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

Rigidity for geometric ideals in uniform Roe algebras

Baojie Jiang and Jiawen Zhang School of Mathematical Sciences, Chongqing Normal University, Chongqing 401331, China. jiangbaojie@cqnu.edu.cn School of Mathematical Sciences, Fudan University, 220 Handan Road, Shanghai, 200433, China. jiawenzhang@fudan.edu.cn
Abstract.

In this paper, we investigate the rigidity problems for geometric ideals in uniform Roe algebras associated to discrete metric spaces of bounded geometry. These ideals were introduced by Chen and Wang, and can be fully characterised in terms of ideals in the associated coarse structures. Our main result is that if two geometric ideals in uniform Roe algebras are stably isomorphic, then the coarse spaces associated to these ideals are coarsely equivalent. We also discuss the case of ghostly ideals and pose some open questions.

Mathematics Subject Classification (2020): 47L20, 46L80, 51F30, 47L40.
Keywords: Uniform Roe algebras, Geometric ideals, Rigidity, Coarse equivalences

1. Introduction

Roe algebras are CC^{*}-algebras associated to metric spaces, which encode the coarse geometry of the underlying spaces. They were introduced by Roe in his pioneering work [16] to study higher indices of differential operators on open manifolds. There is also a uniform version of the Roe algebra, which has found applications in index theory (e.g.,[20]), CC^{*}-algebra theory (e.g., [14]), single operator theory (e.g., [22]) and even mathematical physics (e.g., [9]).

To provide a formal definition, consider a discrete metric space (X,d)(X,d) of bounded geometry (see Section 2.2). Thinking of an operator TT on 2(X)\ell^{2}(X) as an XX-by-XX matrix (T(x,y))x,yX(T(x,y))_{x,y\in X}, we say that TT has finite propagation if sup{d(x,y)|T(x,y)0}<\sup\{d(x,y)~{}|~{}T(x,y)\neq 0\}<\infty. The set of all finite propagation operators forms a \ast-subalgebra of 𝔅(2(X))\mathfrak{B}(\ell^{2}(X)), and its norm closure is called the uniform Roe algebra of XX and denoted by Cu(X)C^{*}_{u}(X).

Uniform Roe algebras have nice behaviour in coarse geometry. More precisely, recall that two metric spaces (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}) are coarsely equivalent if there exist a map f:XYf:X\to Y and functions ρ+,ρ:[0,)\rho_{+},\rho_{-}:[0,\infty)\to\mathbb{R} with limt+ρ±(t)=+\lim\limits_{t\to+\infty}\rho_{\pm}(t)=+\infty such that for any x1,x2Xx_{1},x_{2}\in X, we have

ρ(dX(x1,x2))dY(f(x1),f(x2))ρ+(dX(x1,x2)),\rho_{-}(d_{X}(x_{1},x_{2}))\leq d_{Y}(f(x_{1}),f(x_{2}))\leq\rho_{+}(d_{X}(x_{1},x_{2})),

and there exists R>0R>0 such that the RR-neighbourhood of f(X)f(X) equals YY. It is known (see, e.g., [4, Theorem 4]) that if (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}) are coarsely equivalent, then their uniform Roe algebras are stably isomorphic in the sense that Cu(X)𝔎()Cu(Y)𝔎()C^{*}_{u}(X)\otimes\mathfrak{K}(\mathcal{H})\cong C^{*}_{u}(Y)\otimes\mathfrak{K}(\mathcal{H}) (where 𝔎()\mathfrak{K}(\mathcal{H}) is the CC^{*}-algebra of compact operators on a separable infinite dimensional Hilbert space \mathcal{H}).

Conversely, the rigidity problem concerns whether the coarse geometry of a metric space can be fully determined by the associated uniform Roe algebra, i.e., whether two metric spaces are coarsely equivalent if their uniform Roe algebras are (stably) isomorphic. This problem was initially studied by Špakula and Willett in [21], followed by a series of works in the last decade (e.g., [2, 3, 11]), and recently is completely solved by the profound work [1].

Due to the importance of uniform Roe algebras, Chen and Wang initiated the study of their ideal structures ([6, 7, 23]). They managed to obtain a full description for the ideal structure of the uniform Roe algebra when the underlying space has Yu’s Property A (from [25]). More precisely, they introduced a notion of geometric ideal and provided a detailed picture for these ideals. Let us recall the definition:

Definition A ([6, 23]).

An ideal II in the uniform Roe algebra Cu(X)C^{*}_{u}(X) of a discrete metric space (X,d)(X,d) of bounded geometry is called geometric if the set of finite propagation operators in II is dense in II.

To state the characterisations for geometric ideals in [6], it is convenient to consult the notion of coarse spaces (see, e.g., [17, 18]). Recall that a coarse structure on a set XX is a collection \mathscr{E} of subsets of X×XX\times X which is closed under the formation of subsets, inverses, products and finite unions (see Definition 2.1). For (X,)(X,\mathscr{E}), we can also define the uniform Roe algebra Cu(X,)C^{*}_{u}(X,\mathscr{E}) similarly as in the metric space case while replacing finite propagation operators with those whose supports belong to \mathscr{E} (see Definition 2.5). In the case of a metric space (X,d)(X,d), there is an associated coarse structure d\mathscr{E}_{d} which is the smallest one containing the sets ER:={(x,y)|d(x,y)R}E_{R}:=\{(x,y)~{}|~{}d(x,y)\leq R\} for all R0R\geq 0, and clearly we have Cu(X,d)=Cu(X)C^{*}_{u}(X,\mathscr{E}_{d})=C^{*}_{u}(X).

In [6], the authors discovered that geometric ideals in Cu(X)C^{*}_{u}(X) for a metric space (X,d)(X,d) can be described in terms of ideals in the coarse structure d\mathscr{E}_{d}. Recall that an ideal in d\mathscr{E}_{d} is a coarse structure d\mathscr{I}\subseteq\mathscr{E}_{d} on XX which is closed under products by elements in d\mathscr{E}_{d} (see Definition 2.6). The geometric ideal in Cu(X)C^{*}_{u}(X) associated to \mathscr{I} is Cu(X,)C^{*}_{u}(X,\mathscr{I}), and this procedure provides an isomorphism between the lattice of geometric ideals in Cu(X)C^{*}_{u}(X) and the lattice of ideals in d\mathscr{E}_{d} ([6, Theorem 6.3]). For convenience, we denote (I)\mathscr{I}(I) the ideal in d\mathscr{E}_{d} associated to a geometric ideal II in Cu(X)C^{*}_{u}(X). Moreover, [6, Theorem 6.3] also shows that ideals in d\mathscr{E}_{d} can be described using ideals in (X,d)(X,d) (see Definition 2.8). The ideal in (X,d)(X,d) associated to an ideal \mathscr{I} in d\mathscr{E}_{d} is 𝐋():={r(E)|E}\mathbf{L}(\mathscr{I}):=\{\mathrm{r}(E)~{}|~{}E\in\mathscr{I}\}, where r:X×XX\mathrm{r}:X\times X\to X is the projection onto the first coordinate.

The main focus of this paper is to study the rigidity problem for geometric ideals in uniform Roe algebras of metric spaces. More precisely, we ask the following:

Question B (Rigidity for geometric ideals).

Let (X,dX),(Y,dY)(X,d_{X}),(Y,d_{Y}) be discrete metric spaces of bounded geometry, and IX,IYI_{X},I_{Y} be geometric ideals in the uniform Roe algebras Cu(X)C^{*}_{u}(X) and Cu(Y)C^{*}_{u}(Y), respectively. If IXI_{X} and IYI_{Y} are (stably) isomorphic, do (X,(IX))(X,\mathscr{I}(I_{X})) and (Y,(IY))(Y,\mathscr{I}(I_{Y})) have the same structure?

Our first task is to make the phrase “have the same structure” in a more precise way. Readers might wonder whether it is possible to use a similar notion of coarse equivalence as in the metric space case recalled above. This works well if the coarse structures contain the diagonals (see, e.g., Definition 2.4 and the paragraph thereafter). (Note that in [18, Definition 2.3], one requires that a coarse structure always contains the diagonal. While in [6, 19], it is necessary to consider the more general case.) However in the general case, there would be some issue if we only consider a single map (as in Definition 2.4) due to the lack of units in geometric ideals. More precisely, note that a nontrivial geometric ideal II in Cu(X)C^{*}_{u}(X) does not bare a unit, and hence the associated ideal (I)\mathscr{I}(I) does not contain the diagonal.

To overcome this issue, we need to consult the notion of coarse equivalence for general coarse spaces introduced by Skandalis, Tu and Yu:

Definition C ([19, Definition 2.2]).

Let (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) be coarse spaces. A coarse correspondence from (X,X)(X,\mathscr{E}_{X}) to (Y,Y)(Y,\mathscr{E}_{Y}) is a coarse structure \mathscr{E} on XYX\sqcup Y which restricts to Y\mathscr{E}_{Y} on YY, contains X\mathscr{E}_{X}, and is generated by the elements contained in Y×(XY)Y\times(X\sqcup Y). A coarse equivalence between (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) is a coarse structure on XYX\sqcup Y which is a coarse correspondence from XX to YY and from YY to XX.

As noted in [19, Proposition 2.3] (see also Proposition 3.6 and Corollary 3.7), Definition C coincides with the notion of coarse equivalence recalled above in the metric space case. However, it seems inconvenient to use directly the language of coarse correspondence to treat the rigidity problem. Hence we unpack Definition C by means of families of maps and prove the following:

Proposition D (Corollary 3.24).

Two coarse spaces (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) are coarsely equivalent if and only if there exist coarse families (see Definition 3.14) of maps

={fL:LY|L𝐋(X)}and𝒢={gL:LX|L𝐋(Y)}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}(\mathscr{E}_{X})\}\quad\text{and}\quad\mathcal{G}=\{g_{L^{\prime}}:L^{\prime}\to X~{}|~{}L^{\prime}\in\mathbf{L}(\mathscr{E}_{Y})\}

such that {(gfL(L)fL(x),x)|xL}X\{(g_{f_{L}(L)}\circ f_{L}(x),x)~{}|~{}x\in L\}\in\mathscr{E}_{X} and {(fgL(L)gL(y),y)|yL}Y\{(f_{g_{L^{\prime}}(L^{\prime})}\circ g_{L^{\prime}}(y),y)~{}|~{}y\in L^{\prime}\}\in\mathscr{E}_{Y} for any L𝐋(X)L\in\mathbf{L}(\mathscr{E}_{X}) and L𝐋(Y)L^{\prime}\in\mathbf{L}(\mathscr{E}_{Y}).

Having established Proposition D, we manage to answer Question B completely. The following is the main result of this paper (see Section 4 and Section 6 for the missing definitions):

Theorem E.

Let (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}) be discrete metric spaces of bounded geometry, and IX,IYI_{X},I_{Y} be geometric ideals in the uniform Roe algebras Cu(X)C^{*}_{u}(X) and Cu(Y)C^{*}_{u}(Y), respectively. Consider the following conditions:

  1. (1)

    IXI_{X} and IYI_{Y} are stably isomorphic;

  2. (2)

    (X,(IX))(X,\mathscr{I}(I_{X})) and (Y,(IY))(Y,\mathscr{I}(I_{Y})) are coarsely equivalent;

  3. (3)

    IXI_{X} and IYI_{Y} are Morita equivalent.

Then we have (1) \Rightarrow (2) \Rightarrow (3). Additionally if the associated ideals 𝐋((IX))\mathbf{L}(\mathscr{I}(I_{X})) and 𝐋((IY))\mathbf{L}(\mathscr{I}(I_{Y})) are countably generated, then the three conditions above are all equivalent.

Note that “(2) \Rightarrow (3)” is known to experts, and the key step is to prove “(1) \Rightarrow (2)”. The strategy follows from that for [1, Theorem 1.4], which is the standard approach originated in [21]. However, there are technical issues due to the lack of units in geometric ideals. To overcome this, we discover concrete approximate units in geometric ideals (see Lemma 2.12), which have close relations to the coarse geometry of the underlying spaces and the associated ideals. Hence we can approximate a geometric ideal by a net of uniform Roe algebras of subspaces, and therefore manage to consult the proof in the unital case.

At this point, we would also like to highlight that Theorem E is not just an easy combination of results in [3, Section 4] and [1]. Recall that in [3], the rigidity problem was studied for general coarse spaces. Note that there is an extra hypothesis that coarse spaces are small in the sense of [3, Definition 4.2] to prove that rigid isomorphisms induce coarse equivalences. However in the current setting, coarse structures associated to geometric ideals might not be small in general. This obstructs us from directly using the techniques from [3, Section 4]. To overcome the issue, again we make use of the concrete approximate units to reduce to the unital case.

Also note that in a recent work [24], Wang and the second-named author introduced a notion of ghostly ideals and studied the ideal structure of uniform Roe algebras beyond the scope of Property A. We provide some discussions on the rigidity problems for ghostly ideals and pose some open questions.

The paper is organised as follows. In Section 2, we recall necessary background knowledge in coarse geometry and uniform Roe algebras. In Section 3, we study the notion of coarse equivalence for general coarse spaces (Definition C) and prove Proposition D. Section 4 to Section 6 are devoted to the proof for Theorem E. We record the proof for the easy direction “(2) \Rightarrow (3)” in Section 4 and prove “(1) \Rightarrow (2)” in Section 6. To make the argument in Section 6 more transparent, we prove a weak version of “(1) \Rightarrow (2)” in Section 5. Finally, we provide some discussion on ghostly ideals in Section 7.

Acknowledgement

This project began during BJ’s visit to the School of Mathematical Sciences at Fudan University. BJ would like to express his gratitude to Prof. Xiaoman Chen and Yijun Yao for the invitation and coordination, and also to the faculty members for their hospitality and engaging discussions during his stay. We would also like to thank Rufus Willett for his comments after reading an early draft of this paper.

BJ was supported by NSFC12001066 and NSFC12071183. JZ was partly supported by National Key R&D Program of China 2022YFA100700.

2. Preliminaries

2.1. Standard notation

Here we collect the notation used throughout the paper.

For a set XX, denote |X||X| the cardinality of XX and 𝒫(X)\mathcal{P}(X) the set of all subsets in XX. For AXA\subseteq X, denote χA\chi_{A} the characteristic function of AA, and set δx:=χ{x}\delta_{x}:=\chi_{\{x\}} for xXx\in X.

For a discrete space XX, denote (X)\ell^{\infty}(X) the CC^{*}-algebra of bounded functions on XX with the supremum norm f:=supxX|f(x)|\|f\|_{\infty}:=\sup_{x\in X}|f(x)|. The support of f(X)f\in\ell^{\infty}(X) is defined to be {xX|f(x)0}\{x\in X~{}|~{}f(x)\neq 0\}, denoted by supp(f)\mathrm{supp}(f). Given a Hilbert space \mathcal{H}, denote 2(X;)\ell^{2}(X;\mathcal{H}) the Hilbert space of square-summable functions from XX to \mathcal{H}. When =\mathcal{H}=\mathbb{C}, we denote 2(X):=2(X;)\ell^{2}(X):=\ell^{2}(X;\mathbb{C}), which has an orthonormal basis {δx}xX\{\delta_{x}\}_{x\in X}.

Given a Hilbert space \mathcal{H}, denote 𝔅()\mathfrak{B}(\mathcal{H}) the CC^{*}-algebra of all bounded linear operators on \mathcal{H}, and 𝔎()\mathfrak{K}(\mathcal{H}) the CC^{*}-subalgebra of all compact operators on \mathcal{H}.

2.2. Notions from coarse geometry

Here we collect necessary notions from coarse geometry, and guide readers to [13, 18] for more details.

First recall that for a set XX and E,FX×XE,F\subseteq X\times X, denote

E1\displaystyle E^{-1} :={(y,x)|(x,y)E},\displaystyle:=\{(y,x)~{}|~{}(x,y)\in E\},
EF\displaystyle E\circ F :={(x,z)|yX such that (x,y)E and (y,z)F}.\displaystyle:=\{(x,z)~{}|~{}\exists~{}y\in X\text{ such that }(x,y)\in E\text{ and }(y,z)\in F\}.

Also denote r,s:X×XX\mathrm{r},\mathrm{s}:X\times X\to X the projection onto the first and the second coordinate, respectively. Given EX×XE\subseteq X\times X and A,BXA,B\subseteq X, we say that AA and BB are EE-separated if (A×B)E=(A\times B)\cap E=\emptyset and (B×A)E=(B\times A)\cap E=\emptyset.

Definition 2.1.

A (connected) coarse structure on a set XX is a collection 𝒫(X×X)\mathscr{E}\subseteq\mathcal{P}(X\times X), called entourages, satisfying the following:

  1. (1)

    For any entourages AA and BB, then A1A^{-1}, ABA\circ B and ABA\cup B are entourages;

  2. (2)

    Every finite subset of X×XX\times X is an entourage;

  3. (3)

    Any subset of an entourage is an entourage.

In this case, (X,)(X,\mathscr{E}) is called a coarse space. If additionally the diagonal

ΔX:={(x,x)|xX}\Delta_{X}:=\{(x,x)~{}|~{}x\in X\}

is an entourage, then \mathscr{E} (also the pair (X,)(X,\mathscr{E})) is called unital.

For YXY\subseteq X and EE\in\mathscr{E}, denote the EE-neighbourhood 𝒩E(Y)\mathcal{N}_{E}(Y) of YY by

𝒩E(Y):=Y{xX|yY such that (x,y)E}\mathcal{N}_{E}(Y):=Y\cup\{x\in X~{}|~{}\exists~{}y\in Y\text{ such that }(x,y)\in E\}

and

n(E):=supxXmax{|𝒩E({x})|,|𝒩E1({x})|}.n(E):=\sup_{x\in X}\mathrm{max}\{~{}|\mathcal{N}_{E}(\{x\})|,|\mathcal{N}_{E^{-1}}(\{x\})|~{}\}.
Definition 2.2.

A coarse structure \mathscr{E} on a set XX is said to have bounded geometry (or to be uniformly locally finite) if n(E)n(E) is finite for any entourage EE\in\mathscr{E}. In this case, we also say that the coarse space (X,)(X,\mathscr{E}) has bounded geometry.

For a set XX and a collection 𝒮𝒫(X×X)\mathcal{S}\subseteq\mathcal{P}(X\times X), the smallest coarse structure on XX containing 𝒮\mathcal{S} is called the coarse structure generated by 𝒮\mathcal{S}.

When (X,d)(X,d) is a discrete metric space, there is an associated unital coarse structure d\mathscr{E}_{d} (called the bounded coarse structure) generated by all the RR-entourages defined as ER:={(x,y)X×X|d(x,y)R}E_{R}:=\{(x,y)\in X\times X~{}|~{}d(x,y)\leq R\} for all R0R\geq 0. In this case, we denote the closed ball by BX(x,r):=𝒩Er({x})B_{X}(x,r):=\mathcal{N}_{E_{r}}(\{x\}) for xXx\in X and r0r\geq 0, and 𝒩r(Y):=𝒩Er(Y)\mathcal{N}_{r}(Y):=\mathcal{N}_{E_{r}}(Y) for YXY\subseteq X and r0r\geq 0. We say that (X,d)(X,d) has bounded geometry if d\mathscr{E}_{d} has bounded geometry, i.e., the number supxX|BX(x,r)|\sup_{x\in X}|B_{X}(x,r)| is finite for any r0r\geq 0.

Definition 2.3.

Let X,YX,Y be sets, f:XYf:X\to Y be a map and Y\mathscr{E}_{Y} be a coarse structure on YY.

  1. (1)

    Denote fY:={EX×X|(f×f)(E)Y}f^{*}\mathscr{E}_{Y}:=\{E\subseteq X\times X~{}|~{}(f\times f)(E)\in\mathscr{E}_{Y}\}, which is a coarse structure on XX. Here (f×f)(E):={(f(x),f(y))|(x,y)E}(f\times f)(E):=\{(f(x),f(y))~{}|~{}(x,y)\in E\}.

  2. (2)

    If X\mathscr{E}_{X} is a coarse structure on XX, we say that ff is coarse if XfY\mathscr{E}_{X}\subseteq f^{*}\mathscr{E}_{Y}. To specify the underlying coarse structures, we also write f:(X,X)(Y,Y)f:(X,\mathscr{E}_{X})\to(Y,\mathscr{E}_{Y}).

  3. (3)

    Two coarse maps f,g:(X,X)(Y,Y)f,g:(X,\mathscr{E}_{X})\to(Y,\mathscr{E}_{Y}) are said to be close if for any EXE\in\mathscr{E}_{X}, we have (f×g)(E):={(f(x),g(y))|(x,y)E}Y(f\times g)(E):=\{(f(x),g(y))~{}|~{}(x,y)\in E\}\in\mathscr{E}_{Y}.

Definition 2.4.

Let f:(X,X)(Y,Y)f:(X,\mathscr{E}_{X})\to(Y,\mathscr{E}_{Y}) be a map between unital coarse spaces. We say that ff is a coarse equivalence if ff is coarse and there exists a coarse map g:(Y,Y)(X,X)g:(Y,\mathscr{E}_{Y})\to(X,\mathscr{E}_{X}) (called a coarse inverse to ff) such that fgf\circ g is close to IdY\mathrm{Id}_{Y} and gfg\circ f is close to IdX\mathrm{Id}_{X}. In this case, we say that (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) are coarsely equivalent.

When (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) come from metric spaces, Definition 2.4 coincides with the one recalled in Section 1. Note that Definition 2.4 also makes sense for the general (non-unital) case, however, it does not work well for the rigidity problems (see Proposition D and Theorem E). It turns out that the suitable setting for the notion of coarse equivalence in the general case is Definition C. To obtain an appropriate picture, we will focus on Definition C in Section 3 based on an alternative version of coarse maps from [19].

2.3. Uniform Roe algebras and geometric ideals

Let XX be a set. Each operator T𝔅(2(X))T\in\mathfrak{B}(\ell^{2}(X)) can be written in the matrix form T=(T(x,y))x,yXT=(T(x,y))_{x,y\in X}, where T(x,y)=Tδy,δxT(x,y)=\langle T\delta_{y},\delta_{x}\rangle\in\mathbb{C}. Denote by T\|T\| the operator norm of TT in 𝔅(2(X))\mathfrak{B}(\ell^{2}(X)). Similarly for an operator T𝔅(2(X;))T\in\mathfrak{B}(\ell^{2}(X;\mathcal{H})), we can also write T=(T(x,y))x,yXT=(T(x,y))_{x,y\in X} for T(x,y)𝔅()T(x,y)\in\mathfrak{B}(\mathcal{H}).

Given an operator T𝔅(2(X))T\in\mathfrak{B}(\ell^{2}(X)), we define the support of TT to be

supp(T):={(x,y)X×X|T(x,y)0}.\mathrm{supp}(T):=\{(x,y)\in X\times X~{}|~{}T(x,y)\neq 0\}.

Given ε>0\varepsilon>0, define the ε\varepsilon-support of TT to be

suppε(T):={(x,y)X×X||T(x,y)|ε}.\mathrm{supp}_{\varepsilon}(T):=\big{\{}(x,y)\in X\times X~{}\big{|}~{}|T(x,y)|\geq\varepsilon\big{\}}.

When XX is equipped with a metric dd, we define the propagation of TT to be

prop(T):=sup{d(x,y)|(x,y)supp(T)}.\mathrm{prop}(T):=\sup\{d(x,y)~{}|~{}(x,y)\in\mathrm{supp}(T)\}.
Definition 2.5.

Let (X,)(X,\mathscr{E}) be a coarse space of bounded geometry. The set of all operators in 𝔅(2(X))\mathfrak{B}(\ell^{2}(X)) whose supports belong to \mathscr{E} forms a \ast-algebra, called the algebraic uniform Roe algebra of (X,)(X,\mathscr{E}) and denoted by u[X,]\mathbb{C}_{u}[X,\mathscr{E}]. The uniform Roe algebra of (X,)(X,\mathscr{E}) is defined to be the operator norm closure of u[X,]\mathbb{C}_{u}[X,\mathscr{E}] in 𝔅(2(X))\mathfrak{B}(\ell^{2}(X)), which forms a CC^{*}-algebra and is denoted by Cu(X,)C^{*}_{u}(X,\mathscr{E}).

When (X,d)(X,d) is a discrete metric space of bounded geometry, we simply write u[X]:=u[X,d]\mathbb{C}_{u}[X]:=\mathbb{C}_{u}[X,\mathscr{E}_{d}] and Cu(X):=Cu(X,d)C^{*}_{u}(X):=C^{*}_{u}(X,\mathscr{E}_{d}).

For a coarse space (X,)(X,\mathscr{E}) of bounded geometry and x,yXx,y\in X, denote the rank-one operator ξξ,δyδx\xi\mapsto\langle\xi,\delta_{y}\rangle\delta_{x} by exye_{xy}. It is clear that exyu[X,]e_{xy}\in\mathbb{C}_{u}[X,\mathscr{E}]. This kind of operators will play an important role when we study the rigidity problems later.

In [6, 23], Chen and Wang introduced the notion of geometric ideals (see Definition A) in uniform Roe algebras and provided a full description for these ideals. Here we only focus on the case of metric spaces, which are the main objects we are interested in for the rigidity problem. (Note that geometric ideals for general coarse spaces were also studied in [6].) In the following, we always assume that (X,d)(X,d) is a discrete metric space of bounded geometry and d\mathscr{E}_{d} is the associated bounded coarse structure.

Definition 2.6 ([6, Definition 4.1]).

A coarse structure d\mathscr{I}\subseteq\mathscr{E}_{d} is called an ideal in d\mathscr{E}_{d} if for any EdE\in\mathscr{E}_{d} and AA\in\mathscr{I}, we have EAE\circ A\in\mathscr{I} and AEA\circ E\in\mathscr{I}.

Proposition 2.7 ([6, Proposition 4.2]).

Given an ideal \mathscr{I} in d\mathscr{E}_{d}, the uniform Roe algebra Cu(X,)C^{*}_{u}(X,\mathscr{I}) is a geometric ideal in Cu(X)C^{*}_{u}(X). Conversely, given a geometric ideal II in Cu(X)C^{*}_{u}(X), the collection (I):={suppε(T)|TI,ε>0}\mathscr{I}(I):=\{\mathrm{supp}_{\varepsilon}(T)~{}|~{}T\in I,\varepsilon>0\} is an ideal in d\mathscr{E}_{d}. Moreover, we have (Cu(X,))=\mathscr{I}(C^{*}_{u}(X,\mathscr{I}))=\mathscr{I} and Cu(X,(I))=IC^{*}_{u}(X,\mathscr{I}(I))=I.

We also need the following notion of ideals in space:

Definition 2.8 ([6, Definition 6.1]).

An ideal in (X,d)(X,d) is a collection 𝐋𝒫(X)\mathbf{L}\subseteq\mathcal{P}(X) satisfying the following:

  1. (1)

    If Y𝐋Y\in\mathbf{L} and ZYZ\subseteq Y, then Z𝐋Z\in\mathbf{L};

  2. (2)

    If Y𝐋Y\in\mathbf{L} and r>0r>0, then 𝒩r(Y)𝐋\mathcal{N}_{r}(Y)\in\mathbf{L};

  3. (3)

    If Y,Z𝐋Y,Z\in\mathbf{L}, then YZ𝐋Y\cup Z\in\mathbf{L}.

For 𝒮𝒫(X)\mathcal{S}\subseteq\mathcal{P}(X), we say that 𝐋\mathbf{L} is the ideal generated by 𝒮\mathcal{S} if 𝐋\mathbf{L} is the smallest ideal in (X,d)(X,d) containing 𝒮\mathcal{S}.

Proposition 2.9 ([6, Proposition 6.2]).

Given an ideal \mathscr{I} in d\mathscr{E}_{d}, the collection 𝐋():={r(E)|E}\mathbf{L}(\mathscr{I}):=\{\mathrm{r}(E)~{}|~{}E\in\mathscr{I}\} is an ideal in (X,d)(X,d). Conversely, given an ideal 𝐋\mathbf{L} in (X,d)(X,d), the collection (𝐋):={Ed|L𝐋 such that EL×L}\mathscr{I}(\mathbf{L}):=\{E\in\mathscr{E}_{d}~{}|~{}\exists~{}L\in\mathbf{L}\text{ such that }E\subseteq L\times L\} is an ideal in d\mathscr{E}_{d}. Moreover, we have (𝐋())=\mathscr{I}(\mathbf{L}(\mathscr{I}))=\mathscr{I} and 𝐋((𝐋))=𝐋\mathbf{L}(\mathscr{I}(\mathbf{L}))=\mathbf{L}.

Combining Proposition 2.7 and Proposition 2.9, we know that the lattice of geometric ideals in Cu(X)C^{*}_{u}(X) is isomorphic to the lattice of ideals in d\mathscr{E}_{d}, which is also isomorphic to the lattice of ideals in (X,d)(X,d). Hence we will drift among these three objects freely in the sequel.

Example 2.10.

As a special case, we consider the ideal 𝔎(2(X))\mathfrak{K}(\ell^{2}(X)) of compact operators in the uniform Roe algebra Cu(X)C^{*}_{u}(X) for a discrete metric space (X,d)(X,d). Note that 𝔎(2(X))\mathfrak{K}(\ell^{2}(X)) is a geometric ideal, and it is easy to see that (𝔎(2(X)))\mathscr{I}(\mathfrak{K}(\ell^{2}(X))) consists of all finite subsets of X×XX\times X, and 𝐋(𝔎(2(X)))\mathbf{L}(\mathfrak{K}(\ell^{2}(X))) consists of all finite subsets of XX.

Remark 2.11.

We remark that the definition (I)\mathscr{I}(I) works for any (not necessarily geometric) ideal (see [6]) in the uniform Roe algebra Cu(X)C^{*}_{u}(X). However, the one-to-one correspondence in Proposition 2.7 only works for geometric ideals. As an easy example, one can consider the ideal IGI_{G} consisting of all ghost operators in Cu(X)C^{*}_{u}(X). (Recall that an operator T𝔅(2(X))T\in\mathfrak{B}(\ell^{2}(X)) is a ghost operator if for any ε>0\varepsilon>0, there exists a finite subset FXF\subseteq X such that for any (x,y)F×F(x,y)\notin F\times F, then |T(x,y)|<ε|T(x,y)|<\varepsilon.) One can calculate directly that (IG)=(𝔎(2(X)))\mathscr{I}(I_{G})=\mathscr{I}(\mathfrak{K}(\ell^{2}(X))), which consists of all finite subsets in X×XX\times X.

Finally, we record an elementary observation which will be used later.

Lemma 2.12.

Let 𝐋\mathbf{L} be an ideal in (X,d)(X,d). Then the family {χL|L𝐋}\{\chi_{L}~{}|~{}L\in\mathbf{L}\} is an approximate unit for the ideal Cu(X,(𝐋))C^{*}_{u}(X,\mathscr{I}(\mathbf{L})).

Proof.

Note that Cu(X,(𝐋))C^{*}_{u}(X,\mathscr{I}(\mathbf{L})) is the direct limit of {Cu(L)|L𝐋}\{C^{*}_{u}(L)~{}|~{}L\in\mathbf{L}\} where Cu(L)C^{*}_{u}(L) is the uniform Roe algebra of the metric space LL with the induced metric of dd and 𝐋\mathbf{L} is a direct set under inclusion. Since χL\chi_{L} is the unit of Cu(L)C^{*}_{u}(L), we conclude the proof. ∎

3. Coarse correspondence

In this section, we study the notion of coarse correspondence (Definition C) from [19] and provide a detailed picture for coarse equivalences between general coarse spaces (Proposition D).

Let XX be a set and 𝒫(X×X)\mathscr{E}\subseteq\mathcal{P}(X\times X). For A,BXA,B\subseteq X, denote the restriction of \mathscr{E} on A×BA\times B by

|A×B:={E(A×B)|E}.\mathscr{E}|_{A\times B}:=\{E\cap(A\times B)~{}|~{}E\in\mathscr{E}\}.

Recall from Definition C that for coarse spaces (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}), a coarse correspondence from (X,X)(X,\mathscr{E}_{X}) to (Y,Y)(Y,\mathscr{E}_{Y}) (or simply from XX to YY) is a coarse structure \mathscr{E} on the disjoint union XYX\sqcup Y satisfying the following:

  1. (1)

    |Y×Y=Y\mathscr{E}|_{Y\times Y}=\mathscr{E}_{Y};

  2. (2)

    |X×XX\mathscr{E}|_{X\times X}\supseteq\mathscr{E}_{X};

  3. (3)

    \mathscr{E} is generated by |Y×(XY)\mathscr{E}|_{Y\times(X\sqcup Y)}.

The following fact was implicitly stated in the proof of [19, Proposition 2.3]. It follows directly from condition (3) above, and hence we omit the proof.

Lemma 3.1.

Let (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) be coarse spaces, and \mathscr{E} be a coarse correspondence from (X,X)(X,\mathscr{E}_{X}) to (Y,Y)(Y,\mathscr{E}_{Y}). Then for any E|X×XE\in\mathscr{E}|_{X\times X}, there exists F|Y×XF\in\mathscr{E}|_{Y\times X} such that EF1FE\subseteq F^{-1}\circ F.

We also record the following elementary fact, whose proof is straightforward.

Lemma 3.2.

For a set XX and EX×XE\subseteq X\times X, we have E(E)1EE\subseteq(E^{\prime})^{-1}\circ E^{\prime} for E:=E(EE1)E^{\prime}:=E\cup(E\circ E^{-1}).

3.1. The unital case

It was shown in [19] that for unital coarse spaces, the notion of coarse correspondence can be identified with coarse maps in Definition 2.3(2). Here we recall the outline of the proof, and divide it into several pieces to clarify the dependence on the assumption of unitalness. Some of the results will also be used in the general case.

Let XX be a set, (Y,Y)(Y,\mathscr{E}_{Y}) be a coarse space and f:XYf:X\to Y be a map. Denote

YX(f):={EY×X|(IdY×f)(E)Y},\mathscr{E}_{YX}(f):=\{E\subseteq Y\times X~{}|~{}(\mathrm{Id}_{Y}\times f)(E)\in\mathscr{E}_{Y}\},

and

(f):={EXEYEXYEYX|EXfY,EYY and EYXEXY1YX(f)}.\mathscr{E}(f):=\{E_{X}\cup E_{Y}\cup E_{XY}\cup E_{YX}~{}|~{}E_{X}\in f^{*}\mathscr{E}_{Y},E_{Y}\in\mathscr{E}_{Y}\text{ and }E_{YX}\cup E_{XY}^{-1}\in\mathscr{E}_{YX}(f)\}.

The following result is contained in the proof of [19, Proposition 2.3]. For the reader’s convenience, here we recall the proof.

Lemma 3.3.

Let XX be a set, (Y,Y)(Y,\mathscr{E}_{Y}) be a coarse space and f:XYf:X\to Y be a map. Then (f)\mathscr{E}(f) is a coarse structure on XYX\sqcup Y and generated by (f)|Y×(XY)=YYX(f)\mathscr{E}(f)|_{Y\times(X\sqcup Y)}=\mathscr{E}_{Y}\cup\mathscr{E}_{YX}(f).

Proof.

Consider the map h:XYYh:X\sqcup Y\to Y which restricts to ff on XX and IdY\mathrm{Id}_{Y} on YY. It is easy to see that (f)=hY\mathscr{E}(f)=h^{*}\mathscr{E}_{Y}, and hence (f)\mathscr{E}(f) is a coarse structure on XYX\sqcup Y. For the second statement, it suffices to show that (f)|X×X=fY\mathscr{E}(f)|_{X\times X}=f^{*}\mathscr{E}_{Y} can be generated by (f)|Y×X=YX(f)\mathscr{E}(f)|_{Y\times X}=\mathscr{E}_{YX}(f). Given EfYE\in f^{*}\mathscr{E}_{Y}, it follows from Lemma 3.2 that E(E)1EE\subseteq(E^{\prime})^{-1}\circ E^{\prime} for E:=E(EE1)E^{\prime}:=E\cup(E\circ E^{-1}). It is clear that EfYE^{\prime}\in f^{*}\mathscr{E}_{Y}, and hence F:=(f×IdX)(E)YX(f)F:=(f\times\mathrm{Id}_{X})(E^{\prime})\in\mathscr{E}_{YX}(f). Finally, note that E(E)1EF1FE\subseteq(E^{\prime})^{-1}\circ E^{\prime}\subseteq F^{-1}\circ F, which concludes the proof. ∎

Consequently, we have the following:

Corollary 3.4.

Let (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) be coarse spaces, and f:XYf:X\to Y be a map. Then ff is coarse if and only if (f)\mathscr{E}(f) is a coarse correspondence. In this case, we say that (f)\mathscr{E}(f) is the coarse correspondence associated to ff.

Recall that for a map f:XYf:X\to Y, the graph of ff is defined to be

Gr(f):={(f(x),x)Y×X|xX}.\mathrm{Gr}(f):=\{(f(x),x)\in Y\times X~{}|~{}x\in X\}.

We have the following:

Lemma 3.5.

Let XX be a set, (Y,Y)(Y,\mathscr{E}_{Y}) be a coarse space and f:XYf:X\to Y be a map. Assume that the set {(f(x),f(x))Y×Y|xX}\{(f(x),f(x))\in Y\times Y~{}|~{}x\in X\} belongs to Y\mathscr{E}_{Y}. Then (f)\mathscr{E}(f) is generated by Y\mathscr{E}_{Y} and Gr(f)\mathrm{Gr}(f).

Proof.

Since {(f(x),f(x))Y×Y|xX}Y\{(f(x),f(x))\in Y\times Y~{}|~{}x\in X\}\in\mathscr{E}_{Y}, we have Gr(f)YX(f)\mathrm{Gr}(f)\in\mathscr{E}_{YX}(f). On the other hand, given EYX(f)E\in\mathscr{E}_{YX}(f), we have E(EGr(f)1)Gr(f)E\subseteq(E\circ\mathrm{Gr}(f)^{-1})\circ\mathrm{Gr}(f). Since EGr(f)1=(IdY×f)(E)YE\circ\mathrm{Gr}(f)^{-1}=(\mathrm{Id}_{Y}\times f)(E)\in\mathscr{E}_{Y}, we conclude the proof thanks to Lemma 3.3. ∎

Now we present the following result from [19] that for unital coarse spaces, each coarse correspondence is determined by a coarse map. For the reader’s convenience, here we recall the proof.

Proposition 3.6 ([19, Proposition 2.3]).

Let (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) be coarse spaces, and \mathscr{E} be a coarse correspondence. Assume that X\mathscr{E}_{X} is unital. Then there exists a unique (up to closeness) coarse map f:XYf:X\to Y such that =(f)\mathscr{E}=\mathscr{E}(f).

Proof.

By Lemma 3.1 and the assumption that X\mathscr{E}_{X} is unital, there exists E|Y×XE\in\mathscr{E}|_{Y\times X} such that ΔXE1E\Delta_{X}\subseteq E^{-1}\circ E. It is easy to see that EE contains Gr(f)\mathrm{Gr}(f) for some map f:XYf:X\to Y. Note that {(f(x),f(x))Y×Y|xX}Gr(f)Gr(f)1EE1Y\{(f(x),f(x))\in Y\times Y~{}|~{}x\in X\}\subseteq\mathrm{Gr}(f)\circ\mathrm{Gr}(f)^{-1}\subseteq E\circ E^{-1}\in\mathscr{E}_{Y}. Hence it follows from Lemma 3.5 that (f)\mathscr{E}(f)\subseteq\mathscr{E}.

On the other hand, given E|Y×XE\in\mathscr{E}|_{Y\times X}, we have E(EGr(f)1)Gr(f)E\subseteq(E\circ\mathrm{Gr}(f)^{-1})\circ\mathrm{Gr}(f). Since EGr(f)1|Y×Y=YE\circ\mathrm{Gr}(f)^{-1}\in\mathscr{E}|_{Y\times Y}=\mathscr{E}_{Y}, then we obtain E(f)E\in\mathscr{E}(f). Hence =(f)\mathscr{E}=\mathscr{E}(f), which implies that ff is coarse thanks to Corollary 3.4.

Finally, if (f)=(g)\mathscr{E}(f)=\mathscr{E}(g) for coarse maps f,g:XYf,g:X\to Y, then Gr(f)Gr(g)1Y\mathrm{Gr}(f)\circ\mathrm{Gr}(g)^{-1}\in\mathscr{E}_{Y}. Hence ff and gg are close. ∎

The following result shows that the notion of coarse equivalence from Definition C coincides with the one from Definition 2.4 for unital coarse spaces.

Corollary 3.7.

Let (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) be unital coarse spaces.

  1. (1)

    Let f:XYf:X\to Y be a coarse equivalence with a coarse inverse g:YXg:Y\to X. Then we have (f)=(g)\mathscr{E}(f)=\mathscr{E}(g), which is a coarse correspondence both from XX to YY and from YY to XX. In this case, we have (f)|X×X=X\mathscr{E}(f)|_{X\times X}=\mathscr{E}_{X}.

  2. (2)

    Let \mathscr{E} be a coarse correspondence from XX to YY and YY to XX. Let f:XYf:X\to Y be a coarse map from Proposition 3.6 such that =(f)\mathscr{E}=\mathscr{E}(f), and similarly g:YXg:Y\to X be a coarse map such that =(g)\mathscr{E}=\mathscr{E}(g). Then ff is a coarse equivalence with a coarse inverse gg.

Proof.

(1) is straightforward, and hence we omit the proof. For (2), it suffices to note that {(fg(y),y)|yY}Gr(f)Gr(g)|Y×Y=Y\{(f\circ g(y),y)~{}|~{}y\in Y\}\subseteq\mathrm{Gr}(f)\circ\mathrm{Gr}(g)\in\mathscr{E}|_{Y\times Y}=\mathscr{E}_{Y} and similarly, {(gf(x),x)|xX}=Gr(g)Gr(f)|X×X=X\{(g\circ f(x),x)~{}|~{}x\in X\}=\mathrm{Gr}(g)\circ\mathrm{Gr}(f)\in\mathscr{E}|_{X\times X}=\mathscr{E}_{X}. Hence we obtain the result. ∎

3.2. The general case

Now we move to the general case. Firstly, let us introduce the following notion:

Definition 3.8.

Let XX be a set and 𝐋𝒫(X)\mathbf{L}\subseteq\mathcal{P}(X). We say that 𝐋\mathbf{L} is admissible if the following holds:

  1. (1)

    Every finite subset of XX belongs to 𝐋\mathbf{L};

  2. (2)

    For any Y𝐋Y\in\mathbf{L} and ZYZ\subseteq Y, then Z𝐋Z\in\mathbf{L};

  3. (3)

    For any Y,Z𝐋Y,Z\in\mathbf{L}, then YZ𝐋Y\cup Z\in\mathbf{L}.

The example of admissible collection we are interested in comes from coarse structures:

Definition 3.9.

Let (X,X)(X,\mathscr{E}_{X}) be a coarse space. The collection

𝐋(X):={r(E)|EX}={s(E)|EX}\mathbf{L}(\mathscr{E}_{X}):=\{\mathrm{r}(E)~{}|~{}E\in\mathscr{E}_{X}\}=\{\mathrm{s}(E)~{}|~{}E\in\mathscr{E}_{X}\}

is called associated to (X,X)(X,\mathscr{E}_{X}), also denoted by 𝐋X\mathbf{L}_{X}, if the coarse structure is clear from the context.

Clearly 𝐋(X)\mathbf{L}(\mathscr{E}_{X}) associated to a coarse space (X,X)(X,\mathscr{E}_{X}) is always admissible. When (X,d)(X,d) is a discrete metric space and d\mathscr{I}\subseteq\mathscr{E}_{d} is an ideal (in the sense of Definition 2.6), the definition for 𝐋()\mathbf{L}(\mathscr{I}) above coincides with the one in Proposition 2.9. In particular, any ideal 𝐋\mathbf{L} in (X,d)(X,d) (in the sense of Definition 2.8) is admissible.

The following is straightforward, and hence we omit the proof.

Lemma 3.10.

Let (X,X)(X,\mathscr{E}_{X}) be a coarse space, and 𝐋X\mathbf{L}_{X} be the associated collection. For LXL\subseteq X, we have that L𝐋XL\in\mathbf{L}_{X} if and only if the set ΔL={(x,x)|xL}\Delta_{L}=\{(x,x)~{}|~{}x\in L\} belongs to X\mathscr{E}_{X}.

To study the non-unital case, we need to consider a family of maps instead of a single map:

Definition 3.11.

Let XX be a set, 𝐋\mathbf{L} be an admissible collection in 𝒫(X)\mathcal{P}(X) and (Y,Y)(Y,\mathscr{E}_{Y}) be a coarse space.

  1. (1)

    We say that a family of maps ={fL:LY|L𝐋}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}\} is admissible if for any L1,L2𝐋L_{1},L_{2}\in\mathbf{L}, the set {(fL1(x),fL2(x))|xL1L2}\{(f_{L_{1}}(x),f_{L_{2}}(x))~{}|~{}x\in L_{1}\cap L_{2}\} belongs to Y\mathscr{E}_{Y}.

  2. (2)

    For an admissible family ={fL:LY|L𝐋}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}\}, denote

    Y:={EX×X|L𝐋 such that EL×L and (fL×fL)(E)Y}.\mathcal{F}^{*}\mathscr{E}_{Y}:=\{E\in X\times X~{}|~{}\exists~{}L\in\mathbf{L}\text{ such that }E\subseteq L\times L\text{ and }(f_{L}\times f_{L})(E)\in\mathscr{E}_{Y}\}.

We also need to consider the relation of closeness between families:

Definition 3.12.

Let XX be a set, 𝐋\mathbf{L} be an admissible collection in 𝒫(X)\mathcal{P}(X) and (Y,Y)(Y,\mathscr{E}_{Y}) be a coarse space. Two admissible families ={fL:LY|L𝐋}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}\} and 𝒢={gL:LY|L𝐋}\mathcal{G}=\{g_{L}:L\to Y~{}|~{}L\in\mathbf{L}\} are called close if for any L𝐋L\in\mathbf{L}, the set {(fL(x),gL(x))|xL}\{(f_{L}(x),g_{L}(x))~{}|~{}x\in L\} belongs to Y\mathscr{E}_{Y}.

Lemma 3.13.

Let XX be a set and 𝐋\mathbf{L} be an admissible collection in 𝒫(X)\mathcal{P}(X). Let (Y,Y)(Y,\mathscr{E}_{Y}) be a coarse space, and ={fL:LY|L𝐋}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}\} be an admissible family. Then

  1. (1)

    For any L𝐋L\in\mathbf{L}, we have fL(L)𝐋Yf_{L}(L)\in\mathbf{L}_{Y}.

  2. (2)

    Y\mathcal{F}^{*}\mathscr{E}_{Y} is a coarse structure on XX.

Proof.

(1). Since 𝐋\mathbf{L} is admissible, the set {(fL(x),fL(x))|xL}Y\{(f_{L}(x),f_{L}(x))~{}|~{}x\in L\}\in\mathscr{E}_{Y} for any L𝐋L\in\mathbf{L}. Hence we obtain fL(L)𝐋Yf_{L}(L)\in\mathbf{L}_{Y}.

(2). Given E1,E2YE_{1},E_{2}\in\mathcal{F}^{*}\mathscr{E}_{Y}, assume that there exist L1,L2𝐋L_{1},L_{2}\in\mathbf{L} such that EiLi×LiE_{i}\subseteq L_{i}\times L_{i} and (fLi×fLi)(Ei)Y(f_{L_{i}}\times f_{L_{i}})(E_{i})\in\mathscr{E}_{Y} for i=1,2i=1,2. Take L=L1L2𝐋L=L_{1}\cup L_{2}\in\mathbf{L}, then we have E1E2L×LE_{1}\cup E_{2}\subseteq L\times L and E1E2L×LE_{1}\circ E_{2}\subseteq L\times L. Note that Fi:={(fL(x),fLi(x))|xLi}YF_{i}:=\{(f_{L}(x),f_{L_{i}}(x))~{}|~{}x\in L_{i}\}\in\mathscr{E}_{Y} for i=1,2i=1,2. Hence we have

(fL×fL)(Ei)Fi(fLi×fLi)(Ei)Fi1Y(f_{L}\times f_{L})(E_{i})\subseteq F_{i}\circ(f_{L_{i}}\times f_{L_{i}})(E_{i})\circ F_{i}^{-1}\in\mathscr{E}_{Y}

for i=1,2i=1,2, which implies that (fL×fL)(E1E2)Y(f_{L}\times f_{L})(E_{1}\cup E_{2})\in\mathscr{E}_{Y}. Also note that

(fL×fL)(E1E2)(fL×fL)(E1)(fL×fL)(E2)Y.(f_{L}\times f_{L})(E_{1}\circ E_{2})\subseteq(f_{L}\times f_{L})(E_{1})\circ(f_{L}\times f_{L})(E_{2})\in\mathscr{E}_{Y}.

The rest is trivial, and hence we omit the details. ∎

Now we introduce the notion of coarse family, which is the replacement of coarse maps in the general case.

Definition 3.14.

Let (X,X),(Y,Y)(X,\mathscr{E}_{X}),(Y,\mathscr{E}_{Y}) be coarse spaces and 𝐋\mathbf{L} be an admissible collection in 𝒫(X)\mathcal{P}(X). A family of admissible maps ={fL:LY|L𝐋}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}\} is called coarse if XY\mathscr{E}_{X}\subseteq\mathcal{F}^{*}\mathscr{E}_{Y}.

It is clear from the definition that for a coarse family ={fL:LY|L𝐋}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}\}, we have 𝐋X𝐋\mathbf{L}_{X}\subseteq\mathbf{L}. Note that we do not require 𝐋X=𝐋\mathbf{L}_{X}=\mathbf{L} in general.

Definition 3.15.

Let XX be a set and 𝐋\mathbf{L} be an admissible collection in 𝒫(X)\mathcal{P}(X). Let (Y,Y)(Y,\mathscr{E}_{Y}) be a coarse space and ={fL:LY|L𝐋}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}\} be an admissible family. Denote

YX():={EY×X|L𝐋 such that EY×L and (IdY×fL)(E)Y},\mathscr{E}_{YX}(\mathcal{F}):=\{E\subseteq Y\times X~{}|~{}\exists~{}L\in\mathbf{L}\text{ such that }E\subseteq Y\times L\text{ and }(\mathrm{Id}_{Y}\times f_{L})(E)\in\mathscr{E}_{Y}\},

and

():={EXEYEXYEYX|EXY,EYY and EYXEXY1YX()}.\mathscr{E}(\mathcal{F}):=\{E_{X}\cup E_{Y}\cup E_{XY}\cup E_{YX}~{}|~{}E_{X}\in\mathcal{F}^{*}\mathscr{E}_{Y},E_{Y}\in\mathscr{E}_{Y}\text{ and }E_{YX}\cup E_{XY}^{-1}\in\mathscr{E}_{YX}(\mathcal{F})\}.

We record the following elementary observation, whose proof is straightforward and hence omitted.

Lemma 3.16.

Given X,𝐋,(Y,Y)X,\mathbf{L},(Y,\mathscr{E}_{Y}) and \mathcal{F} as in Definition 3.15, we have the following:

  1. (1)

    For any EYX()E\in\mathscr{E}_{YX}(\mathcal{F}), there exist L𝐋L\in\mathbf{L} and B𝐋YB\in\mathbf{L}_{Y} such that EB×LE\subseteq B\times L.

  2. (2)

    For L1,L2𝐋L_{1},L_{2}\in\mathbf{L} and EY×XE\subseteq Y\times X, if EY×LiE\subseteq Y\times L_{i} for i=1,2i=1,2 and (IdY×fL1)(E)Y(\mathrm{Id}_{Y}\times f_{L_{1}})(E)\in\mathscr{E}_{Y}, then (IdY×fL2)(E)Y(\mathrm{Id}_{Y}\times f_{L_{2}})(E)\in\mathscr{E}_{Y}.

The following result is analogous to Lemma 3.3.

Lemma 3.17.

Given X,𝐋,(Y,Y)X,\mathbf{L},(Y,\mathscr{E}_{Y}) and \mathcal{F} as in Definition 3.15, then ()\mathscr{E}(\mathcal{F}) is a coarse structure on XYX\sqcup Y generated by ()|Y×(XY)=YYX()\mathscr{E}(\mathcal{F})|_{Y\times(X\sqcup Y)}=\mathscr{E}_{Y}\cup\mathscr{E}_{YX}(\mathcal{F}).

Proof.

Firstly, we show that ()\mathscr{E}(\mathcal{F}) is a coarse structure on XYX\sqcup Y. Consider the admissible collection 𝐋:={AB|A𝐋 and B𝐋Y}\mathbf{L}^{\prime}:=\{A\sqcup B~{}|~{}A\in\mathbf{L}\text{ and }B\in\mathbf{L}_{Y}\} in 𝒫(XY)\mathcal{P}(X\sqcup Y). For any L=AB𝐋L=A\sqcup B\in\mathbf{L}^{\prime} with A𝐋A\in\mathbf{L} and B𝐋YB\in\mathbf{L}_{Y}, define a map hL:LYh_{L}:L\to Y which restricts to fAf_{A} on AA and IdB\mathrm{Id}_{B} on BB. Given Li=AiBiL_{i}=A_{i}\sqcup B_{i} with Ai𝐋A_{i}\in\mathbf{L} and Bi𝐋YB_{i}\in\mathbf{L}_{Y} for i=1,2i=1,2, we have

{(hL1(z),hL2(z))|zL1L2}={(fA1(x),fA2(x))|xA1A2}ΔB1B2,\{(h_{L_{1}}(z),h_{L_{2}}(z))~{}|~{}z\in L_{1}\cap L_{2}\}=\{(f_{A_{1}}(x),f_{A_{2}}(x))~{}|~{}x\in A_{1}\cap A_{2}\}\cup\Delta_{B_{1}\cap B_{2}},

which belongs to Y\mathscr{E}_{Y} due to Lemma 3.10. Hence :={hL|L𝐋}\mathcal{H}:=\{h_{L}~{}|~{}L\in\mathbf{L}^{\prime}\} is admissible.

Moreover, E=EXEYEXYEYXE=E_{X}\cup E_{Y}\cup E_{XY}\cup E_{YX} belongs to Y\mathcal{H}^{*}\mathscr{E}_{Y} if and only if there exists L=AB𝐋L=A\sqcup B\in\mathbf{L}^{\prime} with A𝐋A\in\mathbf{L} and B𝐋YB\in\mathbf{L}_{Y} such that EL×LE\subseteq L\times L and (hL×hL)(E)Y(h_{L}\times h_{L})(E)\in\mathscr{E}_{Y}, which is equivalent to that E()E\in\mathscr{E}(\mathcal{F}) thanks to Lemma 3.16. Hence we conclude that ()=Y\mathscr{E}(\mathcal{F})=\mathcal{H}^{*}\mathscr{E}_{Y}, which is a coarse structure on XYX\sqcup Y due to Lemma 3.13(2).

The rest of the proof is similar to that for Lemma 3.3, and hence we omit the details. ∎

Consequently, we obtain the following:

Corollary 3.18.

Let (X,X),(Y,Y)(X,\mathscr{E}_{X}),(Y,\mathscr{E}_{Y}) be coarse spaces, 𝐋\mathbf{L} be an admissible collection in 𝒫(X)\mathcal{P}(X) and ={fL:LY|L𝐋}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}\} be a family of admissible maps. Then \mathcal{F} is coarse if and only if ()\mathscr{E}(\mathcal{F}) is a coarse correspondence. In this case, we say that ()\mathscr{E}(\mathcal{F}) is the coarse correspondence associated to \mathcal{F}.

Similar to Lemma 3.5, we have the following:

Lemma 3.19.

Given X,𝐋,(Y,Y)X,\mathbf{L},(Y,\mathscr{E}_{Y}) and \mathcal{F} as in Definition 3.15, then ()\mathscr{E}(\mathcal{F}) is generated by Y\mathscr{E}_{Y} and {Gr(fL)|L𝐋}\{\mathrm{Gr}(f_{L})~{}|~{}L\in\mathbf{L}\}.

Proof.

Note that for any L𝐋L\in\mathbf{L}, we have {(fL(x),fL(x))|xL}Y\{(f_{L}(x),f_{L}(x))~{}|~{}x\in L\}\in\mathscr{E}_{Y} and hence Gr(fL)YX()\mathrm{Gr}(f_{L})\in\mathscr{E}_{YX}(\mathcal{F}). Now given EYX()E\in\mathscr{E}_{YX}(\mathcal{F}), there exists L𝐋L\in\mathbf{L} such that EY×LE\subseteq Y\times L and (IdY×fL)(E)Y(\mathrm{Id}_{Y}\times f_{L})(E)\in\mathscr{E}_{Y}. Hence we have E(EGr(fL)1)Gr(fL)E\subseteq(E\circ\mathrm{Gr}(f_{L})^{-1})\circ\mathrm{Gr}(f_{L}), where EGr(fL)1YE\circ\mathrm{Gr}(f_{L})^{-1}\in\mathscr{E}_{Y}. Therefore we conclude the proof thanks to Lemma 3.17. ∎

Now we are in the position to provide a detailed picture for coarse correspondence in the general case. This is a crucial step to achieve Proposition D.

Proposition 3.20.

Let (X,X),(Y,Y)(X,\mathscr{E}_{X}),(Y,\mathscr{E}_{Y}) be coarse spaces, and \mathscr{E} be a coarse correspondence from XX to YY. Then there exist an admissible collection 𝐋𝒫(X)\mathbf{L}\subseteq\mathcal{P}(X) and an admissible family of maps ={fL:LY|L𝐋}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}\} such that =()\mathscr{E}=\mathscr{E}(\mathcal{F}). Moreover, such a family \mathcal{F} is unique up to closeness.

Proof.

Set 𝐋:={r(E)|E|X×X}\mathbf{L}:=\{\mathrm{r}(E)~{}|~{}E\in\mathscr{E}|_{X\times X}\}, which is clearly admissible. For any L𝐋L\in\mathbf{L}, Lemma 3.10 shows that ΔLY\Delta_{L}\in\mathscr{E}_{Y}. By Lemma 3.1, there exists EL|Y×XE_{L}\in\mathscr{E}|_{Y\times X} such that ΔLEL1EL\Delta_{L}\subseteq E_{L}^{-1}\circ E_{L}. Hence there exists a map fL:LYf_{L}:L\to Y such that ELGr(fL)E_{L}\supseteq\mathrm{Gr}(f_{L}). For L1,L2𝐋L_{1},L_{2}\in\mathbf{L}, we have

{(fL1(x),fL2(x))|xL1L2}Gr(fL1)Gr(fL2)1EL1EL21Y,\{(f_{L_{1}}(x),f_{L_{2}}(x))~{}|~{}x\in L_{1}\cap L_{2}\}\subseteq\mathrm{Gr}(f_{L_{1}})\circ\mathrm{Gr}(f_{L_{2}})^{-1}\subseteq E_{L_{1}}\circ E_{L_{2}}^{-1}\in\mathscr{E}_{Y},

which implies that the family ={fL:LY|L𝐋}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}\} is admissible. Applying Lemma 3.19, we obtain that ()\mathscr{E}(\mathcal{F})\subseteq\mathscr{E}.

Conversely, we claim that 𝐋={s(F)|F|Y×X}\mathbf{L}=\{\mathrm{s}(F)~{}|~{}F\in\mathscr{E}|_{Y\times X}\}. In fact, for any F|Y×XF\in\mathscr{E}|_{Y\times X}, note that F1F|X×XF^{-1}\circ F\in\mathscr{E}|_{X\times X} and r(F1F)=s(F)𝐋\mathrm{r}(F^{-1}\circ F)=\mathrm{s}(F)\in\mathbf{L}. For any E|X×XE\in\mathscr{E}|_{X\times X}, Lemma 3.1 implies that there exists F|Y×XF\in\mathscr{E}|_{Y\times X} such that EF1FE\subseteq F^{-1}\circ F, and hence r(E)r(F1)=s(F)\mathrm{r}(E)\subseteq\mathrm{r}(F^{-1})=\mathrm{s}(F). This concludes the claim. Now given E|Y×XE\in\mathscr{E}|_{Y\times X}, the claim implies that there exists L𝐋L\in\mathbf{L} such that EY×LE\subseteq Y\times L. Then E(EGr(fL)1)Gr(fL)()E\subseteq(E\circ\mathrm{Gr}(f_{L})^{-1})\circ\mathrm{Gr}(f_{L})\in\mathscr{E}(\mathcal{F}). Hence we obtain that =()\mathscr{E}=\mathscr{E}(\mathcal{F}).

The last statement is easy to prove, and hence we omit the details. ∎

Finally, we consider the notion of coarse equivalence from Definition C. Recall that a coarse equivalence between two coarse spaces (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) is a coarse correspondence \mathscr{E} from both XX to YY and YY to XX. We say that (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) are coarsely equivalent if there exists a coarse equivalence between XX and YY.

Now we would like to apply Proposition 3.20 to unpack Definition C using families of maps, and our main target is to prove Proposition D. To make it more clear, we divide into two parts.

Corollary 3.21.

Let (X,X),(Y,Y)(X,\mathscr{E}_{X}),(Y,\mathscr{E}_{Y}) be coarse spaces, and \mathscr{E} be a coarse equivalence between them. Then there exist coarse families ={fL:LY|L𝐋X}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}_{X}\} and 𝒢={gL:LX|L𝐋Y}\mathcal{G}=\{g_{L^{\prime}}:L^{\prime}\to X~{}|~{}L^{\prime}\in\mathbf{L}_{Y}\} such that =()=(𝒢)\mathscr{E}=\mathscr{E}(\mathcal{F})=\mathscr{E}(\mathcal{G}) and satisfy the following:

(3.1) {(gfL(L)fL(x),x)|xL}Xfor anyL𝐋X,\{(g_{f_{L}(L)}\circ f_{L}(x),x)~{}|~{}x\in L\}\in\mathscr{E}_{X}\quad\text{for any}\quad L\in\mathbf{L}_{X},

and

(3.2) {(fgL(L)gL(y),y)|yL}Yfor anyL𝐋Y.\{(f_{g_{L^{\prime}}(L^{\prime})}\circ g_{L^{\prime}}(y),y)~{}|~{}y\in L^{\prime}\}\in\mathscr{E}_{Y}\quad\text{for any}\quad L^{\prime}\in\mathbf{L}_{Y}.

In this case, we have Y=X\mathcal{F}^{*}\mathscr{E}_{Y}=\mathscr{E}_{X} and 𝒢X=Y\mathcal{G}^{*}\mathscr{E}_{X}=\mathscr{E}_{Y}.

The proof is similar to that for Corollary 3.7 using Proposition 3.20 instead, and hence we omit the proof.

Corollary 3.22.

Let (X,X),(Y,Y)(X,\mathscr{E}_{X}),(Y,\mathscr{E}_{Y}) be coarse spaces, and ={fL:LY|L𝐋X}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}_{X}\}, 𝒢={gL:LX|L𝐋Y}\mathcal{G}=\{g_{L^{\prime}}:L^{\prime}\to X~{}|~{}L^{\prime}\in\mathbf{L}_{Y}\} be coarse families satisfying (3.1) and (3.2). Then Y=X,𝒢X=Y\mathcal{F}^{*}\mathscr{E}_{Y}=\mathscr{E}_{X},\mathcal{G}^{*}\mathscr{E}_{X}=\mathscr{E}_{Y} and ()=(𝒢)\mathscr{E}(\mathcal{F})=\mathscr{E}(\mathcal{G}) is a coarse equivalence between XX and YY.

Proof.

Firstly, we show that Y=X\mathcal{F}^{*}\mathscr{E}_{Y}=\mathscr{E}_{X}. Since \mathcal{F} is a coarse family, we have XY\mathscr{E}_{X}\subseteq\mathcal{F}^{*}\mathscr{E}_{Y}. Conversely given EYE\in\mathcal{F}^{*}\mathscr{E}_{Y}, there exists L𝐋XL\in\mathbf{L}_{X} such that EL×LE\subseteq L\times L and (fL×fL)(E)Y(f_{L}\times f_{L})(E)\in\mathscr{E}_{Y}. Take L:=fL(L)L^{\prime}:=f_{L}(L), which belongs to 𝐋Y\mathbf{L}_{Y} by Lemma 3.13. Then 𝒢\mathcal{G} being coarse implies:

F:={(gLfL(x),gLfL(y))|(x,y)E}=(gL×gL)(fL×fL)(E)X.F:=\{(g_{L^{\prime}}\circ f_{L}(x),g_{L^{\prime}}\circ f_{L}(y))~{}|~{}(x,y)\in E\}=(g_{L^{\prime}}\times g_{L^{\prime}})\circ(f_{L}\times f_{L})(E)\in\mathscr{E}_{X}.

Hence

E{(gLfL(x),x)|xL}1F{(gLfL(y),y)|yL}X,E\subseteq\{(g_{L^{\prime}}\circ f_{L}(x),x)~{}|~{}x\in L\}^{-1}\circ F\circ\{(g_{L^{\prime}}\circ f_{L}(y),y)~{}|~{}y\in L\}\in\mathscr{E}_{X},

which concludes that Y=X\mathcal{F}^{*}\mathscr{E}_{Y}=\mathscr{E}_{X}. Similarly, we have 𝒢X=Y\mathcal{G}^{*}\mathscr{E}_{X}=\mathscr{E}_{Y}.

To see ()=(𝒢)\mathscr{E}(\mathcal{F})=\mathscr{E}(\mathcal{G}), it suffices to show that YX()=XY(𝒢)1\mathscr{E}_{YX}(\mathcal{F})=\mathscr{E}_{XY}(\mathcal{G})^{-1}. For EYX()E\in\mathscr{E}_{YX}(\mathcal{F}), Lemma 3.16(1) shows that there exist BLYB\in L_{Y} and L𝐋XL\in\mathbf{L}_{X} such that EL×BE\subseteq L\times B and (IdY×fL)(E)Y(\mathrm{Id}_{Y}\times f_{L})(E)\in\mathscr{E}_{Y}. Taking L:=BfL(L)𝐋YL^{\prime}:=B\cup f_{L}(L)\in\mathbf{L}_{Y}, we have

F:={(gL(y),gLfL(x))|(y,x)E}=(gL×gL)(IdY×fL)(E)X.F^{\prime}:=\{(g_{L^{\prime}}(y),g_{L^{\prime}}f_{L}(x))~{}|~{}(y,x)\in E\}=(g_{L^{\prime}}\times g_{L^{\prime}})\circ(\mathrm{Id}_{Y}\times f_{L})(E)\in\mathscr{E}_{X}.

Note that

(gB×IdX)(E)F1FF2,(g_{B}\times\mathrm{Id}_{X})(E)\subseteq F_{1}\circ F^{\prime}\circ F_{2},

where F1:={(gB(y),gL(y))|yB}F_{1}:=\{(g_{B}(y),g_{L^{\prime}}(y))~{}|~{}y\in B\} and F2:={(gLfL(x),x)|xL}F_{2}:=\{(g_{L^{\prime}}f_{L}(x),x)~{}|~{}x\in L\}. Since 𝒢\mathcal{G} is admissible, then F1XF_{1}\in\mathscr{E}_{X}. Also (3.2) implies that {(gf(L)fL(x),x)|xL}X\{(g_{f(L)}f_{L}(x),x)~{}|~{}x\in L\}\in\mathscr{E}_{X}. By Lemma 3.16, we obtain F2XF_{2}\in\mathscr{E}_{X}. Hence (gB×IdX)(E)X(g_{B}\times\mathrm{Id}_{X})(E)\in\mathscr{E}_{X}, which means that EXY(𝒢)1E\in\mathscr{E}_{XY}(\mathcal{G})^{-1}. Similarly, we have XY(𝒢)YX()1\mathscr{E}_{XY}(\mathcal{G})\subseteq\mathscr{E}_{YX}(\mathcal{F})^{-1}, which concludes the proof. ∎

Motivated by Corollary 3.21 and Corollary 3.22, we introduce the following:

Definition 3.23.

Let (X,X),(Y,Y)(X,\mathscr{E}_{X}),(Y,\mathscr{E}_{Y}) be coarse spaces, and 𝐋X,𝐋Y\mathbf{L}_{X},\mathbf{L}_{Y} be the associated admissible families. A coarse family ={fL:LY|L𝐋X}\mathcal{F}=\{f_{L}:L\to Y~{}|~{}L\in\mathbf{L}_{X}\} is called a coarse equivalence if there exists a coarse family 𝒢={gL:LX|L𝐋Y}\mathcal{G}=\{g_{L^{\prime}}:L^{\prime}\to X~{}|~{}L^{\prime}\in\mathbf{L}_{Y}\} (which is called a coarse inverse to \mathcal{F}) satisfying (3.1) and (3.2).

Consequently, Corollary 3.21 and Corollary 3.22 can be rewritten in the following form, which concludes Proposition D.

Corollary 3.24.

Let (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) be coarse spaces. Then there exists a coarse equivalence \mathscr{E} between XX and YY in the sense of Definition C if and only if there exists a coarse equivalence \mathcal{F} in the sense of Definition 3.23.

Finally, we discuss the condition of bounded geometry, which will be used in the next section.

Lemma 3.25.

Let (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) be coarse spaces of bounded geometry, and \mathscr{E} be a coarse equivalence between them. Then \mathscr{E} also has bounded geometry as a coarse structure on XYX\sqcup Y.

Proof.

For FX×YF\subseteq X\times Y, xXx\in X and yYy\in Y, denote

Fx:={yY|(x,y)F}andFy:={xX|(x,y)F}.F_{x}:=\{y\in Y~{}|~{}(x,y)\in F\}\quad\text{and}\quad F^{y}:=\{x\in X~{}|~{}(x,y)\in F\}.

It suffices to show that for F|X×YF\in\mathscr{E}|_{X\times Y}, the number supxX|Fx|\sup_{x\in X}|F_{x}| and supyY|Fy|\sup_{y\in Y}|F^{y}| are finite. Note that

F1F=xX(Fx×Fx)andFF1=yY(Fy×Fy).F^{-1}\circ F=\bigcup_{x\in X}(F_{x}\times F_{x})\quad\text{and}\quad F\circ F^{-1}=\bigcup_{y\in Y}(F^{y}\times F^{y}).

Hence we have

supxX|Fx|n(F1F)andsupyY|Fy|n(FF1).\sup_{x\in X}|F_{x}|\leq n(F^{-1}\circ F)\quad\text{and}\quad\sup_{y\in Y}|F^{y}|\leq n(F\circ F^{-1}).

Note that F1FXF^{-1}\circ F\in\mathscr{E}_{X} and FF1YF\circ F^{-1}\in\mathscr{E}_{Y}. Hence we conclude the proof since X\mathscr{E}_{X} and Y\mathscr{E}_{Y} have bounded geometry. ∎

4. Coarse equivalences induce Morita equivalences

In this section, we recall the known result that a coarse equivalence between coarse spaces induces a Morita equivalence between the associated uniform Roe algebras. As a special case, we obtain the proof for Theorem E “(2) \Rightarrow (3)”. Here we provide a detailed proof since similar idea will be used later to treat ghostly ideals (see Section 7).

Firstly, let us recall the following notion of Morita equivalence for CC^{*}-algebras due to Rieffel ([15], see also [8, Definition 2.5.2]).

Definition 4.1.

Let AA and BB be CC^{*}-algebras. An AA-BB imprimitivity bimodule \mathcal{M} is a Banach space that carries the structure of both a right Hilbert BB-module with BB-inner product ,B\langle\cdot,\cdot\rangle_{B} and a left Hilbert AA-module with AA-inner product ,A{}_{A}\langle\cdot,\cdot\rangle satisfying the following:

  1. (1)

    Both inner products on \mathcal{M} are full, i.e., ,A=A{}_{A}{\langle\mathcal{M},\mathcal{M}\rangle}=A and ,B=B\langle\mathcal{M},\mathcal{M}\rangle_{B}=B;

  2. (2)

    ξ,ηAζ=ξη,ζB{}_{A}{\langle\xi,\eta\rangle}\cdot\zeta=\xi\cdot\langle\eta,\zeta\rangle_{B} for all ξ,η,ζ\xi,\eta,\zeta\in\mathcal{M};

  3. (3)

    The actions of AA and BB on \mathcal{M} commutes.

Say that AA and BB are Morita equivalent if such an AA-BB imprimitivity bimodule \mathcal{M} exists.

Also recall from Section 1 that two CC^{*}-algebras AA and BB are stably isomorphic if A𝔎()A\otimes\mathfrak{K}(\mathcal{H}) is \ast-isomorphic to B𝔎()B\otimes\mathfrak{K}(\mathcal{H}), where \mathcal{H} is a separable infinite dimensional Hilbert space. It was proved in [5] (see also [10, Chapter 7]) that if AA and BB are σ\sigma-unital (i.e., they admit countable approximate units), then AA and BB are Morita equivalent if and only if they are stably isomorphic.

Let us restate Theorem E “(2) \Rightarrow (3)” in a more general form as follows:

Proposition 4.2.

Let (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}) be coarse spaces of bounded geometry which are coarsely equivalent. Then the uniform Roe algebras Cu(X,X)C^{*}_{u}(X,\mathscr{E}_{X}) and Cu(Y,Y)C^{*}_{u}(Y,\mathscr{E}_{Y}) are Morita equivalent. Additionally if X\mathscr{E}_{X} and Y\mathscr{E}_{Y} are unital, then Cu(X,X)C^{*}_{u}(X,\mathscr{E}_{X}) and Cu(Y,Y)C^{*}_{u}(Y,\mathscr{E}_{Y}) are stably isomorphic.

This can be proved essentially by combining [19, Corollary 3.6] and [12, Theorem 2.8] (where second countability is assumed) using the notion of Morita equivalence for groupoids. Also note that the metric space case was proved directly in [4, Theorem 4]. For the reader’s convenience, here we provide a direct proof for the general case.

Proof of Proposition 4.2.

Let \mathscr{E} be a coarse equivalence between (X,X)(X,\mathscr{E}_{X}) and (Y,Y)(Y,\mathscr{E}_{Y}), which is a coarse structure on XYX\sqcup Y. For simplicity, denote A:=Cu(X,X)A:=C^{*}_{u}(X,\mathscr{E}_{X}) and B:=Cu(Y,Y)B:=C^{*}_{u}(Y,\mathscr{E}_{Y}). Define \mathcal{M} to be the norm closure of

{Tu[XY,]|supp(T)X×Y}\{T\in\mathbb{C}_{u}[X\sqcup Y,\mathscr{E}]~{}|~{}\mathrm{supp}(T)\subseteq X\times Y\}

in 𝔅(2(XY))\mathfrak{B}(\ell^{2}(X\sqcup Y)), which is a left AA-module as well as a right BB-module under composition of operators. Also define the inner products

,A:×AbyAT,S:=TSforT,S,{}_{A}\langle\cdot,\cdot\rangle:\mathcal{M}\times\mathcal{M}\longrightarrow A\quad\text{by}\quad_{A}\langle T,S\rangle:=TS^{*}\quad\text{for}\quad T,S\in\mathcal{M},

and

,B:×BbyT,SB:=TSforT,S.\langle\cdot,\cdot\rangle_{B}:\mathcal{M}\times\mathcal{M}\longrightarrow B\quad\text{by}\quad\langle T,S\rangle_{B}:=T^{*}S\quad\text{for}\quad T,S\in\mathcal{M}.

It is straightforward to check that \mathcal{M} is a right Hilbert BB-module as well as a left Hilbert AA-module with commuting actions of AA and BB on \mathcal{M}. Moreover, we have

T,SAR=(TS)R=T(SR)=TS,RB{}_{A}{\langle T,S\rangle}\cdot R=(TS^{*})R=T(S^{*}R)=T\cdot\langle S,R\rangle_{B}

for all T,S,RT,S,R\in\mathcal{M}.

It remains to show that both inner products are full. Given au[X,X]a\in\mathbb{C}_{u}[X,\mathscr{E}_{X}], denote E:=supp(a)E:=\mathrm{supp}(a). It follows from Lemma 3.1 that there exists F|X×YF\in\mathscr{E}|_{X\times Y} such that Δs(E)FF1\Delta_{\mathrm{s}(E)}\subseteq F\circ F^{-1}. Hence F1F^{-1} contains a graph of a map ϕ:s(E)Y\phi:\mathrm{s}(E)\to Y. Thanks to Lemma 3.25, we can apply [18, Lemma 4.10] to write Gr(ϕ)\mathrm{Gr}(\phi) as a finite union of Gr(ϕi)\mathrm{Gr}(\phi_{i}) for i=1,2,,ni=1,2,\cdots,n where each ϕi:DiY\phi_{i}:D_{i}\rightarrow Y is an injective map. Moreover, we can further assume that {Di}i=1n\{D_{i}\}_{i=1}^{n} are mutually disjoint.

For each i=1,2,,ni=1,2,\cdots,n, we set Ti=χGr(ϕi)T_{i}=\chi_{\mathrm{Gr}(\phi_{i})}, which is regarded as a bounded operator from 2(X)\ell^{2}(X) to 2(Y)\ell^{2}(Y). Note that supp(Ti)F1\mathrm{supp}(T_{i})\subseteq F^{-1}, and hence TiT^{*}_{i}\in\mathcal{M}. Moreover, for each i=1,2,,ni=1,2,\cdots,n, we have aTi,TiA=aTiTi=aχΔDi{}_{A}{\langle aT^{*}_{i},T^{*}_{i}\rangle}=aT^{*}_{i}T_{i}=a\chi_{\Delta_{D_{i}}}. Hence

i=1naTi,TiA=iaχΔDi=aχΔs(E)=a,{\sum_{i=1}^{n}}~{}{{}_{A}{\langle aT^{*}_{i},T^{*}_{i}\rangle}}=\sum_{i}a\chi_{\Delta_{D_{i}}}=a\chi_{\Delta_{\mathrm{s}(E)}}=a,

which concludes that the inner AA-product is full. Similarly, we obtain that the inner BB-product is also full. Hence \mathcal{M} is an AA-BB imprimitivity bimodule, which concludes that AA and BB are Morita equivalent. ∎

5. Rigidity for geometric ideals

In this section and the next, we focus on the rigidity problem for geometric ideals and prove Theorem E “(1) \Rightarrow (2)”, which is the main task of this paper. To make the proof more transparent, we prove a weak version in this section. The proof is relatively easier, while contains almost all the necessary ingredients to treat the general case.

Recall that for an ideal II in the uniform Roe algebra of a discrete metric space (X,d)(X,d) of bounded geometry, the associated coarse structure is (I):={suppε(T)|TI,ε>0}\mathscr{I}(I):=\{\mathrm{supp}_{\varepsilon}(T)~{}|~{}T\in I,\varepsilon>0\} from Proposition 2.7. We aim to prove the following:

Theorem 5.1.

Let (X,dX),(Y,dY)(X,d_{X}),(Y,d_{Y}) be discrete metric spaces of bounded geometry, and IX,IYI_{X},I_{Y} be geometric ideals in the uniform Roe algebras Cu(X)C^{*}_{u}(X) and Cu(Y)C^{*}_{u}(Y), respectively. Assume that IXI_{X} and IYI_{Y} are isomorphic. Then the coarse spaces (X,(IX))(X,\mathscr{I}(I_{X})) and (Y,(IY))(Y,\mathscr{I}(I_{Y})) are coarsely equivalent.

The proof of Theorem 5.1 is divided into two parts. We follow the outline of [1], while requiring more techniques to overcome the issue of lacking units.

Throughout the rest of this section, let (X,dX),(Y,dY)(X,d_{X}),(Y,d_{Y}) be discrete metric spaces of bounded geometry, and IX,IYI_{X},I_{Y} be geometric ideals in the uniform Roe algebras Cu(X)C^{*}_{u}(X) and Cu(Y)C^{*}_{u}(Y), respectively. Denote the associated ideal in dX\mathscr{E}_{d_{X}} by X:=(IX)\mathscr{I}_{X}:=\mathscr{I}(I_{X}), and the associated ideal in (X,dX)(X,d_{X}) by 𝐋X:=𝐋(X)\mathbf{L}_{X}:=\mathbf{L}(\mathscr{I}_{X}) (from Proposition 2.7 and 2.9). Similarly, denote Y:=(IY)\mathscr{I}_{Y}:=\mathscr{I}(I_{Y}) and 𝐋Y:=𝐋(Y)\mathbf{L}_{Y}:=\mathbf{L}(\mathscr{I}_{Y}). Also let Φ:IXIY\Phi\colon I_{X}\to I_{Y} be an isomorphism.

5.1. Rigid isomorphisms for geometric ideals

Firstly, recall that the isomorphism Φ:IXIY\Phi\colon I_{X}\to I_{Y} can always be spatially implemented. The proof is similar to that for [21, Lemma 3.1] based on the representation theory for compact operators together with the fact that finite subsets are entourages, and hence we omit the details.

Lemma 5.2.

There exists a unitary U:2(X)2(Y)U:\ell^{2}(X)\to\ell^{2}(Y) such that Φ(a)=UaU\Phi(a)=UaU^{*} for any aIXa\in I_{X}.

In [3], the authors introduced a notion of rigid isomorphism which plays a key role to attack the rigidity problem. The idea can be traced back to the work [21]. We introduce the following to deal with the non-unital case.

Definition 5.3.

With the same notation as above, the isomorphism Φ:IXIY\Phi\colon I_{X}\rightarrow I_{Y} is called a rigid isomorphism if Φ\Phi can be spatially implemented by a unitary U:2(X)2(Y)U:\ell^{2}(X)\to\ell^{2}(Y) satisfying:

infxLsupyY|Uδx,δy|>0for all L𝐋X,\inf_{x\in L}\sup_{y\in Y}|\langle U\delta_{x},\delta_{y}\rangle|>0\quad\text{for all }\quad L\in\mathbf{L}_{X},

and

infyLsupxX|Uδy,δx|>0for all L𝐋Y.\inf_{y\in L^{\prime}}\sup_{x\in X}|\langle U^{*}\delta_{y},\delta_{x}\rangle|>0\quad\text{for all }\quad L^{\prime}\in\mathbf{L}_{Y}.

In this case, we say that IXI_{X} and IYI_{Y} are rigidly isomorphic.

The first step to attack the rigidity problem is the following (comparing [3, Theorem 4.12]):

Proposition 5.4.

With the same notation as above, assume that IXI_{X} and IYI_{Y} is rigidly isomorphic. Then the coarse spaces (X,X)(X,\mathscr{I}_{X}) and (Y,Y)(Y,\mathscr{I}_{Y}) are coarsely equivalent.

To prove Proposition 5.4, we need some preparations.

Definition 5.5 ([3, Definition 4.3]).

Let (X,d)(X,d) be a discrete metric of bounded geometry, ε>0\varepsilon>0 and r0r\geq 0. An operator a𝔅(2(X))a\in\mathfrak{B}(\ell^{2}(X)) is called ε\varepsilon-rr-approximable if there exists b𝔅(2(X))b\in\mathfrak{B}(\ell^{2}(X)) with propagation at most rr such that abε\|a-b\|\leq\varepsilon.

The following key “equi-approximability lemma” was originally obtained in [3, Section 4] (see also [1, Lemma 1.9]), which will be used frequently in the sequel.

Lemma 5.6.

Let (X,d)(X,d) be a discrete metric of bounded geometry, and {an}n\{a_{n}\}_{n} be a sequence of operators such that SOT\mathrm{SOT}-nMan\sum_{n\in M}a_{n} converges to an element in Cu(X)C^{*}_{u}(X) for all MM\subseteq\mathbb{N}. Then for all ε>0\varepsilon>0, there exists r>0r>0 such that SOT\mathrm{SOT}-nMan\sum_{n\in M}a_{n} is ε\varepsilon-rr-approximable for all MM\subseteq\mathbb{N}.

We also need the following lemma, which is analogous to [3, Lemma 4.11]. As explained in Section 1, it is unclear whether the coarse structures X\mathscr{I}_{X} and Y\mathscr{I}_{Y} are small in the sense of [3, Definition 4.2]. This obstructs us from using the results in [3, Section 4] directly, and hence more techniques are required.

Lemma 5.7.

With the notation as above, for any EXE\in\mathscr{I}_{X} and δ>0\delta>0, the following set

F(E,δ):={(y1,y2)Y×Y| there exists (x1,x2)E such that|Uδx1,δy1|δ and |Uδx2,δy2|δ}F(E,\delta):=\left\{(y_{1},y_{2})\in Y\times Y\;\middle|\;\begin{array}[]{l}\text{ there exists }(x_{1},x_{2})\in E\text{ such that}\\ |\langle U\delta_{x_{1}},\delta_{y_{1}}\rangle|\geq\delta\text{ and }|\langle U\delta_{x_{2}},\delta_{y_{2}}\rangle|\geq\delta\end{array}\right\}

belongs to Y\mathscr{I}_{Y}.

Proof.

Suppose otherwise. Then there exist EXE\in\mathscr{I}_{X} and δ>0\delta>0 such that F(E,δ)YF(E,\delta)\notin\mathscr{I}_{Y}. Hence for any FYF\in\mathscr{I}_{Y}, we have F(E,δ)FF(E,\delta)\neq F. Note that subsets of any entourage are entourages (see Definition 2.1(3)), so we have F(E,δ)FF(E,\delta)\setminus F\neq\emptyset. Therefore, we can find (x1F,x2F)E(x_{1}^{F},x_{2}^{F})\in E and (y1F,y2F)Y×Y(y_{1}^{F},y_{2}^{F})\in Y\times Y such that |Uδx1F,δy1F|δ|\langle U\delta_{x_{1}^{F}},\delta_{y_{1}^{F}}\rangle|\geq\delta and |Uδx2F,δy2F|δ|\langle U\delta_{x_{2}^{F}},\delta_{y_{2}^{F}}\rangle|\geq\delta while (y1F,y2F)F(y_{1}^{F},y_{2}^{F})\notin F. Ordering Y\mathscr{I}_{Y} by inclusion, it follows from [3, Lemma 4.10] that there exist cofinal IYI\subset\mathscr{I}_{Y}, JYJ\subset\mathscr{I}_{Y} and a map φ:IJ\varphi:I\rightarrow J such that

  1. (1)

    x1Fx1Fx_{1}^{F}\neq x_{1}^{F^{\prime}} and x2Fx2Fx_{2}^{F}\neq x_{2}^{F^{\prime}} for all FFJF\neq F^{\prime}\in J;

  2. (2)

    x1F=x1φ(F)x_{1}^{F}=x_{1}^{\varphi(F)} and x2F=x2φ(F)x_{2}^{F}=x_{2}^{\varphi(F)} for all FIF\in I.

Since elements in {x1F}FJX\{x_{1}^{F}\}_{F\in J}\subseteq X are distinct, then JJ is countable. For AJA\subseteq J, condition (1) and X\mathscr{I}_{X} having bounded geometry imply that SOT\mathrm{SOT}-FAex1Fx2Fu[X,X]\sum_{F\in A}e_{x_{1}^{F}x_{2}^{F}}\in\mathbb{C}_{u}[X,\mathscr{I}_{X}] with support in EE. Hence SOT\mathrm{SOT}-FAΦ(ex1Fx2F)=SOT\sum_{F\in A}\Phi(e_{x_{1}^{F}x_{2}^{F}})=\mathrm{SOT}-Φ(FAex1Fx2F)\Phi(\sum_{F\in A}e_{x_{1}^{F}x_{2}^{F}}) belongs to Cu(Y)C^{*}_{u}(Y) since Φ\Phi is SOT\mathrm{SOT}-continuous. Applying Lemma 5.6, for ϵ=δ2/3\epsilon=\delta^{2}/3 there exists r>0r>0 such that SOT\mathrm{SOT}-FAΦ(ex1Fx2F)\sum_{F\in A}\Phi(e_{x_{1}^{F}x_{2}^{F}}) is ϵ\epsilon-rr-approximable for all AJA\subseteq J. In particular, we have

(5.1) χAΦ(ex1Fx2F)χB<ϵ\|\chi_{A}\Phi(e_{x_{1}^{F}x_{2}^{F}})\chi_{B}\|<\epsilon

for all FJF\in J and A,BYA,B\subseteq Y with dY(A,B)rd_{Y}(A,B)\geq r.

By Proposition 2.9, we can assume that EL×LE\subseteq L\times L for some L𝐋XL\in\mathbf{L}_{X}. Since {χL}L𝐋Y\{\chi_{L^{\prime}}\}_{{L^{\prime}\in\mathbf{L}_{Y}}} is an approximate unit for IYI_{Y} due to Lemma 2.12, there exist L𝐋YL^{\prime}\in\mathbf{L}_{Y} such that

Φ(χL)χLΦ(χL)<ϵandχLΦ(χL)Φ(χL)<ϵ.\|\Phi(\chi_{L})\chi_{L^{\prime}}-\Phi(\chi_{L})\|<\epsilon\quad\text{and}\quad\|\chi_{L^{\prime}}\Phi(\chi_{L})-\Phi(\chi_{L})\|<\epsilon.

Thus, we have

χLΦ(ex1Fx2F)χLΦ(ex1Fx2F)χLΦ(ex1Fx2F)χLΦ(ex1Fx2F)χL+Φ(ex1Fx2F)χLΦ(ex1Fx2F)\displaystyle\|\chi_{L^{\prime}}\Phi(e_{x_{1}^{F}x_{2}^{F}})\chi_{L^{\prime}}-\Phi(e_{x_{1}^{F}x_{2}^{F}})\|\leq\|\chi_{L^{\prime}}\Phi(e_{x_{1}^{F}x_{2}^{F}})\chi_{L^{\prime}}-\Phi(e_{x_{1}^{F}x_{2}^{F}})\chi_{L^{\prime}}\|+\|\Phi(e_{x_{1}^{F}x_{2}^{F}})\chi_{L^{\prime}}-\Phi(e_{x_{1}^{F}x_{2}^{F}})\|
χLΦ(χL)Φ(ex1Fx2F)Φ(χL)Φ(ex1Fx2F)+Φ(ex1Fx2F)Φ(χL)χLΦ(ex1Fx2F)Φ(χL)\displaystyle\leq\|\chi_{L^{\prime}}\Phi(\chi_{L})\Phi(e_{x_{1}^{F}x_{2}^{F}})-\Phi(\chi_{L})\Phi(e_{x_{1}^{F}x_{2}^{F}})\|+\|\Phi(e_{x_{1}^{F}x_{2}^{F}})\Phi(\chi_{L})\chi_{L^{\prime}}-\Phi(e_{x_{1}^{F}x_{2}^{F}})\Phi(\chi_{L})\|
2ε.\displaystyle\leq 2\varepsilon.

Setting F1:={(y1,y2)L×L|dY(y1,y2)<r}F_{1}:=\{(y_{1},y_{2})\in L^{\prime}\times L^{\prime}~{}|~{}d_{Y}(y_{1},y_{2})<r\}, it is clear that F1YF_{1}\in\mathscr{I}_{Y}. For all A,BYA,B\subseteq Y which are F1F_{1}-separated, then dY(LA,LB)rd_{Y}(L^{\prime}\cap A,L^{\prime}\cap B)\geq r. Hence

χAΦ(ex1Fx2F)χBχAχLΦ(ex1Fx2F)χLχB+2ε=χALΦ(ex1Fx2F)χLB+2ε<3ε=δ2,\|\chi_{A}\Phi(e_{x_{1}^{F}x_{2}^{F}})\chi_{B}\|\leq\|\chi_{A}\chi_{L^{\prime}}\Phi(e_{x_{1}^{F}x_{2}^{F}})\chi_{L^{\prime}}\chi_{B}\|+2\varepsilon=\|\chi_{A\cap L^{\prime}}\Phi(e_{x_{1}^{F}x_{2}^{F}})\chi_{L^{\prime}\cap B}\|+2\varepsilon<3\varepsilon=\delta^{2},

where the penultimate inequality comes from (5.1). Since II is cofinal in Y\mathscr{I}_{Y}, we can assume that F1IF_{1}\in I.

Then we have

χ{y1F1}Φ(ex1F1x2F1)χ{y2F1}=χ{y1F1}Φ(ex1φ(F1)x2φ(F1))χ{y2F1}<δ2.\left\|\chi_{\left\{y_{1}^{F_{1}}\right\}}\Phi(e_{x_{1}^{F_{1}}x_{2}^{F_{1}}})\chi_{\left\{y_{2}^{F_{1}}\right\}}\right\|=\left\|\chi_{\left\{y_{1}^{F_{1}}\right\}}\Phi(e_{x_{1}^{\varphi(F_{1})}x_{2}^{\varphi(F_{1})}})\chi_{\left\{y_{2}^{F_{1}}\right\}}\right\|<\delta^{2}.

On the other hand, a direct calculation shows that

χ{y1F1}Φ(ex1F1x2F1)χ{y2F1}=|δy1F1,U(δx1F1)||U(δx2F1),δy2F1|δ2,\left\|\chi_{\left\{y_{1}^{F_{1}}\right\}}\Phi(e_{x_{1}^{F_{1}}x_{2}^{F_{1}}})\chi_{\left\{y_{2}^{F_{1}}\right\}}\right\|=\left|\left\langle\delta_{y_{1}^{F_{1}}},U\left(\delta_{x_{1}^{F_{1}}}\right)\right\rangle\right|\cdot\left|\left\langle U\left(\delta_{x_{2}^{F_{1}}}\right),\delta_{y_{2}^{F_{1}}}\right\rangle\right|\geq\delta^{2},

which leads to a contradiction. ∎

Now we are in the position to prove Proposition 5.4.

Proof of Proposition 5.4.

Let U:2(X)2(Y)U\colon\ell^{2}(X)\rightarrow\ell^{2}(Y) be a unitary operator which spatially implements the rigid isomorphism Φ:IXIY\Phi:I_{X}\to I_{Y}. Hence:

  • for any L𝐋XL\in\mathbf{L}_{X}, there exist δL>0\delta_{L}>0 and fL:LYf_{L}:L\rightarrow Y such that

    |Uδx,δfL(x)|δLfor anyxL;|\langle U\delta_{x},\delta_{f_{L}(x)}\rangle|\geq\delta_{L}\quad\text{for any}\quad x\in L;
  • for any L𝐋YL^{\prime}\in\mathbf{L}_{Y}, there exist δL>0\delta_{L^{\prime}}>0 and gL:LXg_{L^{\prime}}:L^{\prime}\rightarrow X such that

    |Uδy,δgL(y)|δLfor anyyL.|\langle U^{*}\delta_{y},\delta_{g_{L^{\prime}}(y)}\rangle|\geq\delta_{L^{\prime}}\quad\text{for any}\quad y\in L^{\prime}.

We aim to show that the families :={fL|L𝐋X}\mathcal{F}:=\{f_{L}~{}|~{}L\in\mathbf{L}_{X}\} and 𝒢:={gL|L𝐋Y}\mathcal{G}:=\{g_{L^{\prime}}~{}|~{}L^{\prime}\in\mathbf{L}_{Y}\} provide a coarse equivalence between (X,X)(X,\mathscr{I}_{X}) and (Y,Y)(Y,\mathscr{I}_{Y}) in the sense of Definition 3.23, and hence conclude the proof thanks to Corollary 3.24.

Firstly, we show that the families \mathcal{F} and 𝒢\mathcal{G} are coarse. For L1,L2𝐋XL_{1},L_{2}\in\mathbf{L}_{X}, consider the set {(fL1(x),fL2(x))|xL1L2}\{(f_{L_{1}}(x),f_{L_{2}}(x))\;|\;x\in L_{1}\cap L_{2}\}, which is contained in F(ΔL1L2,δL1δL2)F(\Delta_{L_{1}\cap L_{2}},\delta_{L_{1}}\wedge\delta_{L_{2}}). By Lemma 5.7, the set F(ΔL1L2,δL1δL2)F(\Delta_{L_{1}\cap L_{2}},\delta_{L_{1}}\wedge\delta_{L_{2}}) belongs to Y\mathscr{I}_{Y}. Hence \mathcal{F} is admissible.

Given EXE\in\mathscr{I}_{X}, we can assume that EL×LE\subseteq L\times L for some L𝐋XL\in\mathbf{L}_{X} thanks to Proposition 2.9. Then for any (x1,x2)E(x_{1},x_{2})\in E, it follows from the definition that (fL(x1),fL(x2))F(E,δL)(f_{L}(x_{1}),f_{L}(x_{2}))\in F(E,\delta_{L}). In other words, we obtain (fL×fL)(E)F(E,δL)(f_{L}\times f_{L})(E)\subseteq F(E,\delta_{L}), where the latter belongs to Y\mathscr{I}_{Y} due to Lemma 5.7. Hence we have X(Y)\mathscr{I}_{X}\subseteq\mathcal{F}^{*}(\mathscr{I}_{Y}), which concludes that the family \mathcal{F} is coarse. Similarly, we obtain that the family 𝒢\mathcal{G} is also coarse. Moreover by Lemma 3.13, we have fL(L)𝐋Yf_{L}(L)\in\mathbf{L}_{Y} for any L𝐋XL\in\mathbf{L}_{X}, and gL(L)𝐋Yg_{L^{\prime}}(L^{\prime})\in\mathbf{L}_{Y} for any L𝐋YL^{\prime}\in\mathbf{L}_{Y}.

It remains to show that (3.1) and (3.2) holds for \mathcal{F} and 𝒢\mathcal{G}. Fix L𝐋YL^{\prime}\in\mathbf{L}_{Y} and denote L′′:=gL(L)L^{\prime\prime}:=g_{L^{\prime}}(L^{\prime}). By the choice of gLg_{L^{\prime}} and fL′′f_{L^{\prime\prime}}, we have

|UδgL(y),δy|=|δgL(y),Uδy|δL for any yL|\langle U\delta_{g_{L^{\prime}}(y)},\delta_{y}\rangle|=|\langle\delta_{g_{L^{\prime}}(y)},U^{*}\delta_{y}\rangle|\geq\delta_{L^{\prime}}\text{ for any }y\in L^{\prime}

and

|UδgL(y),δfL′′(gL(y))|δL′′ for any yL.|\langle U\delta_{g_{L^{\prime}}(y)},\delta_{f_{L^{\prime\prime}}(g_{L^{\prime}}(y))}\rangle|\geq\delta_{L^{\prime\prime}}\text{ for any }y\in L^{\prime}.

This implies that

{(y,fL′′(gL(y)))|yL}F(ΔL′′,δLδL′′),\{(y,f_{L^{\prime\prime}}(g_{L^{\prime}}(y)))\;|\;y\in L^{\prime}\}\subseteq F(\Delta_{L^{\prime\prime}},\delta_{L^{\prime}}\wedge\delta_{L^{\prime\prime}}),

which concludes (3.2). Similarly, we obtain (3.1) and finish the proof. ∎

5.2. From isomorphisms to rigid isomorphisms

Now we prove that any isomorphism between geometric ideals is always a rigid isomorphism, and hence conclude the proof for Theorem 5.1. We will follow the outline of the proof for [1, Theorem 1.2] with the following extra piece:

Lemma 5.8.

With that notation as above, for any L𝐋XL\in\mathbf{L}_{X} and ϵ>0\epsilon>0, there exist L𝐋YL^{\prime}\in\mathbf{L}_{Y} such that χLΦ1(χL)χL<ϵ\|\chi_{L}-\Phi^{-1}(\chi_{L^{\prime}})\chi_{L}\|<\epsilon.

Proof.

By Lemma 2.12, {χL|L𝐋Y}\{\chi_{L^{\prime}}~{}|~{}L^{\prime}\in\mathbf{L}_{Y}\} is an approximate unit for IYI_{Y}. Hence, {Φ1(χL)|L𝐋Y}\{\Phi^{-1}(\chi_{L^{\prime}})~{}|~{}L^{\prime}\in\mathbf{L}_{Y}\} is an approximate unit for IXI_{X}, which concludes the proof. ∎

Proposition 5.9.

With the same notation as above, assume that IXI_{X} and IYI_{Y} are isomorphic. Then they are rigidly isomorphic.

Proof.

Fixing L𝐋XL\in\mathbf{L}_{X} and ϵ>0\epsilon>0, take L𝐋YL^{\prime}\in\mathbf{L}_{Y} as in Lemma 5.8. Consider {Φ1(eyy)}yLIX\{\Phi^{-1}(e_{yy})\}_{y\in L^{\prime}}\subseteq I_{X}, which is a sequence of orthogonal projections in Cu(X)C^{*}_{u}(X) and satisfies the condition in Lemma 5.6. Hence, there is r>0r>0 such that

Φ1(χA)=Φ1(SOT-yAeyy)=SOT-yAΦ1(eyy)\Phi^{-1}(\chi_{A})=\Phi^{-1}\big{(}\mathrm{SOT}\text{-}\sum_{y\in A}e_{yy}\big{)}=\mathrm{SOT}\text{-}\sum_{y\in A}\Phi^{-1}(e_{yy})

is ϵ\epsilon-rr-approximable for any ALA\subseteq L^{\prime}.

Set Nr:=supxX|BX(x,r)|N_{r}:=\sup_{x\in X}|B_{X}(x,r)|. It follows from Lemma 5.8 that for any xLx\in L, we have

(5.2) χ{x}Φ1(χL)χ{x}=χ{x}χLΦ1(χL)χ{x}χLϵ.\|\chi_{\{x\}}-\Phi^{-1}(\chi_{L^{\prime}})\chi_{\{x\}}\|=\|\chi_{\{x\}}\chi_{L}-\Phi^{-1}(\chi_{L^{\prime}})\chi_{\{x\}}\chi_{L}\|\leq\epsilon.

Taking δϵ2Nr\delta\leq\frac{\epsilon}{2N_{r}}, we denote

M(x,δ):={yL|Φ1(eyy)δxδ}andM(x,δ):=L\M(x,δ).M(x,\delta):=\{y\in L^{\prime}\;|\;\|\Phi^{-1}(e_{yy})\delta_{x}\|\geq\delta\}\quad\text{and}\quad M^{\prime}(x,\delta):=L^{\prime}\backslash M(x,\delta).

Let π:2(X)2(BX(x,r))\pi:\ell^{2}(X)\rightarrow\ell^{2}(B_{X}(x,r)) be the canonical orthogonal projection. Define μ:𝒫(M(x,δ))2(BX(x,r))\mu:\mathcal{P}(M^{\prime}(x,\delta))\rightarrow\ell^{2}(B_{X}(x,r)) by

μ(A)=π(Φ1(χA)δx), for all AM(x,δ).\mu(A)=\pi(\Phi^{-1}(\chi_{A})\delta_{x}),\text{ for all }A\subseteq M^{\prime}(x,\delta).

Then we have μ({y})=π(Φ1(eyy)δx)Φ1(eyy)δx<δ\|\mu(\{y\})\|=\|\pi(\Phi^{-1}(e_{yy})\delta_{x})\|\leq\|\Phi^{-1}(e_{yy})\delta_{x}\|<\delta for all yM(x,δ)y\in M^{\prime}(x,\delta).

Since μ(M(x,δ))2\frac{\mu(M^{\prime}(x,\delta))}{2} belongs to the convex hull of the range of μ\mu and 2(BX(x,r))\ell^{2}(B_{X}(x,r)) has real dimension at most 2Nr2N_{r}, [1, Lemma 2.1] implies that there exists AM(x,δ)A\subseteq M^{\prime}(x,\delta) such that

(5.3) μ(A)μ(M(x,δ))2<2Nrδϵ.\left\|\mu(A)-\frac{\mu(M^{\prime}(x,\delta))}{2}\right\|<2N_{r}\cdot\delta\leq\epsilon.

Moreover,

(5.4) μ(A)μ(M(x,δ))2\displaystyle\left\|\mu(A)-\frac{\mu(M^{\prime}(x,\delta))}{2}\right\| =π((Φ1(χA)12Φ1(χM(x,δ)))δx)\displaystyle=\left\|\pi\left(\left(\Phi^{-1}(\chi_{A})-\frac{1}{2}\Phi^{-1}(\chi_{M^{\prime}(x,\delta)})\right)\delta_{x}\right)\right\|
=π((12Φ1(χA)12Φ1(χM(x,δ)\A))δx).\displaystyle=\left\|\pi\left(\left(\frac{1}{2}\Phi^{-1}(\chi_{A})-\frac{1}{2}\Phi^{-1}(\chi_{M^{\prime}(x,\delta)\backslash A})\right)\delta_{x}\right)\right\|.

Since Φ1(χA)\Phi^{-1}(\chi_{A}) and Φ1(χM(x,δ)\A)\Phi^{-1}(\chi_{M^{\prime}(x,\delta)\backslash A}) are ϵ\epsilon-rr-approximable, the convex combination 12Φ1(χA)12Φ1(χM(x,δ)\A)\frac{1}{2}\Phi^{-1}(\chi_{A})-\frac{1}{2}\Phi^{-1}(\chi_{M^{\prime}(x,\delta)\backslash A}) is also ϵ\epsilon-rr-approximable. Hence we have

(5.5) (1π)((12Φ1(χA)12Φ1(χM(x,δ)\A))δx)<ϵ.\left\|(1-\pi)\left(\left(\frac{1}{2}\Phi^{-1}(\chi_{A})-\frac{1}{2}\Phi^{-1}(\chi_{M^{\prime}(x,\delta)\backslash A})\right)\delta_{x}\right)\right\|<\epsilon.

Using the fact that Φ1(χA)Φ1(χM(x,δ))=Φ1(χA)\Phi^{-1}(\chi_{A})\Phi^{-1}(\chi_{M^{\prime}(x,\delta)})=\Phi^{-1}(\chi_{A}) together with (5.3)-(5.5), we obtain

Φ1(χA)Φ1(χM(x,δ))δx12Φ1(χM(x,δ))δx\displaystyle\left\|\Phi^{-1}(\chi_{A})\Phi^{-1}(\chi_{M^{\prime}(x,\delta)})\delta_{x}-\frac{1}{2}\Phi^{-1}(\chi_{M^{\prime}(x,\delta)})\delta_{x}\right\|
(1π)((12Φ1(χA)12Φ1(χM(x,δ)\A))δx)+μ(A)μ(M(x,δ))2<2ϵ.\displaystyle\leq\left\|(1-\pi)\left(\left(\frac{1}{2}\Phi^{-1}(\chi_{A})-\frac{1}{2}\Phi^{-1}(\chi_{M^{\prime}(x,\delta)\backslash A})\right)\delta_{x}\right)\right\|+\left\|\mu(A)-\frac{\mu(M^{\prime}(x,\delta))}{2}\right\|<2\epsilon.

Hence [1, Lemma 3.1] implies that Φ1(χM(x,δ))δx<4ϵ\|\Phi^{-1}(\chi_{M^{\prime}(x,\delta)})\delta_{x}\|<4\epsilon. Moreover, note that

Φ1(χM(x,δ))δx+Φ1(χM(x,δ))δx=Φ1(χL)δx.\Phi^{-1}(\chi_{M^{\prime}(x,\delta)})\delta_{x}+\Phi^{-1}(\chi_{M(x,\delta)})\delta_{x}=\Phi^{-1}(\chi_{L^{\prime}})\delta_{x}.

Hence combining (5.2), we obtain Φ1(χM(x,δ))δx15ϵ\|\Phi^{-1}(\chi_{M(x,\delta)})\delta_{x}\|\geq 1-5\epsilon.

Taking ϵ=1/10\epsilon=1/10, we have Φ1(χM(x,δ))δx12\|\Phi^{-1}(\chi_{M(x,\delta)})\delta_{x}\|\geq\frac{1}{2}. Set δ=120Nr\delta=\frac{1}{20N_{r}}, then M(x,δ)M(x,\delta) is non-empty for any xLx\in L. Hence we obtain

infxLsupyLΦ1(eyy)δx120Nr.\inf_{x\in L}\sup_{y\in L^{\prime}}\|\Phi^{-1}(e_{yy})\delta_{x}\|\geq\frac{1}{20N_{r}}.

Finally, note that

Φ1(eyy)δx2=|Uδy,δx|2=|Uδx,δy|2.\|\Phi^{-1}(e_{yy})\delta_{x}\|^{2}=|\langle U^{*}\delta_{y},\delta_{x}\rangle|^{2}=|\langle U\delta_{x},\delta_{y}\rangle|^{2}.

Hence Φ\Phi is a rigid isomorphism. ∎

Proof of Theorem 5.1.

By Proposition 5.9, we know that Φ\Phi is a rigid isomorphism. Hence applying Proposition 5.4, we conclude the proof. ∎

6. Rigidity for stable geometric ideals

Now we move to the case of stable isomorphism and finish the proof of Theorem E “(1) \Rightarrow (2)”. Recall that given a coarse space (X,)(X,\mathscr{E}) of bounded geometry, the stable uniform Roe algebra of (X,)(X,\mathscr{E}) is defined to be

Cs(X,):=Cu(X,)𝔎(),C^{*}_{s}(X,\mathscr{E}):=C^{*}_{u}(X,\mathscr{E})\otimes\mathfrak{K}(\mathcal{H}),

where \mathcal{H} is a separable infinite dimensional Hilbert space. Note that Cs(X,)C^{*}_{s}(X,\mathscr{E}) can be regarded as a CC^{*}-subalgebra of 𝔅(2(X;))\mathfrak{B}(\ell^{2}(X;\mathcal{H})). Similar to the case of the uniform Roe algebra, there is a dense \ast-subalgebra s[X,]\mathbb{C}_{s}[X,\mathscr{E}] in Cs(X,)C^{*}_{s}(X,\mathscr{E}) under this viewpoint. More precisely, an operator T𝔅(2(X;))T\in\mathfrak{B}(\ell^{2}(X;\mathcal{H})) belongs to s[X,]\mathbb{C}_{s}[X,\mathscr{E}] if and only if its support (similar to the definition in Section 2.3) belongs to \mathscr{E} and there exists a finite-dimensional subspace \mathcal{H}^{\prime}\subseteq\mathcal{H} such that each matrix entry T(x,y)T(x,y) belongs to 𝔅()\mathfrak{B}(\mathcal{H}^{\prime}).

Hence Theorem E “(1) \Rightarrow (2)” can be rewritten as follows:

Theorem 6.1.

Let (X,dX),(Y,dY)(X,d_{X}),(Y,d_{Y}) be discrete metric spaces of bounded geometry, and IX,IYI_{X},I_{Y} be geometric ideals in the uniform Roe algebras Cu(X)C^{*}_{u}(X) and Cu(Y)C^{*}_{u}(Y), respectively. Assume that the stable geometric ideals Cs(X,(IX))C^{*}_{s}(X,\mathscr{I}(I_{X})) and Cs(Y,(IY))C^{*}_{s}(Y,\mathscr{I}(I_{Y})) are isomorphic. Then the coarse spaces (X,(IX))(X,\mathscr{I}(I_{X})) and (Y,(IY))(Y,\mathscr{I}(I_{Y})) are coarsely equivalent.

The proof is similar to that for Theorem 5.1 but more technical, which follows the outline of that for [1, Theorem 4.1]. Again throughout the rest of this section, let (X,dX),(Y,dY)(X,d_{X}),(Y,d_{Y}) be discrete metric spaces of bounded geometry, and IX,IYI_{X},I_{Y} be geometric ideals in the uniform Roe algebras Cu(X)C^{*}_{u}(X) and Cu(Y)C^{*}_{u}(Y), respectively. Denote X:=(IX),Y:=(IY)\mathscr{I}_{X}:=\mathscr{I}(I_{X}),\mathscr{I}_{Y}:=\mathscr{I}(I_{Y}) and 𝐋X:=𝐋(X),𝐋Y:=𝐋(Y)\mathbf{L}_{X}:=\mathbf{L}(\mathscr{I}_{X}),\mathbf{L}_{Y}:=\mathbf{L}(\mathscr{I}_{Y}). Also let Φ:Cs(X,(IX))Cs(Y,(IY))\Phi\colon C^{*}_{s}(X,\mathscr{I}(I_{X}))\to C^{*}_{s}(Y,\mathscr{I}(I_{Y})) be an isomorphism with the inverse Ψ:=Φ1\Psi:=\Phi^{-1}.

Firstly, we have the following analogue of Proposition 5.9.

Proposition 6.2.

With the same notation as above, for any unit vector ξ\xi\in\mathcal{H} the following holds:

  1. (1)

    for any L𝐋XL\in\mathbf{L}_{X}, there exist fL:LYf_{L}:L\to Y and a finite-rank projection pLp_{L} on \mathcal{H} such that

    infxLΦ(χ{x}pξ)(χ{fL(x)}pL)>0;\inf_{x\in L}\|\Phi(\chi_{\{x\}}\otimes p_{\xi})(\chi_{\{f_{L}(x)\}}\otimes p_{L})\|>0;
  2. (2)

    for any L𝐋YL^{\prime}\in\mathbf{L}_{Y}, there exist gL:LXg_{L^{\prime}}:L^{\prime}\to X and a finite-rank projection pLp_{L^{\prime}} on \mathcal{H} such that

    infyLΦ(χ{y}pξ)(χ{gL(y)}pL)>0.\inf_{y\in L^{\prime}}\|\Phi(\chi_{\{y\}}\otimes p_{\xi})(\chi_{\{g_{L^{\prime}}(y)\}}\otimes p_{L^{\prime}})\|>0.

Here pξp_{\xi} is the orthogonal projection from \mathcal{H} onto ξ\mathbb{C}\xi.

The proof is similar to that for [1, Theorem 4.1] (with the same idea presented in the proof of Proposition 5.9), and hence we only provide a sketch here.

Sketch of proof for Proposition 6.2.

Note that the set

{χLq|L𝐋Y,q is a finite dimensional projection on }\{\chi_{L^{\prime}}\otimes q~{}|~{}L^{\prime}\in\mathbf{L}_{Y},q\text{ is a finite dimensional projection on }\mathcal{H}\}

is an approximate unit for IY𝔎()I_{Y}\otimes\mathfrak{K}(\mathcal{H}). Hence given ε>0\varepsilon>0 and L𝐋XL\in\mathbf{L}_{X}, there exists L𝐋YL^{\prime}\in\mathbf{L}_{Y} and a finite dimensional projection pLp_{L} on \mathcal{H} such that

Ψ(χLpL)(χLpξ)χLpξ<ε.\|\Psi(\chi_{L^{\prime}}\otimes p_{L})(\chi_{L}\otimes p_{\xi})-\chi_{L}\otimes p_{\xi}\|<\varepsilon.

Therefore, we have

(Id2(L;)χLpL)Φ(χ{x}ξ)<ε,xL.\|(\mathrm{Id}_{\ell^{2}(L^{\prime};\mathcal{H})}-\chi_{L^{\prime}}\otimes p_{L})\Phi(\chi_{\{x\}}\otimes\xi)\|<\varepsilon,\quad\forall x\in L.

This provides a similar condition as in the hypothesis of [1, Lemma 4.3], which allows us to apply the same argument therein to obtain a constant δ=δ(ε,L)>0\delta=\delta(\varepsilon,L)>0 and a map fL:LLf_{L}:L\to L^{\prime} satisfying

Φ(χ{x}pξ)(χ{fL(x)}pL)=Ψ(χ{fL(x)}pL)(δxξ)>δ,xL.\|\Phi(\chi_{\{x\}}\otimes p_{\xi})(\chi_{\{f_{L}(x)\}}\otimes p_{L})\|=\|\Psi(\chi_{\{f_{L}(x)\}}\otimes p_{L})(\delta_{x}\otimes\xi)\|>\delta,\quad\forall x\in L.

Hence we obtain (1), and (2) can be treated similarly. ∎

Now we show that the conditions in Proposition 6.2 imply the required coarse equivalence, analogous to Proposition 5.4.

Proposition 6.3.

With the same notation as above, assume that for any unit vector ξ\xi\in\mathcal{H}, condition (1) and (2) in Proposition 6.2 hold with functions fLf_{L} and gLg_{L^{\prime}} for L𝐋XL\in\mathbf{L}_{X} and L𝐋YL^{\prime}\in\mathbf{L}_{Y}. Then the family :={fL|L𝐋X}\mathcal{F}:=\{f_{L}~{}|~{}L\in\mathbf{L}_{X}\} is a coarse equivalence (in the sense of Definition 3.23) from (X,X)(X,\mathscr{I}_{X}) to (Y,Y)(Y,\mathscr{I}_{Y}) with a coarse inverse 𝒢:={gL|L𝐋Y}\mathcal{G}:=\{g_{L^{\prime}}~{}|~{}L^{\prime}\in\mathbf{L}_{Y}\}.

To prove Proposition 6.3, first note that there exists a unitary U:2(X)2(Y)U:\ell^{2}(X)\otimes\mathcal{H}\to\ell^{2}(Y)\otimes\mathcal{H} such that Φ(a)=UaU\Phi(a)=UaU^{*} for any aIX𝔎()a\in I_{X}\otimes\mathfrak{K}(\mathcal{H}), which is similar to Lemma 5.2. Analogous to Lemma 5.7, we also need the following lemma:

Lemma 6.4.

With the notation as above, for any EX,δ>0E\in\mathscr{I}_{X},\delta>0 and a finite dimensional subspace VV\subseteq\mathcal{H}, the following set

F(E,δ,V):={(y1,y2)Y×Y| there exist (x1,x2)E,unit vectors v1,v2V and unit vectors w1,w2 such that|U(δx1v1),δy1w1|δ and |U(δx2v2),δy2w2|δ}F(E,\delta,V):=\left\{(y_{1},y_{2})\in Y\times Y\;\middle|\;\begin{array}[]{l}\text{ there exist }(x_{1},x_{2})\in E,\text{unit vectors }v_{1},v_{2}\in V\\ \text{ and unit vectors }w_{1},w_{2}\in\mathcal{H}\text{ such that}\\ |\langle U(\delta_{x_{1}}\otimes v_{1}),\delta_{y_{1}}\otimes w_{1}\rangle|\geq\delta\text{ and }|\langle U(\delta_{x_{2}}\otimes v_{2}),\delta_{y_{2}}\otimes w_{2}\rangle|\geq\delta\end{array}\right\}

belongs to Y\mathscr{I}_{Y}.

The proof for Lemma 6.4 is similar to that for Lemma 5.7 with minor modifications, and hence we only provide a sketch to explain the difference.

Sketch of proof for Lemma 6.4.

Suppose otherwise. Then there exist EX,δ>0E\in\mathscr{I}_{X},\delta>0 and a finite dimensional subspace VV\subseteq\mathcal{H} such that F(E,δ,V)YF(E,\delta,V)\notin\mathscr{I}_{Y}. Therefore for any FYF\in\mathscr{I}_{Y}, we can find (x1F,x2F)E,(y1F,y2F)Y×Y(x_{1}^{F},x_{2}^{F})\in E,(y_{1}^{F},y_{2}^{F})\in Y\times Y, unit vectors v1F,v2FVv_{1}^{F},v_{2}^{F}\in V and unit vectors w1F,w2Fw_{1}^{F},w_{2}^{F}\in\mathcal{H} such that

|U(δxiFviF),δyiFwiF|δ\left|\left\langle U\left(\delta_{x^{F}_{i}}\otimes v^{F}_{i}\right),\delta_{y^{F}_{i}}\otimes w^{F}_{i}\right\rangle\right|\geq\delta

for i=1,2i=1,2, while (y1F,y2F)F(y_{1}^{F},y_{2}^{F})\notin F. Note that v1F,v2Fv_{1}^{F},v_{2}^{F} are unit vectors in VV and VV is finite dimensional. Hence after a small perturbation, we can assume that the set {v1F,v2F}FY\left\{v_{1}^{F},v_{2}^{F}\right\}_{F\in\mathscr{I}_{Y}} is finite. Ordering Y\mathscr{I}_{Y} by inclusion, it follows from [3, Lemma 4.10] that there exist cofinal IYI\subset\mathscr{I}_{Y}, JYJ\subset\mathscr{I}_{Y} and a map φ:IJ\varphi:I\rightarrow J such that

  1. (1)

    x1Fx1Fx_{1}^{F}\neq x_{1}^{F^{\prime}} and x2Fx2Fx_{2}^{F}\neq x_{2}^{F^{\prime}} for all FFJF\neq F^{\prime}\in J;

  2. (2)

    x1F=x1φ(F)x_{1}^{F}=x_{1}^{\varphi(F)}, x2F=x2φ(F)x_{2}^{F}=x_{2}^{\varphi(F)}, v1F=v1φ(F)v_{1}^{F}=v_{1}^{\varphi(F)} and v2F=v2φ(F)v_{2}^{F}=v_{2}^{\varphi(F)} for all FIF\in I.

The rest is similar to that of Lemma 5.7, and hence we omit the details. ∎

Now we use Lemma 6.4 to prove Proposition 6.3.

Proof of Proposition 6.3.

As remarked above, there exists a unitary U:2(X)2(Y)U:\ell^{2}(X)\otimes\mathcal{H}\to\ell^{2}(Y)\otimes\mathcal{H} such that Φ(a)=UaU\Phi(a)=UaU^{*} for any aIX𝔎()a\in I_{X}\otimes\mathfrak{K}(\mathcal{H}). Fix a unit vector ξ\xi\in\mathcal{H}. Direct calculations show that conditions in Proposition 6.2 can be translated as follows:

  • for any L𝐋XL\in\mathbf{L}_{X}, there exist δL>0\delta_{L}>0, fL:LYf_{L}:L\to Y and a finite-rank projection pLp_{L} on \mathcal{H} such that for any xLx\in L, there exists a unit vector wxpLw_{x}\in p_{L}\mathcal{H} satisfying:

    |U(δxξ),δfL(x)wx|δL;|\langle U(\delta_{x}\otimes\xi),\delta_{f_{L}(x)}\otimes w_{x}\rangle|\geq\delta_{L};
  • for any L𝐋YL^{\prime}\in\mathbf{L}_{Y}, there exist δL>0\delta_{L^{\prime}}>0, gL:LXg_{L^{\prime}}:L^{\prime}\to X and a finite-rank projection pLp_{L^{\prime}} on \mathcal{H} such that for any yLy\in L^{\prime}, there exists a unit vector vypLv_{y}\in p_{L^{\prime}}\mathcal{H} satisfying:

    |U(δyξ),δgL(y)vy|δL.|\langle U^{*}(\delta_{y}\otimes\xi),\delta_{g_{L^{\prime}}(y)}\otimes v_{y}\rangle|\geq\delta_{L^{\prime}}.

We aim to show that the families :={fL|L𝐋X}\mathcal{F}:=\{f_{L}~{}|~{}L\in\mathbf{L}_{X}\} and 𝒢:={gL|L𝐋Y}\mathcal{G}:=\{g_{L^{\prime}}~{}|~{}L^{\prime}\in\mathbf{L}_{Y}\} provide a coarse equivalence between (X,X)(X,\mathscr{I}_{X}) and (Y,Y)(Y,\mathscr{I}_{Y}), and hence conclude the proof thanks to Corollary 3.24.

Firstly, we show that the families \mathcal{F} and 𝒢\mathcal{G} are coarse. For L1,L2𝐋XL_{1},L_{2}\in\mathbf{L}_{X}, consider the set {(fL1(x),fL2(x))|xL1L2}\{(f_{L_{1}}(x),f_{L_{2}}(x))\;|\;x\in L_{1}\cap L_{2}\}, which is contained in F(ΔL1L2,δL1δL2,ξ)F(\Delta_{L_{1}\cap L_{2}},\delta_{L_{1}}\wedge\delta_{L_{2}},\mathbb{C}\xi). By Lemma 6.4, the set F(ΔL1L2,δL1δL2,ξ)F(\Delta_{L_{1}\cap L_{2}},\delta_{L_{1}}\wedge\delta_{L_{2}},\mathbb{C}\xi) belongs to Y\mathscr{I}_{Y}. Hence \mathcal{F} is admissible.

Given EXE\in\mathscr{I}_{X}, we can assume that EL×LE\subseteq L\times L for some L𝐋XL\in\mathbf{L}_{X} thanks to Proposition 2.9. Then for any (x1,x2)E(x_{1},x_{2})\in E, it follows that (fL(x1),fL(x2))F(E,δL,ξ)(f_{L}(x_{1}),f_{L}(x_{2}))\in F(E,\delta_{L},\mathbb{C}\xi). Hence we have X(Y)\mathscr{I}_{X}\subseteq\mathcal{F}^{*}(\mathscr{I}_{Y}), which concludes that the family \mathcal{F} is coarse. Similarly, we obtain that the family 𝒢\mathcal{G} is also coarse.

It remains to show that (3.1) and (3.2) holds for \mathcal{F} and 𝒢\mathcal{G}. Fix L𝐋YL^{\prime}\in\mathbf{L}_{Y} and denote L′′:=gL(L)L^{\prime\prime}:=g_{L^{\prime}}(L^{\prime}). By the choice of gLg_{L^{\prime}} and fL′′f_{L^{\prime\prime}}, for any yLy\in L^{\prime} there exists vypLv_{y}\in p_{L^{\prime}}\mathcal{H} such that

|U(δgL(y)vy),δyξ|δL|\langle U(\delta_{g_{L^{\prime}}(y)}\otimes v_{y}),\delta_{y}\otimes\xi\rangle|\geq\delta_{L^{\prime}}

and there exists wgL(y)pgL(y)w_{g_{L^{\prime}}(y)}\in p_{g_{L^{\prime}}(y)}\mathcal{H} such that

|U(δgL(y)ξ),δfL′′(gL(y))wgL(y)|δL′′.|\langle U(\delta_{g_{L^{\prime}}(y)}\otimes\xi),\delta_{f_{L^{\prime\prime}}(g_{L^{\prime}}(y))}\otimes w_{g_{L^{\prime}}(y)}\rangle|\geq\delta_{L^{\prime\prime}}.

This implies that

{(y,fL′′(gL(y)))|yL}F(ΔL′′,δLδL′′,ξ+pL)Y,\{(y,f_{L^{\prime\prime}}(g_{L^{\prime}}(y)))\;|\;y\in L^{\prime}\}\subseteq F(\Delta_{L^{\prime\prime}},\delta_{L^{\prime}}\wedge\delta_{L^{\prime\prime}},\mathbb{C}\xi+p_{L^{\prime}}\mathcal{H})\in\mathscr{I}_{Y},

which concludes (3.2). Similarly, we obtain (3.1) and finish the proof. ∎

Combining Proposition 6.2 and Proposition 6.3, we conclude the proof of Theorem 6.1.

Finally, we study the σ\sigma-unitalness of geometric ideals and prove the last sentence in Theorem E. Recall from [24, Definition 7.15] that an ideal 𝐋\mathbf{L} in a metric space (X,d)(X,d) is countably generated if there exists a countable subset 𝒮𝐋\mathcal{S}\subseteq\mathbf{L} such that 𝐋\mathbf{L} is generated by 𝒮\mathcal{S}.

Lemma 6.5.

Let (X,d)(X,d) be a discrete metric space of bounded geometry and 𝐋\mathbf{L} be a countably generated ideal in (X,d)(X,d). Then the geometric ideal Cu(X,(𝐋))C^{*}_{u}(X,\mathscr{I}(\mathbf{L})) is σ\sigma-unital.

Proof.

Since 𝐋\mathbf{L} is countably generated, it follows from [24, Lemma 7.17] that there exists a countable subset {Y1,Y2,,Yn,}\{Y_{1},Y_{2},\cdots,Y_{n},\cdots\} of 𝐋\mathbf{L} such that 𝐋={ZX|n such that ZYn}\mathbf{L}=\{Z\subseteq X~{}|~{}\exists~{}n\in\mathbb{N}\text{ such that }Z\subseteq Y_{n}\}. It is then easy to see that {χYn|n}\{\chi_{Y_{n}}~{}|~{}n\in\mathbb{N}\} is an approximate unit for Cu(X,(𝐋))C^{*}_{u}(X,\mathscr{I}(\mathbf{L})), i.e., Cu(X,(𝐋))C^{*}_{u}(X,\mathscr{I}(\mathbf{L})) is σ\sigma-unital. ∎

Combining Proposition 4.2, Theorem 6.1 and Lemma 6.5, we finally conclude the proof for Theorem E.

7. Discussions on ghostly ideals

In [24], Wang and the second-named author introduced a notion of ghostly ideals in uniform Roe algebras. Here we provide some discussions on the rigidity problem for ghostly ideals and pose some open questions. Recall the following:

Definition 7.1 ([24]).

Let (X,d)(X,d) be a discrete metric space of bounded geometry, and 𝐋\mathbf{L} be an ideal in (X,d)(X,d). The associated ghostly ideal of 𝐋\mathbf{L}, denoted by I~(𝐋)\tilde{I}(\mathbf{L}), is defined as follows:

I~(𝐋):={TCu(X)|ε>0,r(suppε(T))𝐋}.\tilde{I}(\mathbf{L}):=\{T\in C^{*}_{u}(X)~{}|~{}\forall~{}\varepsilon>0,\mathrm{r}(\mathrm{supp}_{\varepsilon}(T))\in\mathbf{L}\}.

Similar to the situation for geometric ideals, we would also like to study isomorphisms between ghostly ideals. More precisely, similar to the discussions in Section 4 to Section 6, we pose the following questions:

Question 7.2.

Let (X,dX),(Y,dY)(X,d_{X}),(Y,d_{Y}) be discrete metric spaces of bounded geometry, and 𝐋X,𝐋Y\mathbf{L}_{X},\mathbf{L}_{Y} be ideals in (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}), respectively.

  1. (1)

    If (X,(𝐋X))(X,\mathscr{I}(\mathbf{L}_{X})) and (Y,(𝐋X))(Y,\mathscr{I}(\mathbf{L}_{X})) are coarsely equivalent (in the sense of Definition C), are I~(𝐋X)\tilde{I}(\mathbf{L}_{X}) and I~(𝐋Y)\tilde{I}(\mathbf{L}_{Y}) Morita equivalent?

  2. (2)

    Conversely, if I~(𝐋X)\tilde{I}(\mathbf{L}_{X}) and I~(𝐋Y)\tilde{I}(\mathbf{L}_{Y}) are isomorphic (or stably isomorphic), are (X,(𝐋X))(X,\mathscr{I}(\mathbf{L}_{X})) and (Y,(𝐋X))(Y,\mathscr{I}(\mathbf{L}_{X})) coarsely equivalent?

Unfortunately, currently we are unable to completely answer either of these questions. We merely manage to provide a partial answer to Question 7.2(1). To state our result, we need some preparations.

Let f:(X,dX)(Y,dY)f:(X,d_{X})\to(Y,d_{Y}) be a coarse equivalence with a coarse inverse g:YXg:Y\to X, and 𝐋\mathbf{L} be an ideal in (X,dX)(X,d_{X}). Denote f(𝐋)f_{\ast}(\mathbf{L}) the ideal in (Y,dY)(Y,d_{Y}) generated by the set {f(L)|L𝐋}\{f(L)~{}|~{}L\in\mathbf{L}\}. Then we have:

Lemma 7.3.

With the same notation as above, we have gf(𝐋)=𝐋g_{\ast}f_{\ast}(\mathbf{L})=\mathbf{L}. Hence in this case, the maps ff_{\ast} and gg_{\ast} provide a one-to-one correspondence between ideals in (X,dX)(X,d_{X}) and those in (Y,dY)(Y,d_{Y}).

Proof.

Firstly, note that gf(L)𝐋gf(L)\in\mathbf{L} for any L𝐋L\in\mathbf{L}, which implies that 𝐋gf(𝐋)\mathbf{L}\subseteq g_{\ast}f_{\ast}(\mathbf{L}). Conversely, it suffices to show that g(L)𝐋g(L^{\prime})\in\mathbf{L} for any Lf(𝐋)L^{\prime}\in f_{\ast}(\mathbf{L}). Given Lf(𝐋)L^{\prime}\in f_{\ast}(\mathbf{L}), It follows from [24, Lemma 7.18] that there exist R>0R>0 and L𝐋L\in\mathbf{L} such that L𝒩R(f(L))L^{\prime}\subseteq\mathcal{N}_{R}(f(L)). Hence g(L)g(𝒩R(f(L)))g(L^{\prime})\subseteq g(\mathcal{N}_{R}(f(L))), which belongs to 𝐋\mathbf{L} since gg is coarse. So we conclude the proof. ∎

As a direct corollary, we obtain the following:

Corollary 7.4.

Let f:(X,dX)(Y,dY)f:(X,d_{X})\to(Y,d_{Y}) be a coarse equivalence and 𝐋\mathbf{L} be an ideal in (X,dX)(X,d_{X}). Then the family (f):={f|L:LY|L𝐋}\mathcal{F}(f):=\{f|_{L}:L\to Y~{}|~{}L\in\mathbf{L}\} is a coarse equivalence (in the sense of Definition 3.23) between (X,(𝐋))(X,\mathscr{I}(\mathbf{L})) and (Y,(f(𝐋)))(Y,\mathscr{I}(f_{\ast}(\mathbf{L}))).

Now we present our partial answer to Question 7.2(1):

Proposition 7.5.

Let (X,dX),(Y,dY)(X,d_{X}),(Y,d_{Y}) be discrete metric spaces of bounded geometry, and 𝐋X,𝐋Y\mathbf{L}_{X},\mathbf{L}_{Y} be ideals in (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}), respectively. If f:(X,dX)(Y,dY)f:(X,d_{X})\to(Y,d_{Y}) is a coarse equivalence with f(𝐋X)=𝐋Yf_{\ast}(\mathbf{L}_{X})=\mathbf{L}_{Y}, then I~(𝐋X)\tilde{I}(\mathbf{L}_{X}) and I~(𝐋Y)\tilde{I}(\mathbf{L}_{Y}) are Morita equivalent.

Remark 7.6.

By Corollary 7.4, the assumption in Proposition 7.5 implies that (X,(𝐋X))(X,\mathscr{I}(\mathbf{L}_{X})) and (Y,(𝐋Y))(Y,\mathscr{I}(\mathbf{L}_{Y})) are coarsely equivalent, which is the assumption considered in Question 7.2(1). However, it is unclear to us whether the converse holds or not.

Proof of Proposition 7.5.

Let f:(X,dX)(Y,dY)f:(X,d_{X})\to(Y,d_{Y}) be a coarse equivalence and (f)\mathscr{E}(f) the associated coarse correspondence. Similar to the proof of Proposition 4.2, we define (f)\mathcal{M}(f) to be the norm closure of

{Tu[XY,(f)]|supp(T)X×Y}\{T\in\mathbb{C}_{u}[X\sqcup Y,\mathscr{E}(f)]~{}|~{}\mathrm{supp}(T)\subseteq X\times Y\}

in 𝔅(2(XY))\mathfrak{B}(\ell^{2}(X\sqcup Y)), which is a left Cu(X)C^{*}_{u}(X)-module as well as a right Cu(Y)C^{*}_{u}(Y)-module under composition of operators. Writing I~X:=I~(𝐋X)\tilde{I}_{X}:=\tilde{I}(\mathbf{L}_{X}) and I~Y:=I~(𝐋Y)\tilde{I}_{Y}:=\tilde{I}(\mathbf{L}_{Y}), we set :=I~X(f)I~Y\mathcal{M}:=\tilde{I}_{X}\mathcal{M}(f)\tilde{I}_{Y}. Also define the inner products

,I~X:×I~XbyI~XT,S:=TSforT,S,{}_{\tilde{I}_{X}}\langle\cdot,\cdot\rangle:\mathcal{M}\times\mathcal{M}\longrightarrow\tilde{I}_{X}\quad\text{by}\quad_{\tilde{I}_{X}}\langle T,S\rangle:=TS^{*}\quad\text{for}\quad T,S\in\mathcal{M},

and

,I~Y:×I~YbyT,SI~Y:=TSforT,S.\langle\cdot,\cdot\rangle_{\tilde{I}_{Y}}:\mathcal{M}\times\mathcal{M}\longrightarrow\tilde{I}_{Y}\quad\text{by}\quad\langle T,S\rangle_{\tilde{I}_{Y}}:=T^{*}S\quad\text{for}\quad T,S\in\mathcal{M}.

It is straightforward to check that \mathcal{M} is a right Hilbert I~Y\tilde{I}_{Y}-module as well as a left Hilbert I~X\tilde{I}_{X}-module with commuting actions of I~X\tilde{I}_{X} and I~Y\tilde{I}_{Y} on \mathcal{M}. We will show that \mathcal{M} provides a Morita equivalence between I~X\tilde{I}_{X} and I~Y\tilde{I}_{Y}.

Firstly, we claim that (f)I~Y(f)I~X\mathcal{M}(f)\tilde{I}_{Y}\mathcal{M}(f)^{*}\subseteq\tilde{I}_{X} where (f):={T|T(f)}\mathcal{M}(f)^{*}:=\{T^{*}~{}|~{}T\in\mathcal{M}(f)\}. Given T1,T2u[XY,(f)]T_{1},T_{2}\in\mathbb{C}_{u}[X\sqcup Y,\mathscr{E}(f)] with support in X×YX\times Y and SI~YS\in\tilde{I}_{Y}, there exists R>0R>0 such that

supp(Ti){(x,y)X×Y|dY(f(x),y)R}fori=1,2.\mathrm{supp}(T_{i})\subseteq\{(x,y)\in X\times Y~{}|~{}d_{Y}(f(x),y)\leq R\}\quad\text{for}\quad i=1,2.

Hence for x1,x2Xx_{1},x_{2}\in X, we have

(T1ST2)(x1,x2)=y1BY(f(x1),R)y2BY(f(x2,R))T1(x1,y1)S(y1,y2)T2(y2,x2).(T_{1}ST^{*}_{2})(x_{1},x_{2})=\sum_{\begin{subarray}{c}y_{1}\in B_{Y}(f(x_{1}),R)\\ y_{2}\in B_{Y}(f(x_{2},R))\end{subarray}}T_{1}(x_{1},y_{1})S(y_{1},y_{2})T_{2}^{*}(y_{2},x_{2}).

Denote NR:=supyY|BY(y,R)|<N_{R}:=\sup_{y\in Y}|B_{Y}(y,R)|<\infty. Then for any ε>0\varepsilon>0 and (x1,x2)suppε(T1ST2)(x_{1},x_{2})\in\mathrm{supp}_{\varepsilon}(T_{1}ST^{*}_{2}), there exist y1BY(f(x1),R)y_{1}\in B_{Y}(f(x_{1}),R) and y2BY(f(x2),R)y_{2}\in B_{Y}(f(x_{2}),R) such that

(y1,y2)suppε(S)forε:=εT1T2NR2.(y_{1},y_{2})\in\mathrm{supp}_{\varepsilon^{\prime}}(S)\quad\text{for}\quad\varepsilon^{\prime}:=\frac{\varepsilon}{\|T_{1}\|\cdot\|T_{2}\|\cdot N_{R}^{2}}.

Since SI~YS\in\tilde{I}_{Y}, there exists L𝐋YL\in\mathbf{L}_{Y} such that r(suppε(S))L\mathrm{r}(\mathrm{supp}_{\varepsilon^{\prime}}(S))\subseteq L. Hence we obtain

f(r(suppε(T1ST2)))𝒩R(r(suppε(S)))𝒩R(L)𝐋Y.f\big{(}\mathrm{r}(\mathrm{supp}_{\varepsilon}(T_{1}ST^{*}_{2}))\big{)}\subseteq\mathcal{N}_{R}\big{(}\mathrm{r}(\mathrm{supp}_{\varepsilon^{\prime}}(S))\big{)}\subseteq\mathcal{N}_{R}(L)\in\mathbf{L}_{Y}.

Therefore by Lemma 7.3 and the assumption that f(𝐋X)=𝐋Yf_{\ast}(\mathbf{L}_{X})=\mathbf{L}_{Y}, we obtain that

r(suppε(T1ST2))𝐋X,\mathrm{r}(\mathrm{supp}_{\varepsilon}(T_{1}ST^{*}_{2}))\in\mathbf{L}_{X},

which implies that T1ST2I~XT_{1}ST^{*}_{2}\in\tilde{I}_{X} and concludes the claim.

Now we claim that I~X(f)=(f)I~Y\tilde{I}_{X}\mathcal{M}(f)=\mathcal{M}(f)\tilde{I}_{Y}. Taking a coarse inverse g:YXg:Y\to X to ff, we decompose Gr(g)\mathrm{Gr}(g) into a finite union of Gr(gi)\mathrm{Gr}(g_{i}) for i=1,,ni=1,\cdots,n such that each gi:DiXg_{i}:D_{i}\to X is an injective map with {Di}i=1n\{D_{i}\}_{i=1}^{n} mutually disjoint. For each i=1,,ni=1,\cdots,n, define Ti:=χGr(gi)T^{\prime}_{i}:=\chi_{\mathrm{Gr}(g_{i})}, which is regarded as a bounded operator from 2(Y)\ell^{2}(Y) to 2(X)\ell^{2}(X). Clearly, we have Ti(f)T^{\prime}_{i}\in\mathcal{M}(f), (Ti)Ti=Id2(Di)(T^{\prime}_{i})^{*}T^{\prime}_{i}=\mathrm{Id}_{\ell^{2}(D_{i})} and i=1n(Ti)Ti=Id2(Y)\sum_{i=1}^{n}(T^{\prime}_{i})^{*}T^{\prime}_{i}=\mathrm{Id}_{\ell^{2}(Y)}. Hence for T(f)T\in\mathcal{M}(f) and SI~YS\in\tilde{I}_{Y}, we have

TS=TSi=1n(Ti)Ti((f)I~Y(f))(f)I~X(f),TS=TS\sum_{i=1}^{n}(T^{\prime}_{i})^{*}T^{\prime}_{i}\in\big{(}\mathcal{M}(f)\tilde{I}_{Y}\mathcal{M}(f)^{*}\big{)}\cdot\mathcal{M}(f)\subseteq\tilde{I}_{X}\mathcal{M}(f),

where the last containment follows from the previous paragraph. By symmetry, we conclude I~X(f)=(f)I~Y\tilde{I}_{X}\mathcal{M}(f)=\mathcal{M}(f)\tilde{I}_{Y}. Consequently, we obtain =I~X(f)=(f)I~Y\mathcal{M}=\tilde{I}_{X}\mathcal{M}(f)=\mathcal{M}(f)\tilde{I}_{Y}.

From the proof of Proposition 4.2 together with Corollary 7.4, we know that (f),(f)Cu(Y)=Cu(Y)\langle\mathcal{M}(f),\mathcal{M}(f)\rangle_{C^{*}_{u}(Y)}=C^{*}_{u}(Y). Hence we have

,I~Y=(f)I~Y,(f)I~YI~Y=I~Y(f),(f)Cu(Y)I~Y=I~Y.\langle\mathcal{M},\mathcal{M}\rangle_{\tilde{I}_{Y}}=\langle\mathcal{M}(f)\tilde{I}_{Y},\mathcal{M}(f)\tilde{I}_{Y}\rangle_{\tilde{I}_{Y}}=\tilde{I}_{Y}\cdot\langle\mathcal{M}(f),\mathcal{M}(f)\rangle_{C^{*}_{u}(Y)}\cdot\tilde{I}_{Y}=\tilde{I}_{Y}.

Hence the inner I~Y\tilde{I}_{Y}-product on \mathcal{M} is full. Similarly, we obtain that the inner I~X\tilde{I}_{X}-product on \mathcal{M} is full. Therefore, \mathcal{M} is a I~X\tilde{I}_{X}-I~Y\tilde{I}_{Y} imprimitivity bimodule, which concludes that I~X\tilde{I}_{X} and I~Y\tilde{I}_{Y} are Morita equivalent. ∎

Finally, we remark that it seems that the proof of Proposition 7.5 does not work if we only know that (X,(𝐋X))(X,\mathscr{I}(\mathbf{L}_{X})) and (Y,(𝐋Y))(Y,\mathscr{I}(\mathbf{L}_{Y})) are coarsely equivalent instead of requiring a global coarse equivalence from (X,dX)(X,d_{X}) to (Y,dY)(Y,d_{Y}).

References

  • [1] Florent P. Baudier, Bruno M. Braga, Ilijas Farah, Ana Khukhro, Alessandro Vignati, and Rufus Willett. Uniform Roe algebras of uniformly locally finite metric spaces are rigid. Invent. Math., 230(3):1071–1100, 2022.
  • [2] Bruno M. Braga, Yeong Chyuan Chung, and Kang Li. Coarse Baum-Connes conjecture and rigidity for Roe algebras. J. Funct. Anal., 279(9):108728, 21, 2020.
  • [3] Bruno M. Braga and Ilijas Farah. On the rigidity of uniform Roe algebras over uniformly locally finite coarse spaces. Trans. Amer. Math. Soc., 374(2):1007–1040, 2021.
  • [4] Jacek Brodzki, Graham A. Niblo, and Nick J. Wright. Property A, partial translation structures, and uniform embeddings in groups. J. Lond. Math. Soc. (2), 76(2):479–497, 2007.
  • [5] Lawrence G. Brown, Philip Green, and Marc A. Rieffel. Stable isomorphism and strong Morita equivalence of CC^{*}-algebras. Pacific J. Math., 71(2):349–363, 1977.
  • [6] Xiaoman Chen and Qin Wang. Ideal structure of uniform Roe algebras of coarse spaces. J. Funct. Anal., 216(1):191 – 211, 2004.
  • [7] Xiaoman Chen and Qin Wang. Ghost ideals in uniform Roe algebras of coarse spaces. Arch. Math. (Basel), 84(6):519–526, 2005.
  • [8] Joachim Cuntz, Siegfried Echterhoff, Xin Li, and Guoliang Yu. KK-theory for group CC^{*}-algebras and semigroup CC^{*}-algebras, volume 47 of Oberwolfach Seminars. Birkhäuser/Springer, Cham, 2017.
  • [9] Eske Ellen Ewert and Ralf Meyer. Coarse geometry and topological phases. Comm. Math. Phys., 366(3):1069–1098, 2019.
  • [10] E. Christopher Lance. Hilbert CC^{*}-modules, volume 210 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1995.
  • [11] Kang Li, Ján Špakula, and Jiawen Zhang. Measured asymptotic expanders and rigidity for Roe algebras. Int. Math. Res. Not., (17):15102–15154, 2023.
  • [12] Paul S. Muhly, Jean N. Renault, and Dana P. Williams. Equivalence and isomorphism for groupoid CC^{\ast}-algebras. J. Operator Theory, 17(1):3–22, 1987.
  • [13] Piotr W. Nowak and Guoliang Yu. Large scale geometry. EMS Textbooks in Mathematics. European Mathematical Society (EMS), Zürich, 2012.
  • [14] Narutaka Ozawa. Amenable actions and exactness for discrete groups. C. R. Acad. Sci. Paris Sér. I Math., 330(8):691–695, 2000.
  • [15] Marc A. Rieffel. Morita equivalence for operator algebras. In Operator algebras and applications, Part 1 (Kingston, Ont., 1980), volume 38 of Proc. Sympos. Pure Math., pages 285–298. Amer. Math. Soc., Providence, RI, 1982.
  • [16] John Roe. An index theorem on open manifolds. I, II. J. Differential Geom., 27(1):87–113, 115–136, 1988.
  • [17] John Roe. Coarse cohomology and index theory on complete Riemannian manifolds, volume 497. Amer. Math. Soc., 1993.
  • [18] John Roe. Lectures on coarse geometry, volume 31 of University Lecture Series. Amer. Math. Soc., Providence, RI, 2003.
  • [19] Georges Skandalis, Jean-Louis Tu, and Guoliang Yu. The coarse Baum-Connes conjecture and groupoids. Topology, 41(4):807–834, 2002.
  • [20] Ján Špakula. Uniform KK-homology theory. J. Funct. Anal., 257(1):88–121, 2009.
  • [21] Ján Špakula and Rufus Willett. On rigidity of Roe algebras. Adv. Math., 249:289–310, 2013.
  • [22] Ján Špakula and Rufus Willett. A metric approach to limit operators. Trans. Amer. Math. Soc., 369(1):263–308, 2017.
  • [23] Qin Wang. Remarks on ghost projections and ideals in the Roe algebras of expander sequences. Arch. Math. (Basel), 89(5):459–465, 2007.
  • [24] Qin Wang and Jiawen Zhang. Ghostly ideals in uniform Roe algebras. arXiv preprint arXiv:2301.04921, 2023.
  • [25] Guoliang Yu. The coarse Baum-Connes conjecture for spaces which admit a uniform embedding into Hilbert space. Invent. Math., 139(1):201–240, 2000.