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

Asymptotic dimension and coarse embeddings in the quantum setting

Javier Alejandro Chávez-Domínguez Department of Mathematics, University of Oklahoma, Norman, OK 73019-3103, USA [email protected]  and  Andrew T. Swift Department of Mathematics, University of Oklahoma, Norman, OK 73019-3103, USA [email protected]
Abstract.

We generalize the notions of asymptotic dimension and coarse embeddings from metric spaces to quantum metric spaces in the sense of Kuperberg and Weaver [KW12]. We show that quantum asymptotic dimension behaves well with respect to metric quotients and direct sums, and is preserved under quantum coarse embeddings. Moreover, we prove that a quantum metric space that equi-coarsely contains a sequence of expanders must have infinite asymptotic dimension. This is done by proving a quantum version of a vertex-isoperimetric inequality for expanders, based upon a previously known edge-isoperimetric one from [TKR+10].

Key words and phrases:
Quantum metric spaces; Asymptotic dimension; Quantum expanders.
2010 Mathematics Subject Classification:
Primary: 46L52; Secondary: 54F45, 46L51, 46L65, 81R60
The first-named author was partially supported by NSF grant DMS-1900985.

1. Introduction

In [KW12] the authors explore a generalization of metric spaces, called quantum metric spaces, which is related to the quantum graphs of quantum information theory; and they construct many generalizations of familiar metric space concepts, including a generalization of Lipschtiz map they call a co-Lipschitz morphism. The purpose of this paper is to contribute to the structure theory of quantum metric spaces by applying ideas coming from the large-scale or coarse geometry of classical metric spaces. Specifically, we propose generalizations of coarse embeddings and asymptotic dimension, and then prove some fundamental results about them. Coarse geometry is an important area of mathematics with applications in group theory, Banach space theory, and computer science. It began as a field of study with the polynomial growth theorem of Gromov [Gro81], and the definition of coarse embedding and asymptotic dimension are also both due to Gromov [Gro93], although large-scale geometric ideas appear as early as the late 1960’s in the original proof of Mostow’s rigidity theorem [Mos68]. We refer to [NY12] for an excellent introduction to the subject.

Some notions from coarse geometry have already been explored in the noncommutative setting, see e.g. [Ban15, BM16]. There are two important differences between these works and the present paper: One is that they define and study noncommutative versions of coarse spaces whereas we are studying noncommutative metric spaces in a coarse fashion, but more importantly their approach is CC^{*}-algebraic while ours (as clearly stated by the title of [KW12]) is a von Neumann algebra one. A number of noncommutative notions of topological dimension for CC^{*}-algebras have been also studied in the literature, see e.g. [Rie83, BP91, KW04, Win07, WZ10]. Let us emphasize in particular the nuclear dimension from [WZ10], which is linked to coarse geometry: For a discrete metric space of bounded geometry, the nuclear dimension of the associated uniform Roe algebra is dominated by the asymptotic dimension of the underlying space. Once again, the approach to dimension in the present work is significantly different from the aforementioned ones because we are following the von Neumann algebra path.

In Section 2 we recall the definitions we need from [KW12], including quantum metric, distance, and diameter. In Section 3 we generalize the definition of coarse embedding to quantum metric spaces using alternative versions of the usual moduli of expansion (or uniform continuity) and compression defined for classical functions. It is shown that the moduli for classical functions and their canonically induced quantum functions coincide. In Section 4 we generalize the definition of asymptotic dimension to quantum metric spaces and show that asymptotic dimension is preserved under coarse embeddings. We show as a consequence that the asymptotic dimension of a quotient of a quantum metric space is no greater than the asymptotic dimension of the original space and that the asymptotic dimension of a direct sum of quantum metric spaces is equal to the maximum of the asymptotic dimensions of its summands. We finish Section 4 by establishing the corresponding inequality for arbitrary sums of quantum metric spaces in the case when the sum is a reflexive quantum metric space. In Section 5, we show that quantum expanders satisfy a quantum version of a vertex isoperimetric inequality. This can be used to show that a quantum metric space has infinite asymptotic dimension if it equi-coarsely contains a sequence of reflexive quantum expanders. In particular, this includes the case when the sequence of quantum expanders is induced by a sequence of classical expanders.

2. Definitions

We use the definitions of quantum metric space and related notions found in [KW12]. Just as metrics are defined on sets, quantum metrics are defined on von Neumann algebras, and classical metrics on a set XX are in a natural one-to-one correspondence with quantum metrics on (X)\ell_{\infty}(X). We view von Neumann algebras as subsets of some space ()\mathcal{B}(\mathcal{H}) of bounded linear operators on a Hilbert space \mathcal{H}. Given the von Neumann algebras (1)\mathcal{M}\subseteq\mathcal{B}(\mathcal{H}_{1}), 𝒩(2)\mathcal{N}\subseteq\mathcal{B}(\mathcal{H}_{2}), we denote by ¯𝒩\mathcal{M}\overline{\otimes}\mathcal{N} their normal spatial tensor product, that is, the weak*-closure of 𝒩\mathcal{M}\otimes\mathcal{N} in (122)\mathcal{B}(\mathcal{H}_{1}\otimes_{2}\mathcal{H}_{2}). Given a von Neumann algebra \mathcal{M}, we denote the commutant of \mathcal{M} by \mathcal{M}^{\prime}. An orthogonal projection in a von Neumann algebra will simply be called a projection.

Definition 2.1 ([KW12, Definition 2.3]):

A quantum metric on a von Neumann algebra ()\mathcal{M}\subseteq\mathcal{B}(\mathcal{H}) is a family 𝕍={𝒱t}t[0,)\mathbb{V}=\{\mathcal{V}_{t}\}_{t\in[0,\infty)} of weak*-closed subspaces of ()\mathcal{B}(\mathcal{H}) such that 𝒱0=\mathcal{V}_{0}=\mathcal{M}^{\prime} and for all t[0,)t\in[0,\infty),

  • 𝒱t\mathcal{V}_{t} is self-adjoint.

  • 𝒱s𝒱t𝒱s+t\mathcal{V}_{s}\mathcal{V}_{t}\subseteq\mathcal{V}_{s+t} for all s[0,)s\in[0,\infty).

  • 𝒱t=s>t𝒱s\mathcal{V}_{t}=\bigcap_{s>t}\mathcal{V}_{s}.

A quantum metric space is a pair (,𝕍)(\mathcal{M},\mathbb{V}) of a von Neumann algebra \mathcal{M} with a quantum metric 𝕍\mathbb{V} defined on it. We will often simply call a quantum metric space by its von Neumann algebra if there is no ambiguity regarding the quantum metric being considered.

Given a metric space (X,d)(X,d), the canonical quantum metric space associated to it [KW12, Proposition 2.5] is ((X),{𝒱t}t[0,))(\ell_{\infty}(X),\{\mathcal{V}_{t}\}_{t\in[0,\infty)}), where

𝒱t={A(2(X))Aex,ey=0 for all (x,y)d1[0,t]}\mathcal{V}_{t}=\big{\{}A\in\mathcal{B}(\ell_{2}(X))\mid\langle Ae_{x},e_{y}\rangle=0\mbox{ for all }(x,y)\notin d^{-1}[0,t]\big{\}}

for all t[0,).t\in[0,\infty). Here (ex)xX(e_{x})_{x\in X} is the canonical basis of 2(X)\ell_{2}(X) and (X)\ell_{\infty}(X) is viewed as a subset of (2(X))\mathcal{B}(\ell_{2}(X)) as diagonal operators in the standard way, that is, via the map E:(2(X))E\colon\ell_{\infty}\to\mathcal{B}(\ell_{2}(X)) defined by E(ϕ)[f](x)=ϕ(x)f(x)E(\phi)[f](x)=\phi(x)f(x) for all ϕ(X)\phi\in\ell_{\infty}(X), all f2(X)f\in\ell_{2}(X), and all xXx\in X. It is in this way that quantum metric spaces are generalizations of classical metric spaces. A natural distance function can be defined for projections in a von Neumann algebra that generalizes the notion of distance between subsets of a classical metric space.

Definition 2.2 ([KW12, Definition 2.6]):

Given a quantum metric 𝕍={𝒱t}t[0,)\mathbb{V}=\{\mathcal{V}_{t}\}_{t\in[0,\infty)} on a von Neumann algebra \mathcal{M}, the distance between two projections PP and QQ in ¯(2)\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}) is

dist𝕍(P,Q)=inf{tP(AId)Q0 for some A𝒱t}.\operatorname{dist}_{\mathbb{V}}(P,Q)=\inf\{t\mid P(A\otimes\operatorname{Id})Q\neq 0\mbox{ for some }A\in\mathcal{V}_{t}\}.

Identifying \mathcal{M} with Id¯(2)\mathcal{M}\otimes\operatorname{Id}\subseteq\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}) yields the equivalent formula

dist𝕍(P,Q)=inf{tPAQ0 for some A𝒱t}\operatorname{dist}_{\mathbb{V}}(P,Q)=\inf\{t\mid PAQ\neq 0\mbox{ for some }A\in\mathcal{V}_{t}\}

for projections P,QP,Q in \mathcal{M}. The subscript will usually be omitted if there is no ambiguity regarding the quantum metric being used.

We refer to [KW12, Definition 2.7] and [KW12, Proposition 2.8] for basic properties of quantum distance. In particular, the distance function associated to a quantum metric satisfies an analog of the triangle inequality. If \mathcal{M} is a quantum metric space and P,Q,RP,Q,R are projections in ¯(2)\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}), then dist(P,Q)dist(P,R)+sup{dist(R~,Q)RR~0}\operatorname{dist}(P,Q)\leq\operatorname{dist}(P,R)+\sup\{\operatorname{dist}(\tilde{R},Q)\mid R\tilde{R}\neq 0\}, where R~\tilde{R} ranges over projections in ¯(2)\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}). Note that the same proof found in [KW12, Proposition 2.8] shows that if P,Q,RP,Q,R are projections in \mathcal{M}, then R~\tilde{R} may be taken to range only over projections in \mathcal{M}.

It is not hard to see that if ((X),𝕍)(\ell_{\infty}(X),\mathbb{V}) is the canonical quantum metric space associated to metric space (X,d)(X,d), then dist𝕍(χS,χT)=d(S,T)\operatorname{dist}_{\mathbb{V}}(\chi_{S},\chi_{T})=d(S,T) for any subsets S,TS,T of XX. Likewise, the following definitions for diameter and open ε\varepsilon-neighborhood of a projection generalize the corresponding notions for a subset.

Definition 2.3 ([KW12, Proposition 2.16]):

Given a quantum metric 𝕍={𝒱t}t[0,)\mathbb{V}=\{\mathcal{V}_{t}\}_{t\in[0,\infty)} on a von Neumann algebra \mathcal{M}, the diameter of a nonzero projection PP in \mathcal{M} is

diam𝕍(P)=sup{dist𝕍(Q,R)Q(PAPId)R0 for some A()}.\operatorname{diam}_{\mathbb{V}}(P)=\sup\big{\{}\operatorname{dist}_{\mathbb{V}}(Q,R)\mid Q(PAP\otimes\operatorname{Id})R\neq 0\mbox{ for some }A\in\mathcal{B}(\mathcal{H})\big{\}}.

The diameter of the zero projection is defined to be zero. The subscript will usually be omitted if there is no ambiguity regarding the quantum metric being used.

We remark that [KW12] only considers the case of PP being the identity of \mathcal{M} in the definition above, so our notion is really a generalization of theirs.

Definition 2.4 ([KW12, Proposition 2.17]):

Given a quantum metric 𝕍={𝒱t}t[0,)\mathbb{V}=\{\mathcal{V}_{t}\}_{t\in[0,\infty)} on a von Neumann algebra \mathcal{M}, the open ε\varepsilon-neighborhood of a projection PP in \mathcal{M} is the projection

(P)ε=Id{Qdist𝕍(P,Q)ε}.(P)_{\varepsilon}=\operatorname{Id}-\bigvee\big{\{}Q\in\mathcal{M}\mid\operatorname{dist}_{\mathbb{V}}(P,Q)\geq\varepsilon\big{\}}.

It is easy to see from [KW12, Definition 2.14 (b)] that ((P)ε)δ(P)ε+δ((P)_{\varepsilon})_{\delta}\leq(P)_{\varepsilon+\delta} for all ε,δ>0\varepsilon,\delta>0 and projections PP in a quantum metric space \mathcal{M}.

Our next definition follows [Kor11].

Definition 2.5:

A function ϕ:𝒩\phi\colon\mathcal{M}\to\mathcal{N} between two von Neumann algebras \mathcal{M} and 𝒩\mathcal{N} is called a quantum function if ϕ\phi is a unital weak*-continuous *-homomorphism.

Quantum functions generalize classical functions in the following way: If f:XYf\colon X\to Y is a classical function between sets, then ϕf:(Y)(X)\phi_{f}\colon\ell_{\infty}(Y)\to\ell_{\infty}(X), defined by ϕf(g)=gf\phi_{f}(g)=g\circ f for all g(Y)g\in\ell_{\infty}(Y), is the canonical quantum function associated to ff between the canonical von Neumann algebras associated to YY and XX, and is such that ϕf(χ{y})=χ{f1(y)}\phi_{f}(\chi_{\{y\}})=\chi_{\{f^{-1}(y)\}} for all yYy\in Y.

In what follows, we will often identify a von Neumann algebra \mathcal{M} with Id¯(2)\mathcal{M}\otimes\operatorname{Id}\subseteq\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}) and a quantum function ϕ:𝒩\phi\colon\mathcal{M}\to\mathcal{N} with (ϕId):¯(2)𝒩¯(2)(\phi\otimes\operatorname{Id})\colon\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2})\to\mathcal{N}\overline{\otimes}\mathcal{B}(\ell_{2}).

Recall [KW12, p. 17] that two projections P,Q¯(2)P,Q\in\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}) are said to be unlinkable if there exist P~,Q~Id(2)\tilde{P},\tilde{Q}\in\operatorname{Id}\otimes\mathcal{B}(\ell_{2}) satisfying PP~P\leq\tilde{P}, QQ~Q\leq\tilde{Q} and P~Q~=0\tilde{P}\tilde{Q}=0; otherwise, they are said to be linkable. By [KW12, Prop. 2.13], for a von Neumann algebra (H)\mathcal{M}\subseteq\mathcal{B}(H), two projections P,Q¯(2)P,Q\in\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}) are linkable if and only if there exists A(H)A\in\mathcal{B}(H) such that P(AId)Q0P(A\otimes\operatorname{Id})Q\not=0.

Note that we may generalize these notions to non-projections: two operators S,T¯(2)S,T\in\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}) are said to be linkable if and only if there exists A(H)A\in\mathcal{B}(H) such that S(AId)T0S(A\otimes\operatorname{Id})T\not=0; obviously, this is the same as saying that the source projection S\llbracket S\rrbracket of SS and the range projection [T][T] of TT are linkable. By the aforementioned characterization, two operators S,T¯(2)S,T\in\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}) are not linkable (i.e. unlinkable) if and only if there exist projections S~,T~Id(2)\tilde{S},\tilde{T}\in\operatorname{Id}\otimes\mathcal{B}(\ell_{2}) satisfying SS~\llbracket S\rrbracket\leq\tilde{S}, [T]T~[T]\leq\tilde{T} and S~T~=0\tilde{S}\tilde{T}=0. Note that this is closely related to our definition of diameter of a projection: For a projection PP in \mathcal{M}, its diameter is the supremum of the distances dist(Q,R)\operatorname{dist}(Q,R) where Q,R¯(2)Q,R\in\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}) are projections such that Q(PId)Q(P\otimes\operatorname{Id}) and (PId)R(P\otimes\operatorname{Id})R are linkable.

The following lemma shows that when ϕ\phi is a quantum function, ϕId\phi\otimes\operatorname{Id} maps unlinkable pairs of operators to unlinkable pairs of operators.

Lemma 2.6.

Let ϕ:𝒩\phi:\mathcal{M}\to\mathcal{N} be a quantum function between von Neumann algebras. If S,T¯(2)S,T\in\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}) are unlinkable operators, then (ϕId)S,(ϕId)T𝒩¯(2)(\phi\otimes\operatorname{Id})S,(\phi\otimes\operatorname{Id})T\in\mathcal{N}\overline{\otimes}\mathcal{B}(\ell_{2}) are unlinkable as well.

Proof.

As above, let S~,T~Id(2)\tilde{S},\tilde{T}\in\operatorname{Id}\otimes\mathcal{B}(\ell_{2}) be projections satisfying SS~\llbracket S\rrbracket\leq\tilde{S}, [T]T~[T]\leq\tilde{T}, and S~T~=0\tilde{S}\tilde{T}=0. Note that since ϕId\phi\otimes\operatorname{Id} is a *-homomorphism and thus preserves order, (ϕId)S~,(ϕId)T~Id(2)(\phi\otimes\operatorname{Id})\tilde{S},(\phi\otimes\operatorname{Id})\tilde{T}\in\operatorname{Id}\otimes\mathcal{B}(\ell_{2}) are projections and they satisfy (ϕId)S(ϕId)S~(\phi\otimes\operatorname{Id})\llbracket S\rrbracket\leq(\phi\otimes\operatorname{Id})\tilde{S}, (ϕId)[T](ϕId)T~(\phi\otimes\operatorname{Id})[T]\leq(\phi\otimes\operatorname{Id})\tilde{T}, and (ϕId)S~(ϕId)T~=0(\phi\otimes\operatorname{Id})\tilde{S}(\phi\otimes\operatorname{Id})\tilde{T}=0. The desired result will follow once we prove that (ϕId)S=(ϕId)S(\phi\otimes\operatorname{Id})\llbracket S\rrbracket=\llbracket(\phi\otimes\operatorname{Id})S\rrbracket and (ϕId)[T]=[(ϕId)T](\phi\otimes\operatorname{Id})[T]=[(\phi\otimes\operatorname{Id})T].

By our definition of quantum function, ϕ\phi is a weak* to weak* continuous *-homomorphism. By [Bla06, Prop. III.2.2.2], ϕ\phi is normal. Since the tensor product of normal completely positive contractions is again a normal positive contraction [Bla06, III.2.2.5], ϕId\phi\otimes\operatorname{Id} is normal. By [Bla06, Prop. III.2.2.2] again, ϕId\phi\otimes\operatorname{Id} is σ\sigma-strong to σ\sigma-strong continuous from ¯(2)\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}) to 𝒩¯(2)\mathcal{N}\overline{\otimes}\mathcal{B}(\ell_{2}). It is well-known that the σ\sigma-strong and the strong topologies coincide on bounded sets [Bla06, I.3.1.4], so in particular it follows that ϕId\phi\otimes\operatorname{Id} maps bounded strongly convergent nets to bounded strongly convergent nets.

Now, it is known that (SS)αS(S^{*}S)^{\alpha}\to\llbracket S\rrbracket strongly as α0\alpha\to 0 [Bla06, I.5.2.1]. Since {(SS)α}α(0,1)\{(S^{*}S)^{\alpha}\}_{\alpha\in(0,1)} is norm bounded (which can be easily shown using functional calculus), we conclude that

(ϕId)((SS)α)α0(ϕId)S(\phi\otimes\operatorname{Id})\big{(}(S^{*}S)^{\alpha}\big{)}\xrightarrow[\alpha\to 0]{}(\phi\otimes\operatorname{Id})\llbracket S\rrbracket

strongly. Since (ϕId)(\phi\otimes\operatorname{Id}) is a unital *-homomorphism we have (ϕId)((SS)α)=(((ϕId)S)((ϕId)S))α(\phi\otimes\operatorname{Id})\big{(}(S^{*}S)^{\alpha}\big{)}=\big{(}\big{(}(\phi\otimes\operatorname{Id})S\big{)}^{*}\big{(}(\phi\otimes\operatorname{Id})S\big{)}\big{)}^{\alpha} (again this can be easily shown using functional calculus). Now,

(((ϕId)S)((ϕId)S))αα0(ϕId)S\big{(}\big{(}(\phi\otimes\operatorname{Id})S\big{)}^{*}\big{(}(\phi\otimes\operatorname{Id})S\big{)}\big{)}^{\alpha}\xrightarrow[\alpha\to 0]{}\llbracket(\phi\otimes\operatorname{Id})S\rrbracket

strongly, and therefore (ϕId)S=(ϕId)S\llbracket(\phi\otimes\operatorname{Id})S\rrbracket=(\phi\otimes\operatorname{Id})\llbracket S\rrbracket. The analogous conclusion for the range projection of TT follows from the fact that [T]=TT[T]=\llbracket TT^{*}\rrbracket [Bla06, I.5.2.1]. ∎

Some of our results will only apply to quantum metric spaces that are (operator) reflexive; we recall the definition below.

Definition 2.7 ([KW12, Defns. 1.5 and 2.23]):

A subspace 𝒱(H)\mathcal{V}\subseteq\mathcal{B}(H) is (operator) reflexive if 𝒱={B(H)P𝒱Q={0}PBQ=0}\mathcal{V}=\big{\{}B\in\mathcal{B}(H)\mid P\mathcal{V}Q=\{0\}\Rightarrow PBQ=0\big{\}} with PP and QQ ranging over projections in (H)\mathcal{B}(H). A quantum metric 𝕍={𝒱t}t[0,)\mathbb{V}=\{\mathcal{V}_{t}\}_{t\in[0,\infty)} is called reflexive if 𝒱t\mathcal{V}_{t} is reflexive for each t[0,)t\in[0,\infty).

3. Quantum moduli of expansion and compression

In this section, we define coarse embeddings between quantum metric spaces using moduli and then show how this relates to the definitions of co-Lipschitz and co-isometric morphisms found in [KW12].

Recall that if f:XYf\colon X\to Y is a map between metric spaces, we define its modulus of expansion ωf\omega_{f} by

ωf(t)=sup{dY(f(x),f(y))dX(x,y)t}\omega_{f}(t)=\sup\big{\{}d_{Y}(f(x),f(y))\mid d_{X}(x,y)\leq t\big{\}}

and its modulus of compression ρf\rho_{f} by

ρf(t)=inf{dY(f(x),f(y))dX(x,y)t}\rho_{f}(t)=\inf\big{\{}d_{Y}(f(x),f(y))\mid d_{X}(x,y)\geq t\big{\}}

for all t0t\geq 0. We say that ff is expanding if limtρf(t)=\lim_{t\to\infty}\rho_{f}(t)=\infty, and coarse if ωf(t)<\omega_{f}(t)<\infty for all t0t\geq 0. We say that ff is a coarse embedding if ff is both coarse and expanding.

For our purposes, we will use alternative versions of these moduli. Let

ω~f(t)=inf{dX(x,y)dY(f(x),f(y))t}\tilde{\omega}_{f}(t)=\inf\big{\{}d_{X}(x,y)\mid d_{Y}(f(x),f(y))\geq t\big{\}}

and let

ρ~f(t)=sup{dX(x,y)dY(f(x),f(y))t}.\tilde{\rho}_{f}(t)=\sup\big{\{}d_{X}(x,y)\mid d_{Y}(f(x),f(y))\leq t\big{\}}.

The following observation is surely well-known.

Lemma 3.1.

Let f:XYf\colon X\to Y be a map between metric spaces. Then:

  1. (a)

    ff is coarse if and only if limtω~f(t)=\lim_{t\to\infty}\tilde{\omega}_{f}(t)=\infty.

  2. (b)

    ff is expanding if and only if ρ~f(t)<\tilde{\rho}_{f}(t)<\infty for all t0t\geq 0.

Proof.

Note that ω~\tilde{\omega} is an increasing function, and therefore limtω~f(t)=\lim_{t\to\infty}\tilde{\omega}_{f}(t)=\infty if and only if ω~f\tilde{\omega}_{f} is unbounded.

Suppose first that ff is not coarse so that by definition there exists t0t\geq 0 such that ωf(t)=\omega_{f}(t)=\infty. Then for each nn\in{\mathbb{N}} there exist xn,ynXx_{n},y_{n}\in X such that dY(f(xn),f(yn))nd_{Y}(f(x_{n}),f(y_{n}))\geq n and dX(xn,yn)td_{X}(x_{n},y_{n})\leq t. This implies ω~f(n)t\tilde{\omega}_{f}(n)\leq t, so ω~f\tilde{\omega}_{f} is bounded above by tt.

Suppose now that ω~f\tilde{\omega}_{f} is bounded above by tt. Then for each nn\in{\mathbb{N}}, there exist xn,ynXx_{n},y_{n}\in X such that dX(xn,yn)t+1d_{X}(x_{n},y_{n})\leq t+1 while dY(f(xn),f(yn))nd_{Y}(f(x_{n}),f(y_{n}))\geq n. This implies ωf(t+1)=\omega_{f}(t+1)=\infty. That is, ff is not coarse. This finishes the proof of (a), and the proof for (b) is analogous. ∎

Let us now define corresponding moduli for quantum functions.

Definition 3.2:

Given a quantum function ϕ:𝒩\phi\colon\mathcal{M}\to\mathcal{N} between quantum metric spaces \mathcal{M} and 𝒩\mathcal{N}, we define ω~ϕ\tilde{\omega}_{\phi} and ρ~ϕ\tilde{\rho}_{\phi} by

ω~ϕ(t)=inf{dist(ϕ(P),ϕ(Q))dist(P,Q)t}\tilde{\omega}_{\phi}(t)=\inf\big{\{}\operatorname{dist}(\phi(P),\phi(Q))\mid\operatorname{dist}(P,Q)\geq t\big{\}}

and

ρ~ϕ(t)=sup{diam(ϕ(P))diam(P)t}\tilde{\rho}_{\phi}(t)=\sup\big{\{}\operatorname{diam}(\phi(P))\mid\operatorname{diam}(P)\leq t\big{\}}

for all t0t\geq 0, where P,QP,Q range over projections in \mathcal{M}.

The next proposition shows that the moduli defined above generalize the classical moduli.

Proposition 3.3.

Given metric spaces X,YX,Y and a function f:XYf\colon X\to Y, ω~ϕf=ω~f\tilde{\omega}_{\phi_{f}}=\tilde{\omega}_{f} and ρ~ϕf=ρ~f\tilde{\rho}_{\phi_{f}}=\tilde{\rho}_{f}.

Proof.

Let PP be a projection in (Y)\ell_{\infty}(Y). Then P=χSP=\chi_{S} for some SYS\subseteq Y, and ϕf(P)=χf1[S]\phi_{f}(P)=\chi_{f^{-1}[S]}. Therefore, for t0t\geq 0

ω~ϕf(t)=inf{dist(χf1[S],χf1[T])dist(χS,χT)t}=inf{dX(x,y)dY(f(x),f(y))t}\tilde{\omega}_{\phi_{f}}(t)=\inf\{\operatorname{dist}(\chi_{f^{-1}[S]},\chi_{f^{-1}[T]})\mid\operatorname{dist}(\chi_{S},\chi_{T})\geq t\}\\ =\inf\{d_{X}(x,y)\mid d_{Y}(f(x),f(y))\geq t\}

and

ρ~ϕf(t)=sup{diam(χf1[S])diam(χS)t}=sup{dX(x,y)dY(f(x),f(y))t}.\tilde{\rho}_{\phi_{f}}(t)=\sup\{\operatorname{diam}(\chi_{f^{-1}[S]})\mid\operatorname{diam}(\chi_{S})\leq t\}\\ =\sup\{d_{X}(x,y)\mid d_{Y}(f(x),f(y))\leq t\}.\qed

Proposition 3.3 and Lemma 3.1 justify the following definition. Although it would perhaps be in better keeping with [KW12] to use the terminology “co-coarse”, there are two reasons we do not do this. The first reason is that the inequalities involved concern only a quantum function ϕ\phi and not its amplification ϕId\phi\otimes\operatorname{Id}. The second reason is that we are only exploring a notion of coarseness for functions between quantum metric spaces and not for operators inside of quantum metric spaces.

Definition 3.4:

A quantum function ϕ:𝒩\phi\colon\mathcal{M}\to\mathcal{N} between two quantum metric spaces is called a (quantum) coarse embedding if limtω~ϕ(t)=\lim_{t\to\infty}\tilde{\omega}_{\phi}(t)=\infty and ρ~ϕ(t)<\tilde{\rho}_{\phi}(t)<\infty for all t0t\geq 0.

Remark 3.5:

In [KW12, Definition 2.27], a quantum function ϕ:𝒩\phi\colon\mathcal{M}\to\mathcal{N} is called a co-Lipschitz morphism if there is some C0C\geq 0 such that

dist(P,Q)Cdist((ϕId)(P),(ϕId)(Q))\operatorname{dist}(P,Q)\leq C\operatorname{dist}((\phi\otimes\operatorname{Id})(P),(\phi\otimes\operatorname{Id})(Q))

for all projections P,Q¯(2)P,Q\in\mathcal{M}\overline{\otimes}\mathcal{B}(\ell_{2}). It is easily observed that if ϕ\phi is a co-Lipschitz morphism, then ω~ϕ(t)t/C\tilde{\omega}_{\phi}(t)\geq t/C for all t0t\geq 0.

Remark 3.6:

Also in [KW12, Definition 2.27], a quantum function ϕ:𝒩\phi\colon\mathcal{M}\to\mathcal{N} is called a co-isometric morphism if it is surjective and

dist(P~,Q~)=sup{dist(P,Q)(ϕId)(P)=P~,(ϕId)(Q)=Q~}\operatorname{dist}(\tilde{P},\tilde{Q})=\sup\big{\{}\operatorname{dist}(P,Q)\mid(\phi\otimes\operatorname{Id})(P)=\tilde{P},(\phi\otimes\operatorname{Id})(Q)=\tilde{Q}\big{\}}

for all projections P~,Q~𝒩¯(2)\tilde{P},\tilde{Q}\in\mathcal{N}\overline{\otimes}\mathcal{B}(\ell_{2}). If ϕ\phi is a co-isometric morphism, then in particular, ϕ\phi is a co-Lipschitz morphism with constant 1, and so ω~ϕ(t)t\tilde{\omega}_{\phi}(t)\geq t for all t0t\geq 0; it may be shown that additionally ρ~ϕ(t)t\tilde{\rho}_{\phi}(t)\leq t for all t0t\geq 0. Indeed, if PP is a projection in \mathcal{M}, and Q~,R~\tilde{Q},\tilde{R} are projections in 𝒩¯(2)\mathcal{N}\overline{\otimes}\mathcal{B}(\ell_{2}) such that Q~(ϕ(P)Id)\tilde{Q}(\phi(P)\otimes\operatorname{Id}) and (ϕ(P)Id)R~(\phi(P)\otimes\operatorname{Id})\tilde{R} are linkable, then since ϕ\phi is a co-isometric morphism, by Lemma 2.6,

dist(Q~,R~)=sup{dist(Q,R)(ϕId)(Q)=Q~,(ϕId)(R)=R~}sup{dist(Q,R)(ϕId)(Q(PId)) and (ϕId)((PId)R) are linkable}sup{dist(Q,R)Q(PId) and (PId)R are linkable}=diam(P).\operatorname{dist}(\tilde{Q},\tilde{R})=\sup\big{\{}\operatorname{dist}(Q,R)\mid(\phi\otimes\operatorname{Id})(Q)=\tilde{Q},(\phi\otimes\operatorname{Id})(R)=\tilde{R}\big{\}}\\ \leq\sup\big{\{}\operatorname{dist}(Q,R)\mid(\phi\otimes\operatorname{Id})\big{(}Q(P\otimes\operatorname{Id})\big{)}\text{ and }(\phi\otimes\operatorname{Id})\big{(}(P\otimes\operatorname{Id})R\big{)}\text{ are linkable}\big{\}}\\ \leq\sup\big{\{}\operatorname{dist}(Q,R)\mid Q(P\otimes\operatorname{Id})\text{ and }(P\otimes\operatorname{Id})R\text{ are linkable}\big{\}}\\ =\operatorname{diam}(P).

Thus, diam(ϕ(P))diam(P)\operatorname{diam}(\phi(P))\leq\operatorname{diam}(P), and therefore ρ~ϕ(t)t\tilde{\rho}_{\phi}(t)\leq t.

4. Asymptotic dimension

We will provide a definition of asymptotic dimension that can be applied generally to all quantum metric spaces. Given the definitions that already exist for diameter and ε\varepsilon-neighborhood of a projection, we have chosen to base our generalization on Part 2 of [BD11, Theorem 2.1.2]. We do not explore generalizations of the equivalent formulations of asymptotic dimension found in [BD11, Theorem 2.1.2].

Definition 4.1:

Let \mathcal{M} be a quantum metric space and 𝒫\mathscr{P} a family of projections in \mathcal{M}. We say that 𝒫\mathscr{P} is a cover for \mathcal{M} if Id=P𝒫P\operatorname{Id}=\bigvee_{P\in\mathscr{P}}P. We say that 𝒫\mathscr{P} is rr-disjoint if (P)r(Q)r=0(P)_{r}(Q)_{r}=0 for each P,Q𝒫P,Q\in\mathscr{P} with PQP\not=Q. We say that 𝒫\mathscr{P} is uniformly bounded by RR if supP𝒫diam(P)R\sup_{P\in\mathscr{P}}\operatorname{diam}(P)\leq R, and that 𝒫\mathscr{P} is uniformly bounded if it is uniformly bounded by some R>0R>0.

Definition 4.2:

Let \mathcal{M} be a quantum metric space, and let n{0}n\in{\mathbb{N}}\cup\{0\}. We say that \mathcal{M} has asymptotic dimension less than or equal to nn, written as asdim()n\operatorname{asdim}(\mathcal{M})\leq n, if for every r>0r>0 there exist uniformly bounded, rr-disjoint families of projections 𝒫0,𝒫1,,𝒫n\mathscr{P}^{0},\mathscr{P}^{1},\dotsc,\mathscr{P}^{n} such that i=0n𝒫i\bigcup_{i=0}^{n}\mathscr{P}^{i} is a cover for \mathcal{M}. We say that \mathcal{M} has asymptotic dimension equal to nn, written as asdim()=n\operatorname{asdim}(\mathcal{M})=n, if n=min{m{0}asdim()m}n=\min\{m\in\mathbb{N}\cup\{0\}\mid\operatorname{asdim}(\mathcal{M})\leq m\}.

Remark 4.3:

Let (X,d)(X,d) be a metric space, and consider the von Neumann algebra (X)\ell_{\infty}(X) endowed with the canonical quantum metric induced by dd. It is clear that asdim(X)=asdim((X))\operatorname{asdim}(X)=\operatorname{asdim}(\ell_{\infty}(X)), since the projections in (X)\ell_{\infty}(X) are precisely the indicator functions of subsets of XX.

For classical metric spaces, coarse embeddings are the natural morphisms that preserve asymptotic dimension because for any r>0r>0 they map every RR-disjoint, uniformly bounded family of sets to an rr-disjoint, uniformly bounded family of sets whenever R>0R>0 is large enough. This follows easily from the definition of coarse embedding using the moduli of expansion and compression. We show that the same holds true for coarse embeddings between quantum metric spaces.

Lemma 4.4.

Let ϕ:𝒩\phi\colon\mathcal{M}\to\mathcal{N} be a quantum function between quantum metric spaces. Then for any projection PP\in\mathcal{M},

diam(ϕ(P))ρ~ϕ(diam(P)).\operatorname{diam}(\phi(P))\leq\tilde{\rho}_{\phi}(\operatorname{diam}(P)).

In particular, ϕ\phi maps a family of projections uniformly bounded by RR to a family of projections uniformly bounded by ρ~ϕ(R)\tilde{\rho}_{\phi}(R).

Proof.
diam(ϕ(P))sup{diam(ϕ(Q))diam(Q)diam(P)}=ρ~ϕ(diam(P)).\operatorname{diam}(\phi(P))\leq\sup\{\operatorname{diam}(\phi(Q))\mid\operatorname{diam}(Q)\leq\operatorname{diam}(P)\}=\tilde{\rho}_{\phi}(\operatorname{diam}(P)).\qed
Lemma 4.5.

Let ϕ:𝒩\phi\colon\mathcal{M}\to\mathcal{N} be a quantum function between quantum metric spaces, and let r>0r>0. Then for any projection PP\in\mathcal{M},

(ϕ(P))ω~ϕ(r)ϕ((P)r).\big{(}\phi(P)\big{)}_{\tilde{\omega}_{\phi}(r)}\leq\phi\big{(}(P)_{r}\big{)}.

In particular, ϕ\phi maps an rr-disjoint family of projections to an ω~ϕ(r)\tilde{\omega}_{\phi}(r)-disjoint family of projections.

Proof.

Since ϕ\phi is a quantum function and dist(P,Q)r\operatorname{dist}(P,Q)\geq r implies dist(ϕ(P),ϕ(Q))ω~ϕ(r)\operatorname{dist}(\phi(P),\phi(Q))\geq\tilde{\omega}_{\phi}(r), we have

ϕ((P)r)\displaystyle\phi\big{(}(P)_{r}\big{)} =ϕ(Id{Qdist(P,Q)r})\displaystyle=\phi\big{(}\operatorname{Id}_{\mathcal{M}}-\bigvee\big{\{}Q\in\mathcal{M}\mid\operatorname{dist}(P,Q)\geq r\big{\}}\big{)}
=Id𝒩{ϕ(Q)dist(P,Q)r}\displaystyle=\operatorname{Id}_{\mathcal{N}}-\bigvee\big{\{}\phi(Q)\mid\operatorname{dist}(P,Q)\geq r\big{\}}
Id𝒩{ϕ(Q)dist(ϕ(P),ϕ(Q))ω~ϕ(r)}\displaystyle\geq\operatorname{Id}_{\mathcal{N}}-\bigvee\big{\{}\phi(Q)\mid\operatorname{dist}(\phi(P),\phi(Q))\geq\tilde{\omega}_{\phi}(r)\big{\}}
Id𝒩{R𝒩dist(ϕ(P),R)ω~ϕ(r)}\displaystyle\geq\operatorname{Id}_{\mathcal{N}}-\bigvee\big{\{}R\in\mathcal{N}\mid\operatorname{dist}(\phi(P),R)\geq\tilde{\omega}_{\phi}(r)\big{\}}
=(ϕ(P))ω~ϕ(r).\displaystyle=\big{(}\phi(P)\big{)}_{\tilde{\omega}_{\phi}(r)}.\qed

The next theorem follows immediately. Note that a quantum function is unital and so it maps covers to covers.

Theorem 4.6.

Let ϕ:𝒩\phi\colon\mathcal{M}\to\mathcal{N} be a quantum coarse embedding between quantum metric spaces. Then asdim(𝒩)asdim()\operatorname{asdim}(\mathcal{N})\leq\operatorname{asdim}(\mathcal{M}).

As a consequence of Theorem 4.6, asymptotic dimension plays well with the quotient [KW12, Definition 2.35] and direct sum [KW12, Definition 2.32 (b)] constructions for quantum metric spaces. Compare this to the corresponding results on subspaces and (disjoint) unions of classical metric spaces found in [BD11, Proposition 2.2.6] and [BD11, Corollary 2.3.3]. Note also that these results are related to some of the conditions required of an abstract dimension theory for a class of CC^{*}-algebras from [Thi13].

Corollary 4.7.

Let \mathcal{M} and 𝒩\mathcal{N} be quantum metric spaces.

  1. (a)

    If 𝒩\mathcal{N} is a metric quotient of \mathcal{M}, then asdim(𝒩)asdim()\operatorname{asdim}(\mathcal{N})\leq\operatorname{asdim}(\mathcal{M}).

  2. (b)

    asdim(𝒩)=max{asdim(),asdim(𝒩)}\operatorname{asdim}(\mathcal{M}\oplus\mathcal{N})=\max\{\operatorname{asdim}(\mathcal{M}),\operatorname{asdim}(\mathcal{N})\}.

Proof.

(a): This follows immediately from Theorem 4.6 and Remark 3.6 of this paper and [KW12, Corollary 2.37].

(b): Since each of \mathcal{M} and 𝒩\mathcal{N} is a metric quotient of 𝒩\mathcal{M}\oplus\mathcal{N}, we have asdim(𝒩)max{asdim(),asdim(𝒩)}\operatorname{asdim}(\mathcal{M}\oplus\mathcal{N})\geq\max\{\operatorname{asdim}(\mathcal{M}),\operatorname{asdim}(\mathcal{N})\} from part (a). Now let n=max{asdim(),asdim(𝒩)}n=\max\{\operatorname{asdim}(\mathcal{M}),\operatorname{asdim}(\mathcal{N})\} and take any r>0r>0. By Definition 4.2, there exist uniformly bounded, rr-disjoint families of projections 𝒫0,𝒫1,,𝒫n\mathscr{P}^{0},\mathscr{P}^{1},\dotsc,\mathscr{P}^{n} such that i=0n𝒫i\bigcup_{i=0}^{n}\mathscr{P}^{i} is a cover for \mathcal{M}, and there also exist uniformly bounded, rr-disjoint families of projections 𝒬0,𝒬1,,𝒬n\mathscr{Q}^{0},\mathscr{Q}^{1},\dotsc,\mathscr{Q}^{n} such that i=0n𝒬i\bigcup_{i=0}^{n}\mathscr{Q}^{i} is a cover for 𝒩\mathcal{N}. For 1jn1\leq j\leq n, define j={P0P𝒫j}{0QQ𝒬j}\mathscr{R}^{j}=\{P\oplus 0\mid P\in\mathscr{P}^{j}\}\cup\{0\oplus Q\mid Q\in\mathscr{Q}^{j}\}. It is clear that each j\mathscr{R}^{j} is uniformly bounded, and moreover the union i=0ni\bigcup_{i=0}^{n}\mathscr{R}^{i} is a cover for 𝒩\mathcal{M}\oplus\mathcal{N}. Additionally, each family j\mathscr{R}^{j} is rr-disjoint, since (P0)r=(P)r0(P\oplus 0)_{r}=(P)_{r}\oplus 0 and (0Q)r=0(Q)r(0\oplus Q)_{r}=0\oplus(Q)_{r}, by [KW12, Proposition 2.34]. Therefore, asdim(𝒩)n\operatorname{asdim}(\mathcal{M}\oplus\mathcal{N})\leq n. ∎

An analog of the result concerning asymptotic dimension of possibly nondisjoint unions of classical metric spaces [BD11, Corollary 2.3.3] can also be established, at least for reflexive quantum metric spaces. In a reflexive quantum metric space \mathcal{M}, diameters of projections in \mathcal{M} may be computed using only projections in \mathcal{M}.

Lemma 4.8.

Let (,{𝒱t}t[0,))(\mathcal{M},\{\mathcal{V}_{t}\}_{t\in[0,\infty)}) be a reflexive quantum metric space and let PP be a nonzero projection in \mathcal{M}. Then

diam(P)\displaystyle\operatorname{diam}(P) =sup{dist(Q,R)QPAPR0 for some A()}\displaystyle=\sup\{\operatorname{dist}(Q,R)\mid QPAPR\neq 0\mbox{ for some }A\in\mathcal{B}(\mathcal{H})\}
=sup{dist(Q,R)QP,RP0}\displaystyle=\sup\{\operatorname{dist}(Q,R)\mid QP,RP\neq 0\}
Proof.

It is clear from Definition 2.3 that

diam(P)sup{dist(Q,R)QPAPR0 for some A()}.\operatorname{diam}(P)\geq\sup\{\operatorname{dist}(Q,R)\mid QPAPR\neq 0\mbox{ for some }A\in\mathcal{B}(\mathcal{H})\}.

The result is then trivial when diam(P)=0\operatorname{diam}(P)=0. So suppose diam(P)0\operatorname{diam}(P)\neq 0 and take any 0<ε<diam(P)0<\varepsilon<\operatorname{diam}(P). Let Q,RQ,R be any projections in ()\mathcal{M}\otimes\mathcal{B}(\mathcal{H}) such that dist(Q,R)>diam(P)ε\operatorname{dist}(Q,R)>\operatorname{diam}(P)-\varepsilon while Q(PAPId)R0Q(PAP\otimes\operatorname{Id})R\neq 0 for some A()A\in\mathcal{B}(\mathcal{H}). By [KW12, Proposition 2.10], PAP𝒱diam(P)εPAP\notin\mathcal{V}_{\operatorname{diam}(P)-\varepsilon}. But then by [KW12, Proposition 2.24], there exist projections Q,RQ^{\prime},R^{\prime}\in\mathcal{M} such that dist(Q,R)diam(P)ε\operatorname{dist}(Q^{\prime},R^{\prime})\geq\operatorname{diam}(P)-\varepsilon while QPAPR0QPAPR\neq 0. As ε>0\varepsilon>0 was arbitrary, it follows that

diam(P)sup{dist(Q,R)QPAPR0 for some A()}.\operatorname{diam}(P)\leq\sup\{\operatorname{dist}(Q,R)\mid QPAPR\neq 0\mbox{ for some }A\in\mathcal{B}(\mathcal{H})\}.\qed
Lemma 4.9.

Let \mathcal{M} be a reflexive quantum metric space, let P,Q,RP,Q,R be projections in \mathcal{M}, and take any r,s>0r,s>0. Then:

  1. (a)

    (P)rQ0dist(P,Q)<r(P)_{r}Q\not=0\iff\operatorname{dist}(P,Q)<r.

  2. (b)

    dist(Q,R)dist(Q,P)+dist(R,P)+diam(P)\operatorname{dist}(Q,R)\leq\operatorname{dist}(Q,P)+\operatorname{dist}(R,P)+\operatorname{diam}(P).

  3. (c)

    diam((P)r)diam(P)+2r\operatorname{diam}\big{(}(P)_{r}\big{)}\leq\operatorname{diam}(P)+2r.

  4. (d)

    diam(PQ)dist(P,Q)+diam(P)+diam(Q)\operatorname{diam}(P\vee Q)\leq\operatorname{dist}(P,Q)+\operatorname{diam}(P)+\operatorname{diam}(Q).

  5. (e)

    (P)r(Q)s0diam(PQ)diam(P)+diam(Q)+2(r+s)(P)_{r}(Q)_{s}\neq 0\implies\operatorname{diam}(P\vee Q)\leq\operatorname{diam}(P)+\operatorname{diam}(Q)+2(r+s).

Proof.

(a) The implication \implies follows immediately from Definition 2.4. So suppose QQ is such that dist(P,Q)<r\operatorname{dist}(P,Q)<r and furthermore, that (P)rQ=0(P)_{r}Q=0. Then

QId(P)r={Qdist(P,Q)r}Q\leq\operatorname{Id}-(P)_{r}=\bigvee\{Q^{\prime}\in\mathcal{M}\mid\operatorname{dist}(P,Q^{\prime})\geq r\}

and thus

r>dist(P,Q)dist(P,Id(P)r)=inf{dist(P,Q)dist(P,Q)r}r,r>\operatorname{dist}(P,Q)\geq\operatorname{dist}(P,\operatorname{Id}-(P)_{r})=\inf\{\operatorname{dist}(P,Q^{\prime})\mid\operatorname{dist}(P,Q^{\prime})\geq r\}\geq r,

a contradiction. Therefore (P)rQ0(P)_{r}Q\neq 0 if dist(P,Q)<r\operatorname{dist}(P,Q)<r.

(b) By the remark after Definition 2.2, with SS ranging over projections in \mathcal{M},

dist(Q\displaystyle\operatorname{dist}(Q ,R)\displaystyle,R)
dist(Q,P)+sup{dist(S,R)PS0}\displaystyle\leq\operatorname{dist}(Q,P)+\sup\{\operatorname{dist}(S,R)\mid PS\neq 0\}
dist(Q,P)+dist(R,P)+sup{dist(S,S)PS0,PS0}\displaystyle\leq\operatorname{dist}(Q,P)+\operatorname{dist}(R,P)+\sup\{\operatorname{dist}(S,S^{\prime})\mid PS\neq 0,PS^{\prime}\neq 0\}
dist(Q,P)+dist(R,P)+diam(P),\displaystyle\leq\operatorname{dist}(Q,P)+\operatorname{dist}(R,P)+\operatorname{diam}(P),

where the last inequality follows from the fact that SPAPS0SPAPS^{\prime}\neq 0 for some A()A\in\mathcal{B}(\mathcal{H}) whenever SP,PS0SP,PS^{\prime}\neq 0.

(c) Suppose S,SS,S^{\prime} are projections in \mathcal{M} such that (P)rS0(P)_{r}S\neq 0 and (P)rS0(P)_{r}S^{\prime}\neq 0. By parts (a) and (b), this means dist(S,S)diam(P)+2r\operatorname{dist}(S,S^{\prime})\leq\operatorname{diam}(P)+2r. As S,SS,S^{\prime} were arbitrary, Lemma 4.8 implies that diam((P)r)diam(P)+2r\operatorname{diam}((P)_{r})\leq\operatorname{diam}(P)+2r.

(d) Suppose S,SS,S^{\prime} are projections in \mathcal{M} such that S(PQ)0S(P\vee Q)\neq 0 and S(PQ)0S^{\prime}(P\vee Q)\not=0. If SP0SP\not=0 and SP0S^{\prime}P\not=0, then dist(S,S)diam(P)\operatorname{dist}(S,S^{\prime})\leq\operatorname{diam}(P). If SQ0SQ\not=0 and SQ0S^{\prime}Q\not=0, then dist(S,S)diam(Q)\operatorname{dist}(S,S^{\prime})\leq\operatorname{diam}(Q). Finally, if SP0SP\not=0 and SQ0S^{\prime}Q\not=0, then by part (b),

dist(S,S)\displaystyle\operatorname{dist}(S,S^{\prime}) dist(S,Q)+dist(Q,S)+diam(Q)\displaystyle\leq\operatorname{dist}(S,Q)+\operatorname{dist}(Q,S^{\prime})+\operatorname{diam}(Q)
dist(S,P)+dist(P,Q)+diam(P)+diam(Q)\displaystyle\leq\operatorname{dist}(S,P)+\operatorname{dist}(P,Q)+\operatorname{diam}(P)+\operatorname{diam}(Q)
=dist(P,Q)+diam(P)+diam(Q).\displaystyle=\operatorname{dist}(P,Q)+\operatorname{diam}(P)+\operatorname{diam}(Q).

The same inequality holds if SQ0SQ\neq 0 and SP0S^{\prime}P\neq 0. As S,SS,S^{\prime} were arbitrary, Lemma 4.8 implies that diam(PQ)dist(P,Q)+diam(P)+diam(Q)\operatorname{diam}(P\vee Q)\leq\operatorname{dist}(P,Q)+\operatorname{diam}(P)+\operatorname{diam}(Q).

(e) By parts (c) and (d)

diam(PQ)\displaystyle\operatorname{diam}(P\vee Q) diam((P)r(Q)s)\displaystyle\leq\operatorname{diam}((P)_{r}\vee(Q)_{s})
dist((P)r,(Q)s)+diam((P)r)+diam((Q)s)\displaystyle\leq\operatorname{dist}((P)_{r},(Q)_{s})+\operatorname{diam}((P)_{r})+\operatorname{diam}((Q)_{s})
diam(P)+diam(Q)+2(r+s).\displaystyle\leq\operatorname{diam}(P)+\operatorname{diam}(Q)+2(r+s).\qed

The following is a direct adaptation of [BD11, Prop. 2.3.1] for reflexive quantum metric spaces.

Proposition 4.10.

Let \mathcal{M} be a reflexive quantum metric space, and let 𝒫,𝒬\mathscr{P},\mathscr{Q} be families of projections in \mathcal{M}. Fix r>0r>0 and for each Q𝒬Q\in\mathscr{Q}, let

𝒫Q={P𝒫(P)r(Q)r0} and PQ=P𝒫QP.\mathscr{P}_{Q}=\{P\in\mathscr{P}\mid(P)_{r}(Q)_{r}\not=0\}\mbox{ and }P_{Q}=\bigvee_{P\in\mathscr{P}_{Q}}P.

Suppose that 𝒫\mathscr{P} is rr-disjoint and RR-bounded with R>rR>r, and 𝒬\mathscr{Q} is 7R7R-disjoint and DD-bounded. Then 𝒬r𝒫\mathscr{Q}\cup_{r}\mathscr{P} is rr-disjoint and (D+2(R+D+4r))(D+2(R+D+4r))-bounded, where

𝒬r𝒫={QPQQ𝒬}{P𝒫(P)r(Q)r=0 for all Q𝒬}.\mathscr{Q}\cup_{r}\mathscr{P}=\left\{Q\vee P_{Q}\mid Q\in\mathscr{Q}\right\}\cup\big{\{}P\in\mathscr{P}\mid(P)_{r}(Q)_{r}=0\text{ for all }Q\in\mathscr{Q}\big{\}}.
Proof.

Fix Q𝒬Q\in\mathscr{Q}. By Lemmas 4.8 and 4.9 (b) and (e),

diam(PQ)\displaystyle\operatorname{diam}(P_{Q}) =sup{dist(S,S)SPQ,SPQ0}\displaystyle=\sup\{\operatorname{dist}(S,S^{\prime})\mid SP_{Q},S^{\prime}P_{Q}\neq 0\}
sup{dist(S,Q)+dist(S,Q)SPQ,SPQ0}+diam(Q)\displaystyle\leq\sup\{\operatorname{dist}(S,Q)+\operatorname{dist}(S^{\prime},Q)\mid SP_{Q},S^{\prime}P_{Q}\neq 0\}+\operatorname{diam}(Q)
2supP𝒫Q{diam(PQ)}+diam(Q)\displaystyle\leq 2\sup_{P\in\mathscr{P}_{Q}}\{\operatorname{diam}(P\vee Q)\}+\operatorname{diam}(Q)
2(supP𝒫Q{diam(P)}+diam(Q)+4r)+diam(Q)\displaystyle\leq 2\left(\sup_{P\in\mathscr{P}_{Q}}\{\operatorname{diam}(P)\}+\operatorname{diam}(Q)+4r\right)+\operatorname{diam}(Q)
2(R+D+4r)+D.\displaystyle\leq 2(R+D+4r)+D.

Thus, the bound on the diameter of QPQQ\vee P_{Q} and hence the entire family 𝒫r𝒬\mathscr{P}\cup_{r}\mathscr{Q} is shown.

We now show that 𝒬r𝒫\mathscr{Q}\cup_{r}\mathscr{P} is rr-disjoint. If we take two elements of 𝒬r𝒫\mathscr{Q}\cup_{r}\mathscr{P} coming from 𝒫\mathscr{P}, then they are rr-disjoint by assumption. Using the fact that (RS)r=(R)r(S)r(R\vee S)_{r}=(R)_{r}\vee(S)_{r} for all projections RR and SS, it is also clear that any two elements such that one is of the form QPQQ\vee P_{Q} and the other is in {P𝒫(P)r(Q)r=0 for all Q𝒬}\big{\{}P\in\mathscr{P}\mid(P)_{r}(Q)_{r}=0\text{ for all }Q\in\mathscr{Q}\big{\}} will be rr-disjoint. The only remaining case is to consider two elements of the form QPQQ\vee P_{Q} and QPQQ^{\prime}\vee P_{Q^{\prime}}, where Q,Q𝒬Q,Q^{\prime}\in\mathscr{Q} are distinct. Note that in this case (Q)r(Q)r=0(Q)_{r}(Q^{\prime})_{r}=0. Consider P,PP,P^{\prime} such that P𝒫QP\in\mathscr{P}_{Q} and P𝒫QP^{\prime}\in\mathscr{P}_{Q^{\prime}}. If (P)r(Q)r0(P)_{r}(Q^{\prime})_{r}\neq 0, then by Lemma 4.9 (a), (b), and (c),

dist(Q,Q)dist(Q,(P)r)+dist((P)r,Q)+diam((P)r)<2r+diam(P)+2r<5R.\operatorname{dist}(Q,Q^{\prime})\leq\operatorname{dist}(Q,(P)_{r})+\operatorname{dist}((P)_{r},Q^{\prime})+\operatorname{diam}((P)_{r})\\ <2r+\operatorname{diam}(P)+2r<5R.

By Lemma 4.9 (a), this implies (Q)5RQ0(Q)_{5R}Q^{\prime}\neq 0, a contradiction. Thus (P)r(Q)r=0(P)_{r}(Q^{\prime})_{r}=0 and similarly (P)r(Q)r=0(P^{\prime})_{r}(Q)_{r}=0. And if (P)r(P)r0(P)_{r}(P^{\prime})_{r}\neq 0, then by Lemma 4.9 (a), (b), (c), and (d),

dist(Q,(Q)r)dist(Q,(P)r(P)r)+diam((P)r(P)r)r+diam(P)+2r+diam(P)+2r<7R.\operatorname{dist}(Q,(Q^{\prime})_{r})\leq\operatorname{dist}(Q,(P)_{r}\vee(P^{\prime})_{r})+\operatorname{diam}((P)_{r}\vee(P^{\prime})_{r})\\ \leq r+\operatorname{diam}(P)+2r+\operatorname{diam}(P^{\prime})+2r<7R.

By Lemma 4.9 (a), this implies (Q)7R(Q)R0(Q)_{7R}(Q^{\prime})_{R}\neq 0, a contradiction. Thus (P)r(P)r=0(P)_{r}(P^{\prime})_{r}=0. As PP, PP^{\prime} were arbitrary, it follows that (QPQ)r(QPQ)r=0(Q\vee P_{Q})_{r}(Q^{\prime}\vee P_{Q^{\prime}})_{r}=0. Therefore 𝒬r𝒫\mathscr{Q}\cup_{r}\mathscr{P} is rr-disjoint. ∎

We can now prove the following theorem which provides a bound on the asymptotic dimension of “nondisjoint unions” of quantum metric spaces. Compare this to [BD11, Corollary 2.3.3].

Theorem 4.11.

Let \mathcal{M} be a reflexive quantum metric space. Suppose that 𝒩1\mathcal{N}_{1} and 𝒩2\mathcal{N}_{2} are metric quotients of \mathcal{M}, corresponding to central projections R1R_{1} and R2R_{2}, respectively. If R1R2=IdR_{1}\vee R_{2}=\operatorname{Id}, then

asdim()max{asdim(𝒩1),asdim(𝒩2)}.\operatorname{asdim}(\mathcal{M})\leq\max\{\operatorname{asdim}(\mathcal{N}_{1}),\operatorname{asdim}(\mathcal{N}_{2})\}.

(Note that, in particular, this includes the case =𝒩1𝒩2\mathcal{M}=\mathcal{N}_{1}\oplus\mathcal{N}_{2}).

Proof.

Let n=max{asdim(𝒩1),asdim(𝒩2)}n=\max\{\operatorname{asdim}(\mathcal{N}_{1}),\operatorname{asdim}(\mathcal{N}_{2})\} and fix r>0r>0. Take n+1n+1 uniformly bounded, rr-disjoint families of projections 𝒫0,𝒫1,,𝒫n\mathscr{P}^{0},\mathscr{P}^{1},\dotsc,\mathscr{P}^{n} in 𝒩1\mathcal{N}_{1} such that i=0n𝒫i\bigcup_{i=0}^{n}\mathscr{P}^{i} is a cover for 𝒩1\mathcal{N}_{1} and let R>rR>r be a uniform diameter bound for i=0n𝒫i\bigcup_{i=0}^{n}\mathscr{P}^{i}. Now take n+1n+1 uniformly bounded, 7R7R-disjoint families of projections 𝒬0,𝒬1,,𝒬n\mathscr{Q}^{0},\mathscr{Q}^{1},\dotsc,\mathscr{Q}^{n} in 𝒩2\mathcal{N}_{2} such that i=0n𝒬i\bigcup_{i=0}^{n}\mathscr{Q}^{i} is a cover for 𝒩2\mathcal{N}_{2} and let D>0D>0 be a uniform diameter bound for i=0n𝒬i\bigcup_{i=0}^{n}\mathscr{Q}^{i}. By viewing a projection PP in 𝒩1\mathcal{N}_{1} as the projection P0R1(IdR1)P\oplus 0\in R_{1}\mathcal{M}\oplus(\operatorname{Id}-R_{1})\mathcal{M}\cong\mathcal{M} and a projection Q𝒩2Q\in\mathcal{N}_{2} as the projection Q0R2(IdR2)Q\oplus 0\in R_{2}\mathcal{M}\oplus(\operatorname{Id}-R_{2})\mathcal{M}\cong\mathcal{M}, it follows from [KW12, Proposition 2.34] that the families 𝒫0,𝒫1,,𝒫n\mathscr{P}^{0},\mathscr{P}^{1},\dotsc,\mathscr{P}^{n} and 𝒬0,𝒬1,,𝒬n\mathscr{Q}^{0},\mathscr{Q}^{1},\dotsc,\mathscr{Q}^{n} have the same bounds and disjointedness when viewed as families of projections in \mathcal{M}. Thus, for each 0jn0\leq j\leq n, the families j=𝒬jr𝒫j\mathscr{R}^{j}=\mathscr{Q}^{j}\cup_{r}\mathscr{P}^{j} in \mathcal{M} are rr-disjoint and uniformly bounded by Proposition 4.10. And since R1R2=IdR_{1}\vee R_{2}=\operatorname{Id}, it follows that i=0ni\bigcup_{i=0}^{n}\mathscr{R}^{i} is a cover for \mathcal{M}. Therefore asdim()n\operatorname{asdim}(\mathcal{M})\leq n. ∎

Remark 4.12:

In order to prove the general quantum analog of [BD11, Corollary 2.3.3] found in Theorem 4.11, we had to make the assumption that the quantum metric space is reflexive. Our proof follows [BD11] rather closely and relies on the ability to place an upper bound on the diameter of a neighborhood of a projection in terms of the diameter of the projection itself. This bound is found in Part (c) of Lemma 4.9, which is the first place we use the reflexivity assumption. We do not know whether the reflexivity assumption can be dropped in the statement of Theorem 4.11, but if it can, we expect a method different from that found in [BD11] would be needed to prove it.

5. Asymptotic dimension and quantum expanders

In this section, we will show that a quantum metric space equi-coarsely containing a sequence of classical expanders (or more generally, a sequence of reflexive quantum expanders) has infinite asymptotic dimension. This is a generalization of [DS12, Sec. 2.3] which shows that even for general quantum metric spaces, information about their large-scale structure can be inferred from their bounded metric subspaces. While this statement is quite believable in light of [DS12] and Theorem 4.6, it is not obvious at all that a quantum metric space should coarsely contain a classical metric space of infinite asymptotic dimension even though it equi-coarsely contains a sequence of expander graphs! To prove the statement, we establish a quantum version of a vertex-isoperimetric inequality for expanders from a known edge-isoperimetric inequality.

In what follows, we denote the space of n×nn\times n matrices with complex entries by MnM_{n}. The n×nn\times n identity matrix will be denoted by InI_{n}. The Hilbert-Schmidt norm for matrices will be denoted by HS\|\cdot\|_{\operatorname{HS}}, the trace norm will be denoted by 1\|\cdot\|_{1}, and the operator norm will be denoted by \|\cdot\|_{\infty}. We will use the initialization CPTP for a map Φ:MnMn\Phi\colon M_{n}\to M_{n} to indicate that Φ\Phi is completely positive and trace-preserving. Given a completely positive map Φ:MnMn\Phi\colon M_{n}\to M_{n}, there exist by Choi’s theorem [Cho75] matrices K1,K2,,KNMnK_{1},K_{2},\dots,K_{N}\in M_{n} such that Φ(X)=j=1NKjXKj\Phi(X)=\sum_{j=1}^{N}K_{j}XK_{j}^{*} for all matrices XMnX\in M_{n}. If Φ\Phi is additionally trace-preserving, it may be shown also that i=1NKiKi=In\sum_{i=1}^{N}K_{i}^{*}K_{i}=I_{n}. It is then possible to define a quantum metric 𝕍={𝒱t}t[0,)\mathbb{V}=\{\mathcal{V}_{t}\}_{t\in[0,\infty)} on MnM_{n} by 𝒱0=In\mathcal{V}_{0}=\mathbb{C}\cdot I_{n}, 𝒱1=span{KjKi}1i,jN\mathcal{V}_{1}=\operatorname{span}\{K_{j}^{*}K_{i}\}_{1\leq i,j\leq N}, and 𝒱t=𝒱1t\mathcal{V}_{t}=\mathcal{V}_{1}^{\lfloor t\rfloor} for t>0t>0 [KW12, Sec. 3.2]. There are good information-theoretical reasons [DSW13, Wea15] and metric reasons [KW12] for regarding a CPTP map Φ\Phi (or rather, the operator system 𝒱1\mathcal{V}_{1}) as a quantum analog of a combinatorial graph and the quantum metric 𝕍\mathbb{V} a quantum analog of a graph metric. By an abuse of language, the terminology “quantum graph” will be used for any of Φ\Phi, 𝒱1\mathcal{V}_{1}, and (Mn,𝕍)(M_{n},\mathbb{V}).

Definition 5.1:

Given δ,ε,t>0\delta,\varepsilon,t>0 and nn\in\mathbb{N}, a quantum metric on MnM_{n} is said to satisfy a (δ,ε,t)(\delta,\varepsilon,t)-isoperimetric inequality if

rank((P)δ)(1+ε)rank(P)\operatorname{rank}\big{(}(P)_{\delta}\big{)}\geq(1+\varepsilon)\operatorname{rank}(P)

for all projections PMnP\in M_{n} such that diam(P)t\operatorname{diam}(P)\leq t.

Remark 5.2:

By Lemma 4.9 (c), if a reflexive quantum metric on MnM_{n} satisfies a (δ,ε,t)(\delta,\varepsilon,t)-isoperimetric inequality, it follows from repeated applications of it that, given any mm\in{\mathbb{N}},

rank((P)mδ)(1+ε)mrank(P)\operatorname{rank}\big{(}(P)_{m\delta}\big{)}\geq(1+\varepsilon)^{m}\operatorname{rank}(P)

for all projections PMnP\in M_{n} such that diam(P)+2mδt\operatorname{diam}(P)+2m\delta\leq t.

Definition 5.3:

Given a family of quantum coarse embeddings {ϕα:𝒩α}\{\phi_{\alpha}\colon\mathcal{M}\to\mathcal{N}_{\alpha}\}, we say that the family is equi-coarse if there exist functions f,gf,g satisfying limtf(t)=\lim_{t\to\infty}f(t)=\infty and g(t)<g(t)<\infty for all t0t\geq 0 such that for for each t0t\geq 0 and each α\alpha, f(t)ω~ϕα(t)f(t)\leq\tilde{\omega}_{\phi_{\alpha}}(t) and ρ~ϕα(t)g(t)\tilde{\rho}_{\phi_{\alpha}}(t)\leq g(t). By analogy with the classical setting, in this case we say that \mathcal{M} equi-coarsely contains the family {𝒩α}\{\mathcal{N}_{\alpha}\}.

The strategy of proof in the next Proposition is based on [DS12, Thm. 2.9].

Proposition 5.4.

Let \mathcal{M} be a quantum metric space, and fix δ,ε>0\delta,\varepsilon>0. Suppose that {(Mnt,𝕍t)}t>0\{(M_{n_{t}},\mathbb{V}_{t})\}_{t>0} is a family of reflexive quantum metric spaces and {ϕt:Mnt}t>0\{\phi_{t}\colon\mathcal{M}\to M_{n_{t}}\}_{t>0} is an equi-coarse family of quantum coarse embeddings. If MntM_{n_{t}} satisfies a (δ,ε,t)(\delta,\varepsilon,t)-isoperimetric inequality for every t>0t>0, then asdim()=\operatorname{asdim}(\mathcal{M})=\infty.

Proof.

Suppose \mathcal{M} has finite asymptotic dimension nn, and take any mm\in{\mathbb{N}} such that (1+ε)m1>n(1+\varepsilon)^{m}-1>n. Let f,gf,g be functions satisfying limrf(r)=\lim_{r\to\infty}f(r)=\infty and g(r)<g(r)<\infty for all r0r\geq 0 such that f(r)ω~ϕt(r)f(r)\leq\tilde{\omega}_{\phi_{t}}(r) and ρ~ϕt(r)g(r)\tilde{\rho}_{\phi_{t}}(r)\leq g(r) for all t,r>0t,r>0; and pick r>0r>0 such that f(r)>mδf(r)>m\delta. Let 𝒫0,𝒫1,,𝒫n\mathscr{P}^{0},\mathscr{P}^{1},\dotsc,\mathscr{P}^{n} be uniformly bounded rr-disjoint families of projections in \mathcal{M} such that j=0n𝒫j\bigcup_{j=0}^{n}\mathscr{P}^{j} is a cover for \mathcal{M} and let dd be such that diam(P)d\operatorname{diam}(P)\leq d for every Pj=0n𝒫jP\in\bigcup_{j=0}^{n}\mathscr{P}^{j}. Finally, let t=2mδ+g(d)t=2m\delta+g(d). It follows from Lemmas 4.4 and 4.5 that for each 0jn0\leq j\leq n, the families 𝒬j={ϕt(P)P𝒫j}\mathscr{Q}^{j}=\{\phi_{t}(P)\mid P\in\mathscr{P}^{j}\} are f(r)f(r)-disjoint (and therefore mδm\delta-disjoint) and uniformly bounded by g(d)g(d), and j=0n𝒬j\bigcup_{j=0}^{n}\mathscr{Q}^{j} is a cover for MntM_{n_{t}} since j=0n𝒫j\bigcup_{j=0}^{n}\mathscr{P}^{j} is a cover for \mathcal{M}. Thus, the (δ,ε,2mδ+g(d))(\delta,\varepsilon,2m\delta+g(d))-isoperimetric inequality for MntM_{n_{t}} implies by Lemma 4.9 (c) and Remark 5.2 that for each 0jn0\leq j\leq n,

nt=rank(Int)Q𝒬jrank((Q)mδ)(1+ε)mQ𝒬jrank(Q),n_{t}=\operatorname{rank}(I_{n_{t}})\geq\sum_{Q\in\mathscr{Q}^{j}}\operatorname{rank}((Q)_{m\delta})\geq(1+\varepsilon)^{m}\sum_{Q\in\mathscr{Q}^{j}}\operatorname{rank}(Q),

and adding over jj yields

(n+1)nt(1+ε)mj=0nQ𝒬jrank(Q)(1+ε)mnt,(n+1)\cdot n_{t}\geq(1+\varepsilon)^{m}\sum_{j=0}^{n}\sum_{Q\in\mathscr{Q}^{j}}\operatorname{rank}(Q)\geq(1+\varepsilon)^{m}\cdot n_{t},

where the last inequality follows from the fact that j=0n𝒬j\bigcup_{j=0}^{n}\mathscr{Q}^{j} is a cover for MntM_{n_{t}}. This implies that n(1+ε)m1>nn\geq(1+\varepsilon)^{m}-1>n, a contradiction. Therefore asdim()=\operatorname{asdim}(\mathcal{M})=\infty. ∎

We will show that a sequence of reflexive quantum expanders (which includes the case of classical expanders) satisfies the isoperimetric inequality condition found in Proposition 5.4. We first recall the definition of quantum expander sequence and an associated Cheeger-type inequality below.

Definition 5.5 ([Pis14]):

Given 0<ε<10<\varepsilon<1 and nn\in\mathbb{N}, a CPTP map Φ:MnMn\Phi\colon M_{n}\to M_{n} is said to have an ε\varepsilon-spectral gap if

Φ(X)1ntr(X)InHS(1ε)X1ntr(X)InHS\left\|\Phi(X)-\tfrac{1}{n}\operatorname{tr}(X)I_{n}\right\|_{\operatorname{HS}}\leq(1-\varepsilon)\left\|X-\tfrac{1}{n}\operatorname{tr}(X)I_{n}\right\|_{\operatorname{HS}}

for all XMnX\in M_{n}.

Definition 5.6 ([Pis14]):

A CPTP map Φ:MnMn\Phi\colon M_{n}\to M_{n} is called a dd-regular ε\varepsilon-quantum expander if Φ\Phi has an ε\varepsilon-spectral gap and there exist unitaries U1,,UdMnU_{1},\dotsc,U_{d}\in M_{n} such that Φ(X)=1dj=1dUjXUj\Phi(X)=\frac{1}{d}\sum_{j=1}^{d}U_{j}XU_{j}^{*} for each XMnX\in M_{n}. A sequence of CPTP maps {Φm:MnmMnm}\{\Phi_{m}\colon M_{n_{m}}\to M_{n_{m}}\} is called a sequence of dd-regular ε\varepsilon-quantum expanders if Φm\Phi_{m} is a dd-regular ε\varepsilon-quantum expander for each mm\in\mathbb{N} and nmn_{m}\to\infty as mm\to\infty.

The following is just a restatement of [TKR+10, Lemma 20], which can be described as a quantum Cheeger inequality.

Lemma 5.7.

Let Φ:MnMn\Phi\colon M_{n}\to M_{n} be a CPTP unital map with an ε\varepsilon-spectral gap. Then

tr((InP)ΦΦ(P))tr(P)(1ε)/2\frac{\operatorname{tr}\big{(}(I_{n}-P)\Phi^{*}\Phi(P)\big{)}}{\operatorname{tr}(P)}\geq(1-\varepsilon)/2

for all projections PMnP\in M_{n} such that 0<rank(P)n/20<\operatorname{rank}(P)\leq n/2.

Remark 5.8:

The expression appearing in the preceding lemma can be rewritten in terms of the inner product associated to the Hilbert-Schmidt norm. Indeed,

tr((InP)ΦΦ(P))=tr((InP)ΦΦ(P))=ΦΦ(P),InPHS=Φ(P),Φ(InP)HS.\operatorname{tr}\big{(}(I_{n}-P)\Phi^{*}\Phi(P)\big{)}=\operatorname{tr}\big{(}(I_{n}-P)^{*}\Phi^{*}\Phi(P)\big{)}\\ ={\langle\Phi^{*}\Phi(P),I_{n}-P\rangle}_{\operatorname{HS}}={\langle\Phi(P),\Phi(I_{n}-P)\rangle}_{\operatorname{HS}}.

Thus, Lemma 5.7 says that Φ\Phi maps orthogonal pairs P,InPP,I_{n}-P to nonorthogonal pairs in a uniform way.

The next result says that if the rank of a projection inside an expander is small, then any large neighborhood of the projection has strictly larger rank than the projection itself.

Proposition 5.9.

Let Φ:MnMn\Phi\colon M_{n}\to M_{n} be a dd-regular ε\varepsilon-quantum expander. Then for any δ>1\delta>1,

rank((P)δ)(1+ε)rank(P)\operatorname{rank}\big{(}(P)_{\delta}\big{)}\geq(1+\varepsilon^{\prime})\operatorname{rank}(P)

whenever PMnP\in M_{n} is a projection such that rank(P)n/2\operatorname{rank}(P)\leq n/2, where the quantum metric on MnM_{n} is the one induced by Φ\Phi and ε=(1ε)/2\varepsilon^{\prime}=(1-\varepsilon)/2.

Proof.

Let PMnP\in M_{n} be a projection such that rank(P)n/2\operatorname{rank}(P)\leq n/2. We will show that if QMnQ\in M_{n} is a projection such that dist(P,Q)δ\operatorname{dist}(P,Q)\geq\delta, then rank(Q)n(1+ε)rank(P)\operatorname{rank}(Q)\leq n-(1+\varepsilon^{\prime})\operatorname{rank}(P). The result will then follow from Definition 2.4.

Let U1,,UdMnU_{1},\dotsc,U_{d}\in M_{n} be unitaries such that Φ(X)=1dj=1dUjXUj\Phi(X)=\frac{1}{d}\sum_{j=1}^{d}U_{j}XU_{j}^{*} for each XMnX\in M_{n}. If dist(P,Q)δ>1\operatorname{dist}(P,Q)\geq\delta>1, it follows from the definition of the quantum metric induced by Φ\Phi that PUjUiQ=0PU_{j}^{*}U_{i}Q=0 for all 1i,jd1\leq i,j\leq d. Thus,

Φ(P),Φ(Q)HS=1d2i,j=1dUjPUj,UiQUiHS=1d2i,j=1dtr(UiQUiUjPUj)=1d2i,j=1dtr(UiUjPUjUiQ)=0.{\langle\Phi(P),\Phi(Q)\rangle}_{\operatorname{HS}}=\frac{1}{d^{2}}\sum_{i,j=1}^{d}{\langle U_{j}PU_{j}^{*},U_{i}QU_{i}^{*}\rangle}_{\operatorname{HS}}\\ =\frac{1}{d^{2}}\sum_{i,j=1}^{d}\operatorname{tr}\big{(}U_{i}QU_{i}^{*}U_{j}PU_{j}^{*}\big{)}=\frac{1}{d^{2}}\sum_{i,j=1}^{d}\operatorname{tr}\big{(}U_{i}^{*}U_{j}PU_{j}^{*}U_{i}Q\big{)}=0.

Therefore, by Lemma 5.7 (Definition 5.6 implies that Φ\Phi is unital) and Remark 5.8,

(5.1) εtr(P)Φ(P),Φ(InP)HS=Φ(P),Φ(InPQ)HS.\varepsilon^{\prime}\operatorname{tr}(P)\leq{\langle\Phi(P),\Phi(I_{n}-P)\rangle}_{\operatorname{HS}}={\langle\Phi(P),\Phi(I_{n}-P-Q)\rangle}_{\operatorname{HS}}.

Now, by the trace duality between the trace and operator norms on MnM_{n},

Φ(P),Φ(R)HS=1d2i,j=1dUjPUj,UiRUiHS=1d2i,j=1dtr(UiRUiUjPUj)=1d2i,j=1dtr(RUiUjPUjUi)1d2i,j=1dR1UiUjPUjUiR1{\langle\Phi(P),\Phi(R)\rangle}_{\operatorname{HS}}=\frac{1}{d^{2}}\sum_{i,j=1}^{d}{\langle U_{j}PU_{j}^{*},U_{i}RU_{i}^{*}\rangle}_{\operatorname{HS}}=\frac{1}{d^{2}}\sum_{i,j=1}^{d}\operatorname{tr}\big{(}U_{i}RU_{i}^{*}U_{j}PU_{j}^{*}\big{)}\\ =\frac{1}{d^{2}}\sum_{i,j=1}^{d}\operatorname{tr}\big{(}RU_{i}^{*}U_{j}PU_{j}^{*}U_{i}\big{)}\leq\frac{1}{d^{2}}\sum_{i,j=1}^{d}\left\|R\right\|_{1}\left\|U_{i}^{*}U_{j}PU_{j}^{*}U_{i}\right\|_{\infty}\leq\left\|R\right\|_{1}

for all projections RMnR\in M_{n}. Therefore, using the fact that for projections the rank, the trace, and the trace norm coincide, it follows from (5.1) that

εrank(P)InPQ1=tr(InPQ)=tr(In)tr(P)tr(Q)=nrank(P)rank(Q),\varepsilon^{\prime}\operatorname{rank}(P)\leq\left\|I_{n}-P-Q\right\|_{1}=\operatorname{tr}(I_{n}-P-Q)\\ =\operatorname{tr}(I_{n})-\operatorname{tr}(P)-\operatorname{tr}(Q)=n-\operatorname{rank}(P)-\operatorname{rank}(Q),

which yields the desired inequality. ∎

We would like to use Proposition 5.9 to establish that quantum expanders satisfy a (δ,ε,t)(\delta,\varepsilon,t)-isoperimetric inequality. To do this, we have to first establish a relationship between the rank and the diameter of a projection inside an expander. We do this more generally for projections inside any connected quantum graph and then show that expanders are connected. The importance of the connectedness assumption is that it implies that every projection has finite diameter. A quantum graph (that is, an operator system) 𝒮Mn\mathcal{S}\subseteq M_{n} is said to be connected if there is m𝒩m\in\mathcal{N} such that 𝒮m=Mn\mathcal{S}^{m}=M_{n} [CDS19, Definition 3.1]. See [CDS19] for more information about connected quantum graphs.

Proposition 5.10.

Let 𝒮\mathcal{S} be a connected quantum graph, and RMnR\in M_{n} a projection. If kk\in\mathbb{N} is such that diam(R)k\operatorname{diam}(R)\leq k, then R𝒮kR=RMnRR\mathcal{S}^{k}R=RM_{n}R.

Proof.

Suppose to the contrary that R𝒮kRRMnRR\mathcal{S}^{k}R\subsetneq RM_{n}R, and pick any ARMnRR𝒮kRA\in RM_{n}R\setminus R\mathcal{S}^{k}R. Then by [Wea12, Lemma 2.8], there exist projections P,QRMnR¯(2)P,Q\in RM_{n}R\overline{\otimes}\mathcal{B}(\ell_{2}) such that P(RARId)Q0P(RAR\otimes\operatorname{Id})Q\neq 0, while P(RBRId)Q=0P(RBR\otimes\operatorname{Id})Q=0 for all B𝒮kB\in\mathcal{S}^{k}. Let P~\tilde{P}, Q~\tilde{Q} be the range projections of (RId)P(R\otimes\operatorname{Id})P and (RId)Q(R\otimes\operatorname{Id})Q, respectively. The above implies that P~(RARId)Q~0\tilde{P}(RAR\otimes\operatorname{Id})\tilde{Q}\neq 0, while P~(BId)Q~=0\tilde{P}(B\otimes\operatorname{Id})\tilde{Q}=0 for all B𝒮kB\in\mathcal{S}^{k}. By Definition 2.2 and Definition 2.3, this means that diam(R)>k\operatorname{diam}(R)>k. This is a contradiction, and so RSkR=RMnRRS^{k}R=RM_{n}R. ∎

Lemma 5.11.

Let Φ:MnMn\Phi\colon M_{n}\to M_{n} be CPTP map and let K1,K2,,KNMnK_{1},K_{2},\dots,K_{N}\in M_{n} be such that Φ(X)=j=1NKjXKj\Phi(X)=\sum_{j=1}^{N}K_{j}XK_{j}^{*} for all matrices XMnX\in M_{n}. If the quantum graph associated to Φ\Phi is connected, then for every projection RMnR\in M_{n},

rank(R)Ndiam(R),\operatorname{rank}(R)\leq N^{\operatorname{diam}(R)},

where the diameter is taken with respect to the quantum graph metric associated to Φ\Phi.

Proof.

Let k=diam(R)k=\operatorname{diam}(R) and let 𝒮=span{KjKi1i,jN}\mathcal{S}=\operatorname{span}\{K_{j}^{*}K_{i}\mid 1\leq i,j\leq N\} be the associated quantum graph. It follows from Proposition 5.10 that RMnR=R𝒮kRRM_{n}R=R\mathcal{S}^{k}R and therefore

rank(R)2=dim(RMnR)=dim(R𝒮kR)dim(𝒮k)(N2)k,\operatorname{rank}(R)^{2}=\dim(RM_{n}R)=\dim(R\mathcal{S}^{k}R)\leq\dim(\mathcal{S}^{k})\leq(N^{2})^{k},

which yields the desired inequality. ∎

Proposition 5.12.

Let Φ:MnMn\Phi\colon M_{n}\to M_{n} be a CPTP unital map with an ε\varepsilon-spectral gap. Then the associated quantum graph is connected. In particular, every dd-regular ε\varepsilon-quantum expander is connected.

Proof.

Let K1,,KNMnK_{1},\dotsc,K_{N}\in M_{n} be matrices such that Φ(X)=j=1NKjXKj\Phi(X)=\sum_{j=1}^{N}K_{j}XK_{j}^{*} for each XMnX\in M_{n}. Suppose that the quantum graph 𝒮=span{KjKi1i,jN}\mathcal{S}=\operatorname{span}\{K_{j}^{*}K_{i}\mid 1\leq i,j\leq N\} is disconnected. By [CDS19, Theorem 3.3], there exists a nontrivial projection PMnP\in M_{n} such that P𝒮(InP)=0P\mathcal{S}(I_{n}-P)=0, and without loss of generality, we may assume 0<rank(P)n/20<\operatorname{rank}(P)\leq n/2. In particular, PKjKi(InP)=0PK_{j}^{*}K_{i}(I_{n}-P)=0 for all 1i,jN1\leq i,j\leq N. Therefore

Φ(P),Φ(InP)HS=i,j=1NKjPKj,Ki(InP)KiHS=i,j=1Ntr(Ki(InP)KiKjPKj)=i,j=1Ntr(KiKjPKjKi(InP))=0.{\langle\Phi(P),\Phi(I_{n}-P)\rangle}_{\operatorname{HS}}=\sum_{i,j=1}^{N}{\langle K_{j}PK_{j}^{*},K_{i}(I_{n}-P)K_{i}^{*}\rangle}_{\operatorname{HS}}\\ =\sum_{i,j=1}^{N}\operatorname{tr}\big{(}K_{i}(I_{n}-P)K_{i}^{*}K_{j}PK_{j}^{*}\big{)}=\sum_{i,j=1}^{N}\operatorname{tr}\big{(}K_{i}^{*}K_{j}PK_{j}^{*}K_{i}(I_{n}-P)\big{)}=0.

This contradicts Lemma 5.7, and so the quantum graph associated to Φ\Phi is connected. ∎

Propositions 5.4 and 5.9 and Lemma 5.11 together yield our main theorem.

Theorem 5.13.

If a quantum metric space \mathcal{M} is equi-coarsely embeddable into a sequence of reflexive dd-regular ε\varepsilon-quantum expanders, then asdim()=\operatorname{asdim}(\mathcal{M})=\infty. In particular, this holds whenever \mathcal{M} admits a sequence of reflexive dd-regular ε\varepsilon-quantum expanders as metric quotients.

We point out that, in particular, Theorem 5.13 covers the case when \mathcal{M} equi-coarsely contains a sequence of dd-regular ε\varepsilon-classical expanders, thanks to the following proposition.

Proposition 5.14.

The canonical quantum metric associated to a classical metric is always reflexive.

Proof.

Let (X,d)(X,d) be a classical metric. Let t0t\geq 0. By [KW12, Prop. 2.5], the canonical quantum metric on (X)\ell_{\infty}(X) associated to dd is given by

(5.2) 𝒱t={A(2(X))d(x,y)>tAey,ex=0}.\mathcal{V}_{t}=\big{\{}A\in\mathcal{B}(\ell_{2}(X))\mid d(x,y)>t\Rightarrow{\langle Ae_{y},e_{x}\rangle}=0\big{\}}.

Let us now show that 𝒱t\mathcal{V}_{t} is reflexive. To that end, let B(2(X))B\in\mathcal{B}(\ell_{2}(X)) be such that for any projections P,Q(2(X))P,Q\in\mathcal{B}(\ell_{2}(X)) such that P𝒱tQ={0}P\mathcal{V}_{t}Q=\{0\}, it follows that PBQ=0PBQ=0; we need to show that BB belongs to 𝒱t\mathcal{V}_{t}. Recall that VxyV_{xy} denotes the mapping gg,eyexg\mapsto{\langle g,e_{y}\rangle}e_{x}. Let x,yXx,y\in X satisfy d(x,y)>td(x,y)>t. Note that VxxV_{xx} and VyyV_{yy} are projections, and it follows from (5.2) that Vxx𝒱tVyy={0}V_{xx}\mathcal{V}_{t}V_{yy}=\{0\}. Therefore, VxxBVyy=0V_{xx}BV_{yy}=0. But this implies Bey,ex=0{\langle Be_{y},e_{x}\rangle}=0, and thus B𝒱tB\in\mathcal{V}_{t} by apppealing to (5.2) again. ∎

One final remark is in order regarding Theorem 5.13 and Proposition 5.14. We have shown that equi-coarse containment of reflexive quantum expanders implies infinite asymptotic dimension and we have also shown that quantum expanders induced by classical expanders are reflexive. While this is enough to provide a generalization of [DS12, Sec. 2.3] to the realm of quantum metric spaces, what we have not shown is the existence of a nontrivial reflexive quantum expander. That is, we do not actually know whether every reflexive quantum expander is induced by a classical expander. It would be interesting to know the answer to this question, but it would be more interesting still to know whether the reflexivity assumption in Theorem 5.13 (or more generally Proposition 5.4) can be dropped. As with the proof of Theorem 4.11, it was very important to be able to place an upper bound on the diameter of a neighborhood of a projection in terms of the diameter of the projection (Part (c) of Lemma 4.9). This is what allowed us to repeatedly apply the isoperimetric inequality to derive the inequality found in Remark 5.2.

References

  • [Ban15] Tathagata Banerjee, Coarse geometry for noncommutative spaces, Ph.D. thesis, Georg-August-Universität Göttingen, 2015.
  • [BD11] G. Bell and A. Dranishnikov, Asymptotic dimension in Będlewo, Topology Proc. 38 (2011), 209–236. MR 2725304
  • [Bla06] B. Blackadar, Operator algebras, Encyclopaedia of Mathematical Sciences, vol. 122, Springer-Verlag, Berlin, 2006, Theory of CC^{*}-algebras and von Neumann algebras, Operator Algebras and Non-commutative Geometry, III. MR 2188261
  • [BM16] Tathagata Banerjee and Ralf Meyer, Noncommutative coarse geometry, arXiv preprint arXiv:1610.08969 (2016).
  • [BP91] Lawrence G. Brown and Gert K. Pedersen, CC^{*}-algebras of real rank zero, J. Funct. Anal. 99 (1991), no. 1, 131–149. MR 1120918
  • [CDS19] Javier Alejandro Chávez-Domínguez and Andrew T. Swift, Connectivity for quantum graphs, arXiv preprint arXiv:1911.00446 (2019).
  • [Cho75] Man Duen Choi, Completely positive linear maps on complex matrices, Linear Algebra Appl. 10 (1975), 285–290. MR 0376726
  • [DS12] Alexander Dranishnikov and Mark Sapir, On the dimension growth of groups, arXiv preprint arXiv:1008.3868v4 (2012).
  • [DSW13] Runyao Duan, Simone Severini, and Andreas Winter, Zero-error communication via quantum channels, noncommutative graphs, and a quantum Lovász number, IEEE Trans. Inform. Theory 59 (2013), no. 2, 1164–1174. MR 3015725
  • [Gro81] Mikhael Gromov, Groups of polynomial growth and expanding maps, Inst. Hautes Études Sci. Publ. Math. (1981), no. 53, 53–73. MR 623534
  • [Gro93] by same author, Geometric group theory, vol. 2: Asymptotic invariants of infinite groups, London Math. Soc. Lecture Note Ser. 182, Cambridge University Press, Cambridge, 1993.
  • [Kor11] Andre Kornell, Quantum functions, arXiv preprint arXiv:1101.1694 (2011).
  • [KW04] Eberhard Kirchberg and Wilhelm Winter, Covering dimension and quasidiagonality, Internat. J. Math. 15 (2004), no. 1, 63–85. MR 2039212
  • [KW12] Greg Kuperberg and Nik Weaver, A von Neumann algebra approach to quantum metrics, Mem. Amer. Math. Soc. 215 (2012), no. 1010, v, 1–80. MR 2908248
  • [Mos68] G. D. Mostow, Quasi-conformal mappings in nn-space and the rigidity of hyperbolic space forms, Inst. Hautes Études Sci. Publ. Math. (1968), no. 34, 53–104. MR 236383
  • [NY12] Piotr W. Nowak and Guoliang Yu, Large scale geometry, EMS Textbooks in Mathematics, European Mathematical Society (EMS), Zürich, 2012. MR 2986138
  • [Pis14] Gilles Pisier, Quantum expanders and geometry of operator spaces, J. Eur. Math. Soc. (JEMS) 16 (2014), no. 6, 1183–1219. MR 3226740
  • [Rie83] Marc A. Rieffel, Dimension and stable rank in the KK-theory of CC^{\ast}-algebras, Proc. London Math. Soc. (3) 46 (1983), no. 2, 301–333. MR 693043
  • [Thi13] Hannes Thiel, The topological dimension of type I CC^{*}-algebras, Operator algebra and dynamics, Springer Proc. Math. Stat., vol. 58, Springer, Heidelberg, 2013, pp. 305–328. MR 3142044
  • [TKR+10] K. Temme, M. J. Kastoryano, M. B. Ruskai, M. M. Wolf, and F. Verstraete, The χ2\chi^{2}-divergence and mixing times of quantum Markov processes, J. Math. Phys. 51 (2010), no. 12, 122201, 19. MR 2779170
  • [Wea12] Nik Weaver, Quantum relations, Mem. Amer. Math. Soc. 215 (2012), no. 1010, v–vi, 81–140. MR 2908249
  • [Wea15] by same author, Quantum graphs as quantum relations, arXiv preprint arXiv:1506.03892 (2015).
  • [Win07] Wilhelm Winter, Simple CC^{*}-algebras with locally finite decomposition rank, J. Funct. Anal. 243 (2007), no. 2, 394–425. MR 2289694
  • [WZ10] Wilhelm Winter and Joachim Zacharias, The nuclear dimension of CC^{\ast}-algebras, Adv. Math. 224 (2010), no. 2, 461–498. MR 2609012