diff --git "a/K9E4T4oBgHgl3EQfJgxB/content/tmp_files/2301.04921v1.pdf.txt" "b/K9E4T4oBgHgl3EQfJgxB/content/tmp_files/2301.04921v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/K9E4T4oBgHgl3EQfJgxB/content/tmp_files/2301.04921v1.pdf.txt" @@ -0,0 +1,2734 @@ +arXiv:2301.04921v1 [math.OA] 12 Jan 2023 +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +QIN WANG AND JIAWEN ZHANG +Abstract. In this paper, we investigate the ideal structure of uniform Roe algebras +for general metric spaces beyond the scope of Yu’s property A. Inspired by the +ideal of ghost operators coming from expander graphs and in contrast to the +notion of geometric ideal, we introduce a notion of ghostly ideal in a uniform Roe +algebra, whose elements are locally invisible in certain directions at infinity. We +show that the geometric ideal and the ghostly ideal are respectively the smallest +and the largest element in the lattice of ideals with a common invariant open +subset of the unit space of the coarse groupoid by Skandalis-Tu-Yu, and hence the +study of ideal structure can be reduced to classifying ideals between the geometric +and the ghostly ones. As an application, we provide a concrete description for +the maximal ideals in a uniform Roe algebra in terms of the minimal points in the +Stone- ˇCech boundary of the space. We also provide a criterion to ensure that the +geometric and the ghostly ideals have the same K-theory, which helps to recover +counterexamples to the Baum-Connes type conjectures. Moreover, we introduce +a notion of partial Property A for a metric space to characterise the situation in +which the geometric ideal coincides with the ghostly ideal. +Mathematics Subject Classification (2020): 47L20, 46L80, 51F30. +Keywords: Uniform Roe algebras, Coarse groupoids, Geometric and ghostly ideals, Max- +imal ideals, Partial Property A +1. Introduction +Roe algebras are C∗-algebras associated to metric spaces, which encode the +coarse geometry of the underlying spaces. They were introduced by Roe in his pi- +oneering work on higher index theory [31], where he discovered that the K-theory +of Roe algebras serves as a receptacle for higher indices of elliptic differential oper- +ators on open manifolds. Hence the computation for the K-theory of Roe algebras +becomes crucial in the study of higher index theory, and a pragmatic and prac- +tical approach is to consult the Baum-Connes type conjectures [3, 4, 22]. There +is also a uniform version of the Roe algebra, which equally plays a key role in +higher index theory (see [39, 41]). Over the last four decades, there have been a +number of excellent works around this topic (e.g., [17, 21, 24, 46, 47]), which lead +to significant progresses in topology, geometry and analysis (see, e.g., [32, 33]). +On the other hand, the analytic structure of (uniform) Roe algebras reflects the +coarse geometry of the underlying spaces, and the rigidity problem asks whether +the coarse geometry of a metric space can be fully determined by the associated +(uniform) Roe algebra. This problem was initially studied by ˇSpakula and Willett +in [42], followed by a series of works in the last decade [5, 6, 7, 8, 25]. Recently +this problem is completely solved in the uniform case by the profound work +Date: January 13, 2023. +QW is partially supported by NSFC (No. 11831006, 12171156), and the Science and Technology +Commission of Shanghai Municipality (No. 22DZ2229014). JZ is supported by NSFC11871342. +1 + +2 +QIN WANG AND JIAWEN ZHANG +[2], which again highlights the importance of uniform Roe algebras in coarse +geometry. Meanwhile, uniform Roe algebras have also attained rapidly-growing +interest from researchers in mathematical physics, especially in the theory of +topological materials and topological insulators (see, e.g., [18] and the references +therein). +Due to their importance, Chen and the first-named author initiated the study +of the ideal structure for (uniform) Roe algebras [11, 12, 13, 14, 15, 44]. They +succeeded in obtaining a full description for the ideal structure of the uniform Roe +algebra when the underlying space has Yu’s Property A (see [12, 14]). However, +the general picture is far from clear beyond the scope of Property A. +In the present paper, we aim to provide a systematic study on the ideal structure +of uniform Roe algebras for general discrete metric spaces. To outline our main +results, let us first explain some notions. +Let (X, d) be a discrete metric space of bounded geometry (see Section 2.2 for +precise definitions). Thinking of operators on ℓ2(X) as X-by-X matrices, we say +that such an operator has finite propagation if the non-zero entries appear only in an +entourage of finite width (measured by the metric on X) around the main diagonal +(see Section 2.3 for full details). The set of all finite propagation operators forms +a ∗-subalgebra of B(ℓ2(X)), and its norm closure is called the uniform Roe algebra of +X and denoted by C∗ +u(X). +There is another viewpoint on the uniform Roe algebra based on groupoids. +Recall from [39] that Skandalis, Tu and Yu introduced a notion of coarse groupoid +G(X) associated to a discrete metric space X, and they succeeded in relating +coarse geometry to the theory of groupoids. The coarse groupoid G(X) is a lo- +cally compact, Hausdorff, ´etale and principal groupoid (see Section 2.5 for precise +definitions), and the unit space of G(X) coincides with the Stone- ˇCech compacti- +fication βX of X. Moreover, the uniform Roe algebra C∗ +u(X) can be interpreted as +the reduced groupoid C∗-algebra of G(X) (see also [34, Chapter 10]). +In [12], Chen and the first-named author concentrated on a class of ideals in the +uniform Roe algebra in which finite propagation operators therein are dense, and +they showed that these ideals can be described geometrically using the coarse +groupoid. More precisely, recall that a subset U ⊆ βX is invariant if any element +γ in G(X) with source in U also has its range in U (see Section 2.4). As shown +in [12] (see also Section 4), for any ideal I in C∗ +u(X) one can associate an invariant +open subset U(I) of βX, and conversely for any invariant open subset U ⊆ βX one +can associate an ideal I(U) in C∗ +u(X). Furthermore, these two procedures provide +a one-to-one correspondence between invariant open subsets of βX and ideals in +C∗ +u(X) in which finite propagation operators therein are dense. +Based on [12], the first-named author introduced the following notion in [44, +Definition 1.4]: +Definition A (Definition 4.4). Let (X, d) be a discrete metric space of bounded +geometry. An ideal I in the uniform Roe algebra C∗ +u(X) is called geometric if the set +of all finite propagation operators in I is dense in I. +As explained above, [12, Theorem 6.3] (see also Proposition 4.9) indicates that +the geometric ideals in C∗ +u(X) can be fully determined by invariant open subsets +of βX, which explains the terminology. Consequently, the geometric ideals in + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +3 +C∗ +u(X) are easy to handle and they must have the form of I(U) for some invariant +open subset U ⊆ βX, called the geometric ideal associated to U (see Definition 4.6). +Moreover, it follows from [14, Theorem 4.4] that all ideals in C∗ +u(X) are geometric +when X has Yu’s Property A. +However, things get complicated beyond the context of Property A. As noticed +in [12, Remark 6.5], when X comes from a sequence of expander graphs then the +ideal IG consisting of all ghost operators are not geometric (see also [21]). Recall +that an operator T ∈ B(ℓ2(X)) is a ghost if T ∈ C0(X × X) when regarding T as a +function on X × X. Ghost operators are introduced by Yu, and they are crucial to +provide counterexamples to the coarse Baum-Connes conjecture ([21]). +Direct calculations show that the associated invariant open subsets for IG and +for the ideal of compact operators in B(ℓ2(X)) are the same, both of which equal X +(see also Example 4.8 and 5.5). Hence for a general metric space X and an invariant +open subset U ⊆ βX, there might be more than one ideal I in the uniform Roe +algebra C∗ +u(X) satisfying U(I) = U. Therefore, the study of the ideal structure for +C∗ +u(X) can be reduced to analyse the lattice (where the order is given by inclusion) +(1.1) +IU := {I is an ideal in C∗ +u(X) : U(I) = U} +for each invariant open subset U ⊆ βX. +One of the main contributions of the present paper is to find the smallest and +the largest elements in the lattice IU. Following the discussions in [12], it is easy to +see that I(U(I)) ⊆ I for any ideal I in C∗ +u(X), which implies that the geometric ideal +I(U) is the smallest element in IU (see Proposition 4.10). To explore the largest +element, we have to include every ideal I in C∗ +u(X) with U(I) = U. Inspired by the +definition of U(I) (see Equality (4.1)), we introduce the following key notion: +Definition B (Definition 5.1). Let (X, d) be a discrete metric space of bounded +geometry and U be an invariant open subset of βX. The ghostly ideal associated to +U is defined to be +˜I(U) := {T ∈ C∗ +u(X) : r(suppε(T)) ⊆ U for any ε > 0}, +where suppε(T) := {(x, y) ∈ X ×X : |T(x, y)| ≥ ε} and r : X ×X → X is the projection +onto the first coordinate. +We show that ˜I(U) is indeed an ideal in the uniform Roe algebra C∗ +u(X) (see +Lemma 5.2) and moreover, we obtain the following desired result: +Theorem C (Theorem 5.4). Let (X, d) be a discrete metric space of bounded geometry +and U be an invariant open subset of βX. Then any ideal I in C∗ +u(X) with U(I) = U sits +between I(U) and ˜I(U). More precisely, the geometric ideal I(U) is the smallest element +while the ghostly ideal ˜I(U) is the largest element in the lattice IU in (1.1). +Theorem C draws the border of the lattice IU in (1.1), as an important step to +study the ideal structure of uniform Roe algebras for general metric spaces. More +precisely, once we can bust every ideal between I(U) and ˜I(U) for each invariant +open subset U ⊆ βX, then we will obtain a full description for the ideal structure +of the uniform Roe algebra C∗ +u(X). We pose it as an open question in Section 9 and +hope this will be done in some future work. +Concerning the ghostly ideal ˜I(U), we also provide an alternative picture in +terms of limit operators developed in [43], showing that ˜I(U) consists of operators + +4 +QIN WANG AND JIAWEN ZHANG +which vanish in the (βX \ U)-direction (see Proposition 5.6). +Note that ghost +operators vanish in all directions (see Corollary 5.7), and hence operators in ˜I(U) +can be regarded as “partial” ghosts, which clarifies its terminology. Thanks to +this viewpoint, we discover the deep reason behind the counterexample to the +conjecture in [12], constructed by the first-named author in [44, Section 3] (see +Example 5.10). +As an application, we manage to describe maximal ideals in the uniform Roe +algebra. More precisely, it follows directly from Theorem C that maximal ideals +correspond to minimal invariant closed subsets of the Stone- ˇCech boundary +∂βX := βX \ X. Moreover using the theory of limit spaces1 developed in [43], +we prove the following: +Proposition D (Proposition 6.2, Corollary 6.3 and Lemma 6.5). Let (X, d) be a +strongly discrete metric space of bounded geometry and I be a maximal ideal in the +uniform Roe algebra C∗ +u(X). Then there exists a point ω ∈ ∂βX such that I coincides with +the ghostly ideal ˜I(βX \ X(ω)), where X(ω) is the limit space of ω. +A point ω ∈ ∂βX satisfying the condition in Proposition D is called a minimal +point (see Definition 6.4). We show that there exist a number of non-minimal +points in the boundary even for the simple case of X = Z: +Theorem E (Theorem 6.8). For the integer group Z with the usual metric, there exist +non-minimal points in the boundary ∂βZ. More precisely, for any sequence {hn}n∈N in Z +tending to infinity such that |hn − hm| → +∞ when n + m → ∞ and n � m, and any +ω ∈ ∂βZ with ω({hn}n∈N) = 1, then ω is not a minimal point. +We provide two approaches to prove Theorem E. One is topological, which +makes use of several constructions and properties of ultrafilters (recalled in Ap- +pendix A). The other is C∗-algebraic, which replies on a description of maximal +ideals in terms of limit operators (see Lemma 6.13) together with a recent result +by Roch [30]. +Returning to the lattice IU defined in (1.1), we already notice that generally IU +consists of more than one element. Hence it will be interesting and important to +explore when IU has only a single element, or equivalently (thanks to Theorem +C), when the geometric ideal I(U) coincides with the ghostly ideal ˜I(U). +To study this problem, we start with an extra picture for geometric and ghostly +ideals using the associated groupoid C∗-algebras (see Lemma 4.7 and Proposition +5.9). Based on these descriptions, we show that the amenability of the restriction +G(X)∂βX\U of the coarse groupoid ensures that I(U) = ˜I(U) (see Proposition 7.2). +Meanwhile, we also discuss the K-theory of the geometric and ghostly ideals and +provide a criterion to ensure that K∗(I(U)) = K∗(˜I(U)) for ∗ = 0, 1: +Proposition F (Proposition 7.8). Let X be a discrete metric space of bounded geometry +which can be coarsely embedded into some Hilbert space. Then for any countably generated +invariant open subset U ⊆ βX, we have an isomorphism +(ιU)∗ : K∗(I(U)) −→ K∗(˜I(U)) +1Note that the theory of limit spaces and limit operators developed in [43] only concerns strongly +discrete metric spaces of bounded geometry (see Section 2.2 for precise definitions). Although as +noticed in [43] this will not lose any generality, we put this assumption to simplify proofs. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +5 +for ∗ = 0, 1, where ιU is the inclusion map. +Note that there is a technical condition of countable generatedness (see Def- +inition 7.4) used in Proposition F, which holds for a number of examples (see +Example 7.5) including X itself. However as shown in Example 7.11, there does +exist an invariant open subset which is not countably generated. Applying Propo- +sition F to the case of U = X, we partially recover [19, Proposition 35]. This is +crucial in the constructions of counterexamples to the Baum-Connes type conjec- +tures (see [39] for the coarse version and [20, Section 5] for the boundary version, +which is based on the example considered in [44, Section 3]). Hence presumably +Proposition F will find further applications in higher index theory. +Conversely, it is natural to ask whether I(U) = ˜I(U) implies that the restriction +groupoid G(X)∂βX\U is amenable. Note that when U = X, [36, Theorem 1.3] implies +that I(X) = ˜I(X) is equivalent to that X has Property A (see also Example 4.8 and +5.5), which is further equivalent to that the coarse groupoid G(X) is amenable +thanks to [39, Theorem 5.3]. Inspired by these works, we introduce the following +partial version of Property A: +Definition G (Definition 8.1). Let (X, d) be a discrete metric space of bounded +geometry and U ⊆ βX be an invariant open subset. We say that X has partial +Property A towards ∂βX \ U if G(X)∂βX\U is amenable. +Finally we reach the following, which recovers [36, Theorem 1.3] when U = X: +Theorem H (Theorem 8.16). Let (X, d) be a strongly discrete metric space of bounded +geometry and U ⊆ βX be a countably generated invariant open subset. Then the following +are equivalent: +(1) X has partial Property A towards βX \ U; +(2) ˜I(U) = I(U); +(3) the ideal IG of all ghost operators is contained in I(U). +The proof of Theorem H follows the outline of the case that U = X (cf. [36, +Theorem 1.3]), and is divided into several steps. Firstly, we unpack the groupoid +language of Definition G and provide a concrete geometric description similar to +the definition of Property A (see Proposition 8.2). Then we introduce a notion +of partial operator norm localisation property (Definition 8.9), which is a partial +version of the operator norm localisation property (ONL) introduced in [10]. +Parallel to Sako’s result that Property A is equivalent to ONL ([37]), we show that +partial Property A is equivalent to partial ONL (see Proposition 8.12). Finally +thanks to the assumption of countable generatedness, we conclude Theorem H. +We also remark that in the proof of Theorem H, we make use of the notion of +ideals in spaces introduced by Chen and the first-named author in [12] (see also +Definition 4.12) instead of using invariant open subsets of βX directly. This has the +advantage of playing within the given space rather than going to the mysterious +Stone- ˇCech boundary, which allows us to step over several technical gaps (see, +e.g., Remark 8.18). +The paper is organised as follows. In Section 2, we recall necessary background +knowledge in coarse geometry and groupoid theory. In Section 3, we recall the +theory of limit spaces and limit operators developed in [43], which will be an +important tool used throughout the paper. Section 4 is devoted to the notion + +6 +QIN WANG AND JIAWEN ZHANG +of geometric ideals (Definition A) studied in [12, 44], and we also discuss their +minimality in the lattice of ideals IU from (1.1). We introduce the key notion of +ghostly ideals (Definition B) in Section 5, prove Theorem C and provide several +characterisations for later use. Then we discuss maximal ideals in uniform Roe +algebras in Section 6, and prove Proposition D and Theorem E. In Section 7, we +study the problem when the geometric ideal coincides with the ghostly ideal, +discuss their K-theories and prove Proposition F. Then in Section 8, we introduce +the notion of partial Property A (Definition G) and prove Theorem H. Finally, +we list some open questions in Section 9, and provide Appendix A to record the +notion of ultrafilters and their properties used throughout the paper. +Acknowledgement. We would like to thank Baojie Jiang and J´an ˇSpakula for +some helpful discussions. +2. Preliminaries +2.1. Standard notation. Here we collect the notation used throughout the paper. +For a set X, denote by |X| the cardinality of X. For a subset A ⊆ X, denote by χA +the characteristic function of A, and set δx := χ{x} for x ∈ X. +When X is a locally compact Hausdorff space, we denote by C(X) the set of +complex-valued continuous functions on X, and by Cb(X) the subset of bounded +continuous functions on X. Recall that the support of a function f ∈ C(X) is the +closure of {x ∈ X : f(x) � 0}, written as supp(f), and denote by Cc(X) the set +of continuous functions with compact support. We also denote by C0(X) the set +of continuous functions vanishing at infinity, which is the closure of Cc(X) with +respect to the supremum norm ∥ f∥∞ := sup{| f(x)| : x ∈ X}. +When X is discrete, denote ℓ∞(X) := Cb(X) and ℓ2(X) the Hilbert space of +complex-valued square-summable functions on X. Denote by B(ℓ2(X)) the C∗- +algebra of all bounded linear operators on ℓ2(X), and by K(ℓ2(X)) the C∗-subalgebra +of all compact operators on ℓ2(X). +For a discrete space X, denote by βX its Stone- ˇCech compactification and ∂βX := +βX \ X the Stone- ˇCech boundary. +2.2. Notions from coarse geometry. Here we collect necessary notions from +coarse geometry, and guide readers to [27, 34] for more details. +For a discrete metric space (X, d), denote the closed ball by BX(x, r) := {y ∈ X : +d(x, y) ≤ r} for x ∈ X and r ≥ 0. +For a subset A ⊆ X and r > 0, denote the +r-neighbourhood of A in X by Nr(A) := {x ∈ X : dX(x, A) ≤ r}. For R > 0, denote the +R-entourage by ER := {(x, y) ∈ X × X : d(x, y) ≤ R}. +We saythat(X, d)hasbounded geometry ifforanyr > 0, the numbersupx∈X |BX(x, r)| +is finite. Also say that (X, d) is strongly discrete if the set {d(x, y) : x, y ∈ X} is a dis- +crete subset of R. +Convention. We say that “X is a space” as shorthand for “X is a strongly discrete +metric space of bounded geometry” (as in [43]) throughout the rest of this paper. +We remark that although our results hold without the assumption of strong +discreteness, we choose to add it so as to simplify the proofs. As discussed in [43, +Section 2], this will not lose any generality since one can always modify a discrete +metric space (using a coarse equivalence) to satisfy this assumption. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +7 +Now we recall the notion of Property A introduced by Yu (see, e.g., [45, Propo- +sition 1.2.4] for the equivalence to Yu’s original definition): +Definition 2.1 ([47]). A space (X, d) is said to have Property A if for any ε, R > 0 +there exist an S > 0 and a function f : X × X → [0, +∞) satisfying: +(1) supp(f) ⊆ ES; +(2) for any x ∈ X, we have � +z∈X f(z, x) = 1; +(3) for any x, y ∈ X with d(x, y) ≤ R, then � +z∈X | f(z, x) − f(z, y)| ≤ ε. +Using a standard normalisation argument, we have the following: +Lemma 2.2. A space (X, d) has Property A if and only if for any ε, R > 0 there exist an +S > 0 and a function f : X × X → [0, +∞) satisfying: +(1) supp(f) ⊆ ES; +(2) for any x ∈ X, we have | � +z∈X f(z, x) − 1| ≤ ε; +(3) for any x, y ∈ X with d(x, y) ≤ R, then � +z∈X | f(z, x) − f(z, y)| ≤ ε. +We also need a characterisation for Property A using kernels. Recall that a kernel +on X is a function k: X × X → R. We say that k is of positive type if for any n ∈ N, +x1, . . . , xn ∈ X and λ1, . . . , λn ∈ R, we have: +n +� +i,j=1 +λiλjk(xi, xj) ≥ 0. +The following is well-known (see, e.g., [45, Proposition 1.2.4]): +Lemma 2.3. A space (X, d) has Property A if and only if for any R > 0 and ε > 0, there +exist S > 0 and a kernel k : X × X → R of positive type satisfying the following: +(1) for x, y ∈ X, we have k(x, y) = k(y, x) and k(x, x) = 1; +(2) for x, y ∈ X with d(x, y) ≥ S, we have k(x, y) = 0; +(3) for x, y ∈ X with d(x, y) ≤ R, we have |1 − k(x, y)| ≤ ε. +We also recall the notion of coarse embedding: +Definition 2.4. Let (X, dX) and (Y, dY) be metric spaces and f : X → Y be a map. +We say that f is a coarse embedding if there exist functions ρ± : [0, ∞) → [0, ∞) with +limt→+∞ ρ±(t) = +∞ such that for any x, y ∈ X we have +ρ−(dX(x, y)) ≤ dY(f(x), f(y)) ≤ ρ+(dX(x, y)). +If additionally there exists C > 0 such that Y = NC(f(X)), then we say that f is a +coarse equivalence and (X, dX), (Y, dY) are coarsely equivalent. +2.3. Uniform Roe algebras. Let (X, d) be a discrete metric space. Each operator +T ∈ B(ℓ2(X)) can be written in the matrix form T = (T(x, y))x,y∈X, where T(x, y) = +⟨Tδy, δx⟩ ∈ C. We also regard T ∈ B(ℓ2(X)) as a bounded function on X × X, i.e., +an element in ℓ∞(X × X). Denote by ∥T∥ the operator norm of T in B(ℓ2(X)), and +∥T∥∞ the supremum norm when regarding T as a function in ℓ∞(X × X). It is clear +that ∥T∥∞ ≤ ∥T∥ for any T ∈ B(ℓ2(X)). +Given an operator T ∈ B(ℓ2(X)), we define the support of T to be +supp(T) := {(x, y) ∈ X × X : T(x, y) � 0}, + +8 +QIN WANG AND JIAWEN ZHANG +and the propagation of T to be +prop(T) := sup{d(x, y) : (x, y) ∈ supp(T)}. +Definition 2.5. Let (X, d) be a space. +(1) The set of all finite propagation operators in B(ℓ2(X)) forms a ∗-algebra, +called the algebraic uniform Roe algebra of X and denoted by Cu[X]. For each +R ≥ 0, denote the subset +CR +u[X] := {T ∈ B(ℓ2(X)) : prop(T) ≤ R}. +It is clear that Cu[X] = � +R≥0 CR +u[X]. +(2) The uniform Roe algebra of X is defined to be the operator norm closure of +Cu[X] in B(ℓ2(X)), which forms a C∗-algebra and is denoted by C∗ +u(X). +The following notion was originally introduced by Yu: +Definition 2.6. An operator T ∈ C∗ +u(X) is called a ghost if T ∈ C0(X × X) when +regarding T as a function in ℓ∞(X × X). In other words, for any ε there exists a +finite subset F ⊆ X such that for any (x, y) � F × F, we have |T(x, y)| < ε. +It is easy to see that all the ghost operators in C∗ +u(X) form an ideal in C∗ +u(X), +denoted by IG. Intuitively speaking, a ghost operator is locally invisible at infinity +in all directions. This will be made more precise in the sequel. +2.4. Groupoids and C∗-algebras. We collect here some basic notions and termi- +nology on groupoids. Details can be found in [28], or [38] in the ´etale case. +Recall that a groupoid is a small category, in which every morphism is invertible. +More precisely, a groupoid consists of a set G, a subset G(0) called the unit space, two +maps s, r : G → G(0) called the source and range maps respectively, a composition +law: +G(2) := {(γ1, γ2) ∈ G × G : s(γ1) = r(γ2)} ∋ (γ1, γ2) �→ γ1γ2 ∈ G, +and an inverse map γ �→ γ−1. These operationssatisfya couple ofaxioms, including +associativity law and the fact that elements in G(0) act as units. +For x ∈ G(0), denote Gx := r−1(x) and Gx := s−1(x). For Y ⊆ G(0), denote GY +Y := +r−1(Y) ∩ s−1(Y). Note that GY +Y is a subgroupoid of G (in the sense that it is stable +under multiplication and inverse), called the reduction of G by Y. A subset Y is +said to be invariant if r−1(Y) = s−1(Y), and we write GY instead of GY +Y in this case. +A locally compact Hausdorff groupoid is a groupoid equipped with a locally +compact and Hausdorff topology such that the structure maps (composition and +inverse) are continuous with respect to the induced topologies. Such a groupoid +is called ´etale (also called r-discrete) if the range (hence the source) map is a local +homeomorphism. Clearly in this case, each fibre Gx (and Gx) is discrete with the +induced topology, and G(0) is clopen in G. The notion of ´etaleness for a groupoid +can be regarded as an analogue of discreteness in the group case. +Example 2.7. Let X be a set. The pair groupoid of X is X×X as a set, whose unit space +is {(x, x) ∈ X × X : x ∈ X} and identified with X for simplicity. The source map is +the projection onto the second coordinate and the range map is the projection onto +the first coordinate. The composition is given by (x, y)· (y, z) = (x, z) for x, y, z ∈ X. +When X is a discrete Hausdorff space, then X × X is a locally compact Hausdorff +´etale groupoid. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +9 +Now we introduce the algebras associated to groupoids. Here we only focus +on the case of ´etaleness, and guide readers to [28] for the general case. +Let G be a locally compact, Hausdorff and ´etale groupoid with unit space G(0). +Note that the space Cc(G) is a ∗-involutive algebra with respect to the following +operations: for f, g ∈ Cc(G), +(f ∗ g)(γ) += +� +α∈Gs(γ) +f(γα−1)g(α), +f ∗(γ) += +f(γ−1). +Consider the following algebraic norm on Cc(G) defined by: +∥ f∥I := max +sup +x∈G(0) +� +γ∈Gx +| f(γ)|, sup +x∈G(0) +� +γ∈Gx +| f ∗(γ)| + . +The completion of Cc(G) with respect to the norm ∥ · ∥I is denoted by L1(G). +The maximal (full) groupoid C∗-algebra C∗ +max(G) is defined to be the completion of +Cc(G) with respect to the norm: +∥ f∥max := sup ∥π(f)∥, +where the supremum is taken over all ∗-representations π of L1(G). +In order to define the reduced counterpart, we recall that for each x ∈ G(0) the +regular representation at x, denoted by λx : Cc(G) → B(ℓ2(Gx)), is defined as follows: +(2.1) +� +λx(f)ξ +� +(γ) := +� +α∈Gx +f(γα−1)ξ(α), +where f ∈ Cc(G) and ξ ∈ ℓ2(Gx). +It is routine work to check that λx is a well-defined ∗-homomorphism. The reduced +norm on Cc(G) is +∥ f∥r := sup +x∈G(0) +∥λx(f)∥, +and the reduced groupoid C∗-algebra C∗ +r(G) is defined to be the completion of the +∗-algebra Cc(G) with respect to this norm. Clearly, each regular representation λx +can be extended to a homomorphism λx : C∗ +r(G) → B(ℓ2(Gx)) automatically. It +is also routine to check that there is a canonical surjective homomorphism from +C∗ +max(G) to C∗ +r(G). +2.5. Coarse groupoids. Let (X, d) be a space as in Section 2.2. The coarse groupoid +G(X) on X was introduced by Skandalis, Tu and Yu in [39] (see also [34, Chapter +10]) to relate coarse geometry to the theory of groupoids. As a topological space, +G(X) := +� +r>0 +Er +β(X×X) ⊆ β(X × X). +Recall from Example 2.7 that X × X is the pair groupoid with source and range +maps s(x, y) = y and r(x, y) = x. These maps extend to maps G(X) → βX, still +denoted by r and s. +Now consider (r, s) : G(X) → βX × βX. It was shown in [39, Lemma 2.7] that the +map (r, s) is injective, and hence G(X) can be endowed with a groupoid structure + +10 +QIN WANG AND JIAWEN ZHANG +induced by the pair groupoid βX × βX, called the coarse groupoid of X. Therefore, +G(X) can also be equivalently defined by +G(X) := +� +r>0 +Er +βX×βX ⊆ βX × βX, +with the weak topology. It was also shown in [39, Proposition 3.2] that the coarse +groupoid G(X) is locally compact, Hausdorff, ´etale and principal. Clearly, the unit +space of G(X) can be identified with βX. +Given f ∈ Cc(G(X)), then f is a continuous function supported on Er for some +r > 0; equivalently, we can interpret f as a bounded function on Er. Hence we +define an operator θ(f) on ℓ2(X) by setting its matrix coefficients to be θ(f)(x, y) := +f(x, y) for x, y ∈ X. We have the following: +Proposition 2.8 ([34, Proposition 10.29]). The map θ provides a ∗-isomorphism from +Cc(G(X)) to Cu[X], and extends to a C∗-isomorphism Θ : C∗ +r(G(X)) → C∗ +u(X). Note that +Θ maps the C∗-subalgebra C∗ +r(X × X) onto the compact operators K(ℓ2(X)). +Recall from Section 2.3 that an operator T ∈ B(ℓ2(X)) can be regarded as an +element in ℓ∞(X × X). +Following the notation from [12], we denote by T the +continuous extension of T on β(X × X) when regarding T ∈ ℓ∞(X × X). Then +supp(T) = supp(T). +Note that G(X) is open in β(X × X), hence C0(G(X)) is a subalgebra in C(β(X × X)). +Restricting to G(X), we also regard T as a function on G(X) and hence we can talk +of the value T(α, γ) for (α, γ) ∈ G(X). Moreover, we have the following: +Lemma 2.9. For T ∈ Cu[X], we have T ∈ Cc(G(X)) and θ(T) = T. For T ∈ C∗ +u(X), we +have T ∈ C0(G(X)) and Θ(T) = T. +Proof. The first statement is a direct corollary of Proposition 2.8. Since ∥T∥∞ ≤ ∥T∥ +for any T ∈ B(ℓ2(X)), the second follows from the first. +□ +2.6. Amenability and a-T-menability for groupoids. Amenable groupoids com- +prise a large class of groupoids with relatively nice properties, which are literally +the analogue of amenable groups in the world of groupoids. Here we only focus +on the case of ´etaleness, in which the notion of amenability behaves quite well. +A standard reference is [1] and another reference for just ´etale groupoids is [9, +Chapter 5.6]. +Definition 2.10 ([1]). A locally compact, Hausdorff and ´etale groupoid G is said +to be (topologically) amenable if for any ε > 0 and compact K ⊆ G, there exists +f ∈ Cc(G) with range in [0, 1] such that for any γ ∈ K we have +� +α∈Gr(γ) +f(α) = 1 +and +� +α∈Gr(γ) +| f(α) − f(αγ)| < ε. +Similar to the case of Property A, we have the following by a standard normal- +isation argument: + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +11 +Lemma 2.11. A locally compact, Hausdorff and ´etale groupoid G is amenable if and +only if for any ε > 0 and compact K ⊆ G, there exists f ∈ Cc(G) with range in [0, +∞) +such that for any γ ∈ K we have +��� +� +α∈Gr(γ) +f(α) − 1 +��� < ε +and +� +α∈Gr(γ) +| f(α) − f(αγ)| < ε. +Amenability for ´etale groupoids enjoy similar permanence properties as in the +case of groups. For example, open or closed subgroupoids of amenable ´etale +groupoids are amenable, and amenability is preserved under taking groupoid +extensions. See [1, Section 5] for details. Also recall that we have the following: +Proposition 2.12 ([9, Corollary 5.6.17]). Let G be a locally compact, Hausdorff, ´etale +and amenable groupoid. Then the natural quotient C∗ +max(G) → C∗ +r(G) is an isomorphism. +Now we recall the notion of a-T-menability for groupoids introduced by Tu +[40]. Let G be a locally compact, Hausdorff and ´etale groupoid. A continuous +function f : G → R is said to be of negative type if +(1) f|G(0) = 0; +(2) for any γ ∈ G, f(γ) = f(γ−1); +(3) Given γ1, · · · , γn ∈ G with the same range and λ1, · · · , λn ∈ R with �n +i=1 λi = +0, we have � +i,j λiλj f(γ−1 +i γj) ≤ 0. +A continuous function f : G → R is called locally proper if for any compact subset +K ⊆ G(0), the restriction of f on GK +K is proper. +Definition 2.13 ([40, Section 3.3]). A locallycompact, Hausdorff and ´etale groupoid +G is said to be a-T-menable if there exists a continuous locally proper function +f : G → R of negative type on G. +Analogous to the case of groups, Tu [40] proved that a locally compact, σ- +compact, Hausdorff and ´etale groupoid G is a-T-menable if and only if there exists +a continuous field of Hilbert spaces over G(0) with a proper affine action of G. We +also need the following significant result by Tu: +Proposition 2.14 ([40, Th´eor`eme 0.1]). Let G be a locally compact, σ-compact, Haus- +dorff, ´etale and a-T-menable groupoid. Then G is K-amenable, i.e., the quotient map +induces an isomorphism K∗(C∗ +max(G)) → K∗(C∗ +r(G)) for ∗ = 0, 1. +Finally, we record the following result for coarse groupoids: +Proposition 2.15 ([39, Theorem 5.3 and 5.4]). Let (X, d) be a space and G(X) be the +associated coarse groupoid. Then: +(1) X has Property A if and only if G(X) is amenable; +(2) X can be coarsely embedded into Hilbert space if and only if G(X) is a-T-menable. +3. Limit spaces and limit operators +In this section, we recall the theory of limit spaces and limit operators for metric +spaces developed by ˇSpakula and Willett in [43], which becomes an important +tool for later use. + +12 +QIN WANG AND JIAWEN ZHANG +Throughout the section, we always assume that (X, d) is a space (see “Conven- +tion” in Section 2.2) and G(X) is its coarse groupoid. We will freely use the notion +of ultrafilters on X, and related materials are recalled in Appendix A. +3.1. Limit spaces. First recall that a function t : D → R with D, R ⊆ X is called a +partial translation if t is a bijection from D to R, and supx∈X d(x, t(x)) is finite. The +graph of t is {(t(x), x) : x ∈ D}, denoted by gr(t). It is well-known that each entourage +E on X can be decomposed into finitely many graphs of partial translations (see, +e.g., [34, Lemma 4.10]) thanks to the bounded geometry of X. +Definition 3.1 ([43, Definition 3.2 and 3.6]). Fix an ultrafilter ω ∈ βX. A partial +translation t : D → R on X is compatible with ω if ω(D) = 1. In this case, regarding +t as a function from D to βX, we define the following thanks to Lemma A.2: +t(ω) := lim +ω t ∈ βX. +In other words, consider the extension t : D → R then t(ω) = t(ω). +An ultrafilter α ∈ βX is compatible with ω if there exists a partial translation t +compatible with ω and t(ω) = α. Denote by X(ω) the collection of all ultrafilters on +X compatible with ω. A compatible family for ω is a collection of partial translations +{tα}α∈X(ω) such that each tα is compatible with ω and tα(ω) = α. +Fix an ultrafilter ω on X, and a compatible family {tα}α∈X(ω). Define a function +dω : X(ω) × X(ω) → [0, ∞) by +dω(α, β) := lim +x→ω d(tα(x), tβ(x)). +It is shown in [43, Proposition 3.7] that dω is a uniformly discrete metric of bounded +geometry on X(ω) which does not depend on the choice of {tα}. +This leads to the following: +Definition 3.2 ([43, Definition 3.8]). For each non-principal ultrafilter ω on X, the +metric space (X(ω), dω) is called the limit space of X at ω, which is a space in the +sense of “Convention” in Section 2.2. +It is shown in [43, Proposition 3.9] that for any α ∈ X(ω), we have X(α) = X(ω) +as metric spaces. Also note that when ω is principal, i.e., ω ∈ X, then it is clear +that (X(ω), dω) = (X, d). +We recall the following result, which reveals that the local geometry of X can +be recaptured by those of the limit spaces. +Proposition 3.3 ([43, Proposition 3.10]). Let ω be a non-principal ultrafilter on X, and +{tα : Dα → Rα} a compatible family for ω. Then for each finite F ⊆ X(ω), there exists a +subset Y ⊆ X with ω(Y) = 1 such that for each y ∈ Y, there is a finite subset G(y) ⊆ X +such that the map +fy : F → G(y), α �→ tα(y) +is a surjective isometry. Such a collection {fy}y∈Y is called a local coordinate systerm +for F, and the maps fy are called local coordinates. +Furthermore, if F is a metric ball B(ω, r), then there exist Y ⊆ X with ω(Y) = 1 and a +local coordinate system {fy : F → G(y)}y∈Y such that each G(y) is the ball B(y, r). +As shown in [43, Appendix C], limit spaces can be described in terms of the +coarse groupoid G(X): + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +13 +Lemma 3.4 ([43, Lemma C.3]). Given a non-principal ultrafilter ω ∈ βX, the map +F : X(ω) → G(X)ω, +α �→ (α, ω) +is a bijection. Hence X(ω) is the smallest invariant subset of βX containing ω. Here we +consider G(X) as a subset of βX × βX, as explained in Section 2.5. +Consequently, we obtain the following: +Corollary 3.5. As a set, we have +G(X) = +� +X × X +� +⊔ +� +ω∈∂βX +� +X(ω) × X(ω) +� +. +Now we would like to provide a quantitative version for Corollary 3.5. First +we record the following observation, whose proof is straightforward and almost +identical to that of Lemma 3.4 (originally from [34, discussion in 10.18-10.24]): +Lemma 3.6. Let t : D → R be a partial translation on X. Then we have: +gr(t) +βX×βX = gr(t) ⊔ +� +ω∈∂βX +� +(α, ω) : ω(D) = 1 and α = t(ω) +� +. +In general, we have the following: +Lemma 3.7. For any S ≥ 0, we have: +ES +βX×βX = ES ⊔ +� +ω∈∂βX +� +(α, γ) ∈ X(ω) × X(ω) : dω(α, γ) ≤ S +� +. +Proof. As explained at the beginning of this subsection, we can decompose +ES = gr(t1) ⊔ · · · ⊔ gr(tN) +where each ti : Di → Ri is a partial translation. Hence we have +ES = gr(t1) ∪ · · · ∪ gr(tN). +Applying Lemma 3.6, we obtain that ES is contained in the right hand side in the +statement. +On the other hand, given ω ∈ ∂βX and α, γ ∈ X(ω) with dω(α, γ) ≤ S, then we +have (X(ω), dω) = (X(γ), dγ). Take a partial translation t : D → R such that γ(D) = 1 +and α = t(γ). Note that +dγ(α, γ) = lim +x→γ d(t(x), x) ≤ S. +Hence for D′ := {x ∈ D : d(t(x), x) ≤ S}, we have γ(D′) = 1. Consider the restriction +of t on D′, denoted by t′. Then t′ is also a partial translation and α = t′(γ). By +Lemma 3.6, we obtain that (α, γ) ∈ gr(t′), which is contained in ES as desired. +□ +Now we compute concrete examples of limit spaces. First we recall the case +of groups from [43, Appendix B]. Let Γ be a countable discrete group, equipped +with a left-invariant bounded geometry and strongly discrete metric d. For each +g ∈ Γ, denote +ρg : Γ → Γ, h �→ hg +the right translation map. Each ρg is a partial translation with full domain, and +hence is compatible with every ω ∈ βΓ. Moreover, we have the following: + +14 +QIN WANG AND JIAWEN ZHANG +Lemma 3.8 ([43, Lemma B.1]). For each non-principal ultrafilter ω ∈ βΓ, the map +bω : Γ −→ Γ(ω), g �→ ρg(ω) +is an isometric bijection. +Inspired by Lemma 3.8, we provide the following general method: +Lemma 3.9. Let {tλ : Dλ → Rλ}λ∈Λ be a family of partial translations on X satisfying +the following: for each S > 0, there exists a finite subset ΛS ⊆ Λ such that gr(tλ) ∩ gr(tµ) +is finite for λ � µ in ΛS and ES \ +� � +λ∈ΛS gr(tλ) +� +is finite. Then for any non-principal +ultrafilter ω on X and α ∈ X(ω), there exists λ ∈ Λ such that ω(Dλ) = 1 and α = tλ(ω). +Proof. By definition, we assume that α = t(ω) for some partial translation t : D → R +with ω(D) = 1. For λ ∈ Λ, set �Dλ := {x ∈ D ∩ Dλ : t(x) = tλ(x)}. Choose S > 0 such +that gr(t) ⊆ ES, and hence there exists a finite subset F ⊆ ES such that +gr(t) ⊆ +� � +λ∈ΛS +gr(tλ) +� +⊔ F. +This implies that D ⊆ +� � +λ∈ΛS �Dλ +� +⊔ F′ for some finite F′ ⊆ X. Since ω is non- +principal, ΛS is finite and �Dλ ∩ �Dµ is finite for any λ, µ ∈ ΛS, there exists a unique +λ ∈ ΛS such that ω(�Dλ) = 1. This implies that ω(Dλ) = 1 and α = tλ(ω), which +concludes the proof. +□ +Back to the case of the group Γ, the set {ρg : Γ → Γ}g∈Γ satisfies the condition in +Lemma 3.9, and hence the map bω in Lemma 3.8 is surjective. It is straightforward +to check that bω is isometric, which recovers the proof for Lemma 3.8. +Example 3.10. Consider X = N with the usual metric. For each k ∈ Z with k ≥ 0, +define a partial translation +ρk : N −→ N, +n �→ n + k. +For k ∈ Z with k < 0, define a partial translation +ρk : [−k, ∞) ∩ N −→ N, +n �→ n + k. +Then it is clear that the set {ρk}k∈Z satisfies the condition in Lemma 3.9. Now for a +non-principal ultrafilter ω on N, consider the map +bω : Z −→ N(ω), +k �→ ρk(ω). +Note that for k, l ∈ Z, we have +dω(ρk(ω), ρl(ω)) = lim +n→ω |(k + n) − (l + n)| = |k − l|, +which implies that bω is isometric. Moreover, Lemma 3.9 shows that bω is surjec- +tive. Therefore, every limit space of N is isometric to Z. This provides a detailed +proof for [43, Example 3.14(2)]. +Similar to the analysis in Example 3.10, we can also apply Lemma 3.9 to obtain +proofs for [43, Example 3.14(3)-(5)]. Details are left to readers. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +15 +3.2. Limit operators. Now we recall the notion of limit operators for metric spaces +introduced by ˇSpakula and Willett: +Definition 3.11 ([43, Definition 4.4]). For a non-principal ultrafilter ω on X, fix a +compatible family {tα}α∈X(ω) for ω and let T ∈ C∗ +u(X). The limit operator of T at ω, +denoted by Φω(T), is an X(ω)-by-X(ω) indexed matrix defined by +Φω(T)αγ := lim +x→ω Ttα(x)tγ(x) +for +α, γ ∈ X(ω). +It was studied in [43, Chapter 4] that the above definition does not depend on +the choice of the compatible family {tα}α∈X(ω) for ω. Furthermore, the limit operator +Φω(T) is indeed a bounded operator on ℓ2(X(ω)), and belongs to the uniform Roe +algebra C∗ +u(X(ω)). +Recall from Lemma 2.9 that for an operator T ∈ C∗ +u(X), the continuous extension +T ∈ C0(G(X)). We have the following, which was implicitly mentioned in the +proof of [43, Lemma C.3]. +Lemma 3.12. For a non-principal ultrafilter ω on X and T ∈ C∗ +u(X), we have +Φω(T)αγ = T(α, γ) +for +α, γ ∈ X(ω). +Proof. Choose partial translations tα, tγ compatible with α, γ such that tα(ω) = α, +tγ(ω) = γ. By definition, we have +Φω(T)αγ = lim +x→ω Ttα(x)tγ(x) = lim +x→ω T(tα(x), tγ(x)) = T(α, γ) +where the last equality comes from the discussion before Lemma 2.9. +□ +Since the limit operator Φω(T) contains the information of the asymptotic be- +haviour of T “in the ω-direction”, we introduce the following: +Definition 3.13. For an ω ∈ ∂βX, we say that an operator T ∈ C∗ +u(X) is locally +invisible (or vanishes) in the ω-direction if Φω(T) = 0. For a subset V ⊆ ∂βX, we say +that T is locally invisible (or vanishes) in the V-direction if Φω(T) = 0 for any ω ∈ V. +Finally, we recall from Proposition 2.8 that there is a C∗-isomorphism Θ : +C∗ +r(G(X)) → C∗ +u(X). This allows us to relate limit operators to left regular rep- +resentations of C∗ +r(G(X)): +Lemma 3.14 ([43, Lemma C.3]). For a non-principal ultrafilter ω on X, let Wω : +ℓ2(G(X)ω) → ℓ2(X(ω)) be the unitary representation induced by F in Lemma 3.4. Then +we have the following commutative diagram: +C∗ +r(G(X)) +λω +� +Θ � +� +B(ℓ2(G(X)ω)) +AdWω +� +� +C∗ +u(X) +Φω +� B(ℓ2(X(ω))), +where λω is the left regular representation from (2.1). + +16 +QIN WANG AND JIAWEN ZHANG +4. Geometric ideals +In this section, we recall the notion of geometric ideals, which was originally +introduced by the first-named author in [44] (see also [12]). +Throughout the +section, let X be a space in the sense of “Convention” in Section 2.2. +Definition 4.1 ([12, Definition 3.1, 3.3]). For an operator T ∈ C∗ +u(X) and ε > 0, the +ε-support of T is defined to be +suppε(T) := {(x, y) ∈ X × X : |T(x, y)| ≥ ε}. +Also define the ε-truncation of T to be +Tε(x, y) := +� T(x, y), +if |T(x, y)| ≥ ε; +0, +otherwise. +It is clear that supp(Tε) = suppε(T). We also record the following elementary +result for later use. The proof is straightforward, hence omitted. +Lemma 4.2. Given T ∈ C∗ +u(X) and ε > 0, we have +supp(Tε) ⊆ { ˜ω ∈ β(X × X) : |T( ˜ω)| ≥ ε} ⊆ supp(Tε/2). +The following is a key result in [12]: +Proposition 4.3 ([12, Theorem 3.5]). Let I be an ideal in the uniform Roe algebra C∗ +u(X). +For each T ∈ I and ε > 0, we have Tε ∈ I ∩ Cu[X]. Moreover, we have +I ∩ Cu[X] = {Tε : T ∈ I, ε > 0} +where the closures are taken with respect to the operator norm. +Now we recall the notion of geometric ideals from [44]: +Definition 4.4. An ideal I in the uniform Roe algebra C∗ +u(X) is called geometric if +I ∩ Cu[X] is dense in I. +In [12], Chen and the first-named author provide a full description for geometric +ideals in C∗ +u(X) in terms of invariant open subsets of G(X)(0) = βX. To outline their +work, let us start with the following elementary observation: +Lemma 4.5. Let U be a non-empty invariant open subset of βX, then X ⊆ U. +Proof. Since X is dense in βX and U is open and non-empty, we obtain that +U ∩ X � ∅. Take an x ∈ U ∩ X, then the pair (x, y) ∈ G(X) for any y ∈ X. Thanks to +the invariance of U, we obtain that y ∈ U. This implies that X ⊆ U. +□ +Given an invariant open subset U ⊆ βX, denote G(X)U := G(X) ∩ s−1(U). Fol- +lowing [12], we define +Ic(U) : = {f ∈ Cc(G(X)) : f( ˜ω) = 0 for any ˜ω � G(X)U} += {T ∈ Cu[X] : T( ˜ω) = 0 for any ˜ω � G(X)U}. +Obviously, Ic(U) is a two-sided ideal in Cc(G(X)). Denote its closure in C∗ +r(G(X)) +by I(U), which is a geometric ideal in C∗ +r(G(X)) � C∗ +u(X) from the definition (see +also [12, Lemma 5.1]). This leads to the following: + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +17 +Definition 4.6. For an invariant open subset U ⊆ βX, the ideal I(U) is called the +geometric ideal associated to U. +For later use, we record the following alternative description for the geometric +ideal I(U): +Lemma 4.7. Let U be an invariant open subset of βX. Then the ideal I(U) is isomorphic +to the reduced groupoid C∗-algebra C∗ +r(G(X)U). +Proof. This was implicitly contained in the proof of [12, Proposition 5.5]. +For +convenience to the readers, we include a proof here. By definition, C∗ +r(G(X)U) is +isomorphic to the norm closure of Cc(G(X)U) in C∗ +r(G(X)). Note that +Cc(G(X)U) = {T ∈ Cu[X] : supp(T) ⊆ G(X)U} ⊆ Ic(U). +On the other hand, for T ∈ Ic(U) we have T = limε→0 Tε since T has finite propaga- +tion. Note that supp(Tε) ⊆ G(X)U, which implies Tε ∈ Cc(G(X)U). Hence we obtain +that Cc(G(X)U) and Ic(U) have the same closure in C∗ +r(G(X)), which concludes the +proof. +□ +Example 4.8. For U = X, it follows directly from definition that G(X)X = X × X. +Hence combining Proposition 2.8 and Lemma 4.7, we obtain that the geometric +ideal associated to X is I(X) = K(ℓ2(X)). On the other hand, for U = βX it is clear +that I(βX) = C∗ +u(X). +Conversely, following [12, Section 4 and 5] we can associate an invariant open +subset of βX to any ideal in the uniform Roe algebra. More precisely, let I be an +ideal in the uniform Roe algebra C∗ +u(X). Define: +(4.1) +U(I) := +� +T∈I,ε>0 +r(suppε(T)) = +� +T∈I∩Cu[X],ε>0 +r(suppε(T)), +where the second equality follows directly from Proposition 4.3. Also [12, Lemma +5.2] implies that U(I) is an invariant open subset of βX. Furthermore, as a special +case of [12, Theorem 6.3], we have the following: +Proposition 4.9. For a space (X, d), the map I �→ U(I) provides an isomorphism between +the lattice of all geometric ideals in C∗ +u(X) and the lattice of all invariant open subsets of +βX, with the inverse map given by U �→ I(U). +Proposition 4.9 shows that geometric ideals in C∗ +u(X) can be fully determined +by invariant open subsets of βX. In contrast, general ideals in C∗ +u(X) cannot be +characterised merely by the associated subsets of βX. For example, direct calcu- +lations show that the associated invariant open subsets for the ideal IG defined in +Section 2.3 and for the ideal of compact operators in B(ℓ2(X)) are the same, both +of which equal X (see also Example 4.8 and 5.5 below). +Hence as pointed out in Section 1, the study of the ideal structure for the +uniform Roe algebra can be reduced to analyse the lattice (where the order is +given by inclusion) +IU = {I is an ideal in C∗ +u(X) : U(I) = U} +in (1.1) for each invariant open subset U ⊆ βX. The following result busts the +smallest element in IU: + +18 +QIN WANG AND JIAWEN ZHANG +Proposition 4.10. Let U be an invariant open subset of βX. Then the geometric ideal +I(U) is the smallest element in the lattice IU in (1.1). +The proof of Proposition 4.10 follows directly from the following lemma: +Lemma 4.11. Let (X, d) be a space and I an ideal in C∗ +u(X). Then we have +I(U(I)) = I ∩ Cu[X], +where the closure is taken in C∗ +u(X). Hence we have I(U(I)) ⊆ I. +Proof. Denoting ˚I := I ∩ Cu[X], it is clear that ˚I ∩ Cu[X] = I ∩ Cu[X]. This implies +that ˚I is a geometric ideal in C∗ +u(X), and hence ˚I = I(U(˚I)) by Proposition 4.9. By +definition, we have +U(˚I) = +� +T∈˚I∩Cu[X],ε>0 +r(suppε(T)) = +� +T∈I∩Cu[X],ε>0 +r(suppε(T)) = U(I). +Therefore, we obtain that I(U(I)) = I(U(˚I)) = ˚I = I ∩ Cu[X] as required. +□ +We would like to recall another description for geometric ideals based on the +notion of ideals in spaces introduced in [12]. It has the advantage of playing +within the given metric space, rather than going to the mysterious Stone- ˇCech +boundary, and hence will help us to step over several technical gaps in Section 8. +Definition 4.12 ([12, Definition 6.1]). An ideal in a space (X, d) is a collection L of +subsets of X satisfying the following: +(1) if Y ∈ L and Z ⊆ Y, then Z ∈ L; +(2) if R ≥ 0 and Y ∈ L, then NR(Y) ∈ L; +(3) if Y, Z ∈ L, then Y ∪ Z ∈ L. +For an ideal L in X, we define +U(L) := +� +Y∈L +Y +βX. +Conversely, given an invariant open subset U of βX, we define +L(U) := {Y ⊆ X : Y +βX ⊆ U}. +As a special case of [12, Theorem 6.3], we have the following: +Proposition 4.13. For a space (X, d), the map L �→ U(L) provides an isomorphism +between the lattice of all ideals in X and the lattice of all invariant open subsets of βX, +with the inverse map given by U �→ L(U). +Combining Proposition 4.9 and 4.13, we obtain an isomorphism between the +lattice of all ideals in X and the lattice of all geometric ideals in C∗ +u(X). Direct +calculation shows (see also [12, Theorem 6.4]): +(4.2) +I(U(L)) = {T ∈ Cu[X] : supp(T) ⊆ Y × Y for some Y ∈ L}, +where the closure is taken in C∗ +u(X). +Now we consider a special class of geometric ideals coming from subspaces. +Given a subspace A ⊆ X, recall from [23, Section 5] that there is an associated +ideal IA in C∗ +u(X) whose K-theory is isomorphic to that of the uniform Roe algebra + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +19 +C∗(A). More precisely, recall that an operator T ∈ B(ℓ2(X)) is near A if there exists +R > 0 such that supp(T) ⊆ NR(A) × NR(A), and the ideal IA is defined to be the +operator norm closure of all operators in Cu[X] near A. +The ideal IA is called spatial in [13] since it is related to a subspace in X. However, +as shown in [13, Example 2.1], there exist non-spatial ideals in general. On the +other hand, spatial ideals play an important role in the computation of the K- +theory of Roe algebras via the Mayer-Vietoris sequence argument (see [23]). +To show that spatial ideals are geometric, we observe that the smallest ideal in +X containing A is LA := {Z ⊆ NR(A) : R > 0}. Hence applying Proposition 4.13, we +immediately obtain the following: +Lemma 4.14. Let A be a subset of X. Then the set +(4.3) +UA := U(LA) = +� +R>0 +NR(A) +is an invariant open subset of βX. Moreover, if U is an invariant open subset of βX +containing A, then UA ⊆ U. +Consequently, combining with (4.2) we reach the following: +Corollary 4.15. Let A be a subset of X, then I(UA) = IA. Hence the spatial ideal IA is +geometric. +For later use, we record the following result concerning the set UA defined in +(4.3). Recall from Corollary A.8 that for a subset Z ⊆ X, the closure Z in βX is +homeomorphic to βZ. +Lemma 4.16. Let A be a subset of X. Then we have: +G(X)UA = +� +R>0 +G(NR(A)). +Proof. By definition, we have G(X)UA = � +R>0 G(X)NR(A). Note that for each R > 0 +and (α, ω) ∈ G(X)NR(A), we have ω ∈ NR(A) and there exists S > 0 such that +(α, ω) ∈ ES due to Lemma 3.7. Hence the pair (α, ω) in NR+S(A) × NR+S(A) belongs +to G(NR+S(A)), which concludes the proof. +□ +To end this section, we remark that for a given ideal I in C∗ +u(X), it is usually hard +to compute the associated U(I) directly from definition. However, this is always +achievable for principal ideals: +Lemma 4.17. Let I = ⟨T⟩ be the principal ideal in C∗ +u(X) generated by T ∈ C∗ +u(X). Denote +U := +� +ε>0,R>0 +NR(r(suppε(T))). +Then U is an invariant open subset of βX, and we have U(I) = U. +Proof. By Lemma 4.14, it is clear that U is an invariant open subset of βX. By (4.1), +U(I) contains r(suppε(T)) for any ε > 0. Since U(I) is invariant, we obtain that U(I) +contains U again by Lemma 4.14. + +20 +QIN WANG AND JIAWEN ZHANG +On the other hand, we consider S = �n +i=1 aiTbi where ai, bi are non-zero with +supports being partial translations contained in ER for some R > 0. Hence for any +ε > 0, we have +r +� +suppε(S) +� +⊆ +n +� +i=1 +r +� +supp ε +n(aiTbi) +� +⊆ +n +� +i=1 +NR +� +r +� +supp +ε +n∥ai∥·∥bi∥(T) +�� +, +which implies that r +� +suppε(S) +� +⊆ U. Note that operators of the form �n +i=1 aiTbi as +above are dense in I. Hence for a general element ˜S ∈ I and ε > 0, there exists +S = �n +i=1 aiTbi where ai, bi are non-zero with supports being partial translations +such that ∥ ˜S − S∥ < ε/2. +Hence suppε( ˜S) ⊆ suppε/2(S), which concludes the +proof. +□ +5. Ghostly ideals +In the previous section, we observe that for an invariant open subset U of βX, +the associated geometric ideal I(U) is the smallest element in the lattice +IU = {I is an ideal in C∗ +u(X) : U(I) = U}. +In this section, we would like to explore the largest element in IU. As revealed in +Section 1, a natural idea is to include all operators T ∈ C∗ +u(X) sitting in some ideal +I with U(I) = U, which leads to the following: +Definition 5.1. For a space (X, d) and an invariant open subset U of βX, denote +˜I(U) := {T ∈ C∗ +u(X) : r(suppε(T)) ⊆ U for any ε > 0}. +We call ˜I(U) the ghostly ideal associated to U. +The terminology will become clear later (see Proposition 5.6 and Remark 5.8 +below). First let us verify that ˜I(U) is indeed an ideal in C∗ +u(X). +Lemma 5.2. For an invariant open subset U of βX, ˜I(U) is an ideal in C∗ +u(X). +Proof. For T, S ∈ C∗ +u(X) and ε > 0, we have suppε(T + S) ⊆ suppε/2(T) ∪ suppε/2(S). +It follows that ˜I(U) is a linear space in C∗ +u(X). Given T ∈ ˜I(U) and ε > 0, note that +r(suppε(T∗)) = s(suppε(T)) and hence T∗ ∈ ˜I(U) since U is invariant. Now given a +sequence {Tn}n in ˜I(U) converging to T ∈ C∗ +u(X) and an ε > 0, there exists n ∈ N +such that suppε(T) ⊆ suppε/2(Tn). Hence we obtain that T ∈ ˜I(U). +Finally, given T ∈ ˜I(U) and S ∈ C∗ +u(X) we need show that TS and ST belong to +˜I(U). Note that ST = (T∗S∗)∗ and ˜I(U) is closed under taking the ∗-operation, hence +it suffices to show that TS ∈ ˜I(U). Since ˜I(U) is closed in C∗ +u(X) with respect to +the operator norm, it suffices to consider the case that S ∈ Cu[X]. We can further +assume that S is a partial translation since ˜I(U) is closed under taking addition. +In this case, we have r(suppε(TS)) ⊆ r(suppε/∥S∥(T)) for any ε > 0. Hence we +conclude the proof. +□ +Using the language of ideals in X, we record that for an ideal L in X and +T ∈ C∗ +u(X), then T ∈ ˜I(U(L)) if and only if for any ε > 0 there exist R > 0 and Y ∈ L +such that for any (x, y) � ER ∩ (Y × Y) then |T(x, y)| < ε. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +21 +The following result shows that the ghostly ideal ˜I(U) is indeed the largest +element in the lattice IU: +Lemma 5.3. Given an invariant open subset U of βX, we have the following: +(1) for an ideal I in C∗ +u(X) with U(I) = U, then I ⊆ ˜I(U). +(2) U(˜I(U)) = U. +Proof. (1). Given T ∈ I, the condition U(I) = U implies that r(suppε(T)) ⊆ U for +each ε > 0. Hence by definition, we have T ∈ ˜I(U). +(2). Note that +U(˜I(U)) = +� +T∈˜I(U),ε>0 +r(suppε(T)) ⊆ U. +On the other hand, Proposition 4.9 shows that U(I(U)) = U, which implies that +˜I(U) ⊇ I(U) thanks to (1). +Hence we have U(˜I(U)) ⊇ U(I(U)) = U again by +Proposition 4.9, which concludes the proof. +□ +Combining with Proposition 4.10, we reach the following desired result: +Theorem 5.4. Let (X, d) be a space as in Section 2.2, and U be an invariant open subset of +βX. Then any ideal I in C∗ +u(X) with U(I) = U sits between I(U) and ˜I(U). More precisely, +the geometric ideal I(U) is the smallest element while the ghostly ideal ˜I(U) is the largest +element in the lattice IU in (1.1). +As mentioned in Section 1, Theorem 5.4 draws the border of the lattice IU. +Therefore, once we can bust every ideal between I(U) and ˜I(U) for each invariant +open subset U ⊆ βX, then we will obtain a full description for the ideal structure +of the uniform Roe algebra C∗ +u(X) (see Question 9.1). +Now we aim to provide a geometric description for ghostly ideals, which helps +to explain the terminology. Let us start with an easy example. +Example 5.5. Taking U = X, then ˜I(X) is the ideal IG defined in Section 2.3. Indeed, +T ∈ ˜I(X) if and only if for any ε > 0, r(suppε(T)) is finite. This is equivalent to that +T ∈ C0(X × X) since T ∈ C∗ +u(X). On the other hand, it is clear that ˜I(βX) = C∗ +u(X). +More generally, we have the following result. Note that for any invariant open +subset U of βX, G(X)U is an open subset of β(X × X). Hence both C0(G(X)U) and +Ic(U) can be regarded as subalgebras in C(β(X × X)) � ℓ∞(X × X). +Proposition 5.6. For an invariant open subset U ⊆ βX and T ∈ C∗ +u(X), the following +are equivalent: +(1) T ∈ ˜I(U); +(2) T ∈ C0(G(X)U); +(3) T ∈ Ic(U) +∥·∥∞; +(4) T vanishes in the (βX \ U)-direction, i.e., Φω(T) = 0 for any ω ∈ βX \ U. +Proof. “(1) ⇒ (2)”: By definition, for any ε > 0 we have +r(supp(Tε)) = r(supp(Tε)) = r(suppε(T)) ⊆ U. + +22 +QIN WANG AND JIAWEN ZHANG +Consider the compact set K := { ˜ω ∈ β(X × X) : |T( ˜ω)| ≥ 2ε}. By Lemma 4.2, we +have r(K) ⊆ r(supp(Tε)) ⊆ U, which implies that K ⊆ G(X)U. Moreover, we have +|T( ˜ω)| < 2ε for any ˜ω ∈ β(X × X) \ K, which concludes that T ∈ C0(G(X)U). +“(2) ⇒ (1)”: Assume that T ∈ C0(G(X)U) ⊆ C(β(X × X)). Then for any ε > 0 there +exists a compact subset K ⊆ G(X)U such that for any ˜ω ∈ β(X × X) \ K, we have +|T( ˜ω)| < ε. This implies that { ˜ω ∈ β(X × X) : |T( ˜ω)| ≥ ε} ⊆ K. Using Lemma 4.2, we +obtain that supp(Tε) ⊆ K, which implies that r(suppε(T)) ⊆ U. Hence T ∈ ˜I(U). +“(2) ⇔ (3)”: This is due to the fact that +Ic(U) +∥·∥∞ = {f ∈ C0(G(X)) : f( ˜ω) = 0 for ˜ω � G(X)U}. +“(3) ⇔ (4)”: By Lemma 3.12, (4) holds if and only if T(α, γ) = 0 for any α, γ ∈ X(ω) +and ω ∈ βX \ U. Applying Lemma 2.9 and Lemma 3.4, this holds if and only if +T( ˜ω) = 0 whenever ˜ω � G(X)U, which describes elements in Ic(U) +∥·∥∞. Hence we +conclude the proof. +□ +Note that G(X)X = X × X. Hence as a direct corollary, we recover the following +characterisation for ghost operators: +Corollary 5.7 ([43, Proposition 8.2]). An operator T ∈ C∗ +u(X) is a ghost if and only if +Φω(T) = 0 for any non-principal ultrafilter ω on X. +Remark 5.8. In other words, Corollary 5.7 shows that a ghost in C∗ +u(X) is locally +invisible in all directions. Thissuggestsustoconsideroperatorsin ˜I(U)as“partial” +ghosts, which clarifies the terminology of “ghostly ideals”. +As an application of Proposition 5.6, we now provide another description for +ghostly ideals in terms of operator algebras, which will be used later in Section 7. +Let us start with the short exact sequences studied in [21] (see also [20, Section 2]). +Given an invariant open subset U ⊆ βX, notice that Uc = βX\U is also invariant. +Denote by G(X)Uc := G(X) ∩ s−1(Uc) and clearly, we have a decomposition: +G(X) = G(X)U ⊔ G(X)Uc. +Note that G(X)U is open in G(X), hence the above induces the following short +exact sequence of ∗-algebras: +(5.1) +0 −→ Cc(G(X)U) −→ Cc(G(X)) −→ Cc(G(X)Uc) −→ 0 +where the map Cc(G(X)U) −→ Cc(G(X)) is the inclusion and the map Cc(G(X)) −→ +Cc(G(X)Uc) is the restriction. +We may complete the sequence (5.1) with respect to the maximal groupoid +C∗-norms and obtain the following sequence: +(5.2) +0 −→ C∗ +max(G(X)U) −→ C∗ +max(G(X)) −→ C∗ +max(G(X)Uc) −→ 0, +which is easy to check by definition to be automatically exact (see, e.g., [26, Lemma +2.10]). We may also complete this sequence with respect to the reduced groupoid +C∗-norms and obtain the following sequence: +(5.3) +0 −→ C∗ +r(G(X)U) +iU +−→ C∗ +r(G(X)) +qU +−→ C∗ +r(G(X)Uc) −→ 0. +By construction, iU is injective, qU is surjective and qU ◦ iU = 0. +Also recall +from Lemma 4.7 that iU(C∗ +r(G(X)U)) = I(U), the geometric ideal associated to U. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +23 +However in general, (5.3) fails to be exact at the middle item. This is crucial in [21] +to provide a counterexample to the Baum-Connes conjecture with coefficients. +More precisely when U = X, it is proved in [35, Proposition 2.11] for the group +case and [20, Proposition 4.4] for the Roe algebraic case (see also [43, Proposition +8.2]) that Ker(qX) = IG, i.e., the ideal consisting of all ghost operators in C∗ +u(X). +Hence from Example 4.8 and Example 5.5, the sequence (5.3) is exact for U = X if +and only if I(X) = ˜I(X). More generally, we have the following: +Proposition 5.9. Given an invariant open subset U ⊆ βX, the kernel of qU : C∗ +r(G(X)) → +C∗ +r(G(X)Uc) coincides with the ghostly ideal ˜I(U). Hence the sequence (5.3) is exact if and +only if I(U) = ˜I(U). +Proof. It is easy to see that an operator T ∈ C∗ +u(X) � C∗ +r(G(X)) belongs to the +kernel of qU if and only if λω(T) = 0 for any ω ∈ βX \ U, where λω is the left +regular representation from (2.1). Now Lemma 3.14 implies that Φω(T) = 0 for +any ω ∈ βX \ U. Finally, we conclude the proof thanks to Proposition 5.6. +□ +We end this section with an illuminating example from [44, Section 3] (see also +[20, Section 5]), which is important to construct counterexamples to Baum-Connes +type conjectures: +Example 5.10. Let {Xi}i∈N be a sequence of expander graphs or pertubed expander +graphs (see [44] for the precise definition). Let Yi,j = Xi for all j ∈ N and set +Y := � +i,j Yi,j. We endow Y with a metric d such that it is the graph metric on each +Yi,j and satisfies d(Yi,j, Yk,l) → ∞ as i + j + k + l → ∞. +Let Pi,j ∈ B(ℓ2(Yi,j)) be the orthogonal projection onto constant functions on Yi,j, +and we set P to be the direct sum of Pi,j in the strong operator topology. By the +assumption on the expansion of {Xi}i∈N, it is clear that P ∈ C∗ +u(Y). It is explained +in [44, Section 3] (see also [20, Lemma 5.1]) that P is not a ghost, i.e., P � ˜I(X). +However intuitively, P should vanish “in the i-direction”. We will make it more +precisely in the following. +Recall from [20, Section 5.1] that we have a surjective map βY → βX × βN +induced by the bijection of Y with X×N and the universal property of βY. Define: +f : βY −→ βX × βN −→ βX +where the second map is just the projection onto the first coordinate. Denote +U = f −1(X), which is open in βY. Note that U = � +i f −1(Xi), where each f −1(Xi) is +homeomorphic to Xi × βN. On the other hand, note that +U = +� +i +f −1(Xi) = +� +i +� +j +Yi,j = +� +ε>0 +r(suppε(P)) = +� +ε>0,R>0 +NR(r(suppε(P))). +Hence it follows from Lemma 4.17 that U is invariant (comparing with [20, Lemma +5.2]), and U(⟨P⟩) = U. (Note that P ∈ ˜I(U) was already implicitly proved in [20, +Theorem 5.5], thanks to Proposition 5.9.) Since U contains Y as a proper subset, +we reprove that P is not a ghost. +Moreover, it follows from Proposition 5.6 that P vanishes in the (∂βY \ U)- +direction. In particular, fixing an index j0 ∈ N and taking a sequence {xi ∈ Yi,j0}i∈N, +we choose a cluster point ω ∈ {xi}i. It is clear that ω � U. Intuitively speaking, this +means that P vanishes “in the i-direction”. + +24 +QIN WANG AND JIAWEN ZHANG +We remark that the first-named author proved in [44, Section 3] that the prin- +cipal ideal ⟨P⟩ cannot be decomposed into I(U) + (IG ∩ ⟨P⟩), which provided a +couterexample to the conjecture at the end of [12]. Our explanation above reveals +that the reason behind this counterexample is that the ghostly part of ⟨P⟩ could +not be “exhausted” merely by ghostly elements associated to X (rather than U). +Finally, we remark that the groupoid G(Y)U also plays a key role in constructing +a counterexample to the boundary coarse Baum-Connes conjecture introduced in +[20] (see Section 5.2 therein). +6. Maximal ideals +In this section, we would like to study maximal ideals in uniform Roe algebras +using the tools developed in Section 5. Throughout this section, let (X, d) be a +space as in Section 2.2. +6.1. Minimal points in the boundary. Recall from previous sections that ideals +are closely related to invariant open subsets of the unit space βX. +Hence we +introduce the following: +Definition 6.1. An invariant open subset U ⊆ βX is called maximal if U � βX and U +is not properly contained in any proper invariant open subset of βX. Similarly, an +invariant closed subset K ⊆ βX is called minimal if K � ∅ and K does not properly +contain any non-empty invariant closed subset of βX. +Proposition 6.2. For any maximal invariant open subset U ⊂ βX, the ghostly ideal ˜I(U) +is a maximal ideal in the uniform Roe algebra C∗ +u(X). Conversely for any maximal ideal I +in C∗ +u(X), the associated invariant open subset U(I) is maximal and we have I = ˜I(U(I)). +Proof. For any ideal J in C∗ +u(X) containing ˜I(U), we have U(J) ⊇ U(˜I(U)) = U by +Lemma 5.3(2). Since U is maximal, then either U(J) = βX or U(J) = U. If U(J) = βX, +it follows from Lemma 4.11 that J contains I(U(J)) = C∗ +u(X), which implies that +J = C∗ +u(X). If U(J) = U, then it follows from Lemma 5.3(1) that J ⊆ ˜I(U(J)) = ˜I(U), +which implies that J = ˜I(U). This concludes that ˜I(U) is maximal. +Conversely for any maximal ideal I in C∗ +u(X), we have U(I) � βX. For any open +invariant subset V � βX containing U, we have I ⊆ ˜I(U) ⊆ ˜I(V) by Theorem 5.4, +and ˜I(V) � C∗ +u(X). Hence due to the maximality of I, we obtain that I = ˜I(U) = ˜I(V). +This implies that U = V and also I = ˜I(U) as required. +□ +Taking complements, we obtain the following: +Corollary 6.3. For any minimal invariant closed subset K ⊆ βX, the ghostly ideal +˜I(βX \ K) is a maximal ideal in the uniform Roe algebra C∗ +u(X). Moreover, every maximal +ideal in C∗ +u(X) arises in this form. +Therefore, in order to describe maximal ideals in the uniform Roe algebra, it +suffices to study minimal invariant closed subsets of the unit space βX. Recall from +Lemma 3.4 that for each ω ∈ ∂βX, the limit space X(ω) is the smallest invariant +subset of βX containing ω. However, note that X(ω) might not be closed in general. +Definition 6.4. A point ω ∈ ∂βX is called minimal if the closure of the limit space +X(ω) in βX is minimal in the sense of Definition 6.1. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +25 +The following result is straightforward, hence we omit the proof. It suggests us +to study minimal points in the boundary. +Lemma 6.5. For a minimal invariant closed subset K ⊆ βX, there exists a minimal point +ω ∈ ∂βX such that K = X(ω). Conversely for any minimal point ω ∈ ∂βX, the set X(ω) +is minimal. +One might wonder whether every ω ∈ ∂βX is minimal. However, things become +very complicated after taking closures and we will show later that this does not +hold even in the case of X = Z. Firstly, we notice the following: +Lemma 6.6. Let K be an invariant closed subset of βX. Then K contains a minimal point. +In particular, for any ω ∈ ∂βX there exists a minimal point ω′ such that ω′ ∈ X(ω). +Proof. This follows directly from the Zorn’s lemma together with the fact that βX +is compact. Details are left to readers. +□ +Consequently, we obtain: +Corollary 6.7. For ω ∈ ∂βX, ω is minimal if and only if for any α ∈ X(ω), we have +X(α) = X(ω). Writing X(ω) = � +λ∈Λ X(ωλ) for certain ωλ ∈ ∂βX, then ω is minimal if +and only if X(ωλ) = X(ω) for any λ ∈ Λ. +As promised, now we study the case of X = Z and show that it admits a number +of non-minimal points. The following is the main result: +Theorem 6.8. For the integer group Z with the usual metric, there exist non-minimal +points in the boundary ∂βZ. More precisely, for any sequence {hn}n∈N in Z tending to +infinity such that |hn − hm| → +∞ when n + m → ∞ and n � m, and any ω ∈ ∂βZ with +ω({hn}n∈N) = 1, then ω is not a minimal point. +To prove Theorem 6.8, we need some preparations. For future use, we record +the following result in the context of a countable discrete group Γ equipped with +a left-invariant word length metric. +Lemma 6.9. For ω, α ∈ ∂βΓ, then α ∈ Γ(ω) if and only if for any S ⊆ Γ with α(S) = 1, +there exists gS ∈ Γ such that ω(S · g−1 +S ) = 1. +Proof. Recall from Lemma 3.8 that Γ(ω) = {ρg(ω) : g ∈ Γ}, and from Appendix A +that {S : S ⊆ X} forms a basis for βX. Hence by definition, we obtain that α ∈ X(ω) +if and only if for any S ⊆ X with α ∈ S, there exists gS ∈ Γ such that ρgS(ω) ∈ S. +Equivalently, this means that for any S ⊆ X with α(S) = 1, there exists gS ∈ Γ such +that ρgS(ω)(S) = ω(S · g−1 +S ) = 1, which concludes the proof. +□ +Now we return to the case of Γ = Z and prove Theorem 6.8: +Proof of Theorem 6.8. Fix a subset H = {hn}n∈N ⊆ Z tending to infinity such that +|hn − hm| → +∞ when n + m → ∞ and n � m. For any non-zero g ∈ Z, note that +h ∈ (g + H) ∩ H if and only if there exists h′ ∈ H such that h − h′ = g. Since g is +fixed and distances between different points in H tend to infinity, we obtain that +(g + H) ∩ H is finite. Hence for any g1 � g2 in Z, (g1 + H) ∩ (g2 + H) is finite. + +26 +QIN WANG AND JIAWEN ZHANG +Fixing a non-principal ultrafilter ω ∈ ∂βZ with ω(H) = 1, we denote +U := {B ⊆ H : ω(B) = 1}. +We claim that for each n ∈ N, there exists gn ∈ Z and Bn ∈ U such that {Bn + gn}n∈N +are mutually disjoint. +Indeed, we take g0 = 0 and B0 = H. +Set g1 = 1 and +B1 := H \ (H − g1). Since H ∩ (H − g1) is finite by the previous paragraph, then +ω(B1) = 1, i.e., B1 ∈ U. Similarly for each n ∈ N, we take gn = n and Bn := +H \ +� +(H − g1) ∪ (H − g2) ∪ · · · ∪ (H − gn) +� +, which concludes the claim. +Consider � +H := � +n∈N(Bn + gn), and denote Un := {B ⊆ Bn : ω(B) = 1} for each +n ∈ N. By Lemma A.4, Un is an ultrafilter on Bn. Choose a non-principal ultrafilter +ω0 on N, and we consider: +� +U := +� � +n∈N +(An+gn) ⊆ +� +n∈N +(Bn+gn) = � +H : ∃ J ⊆ N with ω0(J) = 1 s.t. ∀n ∈ J, An ∈ Un +� +. +Lemma A.6 implies that � +U is an ultrafilter on � +H. We define a function α : P(Z) → +{0, 1} by setting α(S) = 1 if and only if S ∩ � +H ∈ � +U, which is indeed an ultrafilter on +Z thanks to Lemma A.5. Also note that α is non-principal and α(� +H) = 1. +Now we show that α ∈ Z(ω) while ω � Z(α), and hence conclude the proof. To +see that α ∈ Z(ω), we will consult Lemma 6.9. For any S ⊆ Z with α(S) = 1, by +definition we have S ∩ � +H ∈ � +U. Writing S ∩ � +H = � +n∈N(An + gn) with An ⊆ Bn, then +there exists J ∈ ω0 such that for any n ∈ J we have An ∈ Un. Hence for any n ∈ J, +we have S ⊇ S ∩ � +H ⊇ An + gn and ω(An) = 1, which implies that ω(S − gn) = 1. +Applying Lemma 6.9, we conclude that α ∈ Z(ω). +Finally, it remains to check that ω � Z(α). Assume the contrary, then Lemma +6.9 implies that there exists g ∈ Z and �B ⊆ � +H with α(�B) = 1 such that H ⊇ �B − g. +Writing �B = � +n∈N(An + gn) with An ⊆ Bn, then there exists J ∈ ω0 such that for any +n ∈ J, ω(An) = 1. This implies that +H + g ⊇ �B ⊇ +� +n∈J +(An + gn) ⊇ An0 + gn0 +for some n0 ∈ J with gn0 � g (this can be achieved since J is infinite). However, +it follows from the first paragraph that (H + g) ∩ (H + gn0) is finite. While this +intersection contains An0 + gn0, which is infinite since ω(An0) = 1. Therefore, we +reach a contradiction and conclude the proof. +□ +6.2. Maximal ideals via limit operators. We already see that maximal ideals in +the uniform Roe algebra correspond to minimal points in the Stone- ˇCech bound- +ary of the underlying space and in Section 6.1, we use topological methods to +show the existence of non-minimal points. Now we turn to a C∗-algebraic view- +point, and use the tool of limit operators to provide an alternative description for +these ideals. +First recall from Corollary 6.3 and Lemma 6.5 that maximal ideals in C∗ +u(X) +arise in the form of ˜I(βX \ X(ω)) for some boundary point ω ∈ ∂βX. Moreover +according to Proposition 5.9, ˜I(βX \ X(ω)) is the kernel of the following surjective +homomorphism: +qβX\X(ω) : C∗ +r(G(X)) −→ C∗ +r(G(X)X(ω)). + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +27 +Hence we obtain the following: +Corollary 6.10. A point ω ∈ ∂βX is minimal if and only if C∗ +r(G(X)X(ω)) is simple. +Example 6.11. Consider the case of a countable discrete group Γ. +For a point +ω ∈ ∂βΓ, it follows from Lemma 3.8 that the limit space Γ(ω) is identical to Γω, +and hence C∗ +r(G(Γ)Γ(ω)) is C∗-isomorphic to the reduced crossed product C(Γω) ⋊ Γ. +Thanks to Corollary 6.10, we obtain that ω is minimal if and only if C(Γω) ⋊ Γ +is simple. +Moreover, note that the action of Γ on βΓ is free (which is also a +consequence of Lemma 3.8). Hence it follows from [29, Corollary 4.6] that C(Γω)⋊Γ +is simple if and only if the action of Γ on Γω is minimal. In conclusion, we reach +the following: +Corollary 6.12. In the case of a countable discrete group Γ, a point ω ∈ ∂βΓ is minimal if +and only if the action of Γ on Γω is minimal. +Corollary 6.10 suggests an approach to distinguish minimal points via the +simplicity of the reduced groupoid C∗-algebra. However, this C∗-algebra is still +not easy to handle since it requires to consider all points in the X(ω)-direction (see +Proposition 5.6). Now we show that this can be simplified by merely considering +the ω-direction: +Lemma 6.13. For ω ∈ ∂βX, an operator T ∈ C∗ +u(X) belongs to the ideal ˜I(βX \ X(ω)) if +and only if T vanishes in the ω-direction, i.e., Φω(T) = 0. +Proof. We assume that Φω(T) = 0, and it suffices to show that Φα(T) = 0 for any +α ∈ X(ω). Fixing such an α, we take a net {ωλ}λ∈Λ in X(ω) such that ωλ → α and +it follows that Φωλ(T) = 0. For any γ1, γ2 ∈ X(α), Lemma 3.4 and the fact that the +coarse groupoid G(X) is ´etale imply that there exist γ1,λ and γ2,λ in X(ωλ) for each +λ ∈ Λ such that γ1,λ → γ1 and γ2,λ → γ2. Now Lemma 3.12 implies that +Φα(T)γ1γ2 = T((γ1, γ2)) = lim +λ∈Λ T((γ1,λ, γ2,λ)) = lim +λ∈Λ Φωλ(T)γ1,λγ2,λ = 0, +which concludes the proof. +□ +Hence for ω ∈ ∂βX, Lemma 6.13 implies that the associated ideal ˜I(β \ X(ω)) +coincides with the kernel of the following limit operator homomorphism (see also +[43, Theorem 4.10]): +(6.1) +Φω : C∗ +u(X) −→ C∗ +u(X(ω)), +T �→ Φω(T). +Consequently, we reach the following: +Corollary 6.14. A point ω ∈ ∂βX is minimal if and only if the image Im(Φω) is simple. +Hence when X(ω) is infinite and Φω is surjective, the point ω is not minimal. +Proof. For the last statement, it suffices to note that the ideal of compact operators +is always contained in C∗ +u(X(ω)), which concludes the proof. +□ +Thanks to Corollary 6.14, a special case of Theorem 6.8 can also be deduced +from a recent work by Roch [30]. More precisely, combining [43, Proposition B.6] +with [30, Lemma 2.1], we have the following: + +28 +QIN WANG AND JIAWEN ZHANG +Proposition 6.15. Let {hn}n∈N be a sequence in ZN tending to infinity such that +∥hn − hk∥∞ ≥ 3k +for any +k > n. +Then for any ω ∈ ∂βZN with ω({hn}n∈N) = 1, the map Φω : C∗ +u(ZN) −→ C∗ +u(ZN(ω)) is +surjective. +Note that the limit space ZN(ω) is bijective to ZN by Lemma 3.8, and hence +infinite. Therefore applying Corollary 6.14, we obtain the following (when N = 1, +it partially recovers Theorem 6.8). +Corollary 6.16. For any sequence {hn}n∈N in ZN tending to infinity such that ∥hn−hk∥∞ ≥ +3k for any k > n, and any ω ∈ ∂βZN with ω({hn}n∈N) = 1, then ω is not a minimal point. +Remark 6.17. We notice from the discussion above that for ω ∈ ∂βX, there is a +C∗-monomorphism: +(6.2) +C∗ +r(G(X)X(ω)) � C∗ +u(X)/˜I(βX \ X(ω)) −→ C∗ +u(X(ω)) +where the first comes from Proposition 5.9 and the second comes from (6.1) +together with Lemma 6.13. +Now we provide another explanation for this map in terms of groupoids. +Firstly, Corollary 3.5 implies that G(X)X(ω) = X(ω) × X(ω) and hence G(X)X(ω) = +X(ω) × X(ω) +G(X). Now we have +G(X)X(ω) = +� +S>0 +� +ES +G(X) ∩ X(ω) × X(ω) +G(X)� += +� +S>0 +ES +G(X) ∩ (X(ω) × X(ω)) +G(X) +, +where the last inequality is due to the fact that ES +G(X) is clopen in G(X). By Lemma +3.7, we have +ES +G(X) ∩ (X(ω) × X(ω)) = {(α, γ) ∈ X(ω) × X(ω) : dω(α, γ) ≤ S} =: ES(X(ω), dω) +i.e., the S-entourage in the limit space (X(ω), dω). Therefore, we conclude that +G(X)X(ω) = +� +S>0 +ES(X(ω), dω) +G(X). +On the other hand, we have (by definition) that +G(X(ω)) = +� +S>0 +ES(X(ω), dω) +β(X(ω)×X(ω)). +By the universal property of the Stone- ˇCech compactification, there is a surjective +continuous map ES(X(ω), dω) +β(X(ω)×X(ω)) −→ ES(X(ω), dω) +G(X), which induces an in- +jective map Cc(G(X)X(ω)) −→ Cc(G(X(ω))). Moreover, it is routine to check that it +induces a C∗-monomorphism +C∗ +r(G(X)X(ω)) −→ C∗ +r(G(X(ω))) � C∗ +u(X(ω)), +which can be verified to coincide with the map (6.2). Details are left to readers. +In particular, we consider a countable discrete group X = Γ. Fixing a point +ω ∈ ∂βΓ, we mentioned in Example 6.11 that C∗ +r(G(Γ)Γ(ω)) � C(Γω)⋊Γ. On the other +hand, we know from Lemma 3.8 that C∗ +u(Γ(ω)) � ℓ∞(Γω) ⋊ Γ. In this case, one can +check that the map (6.2) is induced by the natural embedding C(Γω) ֒→ ℓ∞(Γω). + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +29 +7. Geometric ideals vs ghostly ideals +In this section, we return to the lattice IU in (1.1). Recall from Theorem 5.4 that +for an invariant open subset U ⊆ βX, the geometric ideal I(U) and the ghostly +ideal ˜I(U) are the smallest and the largest elements in IU. Also as noticed before, +generally IU consists of more than one element. +Now we would like to study when I(U) = ˜I(U) or equivalently, when IU consists +of a single element. Moreover, we will discuss their K-theories and provide a +sufficient condition to ensure K∗(I(U)) = K∗(˜I(U)) for ∗ = 0, 1. +First recall from Proposition 5.9 that we already have a characterisation for +I(U) = ˜I(U) using short exact sequences, while the condition therein still seems +hard to check. Now we aim to search for a more practical criterion to ensure +I(U) = ˜I(U). We start with the following result, which combines [14, Theorem 4.4] +and [36, Theorem 1.3]. +Proposition 7.1. For a space (X, d), the following are equivalent: +(1) X has Property A; +(2) G(X) is amenable; +(3) G(X)∂βX is amenable; +(4) I(U) = ˜I(U) for any invariant open subset U ⊆ βX; +(5) I(X) = ˜I(X). +We remark that “(1) ⇒ (4)” was originally proved in [14] using approximations +by kernels. Here we will take a shortcut, and the idea will also be used later. +Proof of Proposition 7.1. “(1) ⇔ (2)” was proved in [39, Theorem 5.3]. “(2) ⇔ (3)” is +due to the permanence properties of amenability (see Section 2.6) together with +the fact that G(X)X � X × X is always amenable. +“(2) ⇒ (4)”: Let U ⊆ βX be an invariant open subset. As open/closed sub- +groupoids, both G(X)U and G(X)Uc are amenable as well. Consider the following +commutative diagram coming from (5.2) and (5.3): +0 +� C∗ +max(G(X)U) +� +� +C∗ +max(G(X)) +� +� +C∗ +max(G(X)Uc) +� +� +0 +0 +� C∗ +r(G(X)U) +� C∗ +r(G(X)) +� C∗ +r(G(X)Uc) +� 0. +By Proposition 2.12, all three vertical lines are isomorphisms. Hence the exactness +of the first row implies that the second row is exact as well. Therefore, we conclude +(4) thanks to Proposition 5.9. +“(4) ⇒ (5)” holds trivially, and “(5) ⇒ (1)” comes from [36, Theorem 1.3] together +with Example 4.8 and Example 5.5. Hence we conclude the proof. +□ +Proposition 7.1 provides a coarse geometric characterisation for I(X) = ˜I(X) +using Property A. However, we notice that assuming Property A is often too +strong to ensure that I(U) = ˜I(U) for merely a specific invariant open subset U ⊆ βX. +A trivial example is that I(βX) = ˜I(βX) holds for any space X. This suggests us to +explore a weaker criterion for I(U) = ˜I(U), and we reach the following: + +30 +QIN WANG AND JIAWEN ZHANG +Proposition 7.2. Let (X, d) be a space and U be an invariant open subset of βX. If the +canonical quotient map C∗ +max(G(X)Uc) → C∗ +r(G(X)Uc) is an isomorphism, then I(U) = +˜I(U). In particular, if the groupoid G(X)Uc is amenable then I(U) = ˜I(U). +Proof. We consider the following commutative diagram: +(7.1) +0 +� C∗ +max(G(X)U) +� +πU +� +C∗ +max(G(X)) +� +� +C∗ +max(G(X)Uc) +� +� +0 +0 +� ˜I(U) +� C∗ +r(G(X)) +� C∗ +r(G(X)Uc) +� 0. +Here the map πU is the composition: +(7.2) +C∗ +max(G(X)U) → C∗ +r(G(X)U) � I(U) ֒→ ˜I(U), +where the middle isomorphism comes from Lemma 4.7. Note that the top hor- +izontal line is automatically exact, while the bottom one is also exact thanks to +Proposition 5.9. Also note that the middle vertical map is always surjective and by +assumption, the right vertical map is an isomorphism. Hence via a diagram chas- +ing argument, we obtain that the left vertical map is surjective. This concludes +that I(U) = ˜I(U) thanks to (7.2). +□ +Remark 7.3. When U = X, Proposition 7.2 recovers “(3) ⇒ (5)” in Proposition 7.1. +Readers might wonder whether the converse of Proposition 7.2 holds as in the +case of U = X. We manage to provide a partial answer in Section 8.3 below. +Now we move to discuss the K-theory of geometric and ghostly ideals. Firstly, +we need an extra notion: +Definition 7.4. For a set S of subsets of X, denote L(S) the smallest ideal in X +containing S, and we say that L(S) is generated by S. An ideal L in X is called +countably generated if there exists a countable set S such that L = L(S). +An invariant open subset U ⊆ βX is called countably generated if the associated +ideal L(U) is countably generated. +Example 7.5. For a space X and a subspace A ⊆ X, it follows from Lemma 4.14 +that the spatial ideal IA is countably generated. On the other hand, it follows from +Lemma 4.17 that principal ideals are always countably generated. In particular, +the ideal ⟨P⟩ considered in [44, Section 3] (see also Example 5.10) is countably +generated. +The property of countable generatedness leads to the following: +Lemma 7.6. Let L be a countably generated ideal in X. Then there exists a countable +subset {Y1, Y2, · · · , Yn, · · · } in L such that +L = {Z ⊆ X : ∃ n ∈ N such that Z ⊆ Yn}. +Consequently for any countably generated open invariant subset U of βX, the subgroupoid +G(X)U is σ-compact. +To prove Lemma 7.6, we need an auxiliary result on the structure of L(S): +Lemma 7.7. For a set S of subsets of X, denote S(1) := {A1 ∪ · · · ∪ An : Ai ∈ S, n ∈ N} +and S(2) := {Nk( ˜A) : ˜A ∈ S(1), k ∈ N}. Then we have: +L(S) = {Z : ∃ Y ∈ S(2) such that Z ⊆ Y}. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +31 +Proof. Denote S(3) := {Z : ∃ Y ∈ S(2) such that Z ⊆ Y}. It is clear that S(3) ⊆ L(S), +and any ideal containing S must contain S(3). Hence it remains to check that S(3) +is an ideal. +For Z ∈ S(3) and W ⊆ Z, there exists Y ∈ S(2) such that Z ⊆ Y. Hence W ⊆ Y, +which implies that W ∈ S(3). Also note that Y ∈ S(2) implies that there exists +˜A ∈ S(1) and k′ ∈ N such that Y = Nk′( ˜A). +Hence for any k ∈ N, we have +Nk(Z) ⊆ Nk(Y) ⊆ Nk+k′( ˜A), which implies that Nk(Z) ∈ S(3). +Finally for Z1, Z2 ∈ S(3), there exist ˜A1, ˜A2 ∈ S(1) and k1, k2 ∈ N such that +Zi ⊆ Nki( ˜Ai) for i = 1, 2. Hence +Z1 ∪ Z2 ⊆ Nk1( ˜A1) ∪ Nk2( ˜A2) ⊆ Nk1+k2( ˜A1 ∪ ˜A2) ∈ S(2), +which implies that Z1 ∪ Z2 ∈ S(3). Therefore, we conclude the proof. +□ +Proof of Lemma 7.6. By assumption, there exists a countable S such that L = L(S). +Using the notation of Lemma 7.7, the set S(2) is countable as well. Hence the first +statement follows directly from Lemma 7.7. +For the second, note that U = � +n∈N Yn and hence G(X)U = � +n∈N G(X)Yn. Since +each Yn is closed, we obtain that G(X)U is σ-compact as desired. +□ +Now we are in the position to discuss the K-theory of geometric and ghostly +ideals: +Proposition 7.8. Let X be a space which can be coarsely embedded into some Hilbert +space. Then for any countably generated invariant open subset U ⊆ βX, we have an +isomorphism +(ιU)∗ : K∗(I(U)) −→ K∗(˜I(U)) +for ∗ = 0, 1, where ιU is the inclusion map. Therefore for any ideal I in C∗ +u(X) with U(I) +countably generated, we have an injective homomorphism +(ιI)∗ : K∗(I ∩ Cu[X]) −→ K∗(I) +for ∗ = 0, 1, where ιI is the inclusion map. +Proof. Fixing such a U ⊆ βX, we consider the commutative diagram (7.1) from the +proof of Proposition 7.2, where both of the horizontal lines are exact. This implies +the following commutative diagram in K-theories: +· · · +� K∗(C∗ +max(G(X)U)) +� +(πU)∗ +� +K∗(C∗ +max(G(X))) +� +� +K∗(C∗ +max(G(X)Uc)) +� +� +K∗+1(C∗ +max(G(X)U)) +� +(πU)∗+1 +� +· · · +· · · +� K∗(˜I(U)) +� K∗(C∗ +r(G(X))) +� K∗(C∗ +r(G(X)Uc)) +� K∗+1(˜I(U)) +� · · · +where both horizontal lines are exact. Recall from [39, Theorem 5.4] that X is +coarsely embeddable if and only if G(X) is a-T-menable. It follows directly from +Definition 2.13 that as subgroupoids, both G(X)U and G(X)Uc are a-T-menable. +Moreover, it is clear that G(X) and G(X)Uc are σ-compact by definition. Hence +Proposition 2.14 implies that G(X) and G(X)Uc are K-amenable. This shows that +the middle two vertical maps in the above diagram are isomorphisms, which +implies that (πU)∗ is an isomorphism by the Five Lemma. Therefore from (7.2), +we obtain that the composition +K∗(C∗ +max(G(X)U)) −→ K∗(C∗ +r(G(X)U)) � K∗(I(U)) +(ιU)∗ +−→ K∗(˜I(U)) + +32 +QIN WANG AND JIAWEN ZHANG +is an isomorphism for ∗ = 0, 1. +On the other hand, it follows from Lemma 7.6 that G(X)U is σ-compact, which +implies that G(X)U is K-amenable again by Proposition 2.14. Hence we conclude +that the map (ιU)∗ is an isomorphism for ∗ = 0, 1. +For the last statement, we assume that U = U(I). By Lemma 4.11, we have that +I ∩ Cu[X] = I(U). Also Lemma 5.3(1) shows that I ⊆ ˜I(U). Hence the inclusion +map ιU can be decomposed as follows: +I(U) = I ∩ Cu[X] +ιI֒→ I ֒→ ˜I(U). +Therefore, (ιU)∗ being an isomorphism implies that (ιI)∗ is injective for ∗ = 0, 1. +□ +Applying Proposition 7.8 to the case of U = X, we partially recover the following +result by Finn-Sell (see [19, Proposition 35]), which is crucial for the counterex- +amples to the coarse Baum-Connes conjecture: +Corollary 7.9. Let X be a space which can be coarsely embedded into some Hilbert space. +Then the inclusion of K(ℓ2(X)) into IG induces an isomorphism on the K-theory level. +Remark 7.10. For a general ideal I in C∗ +u(X), our method in the proof of Proposition +7.8 only provides the injectivity of the induced map (ιI)∗. We wonder whether this +map is indeed an isomorphism under the same assumption (see Question 9.4). +We end this section with an example, which shows that not every ideal in a space +is countably generated. +Example 7.11. Let X = N × N, equipped with the metric induced from the Eu- +clidean metric dE on R2. For each θ ∈ [0, π +2] and k ∈ N, we define +ℓθ := {(x, y) ∈ R × R : y = tan(θ)x} +and +Sθ,k := {(x, y) ∈ X : dE((x, y), ℓθ) ≤ k} = Nk(ℓθ) ∩ X. +Consider S := {Sθ,k : θ ∈ [0, π +2], k ∈ N} and set S(1) := {A1 ∪ · · · ∪ An : Ai ∈ S, n ∈ N} +as in Lemma 7.7. For any R > 0 and A = A1 ∪ · · · ∪ An ∈ S(1) where Ai ∈ S, we +have NR(A) = NR(A1) ∪ · · · ∪ NR(An), which is contained in some element in S. +Hence applying Lemma 7.7, the ideal L(S) generated by S is: +(7.3) +L(S) = {Z : ∃ Y ∈ S(1) such that Z ⊆ Y}. +We claim that L(S) is not countably generated. Otherwise, there exists a count- +able subset S′ generating L(S). Moreover, according to (7.3) we can assume that +S′ = {Yn : n ∈ N} +where +Yn = Sθn,1,kn ∪ · · · ∪ Sθn,pn,kn ∈ S(1). +Choose θ ∈ [0, π +2] \ {θn,i : i = 1, 2, · · · , pn; n ∈ N}, and consider Y = Sθ,1. Since S′ +generates L(S), it follows from Lemma 7.7 that there exist R > 0 and Ym1, · · · , Yml ∈ +S′ such that +Sθ,1 = Y ⊆ NR(Ym1 ∪ · · · ∪ Yml). +Note thatthe righthand side iscontained in a finite union ofsome R′-neighbourhood +of lines (in R2 crossing the origin) with slopes in the set +� +tan(θn,i) : i = 1, 2, · · · , pn; n ∈ N +� +. +This leads to a contradiction due to the choice of θ, which concludes that L(S) +cannot be countably generated. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +33 +8. Partial Property A and partial operator norm localisation property +In the previous section, we find a sufficient condition (Proposition 7.2) to ensure +˜I(U) = I(U) for a given invariant open subset U ⊆ βX. As promised in Remark 7.3, +now we study its converse and show that this is indeed an equivalent condition +under the assumption of countable generatedness. +Our strategy is to follow the outline of the case that U = X. More precisely, we +introduce a notion called partial Property A towards invariant subsets of the boundary +∂βX, and then consider its counterpart in the context of operator norm localisation +property to provide the desired characterisation. +8.1. Partial Property A. Recall from Proposition 7.1 that a space X has Property +A if and only if the groupoid G(X)∂βX is amenable, which characterises I(X) = ˜I(X). +Together with Proposition 7.2, this inspires us to introduce the following: +Definition 8.1. Let (X, d) be a space and U ⊆ βX be an invariant open subset. We +say that X has partial Property A towards ∂βX \ U if G(X)∂βX\U is amenable. +It is clear from definition that X has Property A if and only if it has partial +Property A towards the whole boundary ∂βX. On the other hand, it follows from +Proposition 7.2 that if X has partial Property A towards ∂βX \ U, then we have +I(U) = ˜I(U). The rest of this section is devoted to studying the converse. +Firstly, we aim to unpack the groupoid language and provide a concrete geomet- +ric description for partial Property A, which resembles the definition of Property +A (see Definition 2.1). The following is the main result: +Proposition 8.2. Let (X, d) be a space and U ⊆ βX an invariant open subset. Then X +has partial Property A towards Uc = ∂βX \ U if and only if for any ε, R > 0, there exist +S > 0, a subset D ⊆ X with D ⊇ Uc and a function f : X × X → [0, 1] satisfying: +(1) supp(f) ⊆ ES; +(2) for any x ∈ D, we have � +z∈X f(z, x) = 1; +(3) for any x, y ∈ D with d(x, y) ≤ R, we have � +z∈X | f(z, x) − f(z, y)| ≤ ε. +Comparing Proposition 8.2 with Definition 2.1, it is clear that Property A implies +partial Property A towards any invariant closed subset of ∂βX. +To prove Proposition 8.2, we start with the following lemma: +Lemma 8.3. With the same notation as above, X has partial Property A towards Uc if +and only if for any ε, R > 0, there exist S > 0 and a function f : X × X → [0, 1] +satisfying: +(1) supp(f) ⊆ ES; +(2) for any ω ∈ Uc, we have � +α∈X(ω) f(α, ω) = 1; +(3) for ω ∈ Uc and α ∈ X(ω) with dω(α, ω) ≤ R, then � +γ∈X(ω) | f(γ, α) − f(γ, ω)| ≤ ε, +where f ∈ C0(G(X)) is the continuous extension from Lemma 2.9. +Proof. By definition, X has partial Property A towards Uc if and only if for any +ε > 0 and compact K ⊆ G(X)Uc, there exists g ∈ Cc(G(X)Uc) with range in [0, 1] such + +34 +QIN WANG AND JIAWEN ZHANG +that for any γ ∈ K we have +� +α∈Gr(γ) +g(α) = 1 +and +� +α∈Gr(γ) +|g(α) − g(αγ)| < ε. +Recall from (5.1) that the restriction map Cc(G(X)) → Cc(G(X)Uc) is surjective, and +hence g can be regarded as a function in Cc(G(X)). Taking f to be the restriction +of g on X × X, then f ∈ ℓ∞(X × X) and there exists S > 0 such that supp(f) ⊆ ES +for some S > 0. Using the notation from Lemma 2.9, we have g = f. Note that +G(X)Uc = � +R>0(ER ∩ G(X)Uc), and hence compact subsets of G(X)Uc are always +contained in those of the form ER ∩ G(X)Uc. Furthermore, Lemma 3.7 implies +ER ∩ G(X)Uc = +� +ω∈Uc +{(α, γ) ∈ X(ω) × X(ω) : dω(α, γ) ≤ R}. +Combining with Lemma 3.4, we conclude the proof. +□ +As a direct corollary (together with Lemma 3.7), we obtain: +Corollary 8.4. Assume that X has partial Property A towards Uc. Then the family of +metric spaces {(X(ω), dω)}ω∈Uc has uniform Property A in the sense that the parameters +in Definition 2.1 can be chosen uniformly. +Remark 8.5. It is unclear to us whether the converse of Corollary 8.4 holds. Note +that X(ω) might contain points outside X(ω) as discussed in Theorem 6.8, hence we +do not know whether functions on X(ω) × X(ω) can be glued together to provide +a continuous function on G(X)Uc. +Proof of Proposition 8.2. Sufficiency: For any ε, R > 0, choose S > 0, D ⊆ X and a +function g : X × X → [0, 1] satisfying the conditions (1)-(3) for ε and 3R. Take a +map p : NR(D) → D such that the restriction of p on D is the identity map and +d(p(x), x) ≤ R. Now we define: +f(x, y) = +� g(x, p(y)), +y ∈ NR(D); +g(x, y), +otherwise. +It is clear that supp(f) ⊆ ER+S. Moreover for any y1, y2 ∈ NR(D) with d(y1, y2) ≤ R, +we have d(p(y1), p(y2)) ≤ 3R and hence +� +x∈X +| f(x, y1) − f(x, y2)| = +� +x∈X +|g(x, p(y1)) − g(x, p(y2))| ≤ ε. +Therefore (enlarging S to S + R) we obtain that for any ε, R > 0, there exist S > 0, +a subset D ⊆ X with D ⊇ Uc and a function f : X × X → [0, 1] satisfying: +(1) supp(f) ⊆ ES; +(2) for any x ∈ D, we have � +z∈X f(z, x) = 1; +(3) for any x ∈ D and y ∈ X with d(x, y) ≤ R, we have � +z∈X | f(z, x) − f(z, y)| ≤ ε. +Now we fix ε, R > 0 and take such S, D and function f. +Given ω ∈ Uc, we have ω(D) = 1. Choose {tα : Dα → Rα} to be a compatible +family for ω. Applying Proposition 3.3, there exists Yω ⊆ X with ω(Y) = 1 and a +local coordinate system {hy : B(ω, R + S) → B(y, R + S)}y∈Yω such that the map +hy : B(ω, R + S) → B(y, R + S), +α �→ tα(y) + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +35 +is a surjective isometry for each y ∈ Yω. Replacing Y by Yω ∩ D, we assume that +Yω ⊆ D. Note that supp(f) ⊆ ES, and hence applying Lemma 3.7 we have +� +α∈X(ω) +f(α, ω) = +� +α∈B(ω,S) +f(α, ω) = +� +α∈B(ω,S) +lim +x→ω f(tα(x), x) = +lim +x→ω,x∈Yω +� +α∈B(ω,S) +f(tα(x), x) += +lim +x→ω,x∈Yω +� +z∈B(x,S) +f(z, x) = +lim +x→ω,x∈Yω +� +z∈X +f(z, x), +(8.1) +where the last item equals 1 thanks to the assumption. +On the other hand, for any α ∈ B(ω, R) we have limx→ω d(tα(x), x) ≤ R. Hence +shrinking Yω if necessary, we can assume that d(tα(x), x) ≤ R for any x ∈ Yω. +Therefore, we have +� +γ∈X(ω) +| f(γ, α) − f(γ, ω)| = +� +γ∈B(ω,R+S) +| f(γ, α) − f(γ, ω)| +(8.2) += +lim +x→ω,x∈Yω +� +γ∈B(ω,R+S) +| f(tγ(x), tα(x)) − f(tγ(x), x)| += +lim +x→ω,x∈Yω +� +z∈B(x,R+S) +| f(z, tα(x)) − f(z, x)| += +lim +x→ω,x∈Yω +� +z∈X +| f(z, tα(x)) − f(z, x)|, +where the last item is no more than ε by assumption. Therefore applying Lemma +8.3, we conclude the sufficiency. +Necessity: Given ε, R > 0, Lemma 8.3 provides S > 0 and a function f : X × X → +[0, 1] satisfying the conditions (1)-(3) therein. +Fix an ω ∈ Uc and we choose +{tα : Dα → Rα} to be a compatible family for ω. Applying Proposition 3.3, there +exists Yω ⊆ X with ω(Y) = 1 and a local coordinate system {hy : B(ω, R + 2S) → +B(y, R + S)}y∈Yω such that the map +hy : B(ω, R + S) → B(y, R + S), +α �→ tα(y) +is a surjective isometry for each y ∈ Yω. By the calculations in (8.1), we obtain that +lim +x→ω,x∈Yω +� +z∈X +f(z, x) = 1. +Hence for the given ε, there exists Y′ +ω ⊆ Yω with ω(Y′ +ω) = 1 such that for any x ∈ Y′ +ω +we have +� +z∈X +f(z, x) ∈ (1 − ε, 1 + ε). +On the other hand, for any α ∈ B(ω, R) we apply the calculations in (8.2) and +obtain: +lim +x→ω,x∈Y′ω +� +z∈X +| f(z, tα(x)) − f(z, x)| = +� +γ∈X(ω) +| f(γ, α) − f(γ, ω)| ≤ ε. +Note that for x ∈ Y′ +ω, Proposition 3.3 implies that {tα(x) : α ∈ B(ω, R)} = B(x, R). +Hence there exists Y′′ +ω ⊆ Y′ +ω with ω(Y′′ +ω) = 1 such that for any x ∈ Y′′ +ω and y ∈ B(x, R), +we have +� +z∈X +| f(z, y) − f(z, x)| < 2ε. + +36 +QIN WANG AND JIAWEN ZHANG +Taking D := � +ω∈Uc Y′′ +ω, then it is clear that D ⊇ Uc. Moreover, the analysis above +shows that: +• for any x ∈ D we have � +z∈X f(z, x) ∈ (1 − ε, 1 + ε); +• for any x ∈ D and y ∈ B(x, R), we have � +z∈X | f(z, y) − f(z, x)| < 2ε. +Finallyusinga standard normalisation argument(orequivalently, applyingLemma +2.11 and modifying Lemma 8.3 accordingly), we conclude the proof. +□ +Setting ξy(x) := f(x, y) for the function f in Proposition 8.2, we can rewrite +Proposition 8.2 as follows: +Proposition 8.2′. Let (X, d) be a space and U ⊆ βX be an invariant open subset. Then +X has partial Property A towards Uc if and only if for any ε, R > 0, there exist S > 0, a +subset D ⊆ X with D ⊇ Uc and a function ξ : D → ℓ1(X)1,+, x �→ ξx satisfying: +(1) supp(ξx) ⊆ B(x, S) for any x ∈ D; +(2) for any x, y ∈ D with d(x, y) ≤ R, we have ∥ξx − ξy∥1 ≤ ε. +Remark 8.6. We remark that the function ξ in Proposition 8.2′ can be made such +that ξx ∈ ℓ1(D)1,+. In fact, this is the same trick as in the case of Property A (see, +e.g., [27, Proposition 4.2.5]). Moreover, we can further replace ℓ1(D)1,+ by ℓ2(D)1,+ +using the Mazur map (see, e.g., [45, Proposition 1.2.4] for the same trick). +Now we provide an alternative picture for Proposition 8.2 using the notion of +ideals in spaces (see Definition 4.12). Recall that for an ideal L in X, we denote +U(L) := � +Y∈L Y. We need the following auxiliary lemma: +Lemma 8.7. Let L be an ideal in X and D ⊆ X. Then D ⊇ U(L)c if and only if there +exists Y ∈ L such that D ⊇ Yc. +Proof. Assume Y ∈ L such that D ⊇ Yc. Note that Y ∩ Yc = ∅ and Y ∪ Yc = βX. +Hence we have D ⊇ Yc = βX \ Y ⊇ βX \ U(L) = U(L)c. +Conversely, assume that D ⊇ U(L)c. Then (D)c ⊆ U(L) = � +Y∈L Y. Since D is +clopen in the compact space βX, the set (D)c is compact as well. Hence there exists +Y1, · · · , Yn ∈ L such that (D)c ⊆ Y1 ∪· · ·∪Yn = Y1 ∪ · · · ∪ Yn. Since L is an ideal, the +set Y := Y1∪· · ·∪Yn ∈ L. Then we have (D)c ⊆ Y, which implies that D ⊇ (Y)c = Yc. +Finally we obtain D = D ∩ X ⊇ Yc ∩ X = Yc, which conclude the proof. +□ +Thanks to Lemma 8.7, now we can rewrite Proposition 8.2 (combining with +Remark 8.6) as follows: +Proposition 8.2′′. Let (X, d) be a space, U ⊆ βX an invariant open subset and L = L(U) +the associated ideal in X. Then X has partial Property A towards Uc if and only if for any +ε, R > 0, there exist S > 0, a subset Y ∈ L(U) and a function ξ : Yc → ℓ2(Yc)1,+, x �→ ξx +satisfying: +(1) supp(ξx) ⊆ B(x, S) for any x ∈ Yc; +(2) for any x, y ∈ Yc with d(x, y) ≤ R, we have ∥ξx − ξy∥1 ≤ ε. +Similar to the proof in the case of Property A, we also have the following: + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +37 +Corollary 8.8. Let (X, d) be a space, U ⊆ βX an invariant open subset and L = L(U) the +associated ideal in X. Then X has partial Property A towards Uc if and only if for any +R > 0 and ε > 0, there exist S > 0, a subset Y ∈ L(U) and a kernel k : Yc × Yc → R of +positive type satisfying: +(1) for x, y ∈ Yc, we have k(x, y) = k(y, x) and k(x, x) = 1; +(2) for x, y ∈ Yc with d(x, y) ≥ S, we have k(x, y) = 0; +(3) for x, y ∈ Yc with d(x, y) ≤ R, we have |1 − k(x, y)| ≤ ε. +8.2. Partial operator norm localisation property. Recall that the notion of oper- +ator norm localisation property (ONL) was introduced by Chen, Tessera, Wang +and Yu in [10], and proved by Sako in [37] that ONL is equivalent to Property A. +Here we introduce a partial version of ONL, parallel to Definition 8.1. +Let ν be a positive locally finite Borel measure on X and H be a separable +infinite-dimensional Hilbert space. For an operator T ∈ B(L2(X, ν) ⊗ H), we can +also define its propagation as in Section 2.3. We introduce the following: +Definition 8.9. Let (X, d) be a space and U ⊆ βX an invariant open subset. We +say that X has partial operator norm localisation property (partial ONL) towards Uc = +∂βX \ U if there exists c ∈ (0, 1] such that for any R > 0 there exist S > 0 and +D ⊆ X with D ⊇ Uc satisfying the following: for any positive locally finite Borel +measure ν on X with supp(ν) ⊆ D and any a ∈ B(L2(X, ν) ⊗ H) with propagation +at most R, there exists a non-zero ζ ∈ L2(X, ν) ⊗ H with diam(supp(ζ)) ≤ S such +that c∥a∥ · ∥ζ∥ ≤ ∥aζ∥. +The aim of the rest of this subsection is to show that partial ONL is equivalent +to partial Property A. We will follow the outline of [37]. +To simplify the statement, denote CR +u[X; H] := {T ∈ B(ℓ2(X) ⊗ H) : prop(a) ≤ R} +for R ≥ 0. +For a subspace Y ⊆ X, it is clear that CR +u[Y] � χYCR +u[X]χY (resp. +CR +u[Y; H] � χYCR +u[X; H]χY), and hence can be regarded as a subset of CR +u[X] (resp. +CR +u[X; H]) with support in Y × Y. Similarly, C∗ +u(Y) � χYC∗ +u(X)χY can be regarded +as a C∗-subalgebra in C∗ +u(X). Hence we will not tell the difference in the sequel. +For S > 0, denote +BY +S := +� +x∈X +B(ℓ2(B(x, S) ∩ Y)), +whose elements will be written as b = ([bx(y, z)]y,z∈B(x,S)∩Y)x∈X. We also consider the +map +ψY +S : C∗ +u(Y) � χYC∗ +u(X)χY −→ BY +S +by +a �→ ([a(y, z)]y,z∈B(x,S)∩Y)x∈X. +Recall the notions of completely positive map and completely bounded map: +• A self-adjoint closed subspace F of a unital C∗-algebra B such that 1B ∈ F is +called an operator system. +• A linear map φ from F to a C∗-algebra C is said to be completely positive if +the map φ(n) = φ ⊗ id : F ⊗ Mn(C) → C ⊗ Mn(C) is positive for every n. +• A linear map θ : F → C is said to be completely bounded if the sequence +{∥θ(n) : F⊗Mn(C) → C⊗Mn(C)∥} is bounded. Denote ∥θ∥cb := supn∈N ∥θ(n)∥. +We have the following characterisation for partial ONL, which is analogous to +[37, Proposition 3.1]. The proof is almost identical, hence omitted. + +38 +QIN WANG AND JIAWEN ZHANG +Lemma 8.10. Let (X, d) be a space and U ⊆ βX an invariant open subset. Then the +following are equivalent: +(1) X has partial ONL towards Uc; +(2) there exists c ∈ (0, 1] such that for any R > 0 there exist S > 0 and D ⊆ X with +D ⊇ Uc satisfying condition (α): for any a ∈ CR +u[D; H] there exists a non-zero +ζ ∈ ℓ2(X) ⊗ H with diam(supp(ζ)) ≤ S and c∥a∥ · ∥ζ∥ ≤ ∥aζ∥; +(3) for any c ∈ (0, 1) and R > 0, there exist S > 0 and D ⊆ X with D ⊇ Uc satisfying +condition (α); +(4) for any c ∈ (0, 1) and R > 0, there exist S > 0 and D ⊆ X with D ⊇ Uc +satisfying condition (β): for any a ∈ CR +u[D] there exists a non-zero ξ ∈ ℓ2(X) with +diam(supp(ξ)) ≤ S and c∥a∥ · ∥ξ∥ ≤ ∥aξ∥; +(5) for any ε, R > 0 there exist S > R and D ⊆ X with D ⊇ Uc such that +∥(ψD +S |CR +u[D])−1 : ψD +S (CR +u[D]) −→ CR +u[D]∥ < 1 + ε; +(6) for any ε, R > 0 there exist S > R and D ⊆ X with D ⊇ Uc such that +∥(ψD +S |CR +u[D])−1 : ψD +S (CR +u[D]) −→ CR +u[D]∥cb < 1 + ε; +We record the following, which comes directly from Lemma 8.7 and 8.10. +Lemma 8.11. Let (X, d) be a space, U ⊆ βX an invariant open subset and L = L(U) the +associated ideal in X. Then X has partial ONL towards Uc if and only if for any c ∈ (0, 1) +and R > 0 there exist S > 0 and Y ∈ L(U) satisfying the following: for any a ∈ CR +u[Yc] +there exists a non-zero ξ ∈ ℓ2(X) with diam(supp(ξ)) ≤ S and c∥a∥ · ∥ξ∥ ≤ ∥aξ∥. +Finally, we can mimic the proof of [37, Theorem 4.1] using Proposition 8.2′, +Remark 8.6, Corollary 8.8 and Lemma 8.10 instead, and reach the following. The +proof is almost identical, and hence omitted. +Proposition 8.12. Let (X, d) be a space and U ⊆ βX be an invariant open subset. Then +X has partial Property A towards Uc if and only if X has partial ONL towards Uc. +To end this subsection, we study a permanence property of partial ONL, which +will help to prove the main result in the next subsection. +Let X be a space and L an ideal in X. Assume that X can be decomposed into +X = X1 ∪ X2. Consider +(8.3) +Li := {Y ∩ Xi : Y ∈ L} +for +i = 1, 2. +Then it is routine to check that Li is an ideal in Xi for i = 1, 2, and +L = {Y1 ∪ Y2 : Yi ∈ Li, i = 1, 2}. +With respect to the decomposition above, we now show that partial ONL is +preserved under finite unions. +Proposition 8.13. With the notation as above, assume that Xi has partial ONL towards +βXi \ U(Li) for i = 1, 2. Then X has partial ONL towards βX \ U(L). +One way to prove Proposition 8.13 is to follow the proof of [16, Lemma 3.3] +with minor changes. Here we choose another approach using Proposition 8.12. +Recall from Corollary A.8 that for a subset Z ⊆ X, the closure Z in βX is +homeomorphic to βZ. Hence we will regard them as the same object in the sequel. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +39 +Now we can easily transfer the restriction of ideals in (8.3) to that of invariant +open subsets (the proof is straightforward, hence omitted): +Lemma 8.14. Let X = X1 ∪ X2, L be an ideal in X and U = U(L) ⊆ βX be the associated +invariant open subset of βX. Then U ∩ Xi is an invariant open subset of Xi = βXi, which +corresponds to the ideal Li in (8.3) for i = 1, 2. +Although U ∩ Xi is invariant in βXi, generally it is not invariant in βX. This +coincides with the fact that χXiC∗ +u(X)χXi � C∗ +u(Xi) is a just subalgebra in C∗ +u(X) +rather than an ideal. Instead, we consider the spatial ideal IXi recalled in Section +4, and prove the following permanence property for partial Property A: +Lemma 8.15. Let X = X1 ∪ X2, U ⊆ βX be an invariant open subset and Ui = U ∩ Xi +for i = 1, 2. If Xi has partial Property A towards βXi \ Ui for i = 1, 2, then X has partial +Property A towards βX \ U. +Proof. For any R > 0 and i = 1, 2, set Ui(R) := � +Y∈L(Ui) NR(Y). +Clearly, Ui(R) +is an invariant open subset of NR(Xi) = β(NR(Xi)). Moreover, it follows from +Proposition 8.2 that NR(Xi) has partial Property A towards NR(Xi) \ Ui(R). By +definition, this means that the groupoid G(NR(Xi))NR(Xi)\Ui(R) is amenable. Hence +as a subgroupoid, G(NR(Xi))NR(Xi)\U is amenable, which further implies that the +groupoid � +R>0 G(NR(Xi))NR(Xi)\U is amenable. +Note from Lemma 4.14 that � +R NR(Xi) \ U is invariant in βX and Lemma 4.16 +implies that +G(X)� +R NR(Xi)\U = +� +R>0 +G(NR(Xi))NR(Xi)\U, +which is hence amenable. Note that � +R NR(X1) ∪ � +R NR(X2) = βX, and hence due +to the extension property we obtain that +G(X)βX\U = G(X)� +R NR(X1)\U ∪ G(X)� +R NR(X2)\U +is amenable as required. +□ +Combining Proposition 8.12 and Lemma 8.15, we conclude Proposition 8.13. +8.3. Characterisation for I(U) = ˜I(U). Having established all the necessary ingre- +dients above, now we present the main result of this section: +Theorem 8.16. Let (X, d) be a space as in Section 2.2 and U ⊆ βX be a countably +generated invariant open subset. Then the following are equivalent: +(1) X has partial Property A towards βX \ U; +(2) ˜I(U) = I(U); +(3) the ideal IG of all ghost operators is contained in I(U). +Note that U = X is countably generated, and hence Theorem 8.16 recovers [36, +Theorem 1.3] (see Example 4.8 and Example 5.5). Borrowing the language of +[36], condition (3) in Theorem 8.16 says that all the ghosts can be busted in the +geometric ideal I(U). +We follow the outline of the proof for [36, Theorem 1.3]. Firstly, we need a +modified version of [36, Lemma 4.2]: + +40 +QIN WANG AND JIAWEN ZHANG +Lemma 8.17. Let (X, d) be a space, U ⊆ βX be a countably generated invariant open +subset and L = L(U) the associated ideal in X. Assume that X does not have partial ONL +towards Uc. Then there exist κ ∈ (0, 1), R > 0, a sequence (Tn) in Cu[X], a sequence (Bn) +of finite subsets of X and a sequence (Sn) of positive real numbers such that: +(a) (Sn) is an increasing sequence tending to infinity as n → ∞; +(b) each Tn is positive and has norm 1; +(c) for n � m, then Bn ∩ Bm = ∅; +(d) each Tn is supported in Bn × Bn; +(e) for each n and ξ ∈ ℓ2(X) with ∥ξ∥ = 1 and diam(suppξ) ≤ Sn, then ∥Tnξ∥ ≤ κ; +(f) for each Y ∈ L(U), there exists n such that Bn ∩ Y = ∅. +Proof. Fixing a basepoint x0 ∈ X, consider the decomposition X = X(1) ∪ X(2) with +X(1) := +� +m even +{x ∈ X : m2 ≤ d(x, x0) < (m + 1)2} +and +X(2) := +� +m odd +{x ∈ X : m2 ≤ d(x, x0) < (m + 1)2}. +Set Li := {Y ∩ X(i) : Y ∈ L} for i = 1, 2. By assumption, X does not have partial +Property A towards Uc. Hence without loss of generality, we can assume that +X(1) does not have partial Property A towards βX(1) \ U(L1) thanks to Proposition +8.13. Note that this implies that U(L1) � βX(1). It is clear that L1 is also countably +generated, and hence according to Lemma 7.6 there exists a countable subset +{Y1, Y2, · · · , Yn, · · · } of L1 such that +L1 = {Z ⊆ X(1) : ∃ n ∈ N such that Z ⊆ Yn}. +In the sequel, we fix such a sequence {Y1, Y2, · · · , Yn, · · · }. +Due to Lemma 8.11, we know that there exist c ∈ (0, 1) and R > 0 such that for +any Y ∈ L1 and S > 0, there exists T ∈ CR +u[X(1) \ Y] with ∥T∥ = 1 satisfying: for any +ξ ∈ ℓ2(X(1)) with diam(suppξ) ≤ S and ∥ξ∥ = 1, then ∥Tξ∥ < c. We call such an +operator (R, c, S, Y)-localised. Replacing T by T∗T (and R by 2R and c by √c), we +see that there exist c ∈ (0, 1) and R > 0 such that for any Y ∈ L1 and S > 0, there +exists a positive (R, c, S, Y)-localised operator of norm one. Let us fix such c and R +in the rest of the proof, and set κ := +2c +1+c < 1. +Note that X(1) can be decomposed into: +X(1) := +� +m∈N +Xm +where each Xm is finite and d(Xm, Xn) > R for any n � m. Hence each T ∈ CR +u[X(1)] +splits as a block diagonal sum of finite rank operators T = +� +m T(m) where T(m) ∈ +B(ℓ2(Xm)), with respect to this decomposition. +Take S1 = 1. By assumption, there exists a positive (R, c, S1, Y1)-localised op- +erator T ∈ CR +u[X(1) \ Y1] with norm 1. +Note that ∥T∥ = supm ∥T(m)∥, and then +there exists m1 ∈ N such that ∥T(m1)∥ > 1+c +2 . We set T1 := T(m1)/∥T(m1)∥ and denote +B1 := Xm1 ∩ (X(1) \ Y1), which is nonempty since T has support in B1 × B1 by as- +sumption. Then for any ξ ∈ ℓ2(X(1)) with ∥ξ∥ = 1 and diam(supp(ξ)) ≤ S1, we + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +41 +have +∥T1ξ∥ ≤ +2c +1 + c = κ < 1. +Now we take S2 > max +� +diam +� � +k≤m1 Xk +� +, 2 +� +. +By assumption, there exists a +positive (R, c, S2, Y2)-localised operator T ∈ CR +u[X(1) \ Y2] with norm 1. Again there +exists m2 such that ∥T(m2)∥ > 1+c +2 , which forces m2 > m1. We set T2 := T(m2)/∥T(m2)∥ +and denote B2 := Xm2 ∩(X(1)\Y2), which is nonempty since T has support in B2×B2 +by assumption. Similarly for any ξ ∈ ℓ2(X(1)) with ∥ξ∥ = 1 and diam(supp(ξ)) ≤ S2, +we have ∥T2ξ∥ ≤ κ. +Inductively, we can construct a sequence (Tn) in Cu[X(1)] ⊆ Cu[X], a sequence +(Bn) of finite subsets of X(1) ⊆ X and a sequence (Sn) of positive real numbers +satisfying condition (a)-(e) in the statement. Furthermore for each Z ∈ L(U), there +exists Yn containing Z∩X(1) for some n. By construction, we know that Bn ∩Yn = ∅ +and Bn ⊆ X(1). Hence we have: +Bn ∩ Z = Bn ∩ X(1) ∩ Z ⊆ Bn ∩ Yn = ∅, +which provides condition (f) and concludes the proof. +□ +Remark 8.18. Comparing Lemma 8.17 with [36, Lemma 4.2], we note that condition +(f) is the only extra condition added in Lemma 8.17. +It seems hard to write +condition (f) in the language of the invariant open subset U instead of the ideal +L(U), which indicates the importance of using the notion of ideals in spaces as +mentioned in Section 1. +Now we are in the position to prove Theorem 8.16. +Proof of Theorem 8.16. “(1) ⇒ (2)” is contained in Proposition 7.2, and “(2) ⇒ (3)” +holds trivially since IG = ˜I(X) ⊆ ˜I(U). Hence it suffices to show “(3) ⇒ (1)”, and +we follow the outline of the proof for [36, Theorem 1.3]. +Assume that X does not have partial Property A towards Uc, then it follows +from Proposition 8.12 that X does not have partial ONL towards Uc. Then from +Lemma 8.17, there exist κ ∈ (0, 1), R > 0, a sequence (Tn) in Cu[X], a sequence +(Bn) of finite subsets of X and a sequence (Sn) of positive real numbers satisfying +condition (a)-(f) therein. Now we consider the operator +T := +� +n +Tn, +which is a positive operator in Cu[X] of norm one. +Now we take a continuous function f : [0, 1] → [0, 1] such that supp f ⊆ [1+κ +2 , 1] +and f(1) = 1. Consider the operator f(T) ∈ C∗ +u(X), which is positive, norm one, +and admits a decomposition +f(T) = +� +n +f(Tn), +where each f(Tn) ∈ B(ℓ2(Bn)). We will show that f(T) ∈ ˜I(X) \ I(U), and hence +conclude a contradiction. +First we show that f(T) � I(U). Recall from (4.2) that +I(U) = {T′ ∈ Cu[X] : supp(T′) ⊆ Y × Y for some Y ∈ L(U)}, + +42 +QIN WANG AND JIAWEN ZHANG +Now for any T′ ∈ Cu[X] with supp(T′) ⊆ Y × Y for some Y ∈ L(U), condition (f) in +Lemma 8.17 implies that there exists n such that Bn ∩ Y = ∅. Hence we have: +∥ f(T) − T′∥ ≥ ∥χBn f(T)χBn − χBnT′χBn∥ = ∥ f(Tn) − 0∥ = 1, +which implies that f(T) � I(U). +On the other hand, using the same argument as for [36, Theorem 1.3] (since +Lemma 8.17 provides all the conditions required in [36, Lemma 4.2]), we obtain +that f(T) is a ghost operator. Hence according to Example 5.5, we have f(T) ∈ ˜I(X). +Therefore, we conclude the proof. +□ +9. Open questions +Here we collect several open questions around this topic. +First recall from Theorem 5.4 that for a space (X, d), any ideal I in the uniform +Roe algebra C∗ +u(X) must lie between I(U) and ˜I(U) for U = U(I). However, the +structure of the lattice +IU = {I is an ideal in C∗ +u(X) : U(I) = U} +in (1.1) is still unclear. Note that for any invariant open subset V ⊇ U of βX, the +ideal I(V)∩ ˜I(U) belongs to the lattice IU. Unfortunately, we do not know whether +these ideals can bust every element in IU. Hence we pose the following: +Question 9.1. Let (X, d) be a space and U ⊆ βX be an invariant open subset. Can we +describe elements in the lattice IU = {I is an ideal in C∗ +u(X) : U(I) = U} in details? For +I ∈ IU, can we find an invariant open subset V ⊇ U such that I = I(V) ∩ ˜I(U)? +Note that an answer to the above question together with Theorem 5.4 will +provide a full description for the ideal structure of the uniform Roe algebra. +Our next question concerns minimal points discussed in Section 6.1. Recall that +minimal points in the Stone- ˇCech boundary correspond to maximal ideals in the +uniform Roe algebra. However as shown in Theorem 6.8, there exist a number of +non-minimal points in the boundary. Hence it would be interesting to explore a +practical approach to distinguish minimal points. +Question 9.2. Given a space (X, d), can we find a practical approach to distinguish +minimal points in the Stone- ˇCech boundary ∂βX? +Note from Theorem 6.8 that the answer might not be easy even in the elementary +case that X = Z. +Our last questions concern the assumption of countably generatedness used +in Section 7 and 8. Recall that in Proposition 7.8 we prove that the inclusion +ιU : I(U) ֒→ ˜I(U) induces an isomorphism in K-theory when the space is coarsely +embeddable and U is countably generated. We ask the following: +Question 9.3. Does the inclusion ιU : I(U) ֒→ ˜I(U) induce an isomorphism in K-theory +for coarsely embeddable X without the assumption that U is countably generated? +Also note that even under the assumption of countable generatedness, we are +merely able to show that (ιI)∗ : K∗(I ∩ Cu[X]) −→ K∗(I) is injective in Proposition +7.8. Hence we also pose the following: + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +43 +Question 9.4. For an ideal I in C∗ +u(X) with U(I) countably generated and X coarsely +embeddable, is the map (ιI)∗ : K∗(I ∩ Cu[X]) −→ K∗(I) surjective for ∗ = 0, 1? +Our final question is designed for Theorem 8.16. Recall that the assumption of +countably generatedness plays an important role in Lemma 8.17, which is in turn +crucial in the proof of Theorem 8.16. Hence we ask the following: +Question 9.5. Let (X, d) be a space and U ⊆ βX be an invariant open subset. +If +˜I(U) = I(U), can we deduce that X has partial Property A towards βX \ U? +Appendix A. Ultrafilters +Here we collect some basic knowledge on ultrafilters, which is used throughout +the paper. The material should be fairly well-known (see, e.g., [9, Appendix A], +[34, Chapter 7.4] or [43, Appendix A]). While some of the results might not be +standard and we have not dug into the reference, we include the proofs for +convenience to readers. +Definition A.1. Let X be a set and P(X) be its power set. An ultrafilter on X is a +family U ⊆ P(X) satisfying the following: +(I.1) ∅ � U; +(I.2) for A, B ∈ U, then A ∩ B ∈ U; +(I.3) for A ∈ U and A ⊆ B, then B ∈ U; +(I.4) for any A ⊆ X, either A ∈ U or X \ A ∈ U. +For an ultrafilter U on X, we can associate a function ω : P(X) → {0, 1} by +setting ω(A) = 1 if and only if A ∈ U. It follows from (I.1)-(I.4) above that ω is a +finitely additive {0, 1}-valued probability measure on P(X). Conversely for such +a function ω on P(X), we can associate a family Uω := {A ⊆ X : ω(A) = 1}. It is +clear that Uω is an ultrafilter on X, and these two procedures are inverse to each +other. Therefore throughout the paper, we slide between these two notions freely +without further explanation. +For a ∈ X, it is clear that the family {A ∈ P(X) : a ∈ A} is an ultrafilter on +X. Such an ultrafiler is called principal. An ultrafilter which is not principal is +called non-principal. An argument using Zorn’s lemma shows that non-principal +ultrafilters always exist whenever X is infinite. +The following is well-known (see, e.g., [43, Lemma A.2]): +Lemma A.2. Let ω be an ultrafilter on a set X, and D ⊆ X with ω(D) = 1. Let f : D → Y +be a function from D to a compact Hausdorff topological space Y. Then there exists a +unique point y ∈ Y such that for any open neighbourhood U of y, we have ω(f −1(U)) = 1. +Definition A.3. The unique point in Lemma A.2 is called the ultralimit of f along +ω or the ω-limit of f, denoted by limω f or lima→ω f(a). +We record the following localisation result: +Lemma A.4. Let U be an ultrafilter on a set X, and A ⊆ X with A ∈ U. Then we have: +(1) {S ∩ A : S ∈ U} = {S ⊆ A : S ∈ U} is an ultrafilter on A, denoted by UA. +(2) U = {S ⊆ X : S ∩ A ∈ UA} = {S ⊆ X : ∃ S′ ∈ UA such that S′ ⊆ S}. + +44 +QIN WANG AND JIAWEN ZHANG +Proof. (1). It follows from Definition A.1 that {S ∩ A : S ∈ U} = {S ⊆ A : S ∈ U}, +and hence (I.1)-(I.3) hold for UA. Concerning (I.4): given B ⊆ A, if B ∈ U then +B ∈ UA as well; if B � U then X \ B ∈ U, and hence A \ B = A ∩ (X \ B) ∈ UA. +(2). This is straightforward, hence omitted. +□ +We can also extend an ultrafilter on a subset to the whole space. The proof is +straightforward, hence omitted. +Lemma A.5. Let Y be a subset of a set X, and U0 an ultrafilter on Y. Define +U := {S ⊆ X : S ∩ Y ∈ U0}. +Then U is an ultrafilter on X. +The following result provides an approach to combine a family of ultrafilters +into a single one: +Lemma A.6. Let {Xi}i∈I be a family of sets, and Ui be an ultrafilter on Xi for each i ∈ I. +Let ω0 be an ultrafilter on I. Consider the set X := � +i∈I Xi and define: +U := +� � +i∈I +Ai ⊆ +� +i∈I +Xi : ∃ J ⊆ I with ω0(J) = 1 such that ∀i ∈ J, Ai ∈ Ui +� +. +Then U is an ultrafilter on X. +Proof. Firstly, it is clear that ∅ � U. Assume that � +i∈I Ai and � +i∈I Bi ∈ U, i.e., there +exist JA, JB ⊆ I with ω0(JA) = ω0(JB) = 1 such that Ai ∈ Ui for any i ∈ JA and Bi ∈ Ui +for any i ∈ JB. Consider (� +i∈I Ai) ∩ (� +i∈I Bi) = � +i∈I(Ai ∩ Bi) and J = JA ∩ JB. Then +ω0(J) = 1 and for each i ∈ J, Ai and Bi are in Ui. This implies that Ai ∩ Bi ∈ Ui, +which concludes (I.2). +It is clear that (I.3) holds for U and finally, we consider (I.4). +Given A = +� +i∈I Ai ⊆ X, denote J := {i ∈ I : Ai ∈ Ui}. If ω0(J) = 1, then it follows that A ∈ U. +Otherwise, assume that ω0(J) = 0. Then we consider X \ A = � +i∈I(Xi \ Ai). Then +I \ J = {i ∈ I : Xi \ Ai ∈ Ui} and ω0(I \ J) = 1, which implies that X \ A ∈ U and +concludes the proof. +□ +Recall that ultrafilters can also be characterised by the Stone- ˇCech compactifi- +cation. More precisely, we have the following (see, e.g., [34, Chapter 7.4]): +Lemma A.7. Let X be a set and βX be the Stone- ˇCech compactification of X. +(1) Given ω ∈ βX, the family {A ⊆ X : ω ∈ A} is an ultrafilter on X. +(2) Given an ultrafilter U on X, the intersection � +A∈U A consists of a single point. +The procedures above are inverse to each other, and hence βX can be characterised by +ultrafilters on X. Moreover, points in ∂βX correspond to non-principal ultrafilters. +Thanks to Lemma A.7, we will also use ultrafilters and points in the Stone- ˇCech +compactification freely without further explanation throughout the paper. +For convenience, we also record that for D ⊆ X, its closure D in βX satisfies: +D = {ω ∈ βX : ω(D) = 1} +and D is clopen in βX. The topology of βX is generated by {D : D ⊆ X}. + +GHOSTLY IDEALS IN UNIFORM ROE ALGEBRAS +45 +Finally we recall the following, which can be proved either directly using the +universal property of the Stone- ˇCech compactification or deduced directly from +Lemma A.4, A.5 and A.7: +Corollary A.8. For a subset Z ⊆ X, the closure Z in βX is homeomorphic to βZ. +References +[1] Claire Anantharaman-Delaroche and Jean Renault. Amenable groupoids, volume 36 of Mono- +graphies de L’Enseignement Math´ematique. L’Enseignement Math´ematique, Geneva, 2000. With +a foreword by Georges Skandalis and Appendix B by E. Germain. +[2] 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. +[3] Paul Baum and Alain Connes. Geometric K-theory for Lie groups and foliations. Enseign. +Math., 46(1/2):3–42, 2000 (firstly circulated in 1982). +[4] Paul Baum, Alain Connes, and Nigel Higson. Classifying space for proper actions and K- +theory of group C∗-algebras. Contemporary Mathematics, 167:241–241, 1994. +[5] 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. +[6] 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. +[7] Bruno M. Braga, Ilijas Farah, and Alessandro Vignati. Embeddings of uniform Roe algebras. +Comm. Math. Phys., 377(3):1853–1882, 2020. +[8] Bruno M. Braga, Ilijas Farah, and Alessandro Vignati. General uniform Roe algebra rigidity. +Ann. Inst. Fourier (Grenoble), 72(1):301–337, 2022. +[9] Nathanial P. Brown and Narutaka Ozawa. C∗-algebras and finite-dimensional approximations, +volume 88 of Graduate Studies in Mathematics. Amer. Math. Soc., Providence, RI, 2008. +[10] Xiaoman Chen, Romain Tessera, Xianjin Wang, and Guoliang Yu. Metric sparsification and +operator norm localization. Adv. Math., 218(5):1496–1511, 2008. +[11] Xiaoman Chen and Qin Wang. Notes on ideals of Roe algebras. Q. J. Math., 52(4):437–446, +2001. +[12] Xiaoman Chen and Qin Wang. Ideal structure of uniform Roe algebras of coarse spaces. J. +Funct. Anal., 216(1):191 – 211, 2004. +[13] Xiaoman Chen and Qin Wang. Ideal structure of uniform Roe algebras over simple cores. +Chinese Ann. Math. Ser. B, 25(2):225–232, 2004. +[14] Xiaoman Chen and Qin Wang. Ghost ideals in uniform Roe algebras of coarse spaces. Arch. +Math. (Basel), 84(6):519–526, 2005. +[15] Xiaoman Chen and Qin Wang. Rank distributions of coarse spaces and ideal structure of Roe +algebras. Bull. London Math. Soc., 38(5):847–856, 2006. +[16] Xiaoman Chen, Qin Wang, and Xianjin Wang. Operator norm localization property of metric +spaces under finite decomposition complexity. J. Funct. Anal., 257(9):2938–2950, 2009. +[17] Xiaoman Chen, Qin Wang, and Guoliang Yu. The maximal coarse Baum–Connes conjecture +for spaces which admit a fibred coarse embedding into Hilbert space. Adv. Math., 249:88–130, +2013. +[18] Eske Ellen Ewert and Ralf Meyer. Coarse geometry and topological phases. Comm. Math. +Phys., 366(3):1069–1098, 2019. +[19] Martin Finn-Sell. Fibred coarse embeddings, a-T-menability and the coarse analogue of the +Novikov conjecture. J. Funct. Anal., 267(10):3758–3782, 2014. +[20] Martin Finn-Sell and Nick Wright. Spaces of graphs, boundary groupoids and the coarse +Baum-Connes conjecture. Adv. Math., 259:306–338, 2014. +[21] Nigel Higson, Vincent Lafforgue, and Georges Skandalis. Counterexamples to the Baum- +Connes conjecture. Geom. Funct. Anal., 12(2):330–354, 2002. +[22] Nigel Higson and John Roe. On the coarse Baum-Connes conjecture. In Novikov conjectures, +index theorems and rigidity, Vol. 2 (Oberwolfach, 1993), volume 227 of London Math. Soc. Lecture +Note Ser., pages 227–254. Cambridge Univ. Press, Cambridge, 1995. + +46 +QIN WANG AND JIAWEN ZHANG +[23] Nigel Higson, John Roe, and Guoliang Yu. A coarse Mayer-Vietoris principle. Math. Proc. +Cambridge Philos. Soc., 114(1):85–97, 1993. +[24] Gennadi Kasparov and Guoliang Yu. The coarse geometric Novikov conjecture and uniform +convexity. Adv. Math., 206(1):1–56, 2006. +[25] Kang Li, J´an ˇSpakula, and Jiawen Zhang. Measured asymptotic expanders and rigidity for +roe algebras. arXiv:2010.10749, to appear in Int. Math. Res. Not., 2020. +[26] Paul S. Muhly, Jean. N. Renault, and Dana P. Williams. Continuous-trace groupoid C∗- +algebras. III. Trans. Amer. Math. Soc., 348:3621–3641, 1996. +[27] Piotr W. Nowak and Guoliang Yu. Large scale geometry. EMS Textbooks in Mathematics. +European Mathematical Society (EMS), Z¨urich, 2012. +[28] Jean Renault. A Groupoid Approach to C∗-Algebras, volume 793 of Lecture Notes in Mathematics. +Springer-Verlag, 1980. +[29] Jean Renault. The ideal structure of groupoid crossed product C∗-algebras. J. Operator Theory, +25(1):3–36, 1991. With an appendix by Georges Skandalis. +[30] Steffen Roch. Ideals of band-dominated operators. Complex Var. Elliptic Equ., 67(3):701–715, +2022. +[31] John Roe. An index theorem on open manifolds. I, II. J. Differential Geom., 27(1):87–113, 115– +136, 1988. +[32] John Roe. Coarse cohomology and index theory on complete Riemannian manifolds, volume 497. +Amer. Math. Soc., 1993. +[33] John Roe. Index theory, coarse geometry, and topology of manifolds, volume 90. Amer. Math. Soc., +1996. +[34] John Roe. Lectures on coarse geometry, volume 31 of University Lecture Series. Amer. Math. Soc., +Providence, RI, 2003. +[35] John Roe. Band-dominated Fredholm operators on discrete groups. Integr. Equ. Oper. Theory, +51(3):411–416, 2005. +[36] John Roe and Rufus Willett. Ghostbusting and property A. J. Funct. Anal., 266(3):1674–1684, +2014. +[37] Hiroki Sako. Property A and the operator norm localization property for discrete metric +spaces. J. Reine Angew. Math. (Crelle’s Journal), 2014(690):207–216, 2014. +[38] Aidan Sims. ´Etale groupoids and their C∗-algebras. arxiv preprint arXiv:1710.10897, 2017. +[39] Georges Skandalis, Jean-Louis Tu, and Guoliang Yu. The coarse Baum-Connes conjecture and +groupoids. Topology, 41(4):807–834, 2002. +[40] Jean-Louis Tu. La conjecture de Baum-Connes pour les feuilletages moyennables. K-Theory, +17(3):215–264, 1999. +[41] J´an ˇSpakula. Uniform K-homology theory. J. Funct. Anal., 257(1):88–121, 2009. +[42] J´an ˇSpakula and Rufus Willett. On rigidity of Roe algebras. Adv. Math., 249:289–310, 2013. +[43] J´an ˇSpakula and Rufus Willett. A metric approach to limit operators. Trans. Amer. Math. Soc., +369(1):263–308, 2017. +[44] Qin Wang. Remarks on ghost projections and ideals in the Roe algebras of expander se- +quences. Arch. Math. (Basel), 89(5):459–465, 2007. +[45] Rufus Willett. Some notes on property A. In Limits of graphs in group theory and computer +science, pages 191–281. EPFL Press, Lausanne, 2009. +[46] Rufus Willett and Guoliang Yu. Higher index theory for certain expanders and Gromov +monster groups, I. Adv. Math., 229(3):1380–1416, 2012. +[47] Guoliang Yu. The coarse Baum-Connes conjecture for spaces which admit a uniform embed- +ding into Hilbert space. Invent. Math., 139(1):201–240, 2000. +(Q. Wang) Research Center for Operator Algebras, and Shanghai Key Laboratory of Pure +Mathematics and Mathematical Practice, School of Mathematical Sciences, East China Nor- +mal University, Shanghai, 200241, China. +Email address: qwang@math.ecnu.edu.cn +(J. Zhang) School of Mathematical Sciences, Fudan University, 220 Handan Road, Shanghai, +200433, China. +Email address: jiawenzhang@fudan.edu.cn +