invariants are obtained by pairing the higher rho invariant with certain traces, similar to the way in which the delocalized eta invariant is related to the higher rho invariant through a trace [28].
Overview.
The paper is organized as follows. We begin in section 2 by recalling the geometric and operator-algebraic setup we work with. In section 3 we define the folding maps and establish some of their basic properties. In section 4 we develop the analytical properties of the wave operator in the maximal setting. These tools are then put to use in section 5, where we give a new proof of the functoriality property of the maximal equivariant higher index. This serves as an intermediate step towards Theorem 1.1, which is proved in section 6. In section 7 we provide generalizations of our results to the non-cocompact setting.
Acknowledgements
The authors are grateful to Peter Hochs for pointing us to the reference [17, Theorem 1.1] of Alain Valette for functoriality of the maximal higher index.
2. Preliminaries
In this section, we fix some notation before introducing the necessary operator-algebraic background and geometric setup for our results.
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. A superscript ‘∞’ may be added where appropriate to indicate the additional requirement of smoothness.
We write for the Riemannian distance function on and for the characteristic function of a subset .
For any -algebra , we denote its unitization by , its multiplier algebra by , and view as an ideal of .
The action of a group on naturally induces a -action on spaces of functions on as follows: given a function on and , define by . More generally, for a section of a -vector bundle over , the section is defined by . We say that an operator on sections of a bundle is -equivariant if it commutes with the -action.
2.2. Geometric -algebras
We now recall the notions of geometric modules and their associated -algebras. Throughout this subsection, is a Riemannian manifold equipped with a proper isometric action by a discrete group .
Definition 2.1.
An --module is a separable Hilbert space equipped with a non-degenerate -representation and a unitary representation such that for all and , we have .
For brevity, we will omit from the notation when it is clear from context.
Definition 2.2.
Let be an --module and .
-
•
The support of , denoted , is the complement of all for which there exist such that , , and
-
•
The propagation of is the extended real number
-
•
is locally compact if and for all ;
-
•
is -equivariant if for all ;
The equivariant algebraic Roe algebra for , denoted , is the -subalgebra of consisting of -equivariant, locally compact operators with finite propagation.
We will work with the maximal completion of the equivariant algebraic Roe algebra. To ensure that this completion is well-defined, we require that the module satisfy an additional admissibility condition. To define what this means, we need the following fact: if is a Hilbert space and is a non-degenerate -representation, then extends uniquely to a -representation subject to the property that, for a uniformly bounded sequence in converging pointwise, the corresponding sequence in converges in the strong topology.
Definition 2.3 ([30]).
Let be an --module as in Definition 2.1. We say that is admissible if:
-
(i)
For any non-zero we have ;
-
(ii)
For any finite subgroup of and any -invariant Borel subset , there is a Hilbert space equipped with the trivial -representation such that as -representations, where is defined by extending as above.
If an --module is admissible, we will write in place of , for the reason that is independent of the choice of admissible module – see [26, Chapter 5].
Remark 2.4.
When acts freely and properly on , the Hilbert space is an admissible --module. In the case that the action is not free, can always be embedded into a larger admissible module.
Definition 2.5.
The maximal norm of an operator is
The maximal equivariant Roe algebra of , denoted , is the completion of in the norm .
Remark 2.6.
To make sense of Definition 2.5 for general and , one first needs to establish finiteness of the quantity . It was shown in [8] that if acts on freely and properly with compact quotient, the norm is finite. This was generalized in [11] to the case when has bounded geometry and the -action satisfies a suitable geometric assumption.
Remark 2.7.
Equivalently, one can obtain by taking the analogous maximal completion of the subalgebra of consisting of those operators given by smooth Schwartz kernels.
Definition 2.8.
Consider the -algebra of functions that are uniformly bounded, uniformly continuous, and such that
-
(i)
The maximal equivariant localization algebra, denoted by , is the -algebra obtained by completing with respect to the norm
-
(ii)
The map given by extends to the evaluation map
-
(iii)
The maximal equivariant obstruction algebra is .
2.3. Geometric setup
We will work with the following geometric setup.
Let be a closed Riemannian manifold. Let be a first-order essentially self-adjoint elliptic differential operator on a bundle . We will assume throughout that if is odd-dimensional then is an ungraded operator, while if is even-dimensional then is odd-graded with respect to a -grading on .
Let and be two Galois covers of with deck transformation groups and respectively. We will assume throughout this paper that for some normal subgroup of , so that . Let be the projection map. Note that for , the group acts freely and properly on .
For or , let be the lift of the Riemannian metric to . Let be the pullback of along the covering map , equipped with the natural -action. Since acts locally, it lifts to a -equivariant operator on .
We will apply the notions in subsection 2.2 with and . The Hilbert space , equipped with the natural -action on sections and the -representation defined by pointwise multiplication, is an --module in the sense of Definition 2.2.
Moreover, the fact that the -action on is free and proper implies that is admissible in the sense of Definition 2.3. To see this, choose a compact, Borel fundamental domain for the -action on such that for each , the restriction
is a Borel isomorphism, where the projection maps are as in subsection 2.3. For each we have a map
(2.1) |
This is a -equivariant unitary isomorphism with respect to the tensor product of the left-regular representation on and the trivial representation on . Conjugation by induces a -isomorphism
(2.2) |
Remark 2.9.
3. Folding maps
In this section, we define certain natural -homomorphisms, called folding maps, between geometric -algebras of covering spaces, and discuss the role they play in functoriality for higher invariants. These maps were introduced in [3, Lemma 2.12].
To begin, let us provide some motivation at the level of groups. Observe that the quotient homomorphism induces a natural surjective -homomorphism between group algebras,
where is the class of an element in . If one views elements of as kernel operators on the Hilbert space , then the map takes a kernel to the kernel
There is an analogous map between kernels at the level of the Galois covers and . Given a smooth, -equivariant Schwartz kernel with finite propagation on , one can define a smooth, -equivariant Schwartz kernel with finite propagation on by the formula
(3.1) |
Note that this sum is finite by properness of the -action on . The formula (3.1) defines a map
where the kernel algebras and are as in Remark 2.7. This map is a -homomorphism by [9, Lemma 5.2].
3.1. Definition of the folding map
We now define a more general version of the map (3.1), called the folding map , at the level of finite-propagation operators.
Let denote the -algebra of bounded, -equivariant operators on with finite propagation. The folding map is a -homomorphism with the following properties:
-
•
For any , we have ;
-
•
is surjective;
-
•
restricts to a surjective -homomorphism .
These properties are proved in Propositions 3.3, 3.4, and 3.5 below.
To define , it will be convenient to use partitions of unity on and that are compatible with the and -actions, as follows. Since is compact, there exists such that for each , the ball is evenly covered with respect to both and . Let
(3.2) |
be a finite open cover of such that each has diameter at most . Let be a partition of unity subordinate to .
For each , let be a lift of to via the covering map . Similarly, let be the lift of to . For each , let and be the -translates of and in respectively. Then
(3.3) |
define a locally finite, -invariant open cover of and a subordinate partition of unity. This partition of unity is -equivariant in the sense that for each , , and , we have .
After taking a quotient by the -action, we obtain a -invariant open cover of , together with a subordinate -equivariant partition of unity:
(3.4) |
For each and , we have the following commutative diagram of isomorphisms of local structures involving the covering map :
We will write and for the maps relating a local section of to the corresponding local section of . That is, if is a section of over , and is the corresponding section of over , then we have
If is any vector in the bundle , we will also write for its image under the projection .
We are now ready for the definition of the folding map . Choose a set of coset representatives for . Given an operator , we define to be the operator on that takes a section to the section
(3.5) |
To clarify the presentation of this equation and others like it, we adopt the following notational convention for moving between equivalent local sections of and :
Convention 3.1.
For and , we will use the short-hand
-
•
to denote ;
-
•
to denote .
(Note that the meaning of depends on the representative .)
Using this convention, the formula (3.5) reads
(3.6) |
We call the folding map. The fact that the above definition is independent of the choices of the set and compatible partitions of unity is proved in Proposition 3.3.
Before proceeding further, let us record a few straightfoward identities that will help us navigate through notational clutter:
-
(i)
If is a local section of supported in some neighborhood , then for any we have
(3.7) (3.8) -
(ii)
If is a local section of supported in some neighborhood , then for any and we have
(3.9) (3.10)
The following lemma provides a convenient pointwise formula for :
Lemma 3.2.
For any and , we have
(3.11) |
where is any point in , and the sum on the right-hand side is finite.
Proof.
Observe that for any and , the summand may only be non-zero if . Since has finite propagation and the -action is proper, the set
is finite. Hence the sum in question is finite for every and .
To prove (3.11), let us first rewrite the right-hand side as
(3.12) |
Next, using (3.9), (3.10), and -equivariance of , each summand in (3.12) equals
(3.13) |
where we have used Convention 3.1 when writing . Now observe that if is a section of supported in some open set , then for any , we have
(3.14) |
where is an arbitrary point in the inverse image . Keeping this in mind while summing (3.1) over , , and , we see that (3.12) equals
Proposition 3.3.
The folding map defined in (3.6) is a -homomorphism
and is independent of the choices of the set of coset representatives and compatible partitions of unity used to define it.
Proof.
Boundedness of follows from boundedness and finite propagation of , via the formula (3.11). To see that is -equivariant, note that by (3.11), we have for all and that
Using (3.8) and -equivariance of , this is equal to
where we used a change-of-variable for the last equality.
To see that is a -homomorphism, let be another operator in . Let and . By (3.11), and the fact that both and have finite propagation, one sees that there exists a finite subset such that
One checks directly that respects -operations.
That (3.6) is independent of the choice of coset representatives is implied by the pointwise formula (3.11). To see that (3.6) is independent of partitions of unity, let , , and be another set of compatible partitions of unity for , , and with the same properties as , , and . Let us write
to distinguish the operators defined by (3.6) with respect to these two sets of partitions of unity. Since has finite propagation, there exists a finite subset such that if , then and for any . By (3.11), for each we have
(3.15) |
where is the characteristic function of the set , and we have used that commutes with finite sums. Likewise, we have
(3.16) |
Now observe that
are both equal to the lift of the section to restricted to the compact subset . Thus the right-hand sides of (3.1) and (3.16) are equal, whence
The next proposition shows that the folding map preserves locality of operators. In particular, this means that induces a map at the level of localization algebras (see Definition 3.8), which is a crucial ingredient for our main result, Theorem 1.1 .
Proposition 3.4.
For any operator , we have
Proof.
Take a section , and write it as a finite sum . By (3.11),
where is a lift of to . Now if , then
for any and , since distances between points do not increase under the projection . Thus if , then . Since this holds for any section , we have . ∎
Proposition 3.5.
The map is surjective and preserves local compactness of operators. Moreover, it restricts to a surjective -homomorphism
Proof.
Let be the set of coset representatives used in the definition of . For any , define an operator on by the formula
(3.17) |
for . We claim that and that . Indeed, for each fixed and , the sum of operators
converges strongly in . Since is a finite indexing set, it follows that is bounded. To see that is -equivariant, note that for any , we have, by the identities (3.7) and (3.8), that
where we observe Convention 3.1 as usual. Using -equivariance of and (3.9), one finds this to be equal to
using the definition of and a change of variable for the last equality.
Next, finite propagation of follows from finite propagation of . Indeed, for any , , and , the set
is finite, thus the sum over in equation (3.17) reduces to a finite sum over . By the same equation, and the fact that the -action is isometric, we have
We now show that . By (3.17) and (3.5) we have, for any ,
Finite propagation of and compact support of imply that all of these sums are finite, so by (3.10) this is equal to
Since is a set of coset representatives for for any , the identity (3.10) implies that the above is equal to
Thus as bounded operators on .
Finally, that preserves local compactness of operators follows directly from (3.11), so that restricts to a map . To see that this map is surjective, suppose is locally compact. Then for any , we have
The sum over reduces to a sum over the finite set
hence is equal to
Since the subset is compact, local compactness of means that this is a finite sum of compact operators and hence compact. A similar argument shows that the operator is compact. Thus the operator is locally compact, so restricts to a surjective -homomorphism . ∎
3.2. Description of the folding map on invariant sections
The folding map admits an equivalent description using -invariant sections of . Such sections, while not in general square-integrable over , form a space that is naturally isomorphic to , as we now describe. This allows computations involving to be carried out entirely on . Thus it may be a useful perspective for some applications. The discussion below slightly generalizes the averaging map from [10, subsection 5.2] applied in the context of discrete groups.
Let be a function whose support has compact intersections with every -orbit, and such that for all ,
Note that this sum is finite by properness of the -action.
Let denote the space of -transversally compactly supported sections of , defined as the space of continuous, -invariant sections of whose supports have compact images in under the quotient map. Let denote the Hilbert space of -invariant, -transversally -sections of , defined as the completion of with respect to the inner product
Then one checks that:
Lemma 3.6.
The space is naturally unitarily isomorphic to and is independent of the choice of .
For any , the operator on from (3.6) can be described equivalently by its action on the isomorphic space as follows. Take a partition of unity of as in (3.3). Let be an -transversally -section of . For any , set
(3.18) |
The fact that has finite propagation means that this pointwise sum is finite, and that . Further, is an element of , since for any and we have
where we have used the equivariance properties of and . Finally, a direct comparison shows that this definition is equivalent with that given by equation (3.11).
3.3. Induced maps on geometric -algebras and -theory
By Proposition 3.5, the folding map restricts to a surjective -homomorphism
By the defining property of maximal completions of -algebras, extends to a surjective -homomorphism between maximal equivariant Roe algebras,
In this paper we will make use of a number natural extensions of this map to other geometric -algebras, as well as the induced maps on -theory. We refer to these maps collectively as folding maps.
To begin, we have the following elementary lemma:
Lemma 3.7.
Let and be -algebras and let be a surjective -homomorphism. Then extends to a -homomorphism .
Proof.
Since is surjective, we may define by requiring
for . Concretely, by way of a faithful non-degenerate representation on some Hilbert space , we may identify with a subalgebra of . Since is surjective, the composition is a non-degenerate representation of . Thus the above formulas define a representation of with values in the idealizer of , which one identifies with . (See also [7] Lemma I.9.14.) ∎
By Lemma 3.7 and Proposition 3.5, the map extends to a -homomorphism
Viewing as an ideal in , we will still denote this extended map by . This map will be essential when we apply the functional calculus for the maximal Roe algebra in sections 5 and 6.
Next, suppose we have a path that satisfies
Then by Proposition 3.4, the same is true of the path . This allows the folding map to pass to the level of localization algebras:
Definition 3.8.
Define the map
Then restricts to a -homomorphism between obstruction algebras:
Each of the above maps , , and induces a map at the level of -theory. In this paper, we will make use of two of them:
Remark 3.9.
4. Functional calculus and the wave operator
We now develop the analytical properties of the wave operator on the maximal Roe algebra that will form the basis of our work in subsequent sections.
Let us begin by recalling the functional calculus for the maximal Roe algebra, which was developed in a more general setting in [11]. The discussion here is specialized to the cocompact setting.
Throughout this section we will work in the geometric situation in subsection 2.3. To simplify notation, in this subsection and the next we will write for either or , and for either or . In other words, acts freely and properly on with compact quotient, and is a -equivariant operator on the bundle .
4.1. Functional calculus on the maximal Roe algebra
We shall view the -algebra as a right Hilbert module over itself. The inner product and right action on are defined naturally through multiplication: for ,
(4.1) |
The algebra of compact operators on this Hilbert module can be identified with via left multiplication. Similarly, the algebra of bounded adjointable operators can be identified with the multiplier algebra of .
The operator defines an unbounded operator on this Hilbert module in the following way. First note that acts on smooth sections of the external tensor product by taking a section to the section defined by
(4.2) |
where denotes acting on the -variable.
Let be the -subalgebra of defined in Remark 2.7. That is, an element of is an operator given by a smooth kernel that:
-
(i)
is -equivariant with respect to the diagonal -action;
-
(ii)
has finite propagation, meaning that there exists a constant such that if then .
The operator acts on elements of by acting on the corresponding smooth kernels as in (4.2). In this way, becomes a densely defined operator on the Hilbert module that one verifies is symmetric with respect to the inner product in (4.1).
Theorem 4.1 ([11] Theorem 3.1).
There exists a real number such that the operators
have dense range.
Consequently, the operator on the Hilbert module is regular and essentially self-adjoint, and so admits a continuous functional calculus (see [15, Theorem 10.9] and [14, Proposition 16]):
Theorem 4.2.
For or , there is a -preserving linear map
where denotes the continuous functions , and denotes the regular operators on , such that:
-
(i)
restricts to a -homomorphism ;
-
(ii)
If for all , then ;
-
(iii)
If is a sequence in for which there exists such that for all , and if converge to a limit function uniformly on compact subsets of , then for each ;
-
(iv)
.
4.2. The wave operator
We now discuss the relationship between the wave operator formed using the functional calculus from Theorem 4.2 and the classical wave operator on . Both of these operators can be viewed as bounded multipliers of the maximal Roe algebra , and we will see that:
Proposition 4.3.
For each , we have
Let us begin by explaining both sides of the equation, starting with the right-hand side. For each , the functional calculus from Theorem 4.2 allows one to form a bounded adjointable operator on the Hilbert module . The resulting group of operators is strongly continuous in the sense of Theorem 4.2 (iii) and uniquely solves the wave equation on :
Lemma 4.4.
For any , is the unique solution of the problem
(4.3) |
with a differentiable map taking values in .
Proof.
For each , the function is a unitary in . Hence is bounded adjointable and unitary. Let be a sequence of positive real numbers converging to as . Then for each , the sequence of functions
converges to uniformly on compact subsets of in the limit . Also, each is bounded above by . By Theorem 4.2 (iii), this implies (4.3).
For the uniqueness claim, let be another solution of (4.3) with . For any fixed and , set . Then we have
It follows that is constant for all , hence
On the other hand, we may apply the functional calculus on to the essentially self-adjoint operator
(4.4) |
to form the classical wave operator , not to be confused with with the bounded adjointable operator . The resulting operator group is strongly continuous in and uniquely solves the wave equation on : for every , we have
Via composition, defines a map
(4.5) |
A standard argument involving the Sobolev embedding theorem now shows that for any , the kernel of is smooth, in addition to having finite propagation and being -equivariant. Indeed, for any , the operator is bounded on , as is compact. Together with the fact that
Sobolev theory implies that the image of lies in the smooth sections, so that this operator also has smooth kernel. By Remark 2.9, (4.5) extends uniquely to a bounded multiplier of ,
(4.6) |
This is the operator on the left-hand side of Proposition 4.3.
Observe that if is a smooth, compactly supported Schwartz kernel on , then we have the pointwise equality
(4.7) |
where denotes the smooth Schwartz kernel of the composition . The proof of Proposition 4.3, involves a more general form of this observation for kernels in :
Lemma 4.5.
Let , and let denote the smooth Schwartz kernel of . Then for every and , we have the pointwise equality
(4.8) |
Moreover, the path is continuous with respect to the operator norm, and
(4.9) |
Proof.
Fix , , and , . Take such that if , and if . Since has finite propagation, the smooth kernel is compactly supported in , so by (4.7) we have
(4.10) |
The fact that has propagation at most , together with the fact that
for all , implies that for all we have . It follows that
(4.11) |
To establish norm continuity, it suffices to show that is norm-continuous at . For each , , and in an interval , we have
(4.12) |
by the mean value theorem applied to (4.8). Since the operator is -equivariant with finite propagation, and is cocompact, there exists a compact subset such that
(4.13) |
for some constant . Now since has finite propagation, its operator norm can be estimated using and (4.12) and (4.13):
for some constant depending on the propagation of . We obtain norm continuity by taking a limit . The proof of (7.1) is a straightforward adaptation of this argument. ∎
With these preparations, we now prove Proposition 4.3.
Proof of Proposition 4.3.
Fix . Let with smooth Schwartz kernel . We claim that the kernel satisfies the wave equation in . To see this, first note that by equation (7.1) we have
Now since the kernels and each have propagation at most
by Remark 2.9 there exists a constant such that the norm of
in is bounded above by times its norm in . Thus
so the operator group solves the wave equation (4.3) in . It follows from the uniqueness property in Lemma 4.4 that and coincide on , and hence are equal as elements of . ∎
5. Functoriality for the higher index
In this section, we discuss functoriality in the case of the maximal higher index from the point of view of folding maps and wave operators. This will prepare us for the proof of our main result, Theorem 1.1, in the next section.
Suppose and are two Galois covers of with deck transformation groups and , for some normal subgroup of . Then the quotient homomorphism induces a natural surjective -homomorphism between group algebras,
where is the class of in . In this setting, we have the following functoriality result for the maximal higher index, which follows from a theorem of Valette [17, Theorem 1.1]:
Theorem 5.1.
Let be a closed Riemannian manifold. Let be a first-order, self-adjoint elliptic differential operator acting on a bundle . Let and be Galois covers of with deck transformation groups and respectively, for a normal subgroup of . Let and be the lifts of to and . Then the map on -theory induced by relates the maximal higher indices of and :
This result was proved originally using -theory. To prepare for the proof of Theorem 1.1, we now give a proof of Theorem 5.1 using local properties of the wave operator and the folding map from the previous sections.
5.1. Higher index
We begin by recalling the definition of the maximal higher index of an equivariant elliptic operator. Let , , and be as in section 4.
Let Consider the short exact sequence of -algebras
where is shorthand for the multiplier algebra . This induces the following six-term exact sequence in -theory:
where the connecting maps and , known as index maps, are defined as follows.
Definition 5.2.
-
(i)
: let be an invertible matrix over representing a class in . Let be the inverse of . Let and be lifts of and to a matrix algebra over . Then the matrix
is invertible, and is an idempotent. We define
(5.1) -
(ii)
: let be an idempotent matrix over representing a class in . Let be a lift of to a matrix algebra over . Then is a unitary in the unitized algebra, and we define
(5.2)
This construction is applied to the operator via the functional calculus from Theorem 4.2, as follows. Let be a continuous, odd function such that
known as a normalizing function. Using Theorem 4.2, we obtain an element in . We now have:
Lemma 5.3.
The class of in is invertible and independent of the choice of normalizing function .
Proof.
Let denote the Schwartz space of functions . Then for every with compactly supported Fourier transform , the operator is given by a smooth kernel [18, Proposition 2.10]. Since every function is a uniform limit of such functions, the first part of Theorem 4.2 implies that for every such we have .
Now if is a normalizing function, then . Hence the class of in is invertible. Since any two normalizing functions differ by an element of , this class is independent of the choice of . ∎
Using this lemma, one computes that
is an idempotent modulo and so defines element of . This leads us to the definition of the maximal higher index of :
Definition 5.4.
For , let be the connecting maps from Definition 5.2. The maximal higher index of is the element
5.2. Functoriality
In this subsection, we return to the geometric setup described in subsection 2.3 and give a new proof of Theorem 5.1.
A key idea is to use the local nature of the wave operator to prove:
Proposition 5.5.
For all , we have .
Proof.
For , let be the wave operator on . By (4.6), this operator extends uniquely to a bounded multiplier of . By Proposition 4.3, we have . Thus to prove this proposition, it suffices to show that , as elements of .
As a notational convenience, we will write for . Let the open covers and be as in (3.3) and (3.4), so that by definition, there exists some such that each ball of diameter in is evenly covered with respect to .
Let be another open cover of such that each has diameter at most and such that any compact subset of intersects only finitely many of the . Let be a partition of unity subordinate to .
Choose a positive integer such that . Now since
and is a -homomorphism, it suffices to show that . Noting that any section can be written as a finite sum we have
We first claim that for each we have
(5.3) |
To see this, note that the ball of radius around has diameter at most and so is evenly covered with respect to . Since the definition of is independent of the choice of compatible partitions of unity by Proposition 3.3, we may work with a partition of unity for subordinate to a cover , such that for some open set and on , for some and . By the definition of the folding map (3.6) applied to this choice of open cover, we have
Now the wave equation on reads
Applying to both sides of this equation and using that is the lift of , we obtain
whence (5.3) follows from uniqueness of the solution to the wave equation on .
Taking a sum over now yields
Thus , as bounded operators on . By our previous remarks, this means that . ∎
Applying Fourier inversion together with Proposition 5.5 leads to:
Proposition 5.6.
For any we have
Proof.
Suppose first that with compactly supported Fourier transform. By the Fourier inversion formula, we have
(5.4) |
where the integral converges strongly in . Now for any , Proposition 3.5 implies that there exists such that . Since is a -homomorphism,
By Proposition 5.5, this equals
This proves the claim for . The general claim now follows from density of in .
∎
Proof of Theorem 5.1.
The expressions (5.1) and (5.2) show that, for , the higher index of is represented by a matrix whose entries are operators formed using functional calculus of . More precisely, if the initial operator on is ungraded, as is typically the case for odd-dimensional, then
(5.5) |
When is odd-graded with respect to a -grading on the bundle , as typically occurs when is even, we have a direct sum decomposition . In this case, the index element is represented explicitly by the matrix
(5.6) |
Observe that in either case, each entry of is an operator of the form , for some (modulo grading and the identity operator). By Proposition 5.6, maps each entry of to the corresponding entry of . Hence , which proves the claim. ∎
Remark 5.7.
When and is the trivial group, Theorem 5.1 reduces to the maximal version of Atiyah’s -index theorem mentioned in section 1.
Atiyah’s original -index theorem, which uses the von Neumann trace instead of the folding map, can be proved by an argument along lines similar to the proof of Theorem 5.1.
6. Functoriality for the higher rho invariant
The higher index is a primary obstruction to the existence of positive scalar curvature metrics on a manifold. When the manifold is spin with positive scalar curvature, so that the higher index of the Dirac operator vanishes, one can define a secondary invariant called the higher rho invariant, introduced in [19, 12]. This is an obstruction to the inverse of the Dirac operator being local [4]. In this section we show that the higher rho invariant behaves functorially under the map from Definition 3.8.
6.1. Higher rho invariant
Consider the geometric situation in subsection 2.3, with the additional condition that the Riemannian manifold is spin with positive scalar curvature. The operator is then the Dirac operator acting on the spinor bundle .
As the definition of the higher rho invariant is the same for either or , we will simply write , to mean either , or , . Similarly, will refer to either of the lifted operators or acting on the equivariant spinor bundles or lifted from .
Let be the scalar curvature function of the lifted metric on , which is uniformly positive. Let be the connection on induced by the Levi-Civita connection on . Recall that by the Lichnerowicz formula,
Since is uniformly positive, is strictly positive as an unbounded operator on the Hilbert module . Thus we may use the functional calculus from Theorem 4.2 to form the operator
an element of the multiplier algebra . Observe that is a projection in .
Since is invertible, there exists such that the spectrum of is contained in . Let be a set of normalizing functions satisfying the following conditions:
-
•
has compactly supported distributional Fourier transform for each ;
-
•
as ;
-
•
uniformly on in the limit .
In the limit , the propagations of tend to . By Theorem 4.2 (i), as , the operators converge to in the norm of .
Define a path given by
(6.1) |
where the matrix is defined by (5.5) or (5.6) depending upon the dimension of (the subscript is omitted). Noting that
one sees that is a matrix with entries in .
Definition 6.1.
The higher rho invariant of on the Riemannian manifold is
where .
6.2. Functoriality
We are now ready to complete the proof of our main result, Theorem 1.1, using the tools we developed in sections 3, 4, and 5.
Recall from Definition 3.8 that we have a folding map at level of obstructions algebras,
This map is well-defined because the folding map at the level of maximal equivariant Roe algebras preserves small propagation of operators, by Proposition 3.4. The induced map on -theory,
implements functoriality of the maximal higher rho invariant.
Proof of Theorem 1.1.
For , let the higher rho invariants of be denoted by , as in Definition 6.1. By (6.1), this class is represented the path
where the matrix is as in (5.5) and (5.6). By Definition 3.8, the map takes the class
to the class of the composed path
in . Since each entry of is an operator of the form , for some (up to grading and the identity operator), Proposition 5.6 implies that for each , we have
It follows that . ∎
7. Generalizations to the non-cocompact setting
The methods in this paper can be used to establish analogous results in more general geometric settings. In this final section, we give two such generalizations, both involving non-cocompact actions.
We will work with the non-cocompact analogue of the geometric setup in subsection 2.3, so that the manifold is no longer assumed to be compact. We will assume throughout this section that the operator has unit propagation speed. In place of the finite partition of unity (3.2) used to define the folding map , we take a locally finite partition of unity whose elements are evenly covered with respect to the projections and , and with the property that any compact subset of intersects only finitely many elements of . The equivariant partitions of unity and of and are defined in the same way according to (3.3) and (3.4).
The local nature of the wave operator means that Proposition 5.5 generalizes naturally to this setting:
Proposition 7.1.
Let , , and be as in this section, with not necessarily compact. Then for all , we have .
Proof.
We adapt the proof of Proposition 5.5, indicating only what needs to be changed. Let the open covers and be as above. The difference now is that since may be non-compact, we cannot assume the existence of a uniformly positive covering diameter as in the proof of Proposition 5.5.
Instead, the key point is to observe that for any fixed and , the section is supported within the compact subset . Thus we can find such that for all , the ball is evenly covered with respect to the projection . From here, we proceed precisely as in the proof of Proposition 5.5, with replaced by , to show that . Since and are arbitrary, we conclude. ∎
7.1. Operators invertible at infinity
In this subsection, suppose that the operator is invertible at infinity, meaning that there exists some compact subset on whose complement we have for some . An important special case is when is spin, is the Dirac operator, and the metric has uniformly positive scalar curvature outside of .
For , the lifted operator then satisfies the analogous relation on the complement of a cocompact, -invariant subset . We can define a version of the higher index of , localized around , as follows (see also [20] and [10, section 3]).
For each , let be the maximal equivariant Roe algebra of the -neighborhood of . Since is cocompact, this algebra is isomorphic to by Remark 2.9. One can then show that for any , we have
where is the direct limit of these -algebras. Indeed, this limit algebra is isomorphic to , where denotes the compact operators on a (not necessarily compact) fundamental domain of the -action. The construction from subsection 5.1 then gives an index element
Similar to the cocompact case, we have the following version of Theorem 5.1 for operators that are invertible at infinity:
Theorem 7.2.
Let be a Riemannian manifold and a first-order, self-adjoint elliptic differential operator acting on a bundle . Assume that has unit propagation speed. Let and be Galois covers of with deck transformation groups and respectively, for a normal subgroup of . Let and be the lifts of to and respectively. Then the map on -theory induced by the folding map relates the maximal higher indices of and :
7.2. Manifolds with equivariantly bounded geometry
A second generalization involves non-cocompact coverings that satisfy certain additional geometric conditions. Under these conditions we obtain generalizations of both Theorems 5.1 and 1.1. In contrast to Theorem 7.2, here we do not require invertibility of the operator at infinity, and the Roe algebra we work with does not need to be defined in terms of cocompact sets as in subsection 7.1.
Again, suppose that we are in the setup of subsection 2.3, but with not necessarily compact. We impose two conditions on the geometry of and the -action on (see also [11, subsection 2.1]):
-
A.
The Riemannian manifold has positive injectivity radius, and its curvature tensor is uniformly bounded across along with all of its derivatives.
-
B.
For , there exists a fundamental domain for the action of on such that
where is a fixed length function and is the Riemannian distance on .
It follows from these two assumptions and [11, Proposition 2.14] that the maximal equivariant Roe algebra is well-defined, along with a subalgebra called the maximal equivariant uniform Roe algebra. Using the latter, we constructed in [11] a version of the functional calculus suitable for the maximal setting.
We briefly recall the definition of . Let be the -subalgebra of (see section 4) whose kernels have uniformly bounded derivatives of all orders. In other words, an element of is a bounded operator on given by a Schwartz kernel such that
-
(i)
has finite propagation;
-
(ii)
for all ;
-
(iii)
Each covariant derivative of is uniformly bounded over .
Then is defined to be the closure of in .
Similar to section 4, can be viewed as right Hilbert module over itself, with inner product and multiplication being defined by the same formula (4.1). The operator can be viewed as a densely defined symmetric operator on . By [11, Theorem 3.1], this operator is regular and so admits a functional calculus. The construction from subsection 5.1 then allows one to define the maximal equivariant uniform index of ,
The corresponding versions of the higher rho invariant can be defined as in subsection 6.1, and we denote this by
where is defined analogously to , with in place of . Constructions analogous to those in subsection 3.3 give way to folding maps at -theory level, which we denote by
Further, Proposition 4.3 generalizes naturally to manifolds satisfying conditions A and B above:
Proposition 7.3.
Proof.
Lemma 7.4.
Let , , and be as in this section, with not necessarily compact. Let , and let denote the smooth Schwartz kernel of . Then the path is continuous with respect to the operator norm, and we have
(7.1) |
Proof.
Similar to equation (4.12), we have, for each and ,
Since has bounded Riemannian geometry, it follows from Sobolev theory that the operator has uniformly bounded smooth kernel, along with all derivatives. Indeed, since is unitary, is bounded above by a constant independent of , , and . It follows that we can estimate the -norm of the finite-propagation operator by
for some other constant . We can then finish the proof as in Lemma 4.5. ∎
With this in hand, we arrive (via an obvious analogue of Proposition 5.6) at the following generalizations of Theorems 5.1 and 1.1.
Theorem 7.5.
Let be a Riemannian manifold and and be Galois covers of with deck transformation groups and respectively, for a normal subgroup of . Suppose that the conditions A and B in this subsection are satisfied. Let be a first-order, self-adjoint elliptic differential operator acting on a bundle , and assume that has unit propagation speed. Let and be the lifts of to and respectively. We then have:
Theorem 7.6.
Suppose is a spin Riemannian manifold with uniformly positive scalar curvature. Let and be Galois covers of with deck transformation groups and respectively, for a normal subgroup of . Suppose that the conditions A and B in this subsection are satisfied. Let be the Dirac operator on (with unit propagation speed). Let and be the lifts of to and respectively. We then have:
Remark 7.7.
References
- [1] Paul Baum and Alain Connes. -theory for discrete groups. In Operator algebras and applications, Vol. 1, volume 135 of London Math. Soc. Lecture Note Ser., pages 1–20. Cambridge Univ. Press, Cambridge, 1988.
- [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] Stanley Chang, Shmuel Weinberger, and Guoliang Yu. Positive scalar curvature and a new index theory for noncompact manifolds. J. Geom. Phys., 149:103575, 22, 2020.
- [4] Xiaoman Chen, Hongzhi Liu, and Guoliang Yu. Higher invariant is an obstruction to the inverse being local. J. Geom. Phys., 150:103592, 18, 2020.
- [5] Xiaoman Chen, Jinmin Wang, Zhizhang Xie, and Guoliang Yu. Delocalized eta invariants, cyclic cohomology and higher rho invariants. arXiv:1901.02378.
- [6] Alain Connes. Noncommutative geometry. Academic Press, Inc., San Diego, CA, 1994.
- [7] Kenneth R. Davidson. -algebras by example, volume 6 of Fields Institute Monographs. American Mathematical Society, Providence, RI, 1996.
- [8] Guihua Gong, Qin Wang, and Guoliang Yu. Geometrization of the strong Novikov conjecture for residually finite groups. J. Reine Angew. Math., 621:159–189, 2008.
- [9] Hao Guo, Peter Hochs, and Varghese Mathai. Coarse geometry and Callias quantisation. Trans. Amer. Math. Soc., to appear.
- [10] Hao Guo, Peter Hochs, and Varghese Mathai. Positive scalar curvature and an equivariant Callias-type index theorem for proper actions. Ann. K-theory., to appear.
- [11] Hao Guo, Zhizhang Xie, and Guoliang Yu. A Lichnerowicz vanishing theorem for the maximal Roe algebra. arXiv:1905.12299. Submitted.
- [12] Nigel Higson and John Roe. Mapping surgery to analysis. III. Exact sequences. -Theory, 33(4):325–346, 2005.
- [13] G. G. Kasparov. Equivariant -theory and the Novikov conjecture. Invent. Math., 91(1):147–201, 1988.
- [14] Dan Kucerovsky. A short proof of an index theorem. Proc. Amer. Math. Soc., 129(12):3729–3736, 2001.
- [15] E. C. Lance. Hilbert -modules, volume 210 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1995. A toolkit for operator algebraists.
- [16] Hongzhi Liu, Xiang Tang, Zhizhang Xie, Yi-Jun Yao, and Guoliang Yu. Higher rho invariants and rigidity theorems for relative eta invariants. Preprint.
- [17] Guido Mislin and Alain Valette. Proper group actions and the Baum-Connes conjecture. Advanced Courses in Mathematics. CRM Barcelona. Birkhäuser Verlag, Basel, 2003.
- [18] John Roe. An index theorem on open manifolds. I, II. J. Differential Geom., 27(1):87–113, 115–136, 1988.
- [19] John Roe. Index theory, coarse geometry, and topology of manifolds, volume 90 of CBMS Regional Conference Series in Mathematics. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1996.
- [20] John Roe. Positive curvature, partial vanishing theorems and coarse indices. Proc. Edinb. Math. Soc. (2), 59(1):223–233, 2016.
- [21] Thomas Schick and Mehran Seyedhosseini. On an index theorem of Chang, Weinberger and Yu. Münst. J. Math., to appear.
- [22] Yanli Song and Xiang Tang. Higher orbit integrals, cyclic cocyles, and -theory of reduced group -algebra. arXiv:1910.00175.
- [23] Bai-Ling Wang and Hang Wang. Localized index and -Lefschetz fixed-point formula for orbifolds. J. Differential Geom., 102(2):285–349, 2016.
- [24] Jinmin Wang, Zhizhang Xie, and Guoliang Yu. Approximations of delocalized eta invariants by their finite analogues. arXiv:2003.03401.
- [25] Shmuel Weinberger, Zhizhang Xie, and Guoliang Yu. Additivity of higher rho invariants and nonrigidity of topological manifolds. Comm. Pure Appl. Math., 74(1):3–113, 2021.
- [26] Rufus Willett and Guoliang Yu. Higher Index Theory. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2020.
- [27] Zhizhang Xie and Guoliang Yu. Higher rho invariants and the moduli space of positive scalar curvature metrics. Adv. Math., 307:1046–1069, 2017.
- [28] Zhizhang Xie and Guoliang Yu. Delocalized Eta Invariants, Algebraicity, and K-Theory of Group C*-Algebras. Int. Math. Res. Not. IMRN, 08 2019. rnz170.
- [29] Zhizhang Xie, Guoliang Yu, and Rudolf Zeidler. On the range of the relative higher index and the higher rho-invariant for positive scalar curvature. arXiv:1712.03722.
- [30] Guoliang Yu. A characterization of the image of the Baum-Connes map. In Quanta of maths, volume 11 of Clay Math. Proc., pages 649–657. Amer. Math. Soc., Providence, RI, 2010.