Strongly Quasi-local algebras and their -theories
Abstract.
In this paper, we introduce a notion of strongly quasi-local algebras. They are defined for each discrete metric space with bounded geometry, and sit between the Roe algebra and the quasi-local algebra. We show that strongly quasi-local algebras are coarse invariants, hence encoding coarse geometric information of the underlying spaces. We prove that for a discrete metric space with bounded geometry which admits a coarse embedding into a Hilbert space, the inclusion of the Roe algebra into the strongly quasi-local algebra induces an isomorphism in -theory.
Key words and phrases:
Roe algebras, Quasi-local algebras, Strong quasi-locality, Coarse embeddability1. Introduction
Roe algebras are -algebras associated to metric spaces which encode coarse geometric information of the underlying spaces. They were introduced by J. Roe in his pioneering work of higher index theory on open manifolds [9, 11], in which he showed that the -theory of Roe algebras serves as receptacles for indices of elliptic differential operators. Hence the computation of their K-theories becomes a central problem in higher index theory.
An efficient and practical approach is to employ the coarse Baum-Connes conjecture, which asserts that the coarse assembly map from the coarse -homology of the space to the -theory of the Roe algebra is an isomorphism [10, 18]. The coarse Baum-Connes conjecture has fruitful and significant applications in geometry and topology, for instance on the Novikov conjecture and the bounded Borel rigidity conjecture (see e.g. [2, 20, 21]), and on the non-existence of metrics of positive scalar curvature on open Riemannian manifolds (see e.g. [12, 19]).
By definition, the Roe algebra of a discrete metric space with bounded geometry is defined to be the norm closure of all locally compact operators (where is an infinite-dimensional separable Hilbert space) with finite propagation in following sense: there exists such that for any acting on by amplified pointwise multiplication, we have when their supports are -disjoint (i.e., ). Since general elements in Roe algebras may not have finite propagation, it is usually hard to detect whether a given operator belongs to them or not.
To overcome this issue, J. Roe suggested an asymptotic version of finite propagation called quasi-locality in [9, 11]. More precisely, an operator is quasi-local if for any there exists such that for any with -disjoint supports, we have . We form the quasi-local algebra111Note that a uniform version was already introduced in [7]. of as the -algebra consisting of all locally compact and quasi-local operators in , and show that they are coarse invariants. It is clear that operators with finite propagation are quasi-local, and hence the quasi-local algebra contains the Roe algebra .
A natural question is to ask whether these two algebras coincide, which has been extensively studied over the last few decades [1, 6, 8, 11, 13, 14]. Currently the most general result is due to Špakula and the third author [14], which states that for any discrete metric space with bounded geometry and having Yu’s property A. Here property A is a coarse geometric property introduced by Yu [21] in his study on the coarse Baum-Connes conjecture. However, the question remains widely open outside the world of property A [4, 7].
On the other hand, the property of quasi-locality is also crucial in the work of Engel [1] on index theory of pseudo-differential operators. He discovered that while indices of genuine differential operators on Riemannian manifolds live in the -theory of (appropriate) Roe algebras, the indices of uniform pseudo-differential operators are only known to be in the -theory of quasi-local algebras. Hence it is important to study whether the Roe algebra and the quasi-local algebra have the same -theory.
In this paper, we introduce a notion of strong quasi-locality and study associated strongly quasi-local algebras. Our main focus is to study their -theories, which might be a potential approach to attack the higher indices problem above. To illustrate the idea, let us explain when is uniformly discrete (i.e., there exists such that for ). For the general case, see Section 3.1. Fix an infinite-dimensional separable Hilbert space and denote by the unit ball of the compact operators on . We introduce the following:
Definition A.
Let be a uniformly discrete metric space with bounded geometry and . We say that is strongly quasi-local if for any there exists such that for any -Lipschitz map , we have
where is defined by for .
Definition A is inspired by a characterisation for quasi-locality provided in [13] which states that an operator is quasi-local if and only if for any there exists such that for any -Lipschitz map with , we have . Hence the notion of strong quasi-locality can be regarded as a compact operator valued version of quasi-locality, and undoubtedly strengthens the original notion (as literally suggested).
Analogous to the case of quasi-locality, we form the strongly quasi-local algebra as the -algebra consisting of all locally compact and strongly quasi-local operators in . We show that the strongly quasi-local algebra contains the Roe algebra , and is contained in the quasi-local algebra (see Proposition 3.6). We also study coarse geometric features of strongly quasi-local algebras, and show that they are coarse invariants as in the case of Roe algebras and quasi-local algebras (see Corollary 3.12).
Our motivation of introducing strongly quasi-local algebras is that their -theory is relatively easy to handle when the underlying space is coarsely embeddable. More precisely, we prove the following:
Theorem B.
Let be a discrete metric space of bounded geometry. If admits a coarse embedding into a Hilbert space, then the inclusion of the associated Roe algebra into the strongly quasi-local algebra induces an isomorphism in -theory.
Theorem B is the main result of this paper, which is inspired by the well-known theorem of Yu [21] that the coarse Baum-Connes conjecture holds for discrete bounded geometry spaces admitting a coarse embedding into Hilbert space. The proof of Theorem B follows the outline of [17, Section 12] (which originates in [21]), but is more involved and requires new techniques. We divide the proof into several steps, and here let us explain several key ingredients in the proof.
First we prove a coarse Mayer-Vietoris argument for strongly quasi-local algebras (Proposition 4.5), which allows us to cut the space and decompose the associated algebras. Recall that an analogous result for Roe algebras was already established in [3]. This leads to the reduction of the proof for Theorem B to the case of sequences of finite metric spaces with block-diagonal operators thereon (Lemma 4.7).
We would like to highlight a technical lemma used to achieve the coarse Mayer-Vietoris result. Recall that for a quasi-local operator , it is clear from definition that the restriction belongs to for any subspace . However, this is not obvious in the case of strongly quasi-local algebras due to certain obstructions on Lipschitz extension (see Remark 3.9). To overcome the issue, we provide a characterisation for strong quasi-locality in terms of compact operator valued Higson functions (Proposition 3.7). Note that these functions appeared in [16, Section 4.2] to study the stable Higson corona and the Baum-Connes conjecture. Thanks to the extendability of Higson functions, we obtain a restriction result (Lemma 3.10) as required. Moreover by some delicate analysis, we obtain a “uniform” version (Proposition 3.8) which plays a key role in following steps.
Then we construct a twisted version of strongly quasi-local algebras (Definition 5.10) for sequences of finite metric spaces, and show that the identity map on the -theory of the strongly quasi-local algebra factors through the -theory of its twisted counterpart (Proposition 6.8). To achieve, we replace several propagation requirements for twisted Roe algebras by different versions of (strong) quasi-locality, and construct an index map in terms of the Bott-Dirac operators. We would like to point out that for the original quasi-local algebras, there is a technical issue to define the index map (Lemma 6.4) following the methods either in [21, Lemma 7.6] or in [17, Lemma 12.3.9]. Hence we have to move to the world of strong quasi-locality.
Finally we prove that the inclusion map from the twisted Roe algebra into the twisted strongly quasi-local algebra induces an isomorphism in -theory (Proposition 7.1). Combining with a diagram-chasing argument, we conclude the proof for Theorem B.
Theorem B should be regarded as a first step to attack the problem whether quasi-local algebras have the same -theory as Roe algebras. More precisely, we pose the following open question:
Question C.
Let be a metric space with bounded geometry which admits a coarse embedding into Hilbert space. Then do we have ?
The paper is organised as follows: In Section 2, we collect notions from coarse geometry and recall the definition of Roe algebras. We also define quasi-local algebras and show that they are coarse invariants. In Section 3, we introduce the main concept of this paper—strong quasi-locality and study their permanence property and coarse geometric features. Section 4 is devoted to the coarse Mayer-Vietoris sequence for (strongly) quasi-local algebras, based on which we reduce the proof for Theorem B to the case of sequences of finite metric spaces. We introduce twisted strongly quasi-local algebras in Section 5, and construct the index map in Section 6. In Section 7, we show that twisted Roe algebras and twisted strongly quasi-local algebras have the same -theory, and hence conclude the proof in Section 8. The appendix provides a proof for Proposition 5.4 which slightly strengthens [17, Proposition 12.1.10] and is necessary to achieve the main theorem, hence we give detailed proofs for convenience to readers.
Acknowledgments
We wish to thank Jinmin Wang and Rufus Willett for several helpful discussions, and Yijun Yao for useful comments after reading an early version of this paper.
2. Preliminaries
We start with some notions and definitions.
2.1. Notions in coarse geometry
Here we collect several basic notions.
Definition 2.1.
Let be a metric space, and .
-
(1)
is bounded if its diameter is finite.
-
(2)
The -neighbourhood of in is .
-
(3)
is a net in if there exists some such that .
-
(4)
For , the open -ball of in is .
-
(5)
is said to be proper if every closed bounded subset is compact.
-
(6)
If is discrete, we say that has bounded geometry if for any there exists an such that for any , where denotes the cardinality of the set .
Definition 2.2.
Let be a map between metric spaces.
-
(1)
is uniformly expansive if there exists a non-decreasing function such that for any , we have:
-
(2)
is proper if for any bounded , the pre-image is bounded in .
-
(3)
is coarse if it is uniformly expansive and proper.
-
(4)
is effectively proper if there exists a proper non-decreasing function such that for any , we have:
-
(5)
is a coarse embedding if it is uniformly expansive and effectively proper.
Note that is uniformly expansive is equivalent to that the expansion function of , defined as
(2.1) |
is finite-valued.
Definition 2.3.
Let and be metric spaces.
-
(1)
Two maps are close if the exists such that for all , we have .
-
(2)
A coarse map is called a coarse equivalence if there exists another coarse map such that and are close to identities, where is called a coarse inverse to . It is clear that is a coarse equivalence if and only if it is a coarse embedding and is a net in .
-
(3)
and are said to be coarsely equivalent if there exists a coarse equivalence from to .
For families of metric spaces and maps, we also need the following notions.
Definition 2.4.
Let be a sequence of finite metric spaces. A coarse disjoint union of is a metric space where is the disjoint union of as a set, and is a metric on satisfying:
-
•
the restriction of on coincides with ;
-
•
as .
Note that any two such metric are coarsely equivalent. We say that a sequence has uniformly bounded geometry if its coarse disjoint union has bounded geometry.
Definition 2.5.
A family of maps between metric spaces is called a uniformly coarse embedding if there are non-decreasing proper functions such that
for all and . We say that uniformly coarsely embeds into Hilbert spaces if there exists a uniformly coarse embedding where each is a Hilbert space.
It is clear that a sequence of finite metric spaces uniformly coarsely embeds into Hilbert spaces if and only if its coarse disjoint union coarsely embeds into some Hilbert space.
2.2. Roe algebras and quasi-local algebras
For a proper metric space , recall that an -module is a non-degenerate -representation for some infinite-dimensional separable Hilbert space . We also say that is an -module if the representation is clear from the context. An -module is called ample if no non-zero element of acts as a compact operator on . Note that every proper metric space admits an ample -module.
Let and be ample modules of proper metric spaces and , respectively. Given an operator , the support of is defined to be
When , the propagation of is defined to be
We say that an operator has finite propagation if is finite, and is locally compact if and are compact for all (which is equivalent to that both and are compact for all compact subset ).
Definition 2.6.
For a proper metric space and an ample -module , the algebraic Roe algebra of is defined to be the -algebra of locally compact finite propagation operators on , and the Roe algebra of is defined to be the norm-closure of in .
It is a standard result that the Roe algebra does not depend on the chosen ample module up to -isomorphisms, hence denoted by and called the Roe algebra of . Furthermore, is a coarse invariant of the metric space (up to non-canonical -isomorphisms), and their K-theories are coarse invariants up to canonical isomorphisms (see, e.g., [10]).
Now we move on to the case of quasi-locality.
Definition 2.7.
Given a proper metric space and an ample -module , an operator is said to be quasi-local if for any , there exists such that has -propagation, i.e., for any Borel sets with , we have .
It is clear that the set of all locally compact quasi-local operators on forms a -subalgebra of , which leads to the following:
Definition 2.8.
For a proper metric space and an ample -module , the set of all locally compact quasi-local operators on is called the quasi-local algebra of , denoted by .
As in the case of Roe algebras, we now show that quasi-local algebras do not depend on the chosen ample modules either.
Let and be proper metric spaces and be ample modules, respectively. Let be a coarse map. Recall that a covering isometry for is an isometry such that for some . In this case, we also say that covers . It is shown in [17, Proposition 4.3.4] that covering isometries always exist. Following the case of Roe algebras, we have:
Proposition 2.9.
Let and be ample modules for proper metric spaces and , respectively. Let be a coarse map with a covering isometry . Then induces the following -homomorphism
Furthermore, the induced K-thoeretic map does not depend on the choice of the covering isometry , hence denoted by .
Proof.
Note that there exists a Borel coarse map close to by [17, Lemma A.3.12], hence without loss of generality, we can assume that is Borel and covers .
Following the same argument as in the Roe case (see, e.g., [17, Lemma 5.1.12]), is locally compact. Fix a such that . For any , the quasi-locality of implies that there exists a such that has -propagation. We set where is defined in Equation (2.1). For any Borel sets with , it is clear that and hence . Since covers , we obtain:
Hence
which implies that is quasi-local.
The second statement follows almost the same argument as in the case of Roe algebra (see, e.g., [17, Lemma 5.1.12]), hence omitted. ∎
It is shown in [17, Proposition 4.3.5] that for a coarse equivalence , we can always choose an isometry covering such that is a unitary. Consequently, we obtain the following:
Corollary 2.10.
Let and be ample modules for proper metric spaces and , respectively. If and are coarsely equivalent, then the quasi-local algebra is -isomorphic to . In particular, for a proper metric space the quasi-local algebra does not depend on the chosen ample -module up to -isomorphisms, hence called the quasi-local algebra of and denoted by .
3. Strongly quasi-local algebras
In this section, we introduce a new class of operator algebras which are called the strongly quasi-local algebras. They sit between Roe algebras and quasi-local algebras and their K-theories will be the main focus of the paper. Here we study their basic properties and coarse geometric features.
Let us begin with some more notions:
Definition 3.1.
Let be metric spaces and be a map.
-
(1)
Given , we say that is -Lipschitz if for any .
-
(2)
Given , and , we say that has -variation on if for any with , we have . When , we also say that has -variation.
Definition 3.2.
Let be a continuous function on a metric space .
-
(1)
We say that is bounded if its norm . Denote the set of all bounded continuous functions on by , and by the subset consisting of functions with norm at most .
-
(2)
We say that is a Higson function if and for any and , there exists a compact subset such that has -variation on . Denote the set of all Higson functions on .
Our notion of strong quasi-locality is inspired by the following result partially from [13, Theorem 2.8]. Recall that for operators on some Hilbert space , their commutator is defined to be .
Proposition 3.3.
Let be a proper metric space, an ample -module and be a locally compact operator. Then the following are equivalent:
-
(1)
is quasi-local in the sense of Definition 2.7;
-
(2)
For any , there exists such that for any -Lipschitz function we have ;
-
(3)
For any , there exist such that for any function with -variation we have ;
-
(4)
is a compact operator for any .
Note that the equivalence among (1), (2) and (4) are the “easier” part of [13, Theorem 2.8]. And also note that the equivalence between (1) and (3) can be proved using the same argument therein to show “(1) (2)”, hence omitted.
3.1. Strong quasi-locality
Now we introduce the notion of strong quasi-locality, where we consider compact operator valued functions instead of complex valued ones in Proposition 3.3(3).
Throughout the rest of the paper, we only consider proper discrete metric spaces to simplify the notation. We also fix an infinite-dimensional separable Hilbert space .
Let be a proper discrete metric space and be an ample -module. For each , denote . An operator can be regarded as an -by- matrix , where . Denote the -algebra of compact operators on , and its closed unit ball (with respect to the operator norm).
Recall that a map is bounded if . Given a bounded map , we define an operator by setting its matrix entry as follows:
(3.1) |
Note that this is a block-diagonal operator with respect to the decomposition . We also write instead of when we want to emphasise the module involved.
The following is the main concept of this paper:
Definition 3.4.
Let be a proper discrete metric space and be an ample -module. An operator is called strongly quasi-local if for any there exist such that for any map with -variation, we have
(3.2) |
It is easy to see that the set of all locally compact strongly quasi-local operators on forms a -algebra, hence called the strongly quasi-local algebra of and denoted by .
Remark 3.5.
A direct calculation shows that for , the -matrix entry of the commutator in Inequality (3.2) is given by:
(3.3) |
The following result records the relation amongst Roe algebras, quasi-local algebras and strongly quasi-local algebras.
Proposition 3.6.
Let be a proper discrete metric space and be an ample -module. Then we have:
-
(1)
;
-
(2)
If has bounded geometry, then ;
-
(3)
If has bounded geometry and Property A, then .
Proof.
(1). Fix a rank-one projection . For , we construct by . Since , the conclusion follows from the definition of strong quasi-locality and Proposition 3.3(3).
(2). Assume that has propagation at most . Then for any , the commutator has propagation at most from (3.3). Since has bounded geometry, it is well-known (see, e.g., [17, Lemma 12.2.4]) that there exists an depending on such that for any we have:
This concludes the proof.
(3). It follows from [14, Theorem 3.3] that under the given assumption, which (together with (1) and (2)) concludes the proof. ∎
Our next aim is to explore characterisations for strong quasi-locality as in Proposition 3.3. First note that Definition 3.4 is a compact operator valued version of condition (3) therein. Unfortunately, we cannot find an appropriate substitute for condition (1) in Proposition 3.3. As for condition (2) therein, it is clear that the compact operator valued version is equivalent to strong quasi-locality provided the underlying space is uniformly discrete (i.e., there exists such that for in ). However, it is unclear whether this holds in general.
As for condition (4) in Proposition 3.3, we have the following result concerning compact operator valued Higson functions. Recall that a compact operator valued function on a metric space is a Higson function if is bounded and for any and , there exists a compact subset such that has -variation on .
Proposition 3.7.
Let be a discrete metric space of bounded geometry and an ample -module. Then for a locally compact operator , the following are equivalent:
-
(1)
is strongly quasi-local;
-
(2)
for any Higson function .
3.2. Strong quasi-locality on subspaces
In this subsection, we study the behaviour of strong quasi-locality under taking subspace. First note that in the case of quasi-locality, we have the following observation (which follows directly from Definition 2.7): given a proper discrete metric space and an ample module , for any quasi-local operator and any there exists such that for any the operator has -propagation. In other words, quasi-locality is preserved “uniformly” under taking subspaces.
Now we focus on the case of strongly quasi-local operators, and show that they have similar behaviour when taking subspaces. However, the proof is more involved due to the lack of characterisation in terms of -propagation.
Proposition 3.8.
Let be a discrete metric space with bounded geometry and an ample -module. Assume is locally compact and strongly quasi-local. Then for any , there exist such that for any and with -variation, we have , where is naturally regarded as an operator on .
Remark 3.9.
A natural thought for the proof is to extend a function to and preserve the variation (or at least with controlled variations). However (as pointed out by Rufus Willett [15]), this is at least as hard as finding extensions with values in a Hilbert space. The problem of extending Hilbert space valued functions is fairly well-studied [5], and there are known obstructions. In the following, we will bypass the problem using Proposition 3.7.
First we prove a “subspace-wise” version of Proposition 3.8 (note the difference on orders of quantifiers). To simplify the notation, for we will regard the characteristic function either as the multiplication operator on or the amplified multiplication operator on according to the context.
Lemma 3.10.
Let be a discrete metric space with bounded geometry and an ample -module. Assume that is locally compact and strongly quasi-local. Then for any and , there exist such that for any with -variation, we have .
Proof.
By Proposition 3.7, we know that for any Higson function . Now fix a subspace . For any Higson function , it follows from [16, Lemma 4.3.4] that can be extended to a Higson function . Hence we obtain:
Using Proposition 3.7 again, we obtain that is strongly quasi-local on . This concludes the proof. ∎
Proof of Proposition 3.8.
Fix a base point , and write for where . Assume the contrary, then there exists some such that for each , there exist and with -variation on such that
(3.4) |
Without loss of generality, we can assume that each is finite. For the above , there exists such that has -propagation.
Claim. For any , there exists such that for any we have:
We assume the contrary, i.e., assume that there exist some and an increasing sequence tending to infinity such that
Since we obtain:
Now we cut up the operator as follows:
Combining the above inequalities with (3.4), we obtain:
which is a contradiction since is contained in a fixed finite subset and has -variation on . Hence we prove the Claim.
Now we continue the proof of Proposition 3.8. Set and choose such that . We recursively choose subsets , positive numbers and natural numbers as follows. Suppose that , and are chosen for . The Claim implies that there exists a natural number such that
We take (which is non-empty by the above estimate) and choose such that . In summary, we obtain non-empty subsets and functions with -variation such that
Define and extend each to by zero on the complement (still denoted by ). It is clear from the above construction that , and hence has -variation on . Moreover, we have:
This is a contradiction to Lemma 3.10. Hence we conclude the proof. ∎
3.3. Coarse invariance of strongly quasi-local algebras
In this subsection, we show that strongly quasi-local algebras are coarse invariants provided the underlying spaces have bounded geometry. In particular, this implies that strongly quasi-local algebras are independent of ample modules. The proof follows the outline of that for Proposition 2.9 but is more involved.
Proposition 3.11.
Let be discrete metric spaces with bounded geometry and be ample modules for and , respectively. Let be a coarse map with a covering isometry . Then induces the following -homomorphism
Furthermore, the induced K-thoeretic map does not depend on the choice of the covering isometry , hence denoted by .
Proof.
We only show that if . The “Furthermore” part follows almost the same argument as in the case of Roe algebra, hence omitted.
First note that is locally compact as in Proposition 2.9. To see that is strongly quasi-local, we assume that for some . Since has bounded geometry, there exists such that for any . Hence we can write:
where each satisfies , is empty for any , and for any pair we have . Set . For later use, we denote . It follows that for each there exists a map such that if and only if and .
It suffices to show that each is strongly quasi-local. Given an , there exist such that for any with -variation, we have . Set
where is defined in (2.1). For any with -variation and each , we construct as follows:
It is clear that for each and , which implies that . Hence we obtain
which implies that for each we have:
(3.5) |
On the other hand, direct calculations show that for each we have:
Hence we obtain:
where we use (3.5) in the second inequality. Note that has -variation. Hence , which implies:
Hence each is strongly quasi-local. ∎
As a direct corollary, we obtain:
Corollary 3.12.
Let and be ample modules for discrete metric spaces and of bounded geometry, respectively. If and are coarsely equivalent, then the strongly quasi-local algebra is -isomorphic to . In particular, for a discrete metric space of bounded geometry the strongly quasi-local algebra does not depend on the chosen ample -module up to -isomorphisms, hence called the strongly quasi-local algebra of and denoted by .
3.4. The case for sequences of metric spaces
Here we study the strongly quasi-local algebra for a sequence of metric spaces. This is crucial to analyse the “building blocks” when we prove our main theorem.
Let be a sequence of finite metric spaces and an ample module for . Let be a coarse disjoint union of and . Since , we can compose into a single representation:
It is clear that is an ample module for . In the following, we also regard a sequence as a single operator in .
For a locally compact operator with finite propagation, it follows directly from definition that is block-diagonal upto compact operators. Hence we have the following decomposition for Roe algebras:
Lemma 3.13.
Using the same notation as above, we have:
-
(1)
;
-
(2)
.
In the case of (strong) quasi-locality, we have similar results as follows. We only need those concerning strong quasi-locality for later use, while we collect them here for completion.
Lemma 3.14.
Using the same notation as above, we have:
-
(1)
;
-
(2)
;
-
(3)
;
-
(4)
.
Proof.
The proof is different from that for Roe algebras, and we only prove (3) and (4) since the other two are similar and easier.
For (3): note that , hence the left hand side is contained in the right one. For the converse, it follows from [13, Corollary 4.3] that for any and , there exists some such that
Since is locally compact, then is compact. It suffice to show that . Given , the strong quasi-locality of implies that there exist such that for any with -variation, we have . Now for any such , we have:
Hence we obtain that is strongly quasi-local, which concludes (3).
For (4): note that for each and hence:
Hence we conclude the proof. ∎
For later use, we introduce the following notion of (strong) quasi-locality for a sequence of operators. Note that the definition is nothing but uniform versions of (strong) quasi-locality.
Definition 3.15.
Let be a sequence of finite metric spaces and be ample modules. For a sequence where , we say that:
-
(1)
is uniformly quasi-local if for any there exists such that for any and with , we have .
-
(2)
is uniformly strongly quasi-local if for any there exist such that for any and with -variation, we have .
Lemma 3.16.
Let be a sequence of finite metric spaces, be ample modules and . For a sequence , we have:
-
(1)
if and only if is uniformly quasi-local.
-
(2)
if and only if is uniformly strongly quasi-local.
Hence if is uniformly strongly quasi-local then it is uniformly quasi-local.
The proof is straightforward, hence omitted.
Analogous to the coarse invariance of Roe algebras, we have the following result concerning sequences of spaces. The proof is similar, hence omitted.
Proposition 3.17.
Let be a sequence of finite metric spaces with uniformly bounded geometry, and be an ample module for . Let be a coarse disjoint union of and . Then the K-theories and are independent of up to canonical isomorphisms.
4. The coarse Mayer-Vietoris sequence
The tool of Mayer-Vietories sequences is widely used within different area of mathematics, especially in algebraic topology. It provides a “cutting and pasting” procedure, which allows us to obtain global information from local pieces.
In coarse geometry, Higson, Roe and Yu introduced a coarse Mayer-Vietoris sequence for K-theories of Roe algebras associated to a suitable decomposition of the underlying metric space in [3]. More precisely, recall that a closed cover of a metric space is said to be -excisive if for each there is some such that . Associated to an -excisive closed cover of a metric space , we have the following short exact sequence (called the coarse Mayer-Vietoris sequence):
In this section, we explore a coarse Mayer-Vietoris sequence for strongly quasi-local algebras and use it to reduce the proof of Theorem B to the case of “sparse” spaces. Let be a discrete metric space with bounded geometry and be an ample -module.
Definition 4.1.
Let be a (closed) subset of . Denote by the norm-closure of the set of all operators with support contained in for some .
Lemma 4.2.
is a closed two-sided -ideal in .
Proof.
It suffices to show that for with for some , then and belong to . By Proposition 3.6(1), we know that . Hence for any , there exists such that has -propagation. It follows that
Hence by definition, we obtain that . A similar argument shows that as well, which concludes the proof. ∎
Based on a similar argument as in the proof of [3, Section 5/Lemma 1] together with Corollary 3.12, we have the following:
Lemma 4.3.
For a (closed) subset , take an isometry covering the inclusion . Then the range of is contained in . Furthermore, the map is an isomorphism.
We also have the following result analogous to [3, Section 5/Lemma 2]:
Lemma 4.4.
Let be an -excisive (closed) cover of , then we have
and
Proof.
Given and , it follows from Proposition 3.6(1) that there exists such that has -propagation. Note that since , then is -close to . Hence we obtain that is dense in . It follows from a standard argument in -algebras (e.g., [3, Section 3/Lemma 1]) that .
Concerning the second equation, we only need to show that . Fix with and for some . The assumption of -excision implies that there exists an such that . Hence we have . For any there exists an such that has -propagation and has -propagation. Hence we have:
Therefore we obtain that , which concludes the proof. ∎
Applying the Mayer-Vietoris sequence in -theory for -algebras (see [3, Section 3]) to the ideals in and combining with Lemma 4.3 and Lemma 4.4, we obtain the following coarse Mayer-Vietoris principle for strongly quasi-local algebras:
Proposition 4.5.
Let be a (closed) -excisive cover of . Then there is a six-term exact sequence
For future use, we record that the same argument can be applied to obtain the Mayer-Vietoris principle for quasi-local algebras as follows. However, this will not be used in this paper.
Proposition 4.6.
Let be a (closed) -excisive cover of . Then there is a six-term exact sequence
Now we use Proposition 4.5 to reduce the proof of Thereom B to the case of block-diagonal operators:
Lemma 4.7.
To prove Theorem B for all bounded geometry metric spaces that coarsely embed into Hilbert space, it suffices to prove that for any sequence of finite metric spaces which has uniformly bounded geometry and uniformly coarsely embeds into Hilbert space, the inclusion induces isomorphisms in -theory where is an ample -module, is their direct sum and is a coarse disjoint union of .
Proof.
Lemma 3.13 and 3.14 imply that
Since for each , we obtain the following commutative diagram:
Hence the right vertical map is an isomorphism from the assumption and the Five Lemma. Now consider the following commutative diagram:
we obtain that is an isomorphism by the Five Lemma.
Now for a metric space satisfying the assumption, we follow the argument in [17, Lemma 12.5.3]. Fix a basepoint and for each , we set
Let and . It is obvious that is an -exicisive cover of . Applying the coarse Mayer-Vietoris sequences for the associated Roe algebras ([3]) and strongly quasi-local algebras (Proposition 4.5), we obtain the following commutative diagram
The left and middle vertical maps are isomorphisms according to the previous paragraph, hence we conclude the proof by the Five Lemma again. ∎
5. Twisted strongly quasi-local algebras
In this section, we recall the Bott-Dirac operators which will be used in the next section to construct index maps. We also recall the notion of twisted Roe algebras from [17, Section 12.3] (originally in [21, Section 5]) and introduce their strongly quasi-local analogue.
5.1. The Bott-Dirac operators on Euclidean spaces
Let us start by recalling some elementary properties of the Bott-Dirac operators. Here we only list necessary notions and facts, while guide readers to [17, Section 12.1] for details.
Let be a real Hilbert space (also called a Euclidean space) with even dimension . The Clifford algebra of , denoted by , is the universal unital complex algebra containing as a real subspace and subject to the multiplicative relations for all . It is natural to treat as a graded Hilbert space (see for example [17, Example E.2.12]), and in this case we denote it by .
Denote the graded Hilbert space of square integrable functions from to where the grading is inherited from , and the dense subspace consisting of Schwartz class functions from to . Fix an orthonormal basis of and let be the corresponding coordinates. Recall that the Bott operator and the Dirac operator are unbounded operators on with domain defined as:
for and , where is the operator determined by for any homogeneous element .
Definition 5.1.
The Bott-Dirac operator is the unbounded operator on with domain .
Given , recall that the left Clifford multiplication operator associated to is the bounded operator on defined as the left Clifford multiplication by the fixed vector , and the translation operator associated to is the unitary operator on defined by . Given , recall that the shrinking operator associated to is the unitary operator on defined by .
Definition 5.2.
For and , the Bott-Dirac operator associated to is the unbounded operator on with domain .
Note that and . It is also known that for each and , the operator is unbounded, odd, essentially self-adjoint and maps to itself (see, e.g., [17, Corollary 12.1.4]).
Definition 5.3.
Let , and be the Bott-Dirac operator associated to . Define a bounded operator on by:
We list several important properties of the operator . For simplicity, denote for and .
Proposition 5.4 ([17, Proposition 12.1.10]).
For each there exists an odd function with as , satisfying the following:
-
(1)
For all and , we have .
-
(2)
There exists such that for all and , we have .
-
(3)
For all and , is compact.
-
(4)
For all and , is compact.
-
(5)
For all and , . And there exists such that for all and , we have
-
(6)
For all , the function
is strong- continuous.
-
(7)
The family of functions
is norm equi-continuous as varies over and varies over any fixed compact subset of .
-
(8)
For any , the family of functions
is norm equi-continuous as varies over the elements of with , and varies over any fixed compact subset of .
-
(9)
For any , there exists such that for all , and , we have
-
(10)
For any there exists such that for all , and with , we have
Moreover, we can require the function , constants in (2), in (5), in (9) and in (10) are independent of the dimension of the Euclidean space .
5.2. Twisted Roe and strongly quasi-local algebras
Thanks to Lemma 4.7, we only focus on sequences of finite metric spaces with uniformly bounded geometry.
We fix some notation first. Let be a sequence of finite metric spaces with uniformly bounded geometry which admits a uniformly coarse embedding into Euclidean spaces where each is a Euclidean space of even dimension . Let be a coarse disjoint union of and denote .
Recall that is a fixed infinite-dimensional separable Hilbert space. Denote , which is an ample -module under the amplified multiplication representation. Denote , which is both an ample -module and an ample -module similarly. Also define and , both of which are ample -modules. For , write and for the propagation of with respect to the -module structure and the -module structure, respectively. From Definition 2.6 and Definition 3.4, we form the Roe algebras of and of , and the strongly quasi-local algebras of and of .
To introduce the twisted Roe and strongly quasi-local algebras, we need an extra construction from [17, Definition 12.3.1] which involves the information of uniformly coarse embedding as follows:
Definition 5.6.
Given and , we define a bounded operator on by the formula
for , and , where is the uniformly coarse embedding and is the translation operator defined in Section 5.1.
For each , decompose where for and . Hence can be considered as an -by- matrix operator where is a bounded operator from to . It is clear that for we have:
Hence is a block-diagonal operator with respect to the above decomposition.
Now we introduce the following twisted Roe algebras from [17, Section 12.6].
Definition 5.7.
Let denote the product -algebra of all bounded continuous functions from to with supremum norm. Write elements of this -algebra as a collection for , whose norm is
Let denote the -algebra of consisting of elements satisfying the following conditions:
-
(1)
;
-
(2)
;
-
(3)
.
The twisted Roe algebra of is defined to be the norm-closure of in .
Remark 5.8.
Lemma 5.9.
Given and a compact operator , we have
where is the net of finite rank projections on .
Proof.
Given , it suffices to find a finite rank projection such that . Replacing by its adjoint , we obtain the other equality as well.
Since is compact, there exists a finite rank projection such that . Moreover, we can assume that the image of is contained in the subspace spanned by the finite set:
Hence for each , there exists a finite rank projection such that
where is the orthogonal projection onto . Take an arbitrary finite rank projection with for each . Then we have:
This implies that . Hence we obtain
which concludes the proof. ∎
Definition 5.10.
Let denote the product -algebra of all bounded continuous functions from to with supremum norm. Write elements of this -algebra as a collection for , whose norm is
Let denote the -algebra of consisting of elements satisfying the following conditions:
-
(1)
For any , there exists such that for any and with -variation we have , where is the fixed Hilbert space and is from (3.1).
-
(2)
.
-
(3)
For any , there exists such that for each , and Borel set with we have and for all .
The twisted strongly quasi-local algebra of is defined to be the norm-closure of in .
Remark 5.11.
We provide some explanation on condition (3) in Definition 5.10. Recall that is both an -module and an -module, so we can consider the -support of a given operator defined as
We define the associated -propagation of to be
Definition 5.10(3) says that and are uniformly -quasi-local in the sense that for any there exists such that for each , and Borel set with , we have and for all . It is clear that limits of uniformly finite -propagation operators are uniformly -quasi-local.
Lemma 5.12.
We have .
Proof.
Given , condition (1) in Definition 5.10 follows from Proposition 3.6 and Lemma 3.16. We only need to check condition (3). Given , Remark 5.11 implies that it suffices to show that and have uniformly finite -propagation for and . Now Definition 5.7(1) implies that there exists an such that and for all and . Since is a uniformly coarse embedding, there exists some such that for and . It follows directly from definition that and for all and . ∎
Finally, we introduce the following operators:
Definition 5.13 ([17, Section 12.3 and 12.6]).
For each , and , Definition 5.3 provides a bounded operator , also denoted by . Applying Definition 5.6, we obtain an operator in where is the dimension of . Let be the block diagonal operator in defined by . Finally, we define to be an element in defined by .
Similarly, given let be a function as in Proposition 5.4 and be the bounded diagonal operator on defined by where . Let be the bounded operator on defined by .
6. The index map
Recall that in [17, Secition 12.3 and 12.6], Willett and Yu construct an index map (with notation as in Section 5.2):
where is the operator in Definition 5.13. They use to transfer -theoretic information from Roe algebras to their twisted counterparts, which allow them to reprove the coarse Baum-Connes conjecture via local isomorphisms. More precisely, they prove the following:
Proposition 6.1 ([17, Proposition 12.6.3]).
With notation as in Section 5.2, for each the composition
is an isomorphism, where is the evaluation map at .
In this section, we construct index maps in the strongly quasi-local setting and prove similar results. This allows us to prove certain isomorphisms in -theory to attack Theorem B later. We follow the procedure in [17, Section 12.3], while more technical analysis is required.
We follow the same notation as in Section 5.2. Let be a sequence of finite metric spaces with uniformly bounded geometry which admits a uniformly coarse embedding into Euclidean spaces where each is a Euclidean space of even dimension .
Let us start with several lemmas to analyse relations between the operator from Definition 5.13 and the twisted strongly quasi-local algebra .
Lemma 6.2.
The operator is a self-adjoint, norm one, odd operator in the multiplier algebra of .
Proof.
The operator is self-adjoint, norm one and odd since each is. Given , let be a function as in Proposition 5.4 for this . Then Proposition 5.4(1) implies:
Hence it suffices to show that belongs to for any .
First it follows from [17, Lemma 12.3.5] that for each , the map is norm-continuous. Now we verify conditions (1)-(3) in Definition 5.10 for . Note that condition (2) follows directly from Proposition 5.4(2) and (3) holds since are uniformly finite for all and . For condition (1), note that for any , and , we have
Hence we obtain:
which concludes the proof. ∎
Lemma 6.3.
Considered as represented on via amplification of identity, is a subalgebra of the multiplier algebra of .
Proof.
It suffices to show that for any and .222To be more precise, stands for . It is clear that the map is norm-continuous and bounded for each .
Now we verify conditions (1)-(3) in Definition 5.10 for . First note that for any , and we have
(6.1) |
Hence condition (1) follows from direct calculations together with Lemma 3.16.
Condition (2) follows from the fact that each has zero -propagation. Now we check condition (3). Given , it follows from that there exists such that has -propagation for all . On the other hand, there exists such that for any , , and Borel set with we have and . Now let where comes from the uniformly coarse embedding . For any , and Borel set with we have , which implies that . Therefore, we obtain:
for all . On the other hand, we have:
for all . So we finish the proof. ∎
Regarding as a subalgebra in as in Lemma 6.3, we have the following:
Lemma 6.4.
For any , we have .
Proof.
From Proposition 5.4(1), it suffices to show that the map
belongs to for any as in Proposition 5.4, i.e., to verify conditions (1)-(3) in Definition 5.10.
First note that for any , and we have
which has norm at most according to (6.1). Hence we conclude condition (1) from the strong quasi-locality of . Condition (2) follows from Propostion 5.4(2) and that fact that has zero -propagation.
To check condition (3), we fix an . It follows from Proposition 3.8 that there exist such that for any , and with -variation we have . Moreover since , we assume that has -propagation. Denote by the parameter function from the uniformly coarse embedding .
By Proposition 5.4(10), there exists such that for all and with we have . Set . For any , , and Borel set with , we are going to estimate the norm .
Denote . Since has -propagation, we obtain:
(6.2) |
Consider the function
Proposition 5.4(4) implies that for any . Moreover, we claim that has -variation on . In fact, for any with we have . Note that and , hence . Therefore by the choice of above, we obtain that has -variation on .
Finally, note that each is separable and infinite dimensional, hence isomorphic to the fixed Hilbert space . Fixing an , we define by . It follows from the above analysis that has -variation on . Hence by the choice of at the beginning, we obtain that
has norm at most . Combining with (6.2), we obtain:
Similarly, we have . Hence we conclude the proof. ∎
Lemma 6.5.
For any projection , the function
is in .
Proof.
From Lemma 6.4, it suffices to show that the function is in . Moreover, we only need to show that the function
is in for any as in Proposition 5.4. For , it follows from Proposition 5.4(7) that the function is bounded and continuous.
Now we verify conditions (1)-(3) in Definition 5.10. First note that for any , and we have
which has norm at most according to (6.1). Hence we conclude condition (1) from the strong quasi-locality of . Condition (2) follows from Propostion 5.4(2) and that fact that has zero -propagation. Finally, condition (3) follows from the uniform quasi-locality of together with Proposition 5.4(9). Hence we conclude the proof. ∎
Having obtained the above essential ingredients, we are now in the position to construct the index map. It follows from a standard construction in -theories (see, e.g., [17, Definitoin 2.8.5]):
Definition 6.6.
Let be a graded Hilbert space with grading operator (i.e., is a self-adjoint unitary operator in such that coincides with the -eigenspace of ), and be a -subalgebra of such that is in the multiplier algebra of . Let be an odd operator of the form
for some operators and . Suppose satisfies:
-
•
is in the multiplier algebra of ;
-
•
is in .
Then we define the index class of to be
Combining Lemma 6.2Lemma 6.5, we obtain that for any projection the operator is an odd self-adjoint operator on the graded Hilbert space satisfying:
-
•
is in the multiplier algebra of ;
-
•
is in .
Hence Definition 6.6 produces an index class in . Composing with the -map induced by the inclusion , we get an element in , denoted by . Analogous to [17, Lemma 12.3.11], we obtain the following:
Proposition 6.7.
Through the process above together with a suspension argument, we get well-defined -maps for
which are called the strongly quasi-local index maps.
Finally, we have the follwing result (comparing with Proposition 6.1). The proof is almost identical to that for [17, Proposition 12.3.13 and Proposition 12.6.3], hence omitted.
Proposition 6.8.
Given , let be the evaluation map at . Then the composition
is an isomorphism.
7. Isomorphisms between twisted algebras in -theory
In this section, we study the -theory of the twisted algebras and defined in Section 5.2. The main result is the following:
Proposition 7.1.
The inclusion map from to induces an isomorphism in -theory.
The proof follows the outline of that in [17, Section 12.4], and the main ingredient is to use appropriate Mayer-Vietoris arguments for twisted algebras (Proposition 7.4). This allows us to chop the space into easily-handled pieces, on which we prove the required local isomorphisms (Proposition 7.5).
By saying that is a sequence of closed subsets in , we mean that is a closed subset of for each . Firstly we define the following subalgebras associated to , which is inspired by [17, Definition 6.3.5].
Definition 7.2.
For a sequence of closed subsets in , we define to be the set of elements satisfying the following: for each and there exists such that for we have
Denote by the norm closure of in . Similarly, we define in the case of twisted Roe algebra.
It is easy to see that and are closed two-side ideals in and respectively. Moreover, we have the following:
Lemma 7.3.
Let and be two sequences of compact subsets in . Then
and
The same holds for twisted Roe algebras.
Proof.
We only prove the case of twisted strongly quasi-local algebras, while the Roe algebra case is similar. The first equation follows from a -algebraic fact that two intersections of ideals coincides with their product together with a basic fact for metric space: For a compact metric space , a closed cover of and , there exists such that .
For the second, we fix . By definition, for each there is a strictly increasing sequence in tending to infinity such that for we have
For each , we construct an operator on as follows, where . We set:
Then is in the multiplier algebra of . Now we consider:
It is clear that and are in . Also note that from the construction above, for each and we have:
Hence we obtain that is dense in , which concludes the proof. ∎
Consequently, we obtain the following Mayer-Vietoris sequences for twisted algebras:
Proposition 7.4.
Let and be two sequences of compact subsets in . Then we have the following six-term exact sequence:
The same holds in the case of twisted Roe algebra. Furthermore, we have the following commutative diagram:
where the vertical maps are induced by inclusions.
Proposition 7.4 allows us to chop the space into small pieces, on which we have the following “local isomorphism” result. Recall that a family of subspaces in a metric space is mutually -separated for some if for .
Proposition 7.5.
Let be a sequence of closed subsets in such that for a mutually -separated family and there exist and such that . Then the inclusion map from to induces an isomorphism in -theory.
Before we prove Proposition 7.5, let us use it to finish the proof of Proposition 7.1. To achieve that, we need an extra lemma from [17, Lemma 12.4.5]:
Lemma 7.6.
For any , there exist and decompositions
such that the family is mutually -separated for each and ,.
Proof of Proposition 7.1.
Given , let and be obtained by Lemma 7.6. Setting and , we have . For each applying Proposition 7.5 to the sequence of subsets , we obtain that the inclusion map
induces an isomorphism in -theory. Applying the Mayer-Vietoris sequence from Proposition 7.4 -times (and Proposition 7.5 again to deal with the intersection) together with the Five Lemma, we obtain that the inclusion map
induces an isomorphism in -theory. Finally, note that condition in Definition 5.7 and condition in Definition 5.10 imply that
Hence we conclude the proof. ∎
The rest of this section is devote to the proof of Proposition 7.5. First let us introduce some more notation:
Let and be sequences of closed subsets in . We define:
and
Also define and in a similar way. Moreover, given a sequence of subspaces we define:
and
Also define and in a similar way.
Now we move back to the setting of Proposition 7.5. Let be a sequence of closed subsets in such that for a mutually -separated family . Taking for each and , we define the “restricted product”:
Similarly, we define in the case of twisted Roe algebra.
The following lemma is a key step in the proof of Proposition 7.5:
Lemma 7.7.
Using the same notation as above, the inclusion
induces an isomorphism in -theory. The same holds for the twisted Roe algebra case.
Proof.
We only prove the case of twisted strongly quasi-local algebras, and the Roe case is similar. The proof follows the outline of [17, Theorem 6.4.20].
Consider the following quotient algebras:
where consists of elements such that for each , and is defined in a simialr way. From a standard Eilenberg Swindle argument (see for example [17, Lemma 6.4.11]), and both have trivial -theories. Hence the quotient maps
induce isomophisms in -theory.
It is clear that the inclusion induces a -homomorphism:
We also define a map
which induces a -homomorphism
We can check that the compositions and are both the identity maps. Hence is an isomorphism in -theory, which implies that the inclusion induces an isomorphism in -theory. ∎
Proof of Proposition 7.5.
We use the same notation as above and define for each and . Then there is a commutative diagram
where all maps involved are inclusion maps. It follows from Lemma 7.7 that vertical maps induce isomorphisms in -theory. Hence it suffices to show that the bottom horizontal map induces an isomorphism in -theory.
Note that conditions (3) in Definition 5.7 and 5.10 imply that
and
Hence it suffices to show that for each fixed , the inclusion
induces an isomorphism in -theory.
Note that the inclusion induces a commutative diagram
where the vertical maps are -isomorphisms by standard arguments (see for example Proposition 2.9). Also note that the bottom horizontal inclusion map is a -isomorphism as well, since conditions and in Definition 5.7 and 5.10 are equivalent in this case. Hence we conclude the proof. ∎
8. Proof of Theorem B
In this final section, we finish the proof of the main result.
Proof of Theorem B.
Consider the following commutative diagram
where the horizontal maps come from Proposition 6.1 and Proposition 6.8 and all vertical maps are induced by inclusions. From Proposition 6.1 and Proposition 6.8 again, we know that the compositions of horizontal maps are isomorphisms. The middle vertical map is an isomorphism by Proposition 7.1, and the left vertical map identifies with the right one due to Proposition 3.17. Hence the inclusion map
induces an isomorphism in -theory from diagram chasing. Finally combining with Lemma 4.7, we finish the proof. ∎
Appendix A Proof of Proposition 5.4
This appendix is devoted to the proof of Proposition 5.4. We follow the outline of that for [17, Proposition 12.1.10] and use the same notation as in Section 5.1.
Define a function by , . Also fix a smooth even function of integral one and having compactly supported Fourier transform. It follows from the proof of [17, Proposition 12.1.10] that given there exists such that the convolution satisfies condition (1)-(8) in Proposition 5.4, where for . In the following, we will prove condition (9) and (10) therein.
Let us recall the following two lemmas, which follow from [17, Lemma 12.1.6 and 12.1.8].
Lemma A.1.
For all , and , we have that
where the integral on the right converges in the strong- operator topology.
Moreover for any , and , we have that
where the integral on the right again converges in the strong- topology.
Proof.
The first formula follows from that for any , we have the formula
and functional calculus. And the second formula follows by easy computations as in the proof of [17, Lemma 12.1.6]. ∎
Lemma A.2.
For any , , and , we have that
Proof of Proposition 5.4(9)..
Given , there exists a compact subset and a function of integral one and support in such that . Setting , we have:
which implies . Hence it suffices to show that there exists such that for all and , we have
Now we set by . For any , we have:
where the last inequality comes from Lemma A.2 for . We claim that the function is bounded on . Indeed, since has support on and integral one we have:
Direct calculation shows that
which is uniformly bounded on for . Similarly, is uniformly bounded on for . Hence is bounded on .
On the other hand, note that from functional calculus. Hence we obtain that tends to zero as tends to infinity, which conclude the proof. ∎
Proof of Proposition 5.4(10)..
Given , there exists a compact subset and a function of integral one and support in such that . Setting , we have . Hence it suffices to show that for any there exists such that for any and with , we have
For any , we have:
Combining with Lemma A.1, we have
Then it is suffices to show that each of the three terms on the right side tends to zero as tends to infinity.
For the first term, note that the following constant
is finite since is compact. Hence using Lemma A.2 for , we obtain
which tends to zero as tends to infinity.
For the second term, note that
From functional calculus, for any and we have
and
Also note that the constant
is finite since is compact. Hence using Lemma A.2, we obtain
which tends to zero as tends to infinity.
Finally, let us look at the last term. Note that
It is easy to see that
Hence functional calculus gives that for any ,
Note also that functional calculus give that for any and ,
Then using Lemma A.2, we have
which tends to zero as tends to infinity. Hence we conclude the proof. ∎
References
- [1] Alexander Engel, Rough index theory on spaces of polynomial growth and contractibility, arXiv preprint arXiv:1505.03988 (2015).
- [2] Erik Guentner, Romain Tessera, and Guoliang Yu, A notion of geometric complexity and its application to topological rigidity, Invent. Math. 189 (2012), no. 2, 315–357. MR 2947546
- [3] Nigel Higson, John Roe, and Guoliang Yu, A coarse Mayer-Vietoris principle, Math. Proc. Cambridge Philos. Soc. 114 (1993), no. 1, 85–97.
- [4] Ana Khukhro, Kang Li, Federico Vigolo, and Jiawen Zhang, On the structure of asymptotic expanders, arXiv:1910.13320v2 (2019).
- [5] Mojzesz Kirszbraun, Über die zusammenziehende und lipschitzsche transformationen, Fundamenta Mathematicae 22 (1934), no. 1, 77–108.
- [6] B. V. Lange and V. S. Rabinovich, Noethericity of multidimensional discrete convolution operators, Mat. Zametki 37 (1985), no. 3, 407–421, 462. MR 790430
- [7] Kang Li, Piotr Nowak, Ján Špakula, and Jiawen Zhang, Quasi-local algebras and asymptotic expanders, arXiv:1908.07814, to appear in Groups, Geometry and Dynamics.
- [8] Kang Li, Zhijie Wang, and Jiawen Zhang, A quasi-local characterisation of -Roe algebras, Journal of Mathematical Analysis and Applications 474 (2019), no. 2, 1213–1237.
- [9] John Roe, An index theorem on open manifolds. I, II, J. Differential Geom. 27 (1988), no. 1, 87–113, 115–136. MR 918459
- [10] by same author, Coarse cohomology and index theory on complete riemannian manifolds, vol. 497, American Mathematical Soc., 1993.
- [11] by same author, Index theory, coarse geometry, and topology of manifolds, CBMS Regional Conference Series in Mathematics, vol. 90, Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1996. MR 1399087
- [12] Thomas Schick, The topology of positive scalar curvature, Proceedings of the International Congress of Mathematicians—Seoul 2014. Vol. II, Kyung Moon Sa, Seoul, 2014, pp. 1285–1307. MR 3728662
- [13] Ján Špakula and Aaron Tikuisis, Relative Commutant Pictures of Roe Algebras, Comm. Math. Phys. 365 (2019), no. 3, 1019–1048.
- [14] Ján Špakula and Jiawen Zhang, Quasi-locality and property A, Journal of Functional Analysis 278 (2020), no. 1, 108299.
- [15] R. Willett, personal communications, 2021.
- [16] Rufus Willett, Band-dominated operators and the stable Higson corona, ProQuest LLC, Ann Arbor, MI, 2009, Thesis (Ph.D.)–The Pennsylvania State University.
- [17] Rufus Willett and Guoliang Yu, Higher index theory, vol. 189, Cambridge University Press, 2020.
- [18] Guoliang Yu, Coarse Baum-Connes conjecture, -Theory 9 (1995), no. 3, 199–221. MR 1344138
- [19] by same author, Zero-in-the-spectrum conjecture, positive scalar curvature and asymptotic dimension, Invent. Math. 127 (1997), no. 1, 99–126. MR 1423027
- [20] by same author, The Novikov conjecture for groups with finite asymptotic dimension, Ann. of Math. (2) 147 (1998), no. 2, 325–355. MR 1626745
- [21] by same author, The coarse Baum-Connes conjecture for spaces which admit a uniform embedding into Hilbert space, Invent. Math. 139 (2000), no. 1, 201–240.