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

quantum Gromov-Hausdorff propinquity convergence of Christensen-Ivan quantum metrics on AF algebras

Clay Adams Konrad Aguilar Department of Mathematics and Statistics, Pomona College, 610 N. College Ave., Claremont, CA 91711 konrad.aguilar@pomona.edu https://aguilar.sites.pomona.edu Esteban Ayala Evelyne Knight  and  Chloe Marple
Abstract.

We provide convergence in the quantum Gromov-Hausdorff propinquity of Latrémolière of some sequences of infinite-dimensional Leibniz compact quantum metric spaces of Rieffel given by AF algebras and Christensen-Ivan spectral spaces. The main examples are convergence of Effros-Shen algebras and UHF algebras.

Key words and phrases:
Gromov–Hausdorff propinquity, quantum metric spaces, Effros–Shen algebras, UHF algebras, spectral triples
2000 Mathematics Subject Classification:
Primary: 46L89, 46L30, 58B34.
The second author is supported by NSF grant DMS-2316892

1. Introduction and Background

The first example of convergence of sequences of infinite-dimensional quantum metric spaces was established by Rieffel [22], where he showed that the quantum tori converged with respect their parameters that defined their anti-commutation relation. This was accomplished by the introduction of the theory of quantum metric spaces and a noncommutative analogue to the Gromov-Hausdorff distance both introduce by Rieffel in [21, 22], respectively. This introduced an new field of study known as noncommutative metric geometry, which has its roots from work of Connes in [6, 5] and Gromov [9].

Since the introduction of Rieffel’s noncommutative analogue to the Gromov-Hausdorff distance, there has been much progress in developing noncommutative analogues of the Gromov-Hausdorff distance to capture the C*-algebraic structure of the quantum metric space [19, 11, 24, 15, 14]. In particular, in [13], Latrémolière proved convergence of the quantum tori in this stronger sense. Moreoever, the Gromov-Hausdorff propinquity of Latrémolière first introduced in [15] has been adapted to capture more structure such as module structure [12] and spectral triple structure [18, 17].

Another example of convergence of infinite-dimensional quantum metric spaces appeared in [2], where it was shown that the Effros-Shen algebras [8] are continuous in Gromov-Hausdorff propinquity with respect to their natural parameter space of irrationals in (0,1)(0,1) with the usuual topology. It was also show that UHF algebras are continuous with respect to their natural parameter space of multiplicity sequences metrized by the Baire space. This was accomplished by introducing new quantum metrics on AF algebras equipped with faithful tracial state motivated by work of Christensen and Ivan in [4]. Now, in [4], Christensen and Ivan did introduce quantum metrics on these infinite-dimensional algebras, but at the time it wasn’t clear how to provide convergence of these algebras in any noncommutative analogue to the Gromov-Hausdorff distance, which is one reason why the quantum metrics of [2] were introduced which are not defined using spectral triples. But, in an effort, to bring the realms of noncommutative geometry and noncommutative metric geometry closer, it is important to provide the convergence results of these infinite-dimensional algebras of [2] with the quantum metrics induced by the spectral triples of [4], which is exactly what is accomplished in this article.

The main hurdles to overcome in proving this arise from two issues that were circumvented by the quantum metrics introduced in [2]. First, the spectral triples of [4] are constructed using equivalence constants which are only provided by existence and not explicitly given, which cause an issue when providing continuous fields of LL-seminorms as it is difficult to control these non-explicit constants. Second, providing continuous fields of LL-seminorms provided by faithful tracial states is difficult when relying on convergence in various operator norms given by different GNS represenations for each spectral triple rather than a fixed C*-norm. The first issue is overcome by an application of [1, Lemma 3.9], and the second issue is overcome by a generalization of [1, Lemma 3.9]. Both of these issues are overcome in Section 2, and we apply these results in the last section to provide convergence of these infinite-dimensional algebras using quantum metrics induced by Christensen-Ivan spectral triples. We only define what we mean by a Leibniz compact quantum metric space as things can get quite overwhelming as more definitions are provided, but for references regarding quantum metric spaces, propinquity and propinquity in the context of AF algebras see [23, 15, 2, 16].

Definition 1.1.

[21, 15] Let 𝒜{\mathcal{A}} be a unital C*-algebra with norm 𝒜\|\cdot\|_{\mathcal{A}} and unit 1𝒜1_{\mathcal{A}}. Let L:𝒜[0,)L:{\mathcal{A}}\rightarrow[0,\infty) be a seminnorm (possibly taking value \infty) such that dom(L)={a𝒜:L(a)<}\mathrm{dom}(L)=\{a\in{\mathcal{A}}:L(a)<\infty\} is a dense *-subalgebra of 𝒜{\mathcal{A}}. If

  1. (1)

    L(a)=L(a)L(a)=L(a^{*}) for every a𝒜a\in{\mathcal{A}},

  2. (2)

    {a𝒜:L(a)=0}=1𝒜\{a\in{\mathcal{A}}:L(a)=0\}={\mathds{C}}1_{\mathcal{A}},

  3. (3)

    L(ab)a𝒜L(b)+b𝒜L(a)L(ab)\leqslant\|a\|_{\mathcal{A}}L(b)+\|b\|_{\mathcal{A}}L(a) for every a,b𝒜a,b\in{\mathcal{A}},

  4. (4)

    the metric on the state space S(𝒜)S({\mathcal{A}}) of 𝒜{\mathcal{A}} defined for every ϕ,ψS(𝒜)\phi,\psi\in S({\mathcal{A}}) by

    mkL(ϕ,ψ)=sup{|ϕ(a)ψ(a)|:a𝒜,L(a)1}mk_{L}(\phi,\psi)=\sup\{|\phi(a)-\psi(a)|:a\in{\mathcal{A}},L(a)\leqslant 1\}

    metrizes the weak* topology,

then we call LL an LL-seminorm and (𝒜,L)({\mathcal{A}},L) a Leibniz compact quantum metric space.

2. Finite-dimensional approximations and associated continuous fields of LL-seminorms

In what follows, we use various results from the beginning of [4, Section 2] with some slightly different notation. 𝒜=nAn¯A{\mathcal{A}}=\overline{\cup_{n\in{\mathds{N}}}A_{n}}^{\|\cdot\|_{A}} be a unital AF algebra, where 𝒜0=1𝒜{\mathcal{A}}_{0}={\mathds{C}}1_{\mathcal{A}} equipped with a faithful tracial state τ\tau. Let HτH_{\tau} denote the associated GNS Hilbert space with inner product defined for every a,bHτa,b\in H_{\tau} by

a,bτ=τ(ba)\langle a,b\rangle_{\tau}=\tau(b^{*}a)

and associated norm aτ=a,aτ\|a\|_{\tau}=\sqrt{\langle a,a\rangle_{\tau}}. Since τ\tau is faithful, we can canonically view 𝒜{\mathcal{A}} as a subspace (not necessarily closed) of HτH_{\tau}. Let

πτ:𝒜B(Hτ)\pi_{\tau}:{\mathcal{A}}\longrightarrow B(H_{\tau})

be the associated GNS representation such that πτ(a)(b)=ab\pi_{\tau}(a)(b)=ab for every a,b𝒜a,b\in{\mathcal{A}}.

Let nn\in{\mathds{N}}, since 𝒜n{\mathcal{A}}_{n} is finite dimensional, we have that 𝒜n{\mathcal{A}}_{n} is a closed subspace of HτH_{\tau}. Let

Pnτ:Hτ𝒜nP^{\tau}_{n}:H_{\tau}\rightarrow{\mathcal{A}}_{n}

denote the orthogonal projection of HτH_{\tau} onto 𝒜n{\mathcal{A}}_{n} and define Qnτ=PnτPn1τQ^{\tau}_{n}=P^{\tau}_{n}-P^{\tau}_{n-1}. Let

Enτ:𝒜𝒜nE^{\tau}_{n}:{\mathcal{A}}\rightarrow{\mathcal{A}}_{n}

denote the restriction of PnτP^{\tau}_{n} to 𝒜{\mathcal{A}}, and by [2, Theorem 3.5], we have that EnτE^{\tau}_{n} is the unique τ\tau-preserving conditional expectation onto 𝒜n{\mathcal{A}}_{n}.

Next, since 𝒜n+1{\mathcal{A}}_{n+1} is finite dimensional, there exists a sharp cn+1τ>0c^{\tau}_{n+1}>0 such that

(2.1) a𝒜cn+1τaτ\|a\|_{\mathcal{A}}\leqslant c^{\tau}_{n+1}\cdot\|a\|_{\tau}

for every a𝒜n+1a\in{\mathcal{A}}_{n+1}. Note that cn+1τ1c^{\tau}_{n+1}\geqslant 1 since τ𝒜\|\cdot\|_{\tau}\leqslant\|\cdot\|_{\mathcal{A}} on 𝒜{\mathcal{A}}.

We now prove a crucial fact about these constants.

Proposition 2.1.

Let (τn)n(\tau^{n})_{n\in{\mathds{N}}} be a sequence of faithful tracial states on 𝒜{\mathcal{A}} and let τ\tau be a faithful tracial state on 𝒜{\mathcal{A}}. If (τn)n(\tau^{n})_{n\in{\mathds{N}}} converges to τ\tau in the weak* topology, then for every NN\in{\mathds{N}}, the sequence (cNτn)n(c^{\tau^{n}}_{N})_{n\in{\mathds{N}}} converges to cNτc^{\tau}_{N} in the usual topology on {\mathds{R}}.

Proof.

This is just [1, Proposition 3.10] applied to [1, Proposition 3.6] since norm τ\|\cdot\|_{\tau} is a Frobenius-Rieffel norm. ∎

Let (β(n))n(\beta(n))_{n\in{\mathds{N}}} be a summable sequence of positive reals. Set

aβ,n+1τ=cn+1τβn+1.a^{\tau}_{\beta,n+1}=\frac{c^{\tau}_{n+1}}{\beta_{n+1}}.

Next, we state a main result from [4].

Theorem 2.2.

[4, Theorem 2.1] Let (β(n))n(\beta(n))_{n\in{\mathds{N}}} be a summable sequence of positive reals. Using the above setting, we have that

Dβτ=n=1aβ,nτQnτD^{\tau}_{\beta}=\sum_{n=1}^{\infty}a^{\tau}_{\beta,n}Q^{\tau}_{n}

defines an unbounded self-adjoint operator on HτH_{\tau}. Furthermore, if we define

Lβτ(a)=[Dβτ,πτ(a)]B(Hτ)L^{\tau}_{\beta}(a)=\|[D^{\tau}_{\beta},\pi_{\tau}(a)]\|_{B(H_{\tau})}

for every a𝒜a\in{\mathcal{A}} such that [Dβτ,πτ(a)][D^{\tau}_{\beta},\pi_{\tau}(a)] extends to a bounded operator on HτH_{\tau} denoted by [Dβτ,πτ(a)][D^{\tau}_{\beta},\pi_{\tau}(a)], and set Lβτ(a)=L^{\tau}_{\beta}(a)=\infty if not, then

(𝒜,Lβτ)({\mathcal{A}},L^{\tau}_{\beta})

is a Leibniz compact quantum metric space, and for every nn\in{\mathds{N}}, (𝒜n,Lβτ)({\mathcal{A}}_{n},L^{\tau}_{\beta}) is a Leibniz compact quantum metric space such that dom(Lβτ(a))𝒜n=𝒜n\mathrm{dom}(L^{\tau}_{\beta}(a))\cap{\mathcal{A}}_{n}={\mathcal{A}}_{n}.

The following fact is stated after [3, Expression (4.5)], but we provide a proof here.

Proposition 2.3.

Using the setting of Theorem 2.2, we have for every nn\in{\mathds{N}} and for every a𝒜na\in{\mathcal{A}}_{n} that

Lβτ(a)=[Dβτ,πτ(a)]B(𝒜nτ),L^{\tau}_{\beta}(a)=\|[D^{\tau}_{\beta},\pi_{\tau}(a)]\|_{B({\mathcal{A}}^{\tau}_{n})},

where 𝒜nτ=𝒜n{\mathcal{A}}_{n}^{\tau}={\mathcal{A}}_{n} but for B(𝒜nτ)B({\mathcal{A}}^{\tau}_{n}) we are considering bounded operators with respect to the norm τ\|\cdot\|_{\tau} on 𝒜n{\mathcal{A}}_{n}.

Proof.

Let nn\in{\mathds{N}}. Let a𝒜na\in{\mathcal{A}}_{n}. By definition, we have that

[Dβτ,πτ(a)]B(𝒜n)Lβτ(a).\|[D^{\tau}_{\beta},\pi_{\tau}(a)]\|_{B({\mathcal{A}}_{n})}\leqslant L^{\tau}_{\beta}(a).

Next, let k{1,2,,n}k\in\{1,2,\ldots,n\}. We have since πτ(a)\pi_{\tau}(a) commutes with PnτP^{\tau}_{n} by the proof of [4, Theorem 2.1] or the proof of Step 1 of [2, Theorem 3.5]. Moreover, PnPk=PkPn=PkP_{n}P_{k}=P_{k}P_{n}=P_{k} and PnPk1=Pk1Pn=Pk1P_{n}P_{k-1}=P_{k-1}P_{n}=P_{k-1} by construction. Thus

Pnτ[Qkτ,πτ(a)]Pnτ\displaystyle P^{\tau}_{n}[Q^{\tau}_{k},\pi_{\tau}(a)]P^{\tau}_{n} =Pnτ(Qkτπτ(a)πτ(a)Qkτ)Pnτ\displaystyle=P^{\tau}_{n}(Q^{\tau}_{k}\pi_{\tau}(a)-\pi_{\tau}(a)Q^{\tau}_{k})P^{\tau}_{n}
=Pnτ((PkτPk1τ)πτ(a)πτ(a)(PkτPk1τ))Pnτ\displaystyle=P^{\tau}_{n}((P^{\tau}_{k}-P^{\tau}_{k-1})\pi_{\tau}(a)-\pi_{\tau}(a)(P^{\tau}_{k}-P^{\tau}_{k-1}))P^{\tau}_{n}
=(PkτPk1τ)πτ(a)PnτPnτπτ(a)(PkτPk1τ)\displaystyle=(P^{\tau}_{k}-P^{\tau}_{k-1})\pi_{\tau}(a)P^{\tau}_{n}-P^{\tau}_{n}\pi_{\tau}(a)(P^{\tau}_{k}-P^{\tau}_{k-1})
=(PkτPk1τ)Pnτπτ(a)πτ(a)Pnτ(PkτPk1τ)\displaystyle=(P^{\tau}_{k}-P^{\tau}_{k-1})P^{\tau}_{n}\pi_{\tau}(a)-\pi_{\tau}(a)P^{\tau}_{n}(P^{\tau}_{k}-P^{\tau}_{k-1})
=(PkτPk1τ)πτ(a)πτ(a)(PkτPk1τ)\displaystyle=(P^{\tau}_{k}-P^{\tau}_{k-1})\pi_{\tau}(a)-\pi_{\tau}(a)(P^{\tau}_{k}-P^{\tau}_{k-1})
=Qkτπτ(a)πτ(a)Qkτ\displaystyle=Q^{\tau}_{k}\pi_{\tau}(a)-\pi_{\tau}(a)Q^{\tau}_{k}
=[Qkτ,πτ(a)]\displaystyle=[Q^{\tau}_{k},\pi_{\tau}(a)]

Thus

[Dβτ,πτ(a)]\displaystyle[D^{\tau}_{\beta},\pi_{\tau}(a)] =k=1naβ,kτ[Qkτ,πτ(a)]\displaystyle=\sum_{k=1}^{n}a^{\tau}_{\beta,k}[Q^{\tau}_{k},\pi_{\tau}(a)]
=k=1naβ,kτPnτ[Qkτ,πτ(a)]Pnτ\displaystyle=\sum_{k=1}^{n}a^{\tau}_{\beta,k}P^{\tau}_{n}[Q^{\tau}_{k},\pi_{\tau}(a)]P^{\tau}_{n}
=Pnτ[Dβτ,πτ(a)]Pnτ.\displaystyle=P^{\tau}_{n}[D^{\tau}_{\beta},\pi_{\tau}(a)]P^{\tau}_{n}.

Hence, since (Pnτ)2=Pnτ(P^{\tau}_{n})^{2}=P^{\tau}_{n} and PnτP^{\tau}_{n} is contractive with respect to τ\|\cdot\|_{\tau} and Pnτ(h)𝒜nP^{\tau}_{n}(h)\in{\mathcal{A}}_{n} for every hHτh\in H_{\tau}, we have

Lβτ(a)\displaystyle L^{\tau}_{\beta}(a) =Pnτ[Dβτ,πτ(a)]PnτB(Hτ)\displaystyle=\|P^{\tau}_{n}[D^{\tau}_{\beta},\pi_{\tau}(a)]P^{\tau}_{n}\|_{B(H_{\tau})}
=sup{Pnτ[Dβτ,πτ(a)]Pnτ(h)τ:hHτ,hτ1}\displaystyle=\sup\left\{\|P^{\tau}_{n}[D^{\tau}_{\beta},\pi_{\tau}(a)]P^{\tau}_{n}(h)\|_{\tau}:h\in H_{\tau},\|h\|_{\tau}\leqslant 1\right\}
=sup{Pnτ[Dβτ,πτ(a)](Pnτ)2(h)τ:hHτ,hτ1}\displaystyle=\sup\left\{\|P^{\tau}_{n}[D^{\tau}_{\beta},\pi_{\tau}(a)](P^{\tau}_{n})^{2}(h)\|_{\tau}:h\in H_{\tau},\|h\|_{\tau}\leqslant 1\right\}
=sup{Pnτ[Dβτ,πτ(a)]Pnτ(Pnτ(h))τ:hHτ,hτ1}\displaystyle=\sup\left\{\|P^{\tau}_{n}[D^{\tau}_{\beta},\pi_{\tau}(a)]P^{\tau}_{n}(P^{\tau}_{n}(h))\|_{\tau}:h\in H_{\tau},\|h\|_{\tau}\leqslant 1\right\}
=sup{[Dβτ,πτ(a)](Pnτ(h))τ:hHτ,hτ1}\displaystyle=\sup\left\{\|[D^{\tau}_{\beta},\pi_{\tau}(a)](P^{\tau}_{n}(h))\|_{\tau}:h\in H_{\tau},\|h\|_{\tau}\leqslant 1\right\}
sup{[Dβτ,πτ(a)]τ:h𝒜n,hτ1}\displaystyle\leqslant\sup\left\{\|[D^{\tau}_{\beta},\pi_{\tau}(a)]\|_{\tau}:h\in{\mathcal{A}}_{n},\|h\|_{\tau}\leqslant 1\right\}
=[Dβτ,πτ(a)]B(𝒜n).\displaystyle=\|[D^{\tau}_{\beta},\pi_{\tau}(a)]\|_{B({\mathcal{A}}_{n})}.

Therefore [Dβτ,πτ(a)]B(𝒜n)Lβτ(a)[Dβτ,πτ(a)]B(𝒜n)\|[D^{\tau}_{\beta},\pi_{\tau}(a)]\|_{B({\mathcal{A}}_{n})}\leqslant L^{\tau}_{\beta}(a)\leqslant\|[D^{\tau}_{\beta},\pi_{\tau}(a)]\|_{B({\mathcal{A}}_{n})} as desired. ∎

With this we can provide finite-dimensional approximations, which has been conveniently already proven in [3].

Theorem 2.4.

[3, Theorem 4.8] For every nn\in{\mathds{N}}, it holds that

Λ((𝒜,Lβτ),(𝒜n,Lβτ))k=nβk,{\mathsf{\Lambda}}(({\mathcal{A}},L^{\tau}_{\beta}),({\mathcal{A}}_{n},L^{\tau}_{\beta}))\leqslant\sum_{k=n}^{\infty}\beta_{k},

where Λ{\mathsf{\Lambda}} is the quantum Gromov-Hausdorff propinquity of [15].

The main examples of AF algebras in this article, Effros-Shen algebras and UHF algebras, are given in the setting of inductive limits of finite-dimensional C*-algebras. Thus, we introduce notation to prove results in this setting.

Let (n,αn)n({\mathcal{B}}_{n},\alpha_{n})_{n\in{\mathds{N}}} be an inductive sequence of C*-algebras (see [20, Section 6.1]) such that:

  1. (1)

    0={\mathcal{B}}_{0}={\mathds{C}} and n=k=1nn𝔐dn,k(){\mathcal{B}}_{n}=\bigoplus_{k=1}^{n_{n}}{\mathfrak{M}}_{d_{n,k}}({\mathds{C}}) for all n{0}n\in{\mathds{N}}\setminus\{0\}, where dn,k{0}d_{n,k}\in{\mathds{N}}\setminus\{0\} for each n{0}n\in{\mathds{N}}\setminus\{0\} and k{1,2,,nn}k\in\{1,2,\ldots,n_{n}\};

  2. (2)

    αn:nn+1\alpha_{n}:{\mathcal{B}}_{n}\rightarrow{\mathcal{B}}_{n+1} is a unital *-monomorphism for all nn\in{\mathds{N}};

  3. (3)

    the inductive limit 𝒜=lim(n,αn)n{\mathcal{A}}=\underrightarrow{\lim}\ ({\mathcal{B}}_{n},\alpha_{n})_{n\in{\mathds{N}}} is equipped with a faithful tracial state τ\tau.

For each nn\in{\mathds{N}}, let α(n):n𝒜\alpha^{(n)}:{\mathcal{B}}_{n}\rightarrow{\mathcal{A}} be the canonical unital *-monomorphism satisfying

(2.2) α(n+1)αn=α(n),\alpha^{(n+1)}\circ\alpha_{n}=\alpha^{(n)},

and if for each k{1,2,,n1}k\in\{1,2,\ldots,n-1\}, we define

αk,n=αnαn1αk,\alpha_{k,n}=\alpha_{n}\circ\alpha_{n-1}\circ\cdots\alpha_{k},

then inductively, we have

(2.3) α(n+1)αk,n=α(k)\alpha^{(n+1)}\circ\alpha_{k,n}=\alpha^{(k)}

Note that 𝒜=nα(n)(n)¯𝒜{\mathcal{A}}=\overline{\cup_{n\in{\mathds{N}}}\alpha^{(n)}({\mathcal{B}}_{n})}^{\|\cdot\|_{\mathcal{A}}} and α(n)(n)α(n+1)(n+1)\alpha^{(n)}({\mathcal{B}}_{n})\subseteq\alpha^{(n+1)}({\mathcal{B}}_{n+1}) and α(0)(0)=1𝒜\alpha^{(0)}({\mathcal{B}}_{0})={\mathds{C}}1_{\mathcal{A}} (see [20, Section 6.1]). So, for each nn\in{\mathds{N}}, set

𝒜n=α(n)(n).{\mathcal{A}}_{n}=\alpha^{(n)}({\mathcal{B}}_{n}).

As above, for each nn\in{\mathds{N}}, let

Enτ:𝒜𝒜nE^{\tau}_{n}:{\mathcal{A}}\rightarrow{\mathcal{A}}_{n}

denote the unique τ\tau-preserving faithful conditional expectation onto 𝒜n.{\mathcal{A}}_{n}. For each nn\in{\mathds{N}}, let

(2.4) τn=τα(n),\tau_{n}=\tau\circ\alpha^{(n)},

which is a faithful tracial state on n{\mathcal{B}}_{n} and let πτn\pi_{\tau_{n}} denote the associated GNS representation. Let k{0,1,,n+1}k\in\{0,1,\ldots,n+1\} let

En+1,kτn+1:n+1αk,n(k)E^{\tau_{n+1}}_{n+1,k}:{\mathcal{B}}_{n+1}\rightarrow\alpha_{k,n}({\mathcal{B}}_{k})

be the unique τn+1\tau_{n+1}-preserving faithful conditional expectation onto αk,n(k)\alpha_{k,n}({\mathcal{B}}_{k}). Define

Qn+1,kτn+1=En+1,kτn+1En+1,k1τn+1Q^{\tau_{n+1}}_{n+1,k}=E^{\tau_{n+1}}_{n+1,k}-E^{\tau_{n+1}}_{n+1,k-1}

and let

Dβτn+1=k=1n+1aβ,kτQn+1,kτn+1.D^{\tau_{n+1}}_{\beta}=\sum_{k=1}^{n+1}a^{\tau}_{\beta,k}Q^{\tau_{n+1}}_{n+1,k}.

For every an+1a\in{\mathcal{B}}_{n+1}, define

(2.5) Lβτn+1(a)=[Dβτn+1,πτn+1(a)]B(n+1τn+1).L^{\tau_{n+1}}_{\beta}(a)=\|[D^{\tau_{n+1}}_{\beta},\pi_{\tau_{n+1}}(a)]\|_{B({\mathcal{B}}^{\tau_{n+1}}_{n+1})}.

By finite dimensionality, we have that

(n+1,Lβτn+1)({\mathcal{B}}_{n+1},L^{\tau_{n+1}}_{\beta})

is a Leibniz compact quantum metric space. We can now add to Proposition 2.3 in the inductive limit setting.

Theorem 2.5.

Let nn\in{\mathds{N}}. It holds that

Lβτα(n)(a)=sup{[Dβτ,πτ(α(n)(a))](α(n)(b))τ:bn,bτn1}=Lβτn(a)L^{\tau}_{\beta}\circ\alpha^{(n)}(a)=\sup\{\|[D^{\tau}_{\beta},\pi_{\tau}(\alpha^{(n)}(a))](\alpha^{(n)}(b))\|_{\tau}:b\in{\mathcal{B}}_{n},\|b\|_{\tau_{n}}\leqslant 1\}=L^{\tau_{n}}_{\beta}(a)

for every ana\in{\mathcal{B}}_{n}.

Proof.

Let nn\in{\mathds{N}} and let ana\in{\mathcal{B}}_{n}, then by Proposition 2.3

Lβτα(n)(a)\displaystyle L^{\tau}_{\beta}\circ\alpha^{(n)}(a) =Lβτ(α(n)(a))\displaystyle=L^{\tau}_{\beta}(\alpha^{(n)}(a))
=[Dβτ,πτ(α(n)(a))]B(𝒜nτ)\displaystyle=\|[D^{\tau}_{\beta},\pi_{\tau}(\alpha^{(n)}(a))]\|_{B({\mathcal{A}}^{\tau}_{n})}
=sup{[Dβτ,πτ(α(n)(a))](c)τ:c𝒜n,cτ1}.\displaystyle=\sup\{\|[D^{\tau}_{\beta},\pi_{\tau}(\alpha^{(n)}(a))](c)\|_{\tau}:c\in{\mathcal{A}}_{n},\|c\|_{\tau}\leqslant 1\}.

Consider c𝒜nc\in{\mathcal{A}}_{n}, then there exists a unique bnb\in{\mathcal{B}}_{n} such that α(n)(b)=c\alpha^{(n)}(b)=c. We have

cτ2=τ(cc)=τ(α(n)(b)α(n)(b))=τ(α(n)(bb))=τn(bb)=bτn2.\displaystyle\|c\|_{\tau}^{2}=\tau(c^{*}c)=\tau\left(\alpha^{(n)}(b)^{*}\alpha^{(n)}(b)\right)=\tau\left(\alpha^{(n)}(b^{*}b)\right)=\tau_{n}(b^{*}b)=\|b\|^{2}_{\tau_{n}}.

Hence

Lβτα(n)(a)=sup{[Dβτ,πτ(α(n)(a))](α(n)(b))τ:bn,bτn1}L^{\tau}_{\beta}\circ\alpha^{(n)}(a)=\sup\{\|[D^{\tau}_{\beta},\pi_{\tau}(\alpha^{(n)}(a))](\alpha^{(n)}(b))\|_{\tau}:b\in{\mathcal{B}}_{n},\|b\|_{\tau_{n}}\leqslant 1\}

Let bnb\in{\mathcal{B}}_{n}. Let k{0,1,,n}k\in\{0,1,\ldots,n\}. Then a similar argument as the beginning of the proof of [1, Proposition 3.5] provides

Ekτα(n)=α(n)En,kτnE^{\tau}_{k}\circ\alpha^{(n)}=\alpha^{(n)}\circ E^{\tau_{n}}_{n,k}

and

Ek1τα(n)=α(n)En,k1τnE^{\tau}_{k-1}\circ\alpha^{(n)}=\alpha^{(n)}\circ E^{\tau_{n}}_{n,k-1}

by Expression (2.3). Next, we have

[Qkτ,πτ(α(n)(a))](α(n)(b))\displaystyle[Q^{\tau}_{k},\pi_{\tau}(\alpha^{(n)}(a))](\alpha^{(n)}(b))
=((EkτEk1τ)πτ(α(n)(a))πτ(α(n)(a))(EkτEk1τ))(α(n)(b))\displaystyle\quad=((E^{\tau}_{k}-E^{\tau}_{k-1})\pi_{\tau}(\alpha^{(n)}(a))-\pi_{\tau}(\alpha^{(n)}(a))(E^{\tau}_{k}-E^{\tau}_{k-1}))(\alpha^{(n)}(b))

Now

πτ(α(n)(a))(EkτEk1τ)(α(n)(b))\displaystyle\pi_{\tau}(\alpha^{(n)}(a))(E^{\tau}_{k}-E^{\tau}_{k-1})(\alpha^{(n)}(b)) =πτ(α(n)(a))(α(n)(En,kτn(b))α(n)(En,k1τn(b)))\displaystyle=\pi_{\tau}(\alpha^{(n)}(a))(\alpha^{(n)}(E^{\tau_{n}}_{n,k}(b))-\alpha^{(n)}(E^{\tau_{n}}_{n,k-1}(b)))
=α(n)(a)(α(n)(En,kτn(b))α(n)(En,k1τn(b)))\displaystyle=\alpha^{(n)}(a)(\alpha^{(n)}(E^{\tau_{n}}_{n,k}(b))-\alpha^{(n)}(E^{\tau_{n}}_{n,k-1}(b)))
=α(n)(aEn,kτn(b)aEn,k1τn(b))\displaystyle=\alpha^{(n)}(aE^{\tau_{n}}_{n,k}(b)-aE^{\tau_{n}}_{n,k-1}(b))

and similarly

(EkτEk1τ)πτ(α(n)(a))(α(n)(b))=α(n)(En,kτn(ab)En,k1τn(ab)).(E^{\tau}_{k}-E^{\tau}_{k-1})\pi_{\tau}(\alpha^{(n)}(a))(\alpha^{(n)}(b))=\alpha^{(n)}(E^{\tau_{n}}_{n,k}(ab)-E^{\tau_{n}}_{n,k-1}(ab)).

Thus

[Qkτ,πτ(α(n)(a))](α(n)(b))\displaystyle[Q^{\tau}_{k},\pi_{\tau}(\alpha^{(n)}(a))](\alpha^{(n)}(b))
=α(n)(En,kτn(ab)En,k1τn(ab)(aEn,kτn(b)aEn,k1τn(b))).\displaystyle\quad=\alpha^{(n)}\left(E^{\tau_{n}}_{n,k}(ab)-E^{\tau_{n}}_{n,k-1}(ab)-(aE^{\tau_{n}}_{n,k}(b)-aE^{\tau_{n}}_{n,k-1}(b))\right).

However,

En,kτn(ab)En,k1τn(ab)(aEn,kτn(b)aEn,k1τn(b))\displaystyle E^{\tau_{n}}_{n,k}(ab)-E^{\tau_{n}}_{n,k-1}(ab)-(aE^{\tau_{n}}_{n,k}(b)-aE^{\tau_{n}}_{n,k-1}(b))
=En,kτn(πτn(a)(b))En,k1τn(πτn(a)(b))\displaystyle\quad=E^{\tau_{n}}_{n,k}(\pi_{\tau_{n}}(a)(b))-E^{\tau_{n}}_{n,k-1}(\pi_{\tau_{n}}(a)(b))
(πτn(a)(En,kτn(b))πτn(a)(En,k1τn(b)))\displaystyle\quad\quad-(\pi_{\tau_{n}}(a)(E^{\tau_{n}}_{n,k}(b))-\pi_{\tau_{n}}(a)(E^{\tau_{n}}_{n,k-1}(b)))
=(En,kτnEn,k1τn)(πτn(a)(b))πτn(a)((En,kτnEn,k1τn)(b))\displaystyle\quad=(E^{\tau_{n}}_{n,k}-E^{\tau_{n}}_{n,k-1})(\pi_{\tau_{n}}(a)(b))-\pi_{\tau_{n}}(a)((E^{\tau_{n}}_{n,k}-E^{\tau_{n}}_{n,k-1})(b))
=Qn,kτn(πτn(a)(b))πτn(a)((Qn,kτn)(b))\displaystyle\quad=Q^{\tau_{n}}_{n,k}(\pi_{\tau_{n}}(a)(b))-\pi_{\tau_{n}}(a)((Q^{\tau_{n}}_{n,k})(b))
=(Qn,kτn(πτn(a))πτn(a)(Qn,kτn))(b)\displaystyle\quad=(Q^{\tau_{n}}_{n,k}(\pi_{\tau_{n}}(a))-\pi_{\tau_{n}}(a)(Q^{\tau_{n}}_{n,k}))(b)
=[Qn,kτn,πτn(a)](b).\displaystyle\quad=[Q^{\tau_{n}}_{n,k},\pi_{\tau_{n}}(a)](b).

Hence

[Dβτ,πτ(α(n)(a))](α(n)(b))=α(n)([Dβτn,πτn(a)](b))[D^{\tau}_{\beta},\pi_{\tau}(\alpha^{(n)}(a))](\alpha^{(n)}(b))=\alpha^{(n)}([D^{\tau_{n}}_{\beta},\pi_{\tau_{n}}(a)](b))

and so as above

[Dβτ,πτ(α(n)(a))](α(n)(b))τ=[Dβτn,πτn(a)](b)τn.\|[D^{\tau}_{\beta},\pi_{\tau}(\alpha^{(n)}(a))](\alpha^{(n)}(b))\|_{\tau}=\|[D^{\tau_{n}}_{\beta},\pi_{\tau_{n}}(a)](b)\|_{\tau_{n}}.

Therefore

Lβτα(n)(a)=sup{[Dβτ,πτ(α(n)(a))](α(n)(b))τ:bn,bτn1}=Lβτn(a)L^{\tau}_{\beta}\circ\alpha^{(n)}(a)=\sup\{\|[D^{\tau}_{\beta},\pi_{\tau}(\alpha^{(n)}(a))](\alpha^{(n)}(b))\|_{\tau}:b\in{\mathcal{B}}_{n},\|b\|_{\tau_{n}}\leqslant 1\}=L^{\tau_{n}}_{\beta}(a)

of Expression (2.5) as desired. ∎

Now that we have an expression for the L-seminorms on the terms of a given inductive sequence, we would like to show that these form a continuous field of L-seminorms with respect to weak* convergence of the faithful tracial state. However, since the norms defining our L-seminorms are operator norms this takes some care, which is why we need some tools from metric geometry. The following result might be known in metric geometry, but we cannot find a proof and so we provide one here. The following result also serves as a generalization of [1, Lemma 3.9].

Lemma 2.6.

Let (X,d)(X,d) be a metric space. Let (Cn)n(C_{n})_{n\in{\mathds{N}}} be a sequence of compact subsets of XX that converges in the Hausdorff distance with respect to dd, Hausd,Haus_{d}, to a compact CXC\subseteq X. Let CXC^{\prime}\subseteq X be a compact set such that C(nCn)CC\cup(\cup_{n\in{\mathds{N}}}C_{n})\subseteq C^{\prime}. Let (fn)n(f_{n})_{n\in{\mathds{N}}} be a sequence of real-valued continuous functions on XX and let f:Xf:X\rightarrow{\mathds{R}} be continuous.

If (fn)n(f_{n})_{n\in{\mathds{N}}} converges uniformly to ff on CC^{\prime}, then (supxCnfn(x))n(\sup_{x\in C_{n}}f_{n}(x))_{n\in{\mathds{N}}} converges to supxCf(x)\sup_{x\in C}f(x) in the usual topology on .{\mathds{R}}.

Proof.

Let ε>0\varepsilon>0. By uniform convergence, there exists δ>0\delta>0 such that for every a,bCa,b\in C^{\prime} and nn\in{\mathds{N}}, we have

|fn(a)fn(b)|<ε/2.|f_{n}(a)-f_{n}(b)|<\varepsilon/2.

Let NN\in{\mathds{N}} such that for every nNn\geqslant N

Hausd(Cn,C)<δ/3Haus_{d}(C_{n},C)<\delta/3

and

|supxCfn(x)supxCf(x)|<ε/2|\sup_{x\in C}f_{n}(x)-\sup_{x\in C}f(x)|<\varepsilon/2

by [1, Lemma 3.9].

Let nNn\geqslant N. By compact, there exists xCx^{\prime}\in C such that

supxCfn(x)=fn(x).\sup_{x\in C}f_{n}(x)=f_{n}(x^{\prime}).

Now consider supxCnfn(x).\sup_{x\in C_{n}}f_{n}(x). Assume by way of contradiction that |supxCfn(x)supxCnfn(x)|>ε/2.|\sup_{x\in C}f_{n}(x)-\sup_{x\in C_{n}}f_{n}(x)|>\varepsilon/2. Assume first that

supxCfn(x)supxCnfn(x)>ε/2\sup_{x\in C}f_{n}(x)-\sup_{x\in C_{n}}f_{n}(x)>\varepsilon/2
fn(x)supxCnfn(x)>ε/2f_{n}(x^{\prime})-\sup_{x\in C_{n}}f_{n}(x)>\varepsilon/2
supxCnfn(x)<fn(x)ε/2.\sup_{x\in C_{n}}f_{n}(x)<f_{n}(x^{\prime})-\varepsilon/2.

Hence

fn(x)<fn(x)ε/2f_{n}(x)<f_{n}(x^{\prime})-\varepsilon/2

for every xCnx\in C_{n}. Now there exists xCnx\in C_{n} such that 𝖽(x,x)<δ{\mathsf{d}}(x,x^{\prime})<\delta by definition of the Hausdorff distance. Hence

|fn(x)fn(x)|<ε/2.|f_{n}(x)-f_{n}(x^{\prime})|<\varepsilon/2.

And so

fn(x)ε/2<fn(x)<fn(x)ε/2,f_{n}(x^{\prime})-\varepsilon/2<f_{n}(x)<f_{n}(x^{\prime})-\varepsilon/2,

contradiction.

On the other hand, if supxCnfn(x)supxCfn(x)>ε/2\sup_{x\in C_{n}}f_{n}(x)-\sup_{x\in C}f_{n}(x)>\varepsilon/2. Then

ε/2<supxCnfn(x)fn(x).\varepsilon/2<\sup_{x\in C_{n}}f_{n}(x)-f_{n}(x^{\prime}).

Now, by compact, there exists zCnz\in C_{n} such that supxCnfn(x)=fn(z)\sup_{x\in C_{n}}f_{n}(x)=f_{n}(z). Hence

fn(x)<fn(z)ε/2f_{n}(x^{\prime})<f_{n}(z)-\varepsilon/2

and so

supxCfn(x)<fn(z)ε/2.\sup_{x\in C}f_{n}(x)<f_{n}(z)-\varepsilon/2.

Thus

fn(x)<fn(z)ε/2f_{n}(x)<f_{n}(z)-\varepsilon/2

for every xCx\in C. This leads to a similar contradiction. Hence,

|supxCnfn(x)supxCfn(x)|ε/2.|\sup_{x\in C_{n}}f_{n}(x)-\sup_{x\in C}f_{n}(x)|\leqslant\varepsilon/2.

Finally,

|supxCnfn(x)supxCf(x)|\displaystyle|\sup_{x\in C_{n}}f_{n}(x)-\sup_{x\in C}f(x)| |supxCnfn(x)supxCfn(x)|+|supxCfn(x)supxCf(x)|\displaystyle\leqslant|\sup_{x\in C_{n}}f_{n}(x)-\sup_{x\in C}f_{n}(x)|+|\sup_{x\in C}f_{n}(x)-\sup_{x\in C}f(x)|
ε/2+|supxCfn(x)supxCf(x)|<ε/2+ε/2=ε.\displaystyle\leqslant\varepsilon/2+|\sup_{x\in C}f_{n}(x)-\sup_{x\in C}f(x)|<\varepsilon/2+\varepsilon/2=\varepsilon.\qed

Before providing continuous fields of LL-seminorms, we need one more result so that we can satisfy the hypothesis of the previous Lemma.

Proposition 2.7.

Let ¯={}.\overline{{\mathds{N}}}={\mathds{N}}\cup\{\infty\}. Let {\mathcal{B}} be a finite-dimensional C*-algebra and let (τn)n¯(\tau^{n})_{n\in\overline{{\mathds{N}}}} be a sequence of faithful tracial states on {\mathcal{B}} such that (τn)n(\tau^{n})_{n\in{\mathds{N}}} weak* converges to τ\tau_{\infty}. For each n¯n\in\overline{{\mathds{N}}}, define Cn={b:bτn1}C_{n}=\{b\in{\mathcal{B}}:\|b\|_{\tau^{n}}\leqslant 1\}.

It holds that (Cn)n(C_{n})_{n\in{\mathds{N}}} converges to CC_{\infty} in the Hausdorff distance with respect to \|\cdot\|_{\mathcal{B}}.

Proof.

Since {\mathcal{B}} is finite dimensional there exist N,m1,m2,,mNN\in{\mathds{N}},m_{1},m_{2},\ldots,m_{N}\in{\mathds{N}} and a *-isomorphism α:k=1NMmk()\alpha:\oplus_{k=1}^{N}M_{m_{k}}({\mathds{C}})\rightarrow{\mathcal{B}} onto {\mathcal{B}}. Set k=1NMnk()=𝒜\oplus_{k=1}^{N}M_{n_{k}}({\mathds{C}})={\mathcal{A}}. For each n¯n\in\overline{{\mathds{N}}}, define that σn=τnα\sigma^{n}=\tau^{n}\circ\alpha. We have that (σn)n(\sigma^{n})_{n\in{\mathds{N}}} is s sequence of faithful tracial states that weak* converges to σ\sigma_{\infty}. Let n¯n\in\overline{{\mathds{N}}}, since σn\sigma^{n} is a faithful tracial state there exist μ1n,μ2n,,μNn(0,)\mu^{n}_{1},\mu^{n}_{2},\ldots,\mu^{n}_{N}\in(0,\infty) such that k=1Nμkn=1\sum_{k=1}^{N}\mu^{n}_{k}=1 and

σn((a1,a2,,aN))=k=1NμknmkTr(ak)\sigma^{n}((a_{1},a_{2},\ldots,a_{N}))=\sum_{k=1}^{N}\frac{\mu^{n}_{k}}{m_{k}}\mathrm{Tr}(a_{k})

for every (a1,a2,,aN)(a_{1},a_{2},\ldots,a_{N}). By weak* convergence, we have that ((μ1n,μ2n,,μNn))n((\mu^{n}_{1},\mu^{n}_{2},\ldots,\mu^{n}_{N}))_{n\in{\mathds{N}}} converges to (μ1,μ2,,μN)(\mu^{\infty}_{1},\mu^{\infty}_{2},\ldots,\mu^{\infty}_{N}) in the product topology on N{\mathds{R}}^{N}.

Define Dn={a𝒜:aσn1}D_{n}=\{a\in{\mathcal{A}}:\|a\|_{\sigma^{n}}\leqslant 1\}. Let aDa\in D_{\infty}. Now since σn\sigma^{n} is faithful we may define

y=(μ1μ1na1,μ2μ2na2,,μNμNnaN).y=\left(\frac{\sqrt{\mu^{\infty}_{1}}}{\sqrt{\mu^{n}_{1}}}a_{1},\frac{\sqrt{\mu^{\infty}_{2}}}{\sqrt{\mu^{n}_{2}}}a_{2},\ldots,\frac{\sqrt{\mu^{\infty}_{N}}}{\sqrt{\mu^{n}_{N}}}a_{N}\right).

Thus

yσn2\displaystyle\|y\|_{\sigma^{n}}^{2} =σn(yy)\displaystyle=\sigma^{n}(y^{*}y)
=σn(μ1μ1na1a1,μ2μ2na2a2,,μNμNnaNaN)\displaystyle=\sigma^{n}\left(\frac{\mu^{\infty}_{1}}{\mu^{n}_{1}}a_{1}^{*}a_{1},\frac{\mu^{\infty}_{2}}{\mu^{n}_{2}}a_{2}^{*}a_{2},\ldots,\frac{\mu^{\infty}_{N}}{\mu^{n}_{N}}a_{N}^{*}a_{N}\right)
=k=1NμnmkTr(akak)\displaystyle=\sum_{k=1}^{N}\frac{\mu^{n}_{\infty}}{m_{k}}\mathrm{Tr}(a_{k}^{*}a_{k})
=aσ21.\displaystyle=\|a\|_{\sigma^{\infty}}^{2}\leqslant 1.

Thus yDny\in D_{n}. Next,

ay𝒜\displaystyle\|a-y\|_{\mathcal{A}} =max{a1y1Mn1(),a2y2Mn2(),,aNyNMnN()}\displaystyle=\max\{\|a_{1}-y_{1}\|_{M_{n_{1}}({\mathds{C}})},\|a_{2}-y_{2}\|_{M_{n_{2}}({\mathds{C}})},\ldots,\|a_{N}-y_{N}\|_{M_{n_{N}}({\mathds{C}})}\}

Consider k{1,2,,N}k\in\{1,2,\ldots,N\}. We have that since the operator norm is bounded by the Frobenius norm

akykMnk()\displaystyle\|a_{k}-y_{k}\|_{M_{n_{k}}({\mathds{C}})} =akμkμknakMnk()\displaystyle=\left\|a_{k}-\frac{\sqrt{\mu^{\infty}_{k}}}{\sqrt{\mu^{n}_{k}}}a_{k}\right\|_{M_{n_{k}}({\mathds{C}})}
=|1μkμkn|aMnk()\displaystyle=\left|1-\frac{\sqrt{\mu^{\infty}_{k}}}{\sqrt{\mu^{n}_{k}}}\right|\cdot\|a\|_{M_{n_{k}}({\mathds{C}})}
|1μkμkn|Tr(akak).\displaystyle\leqslant\left|1-\frac{\sqrt{\mu^{\infty}_{k}}}{\sqrt{\mu^{n}_{k}}}\right|\cdot\sqrt{\mathrm{Tr}(a_{k}^{*}a_{k})}.

However, as

k=1NμkmkTr(akak)=aσ21,\sum_{k=1}^{N}\frac{\mu^{\infty}_{k}}{m_{k}}\mathrm{Tr}(a_{k}^{*}a_{k})=\|a\|_{\sigma^{\infty}}^{2}\leqslant 1,

we have that μnmkTr(akak)1\frac{\mu^{n}_{\infty}}{m_{k}}\mathrm{Tr}(a_{k}^{*}a_{k})\leqslant 1, and so

Tr(akak)mkμk\sqrt{\mathrm{Tr}(a_{k}^{*}a_{k})}\leqslant\frac{\sqrt{m_{k}}}{\sqrt{\mu^{\infty}_{k}}}

and thus

akykMnk()|1μkμkn|mkμk\|a_{k}-y_{k}\|_{M_{n_{k}}({\mathds{C}})}\leqslant\left|1-\frac{\sqrt{\mu^{\infty}_{k}}}{\sqrt{\mu^{n}_{k}}}\right|\cdot\frac{\sqrt{m_{k}}}{\sqrt{\mu^{\infty}_{k}}}

Hence

ay𝒜max{|1μkμkn|mkμk:k{1,2,,N}}.\|a-y\|_{\mathcal{A}}\leqslant\max\left\{\left|1-\frac{\sqrt{\mu^{\infty}_{k}}}{\sqrt{\mu^{n}_{k}}}\right|\cdot\frac{\sqrt{m_{k}}}{\sqrt{\mu^{\infty}_{k}}}:k\in\{1,2,\ldots,N\}\right\}.

By a symmetric argument, we have that

Haus𝒜(Dn,D)\displaystyle Haus_{\|\cdot\|_{\mathcal{A}}}(D_{n},D_{\infty}) max{max{|1μkμkn|mkμk:k{1,2,,N}},\displaystyle\leqslant\max\Bigg{\{}\max\left\{\left|1-\frac{\sqrt{\mu^{\infty}_{k}}}{\sqrt{\mu^{n}_{k}}}\right|\cdot\frac{\sqrt{m_{k}}}{\sqrt{\mu^{\infty}_{k}}}:k\in\{1,2,\ldots,N\}\right\}\ ,
max{|1μknμk|mkμkn:k{1,2,,N}}}\displaystyle\quad\quad\quad\quad\quad\ \max\left\{\left|1-\frac{\sqrt{\mu^{n}_{k}}}{\sqrt{\mu^{\infty}_{k}}}\right|\cdot\frac{\sqrt{m_{k}}}{\sqrt{\mu^{n}_{k}}}:k\in\{1,2,\ldots,N\}\right\}\Bigg{\}}

by definition of the Hausdorff distance. Thus as (μkn)n(\mu^{n}_{k})_{n\in{\mathds{N}}} converges to μk\mu^{\infty}_{k} for each k{1,2,,N}k\in\{1,2,\ldots,N\}, we have that limnHaus𝒜(Dn,D)=0.\lim_{n\to\infty}Haus_{\|\cdot\|_{\mathcal{A}}}(D_{n},D_{\infty})=0. By construction of σn\sigma^{n} and since α\alpha is a *-isomorphism, the proof is complete. ∎

We use these results to provide continuous fields of LL-seminorms.

Theorem 2.8.

Let m¯={}m\in\overline{{\mathds{N}}}={\mathds{N}}\cup\{\infty\}. Let (τn)n¯(\tau^{n})_{n\in\overline{{\mathds{N}}}} be a sequence of faithful tracial states on 𝒜{\mathcal{A}}. If (τmn)n(\tau^{n}_{m})_{n\in{\mathds{N}}} of Expression (2.4) weak* converges to τm\tau^{\infty}_{m} on m{\mathcal{B}}_{m}, then for every ama\in{\mathcal{B}}_{m}, we have (Lβτmn(a))n(L^{\tau^{n}_{m}}_{\beta}(a))_{n\in{\mathds{N}}} of Expression (2.5) converges to Lβτm(a)L^{\tau^{\infty}_{m}}_{\beta}(a) in the usual topology on {\mathds{R}}.

Proof.

Let ama\in{\mathcal{B}}_{m}. Let n¯n\in\overline{{\mathds{N}}}, define

fn:mf_{n}:{\mathcal{B}}_{m}\rightarrow{\mathds{R}}

by fn(b)=[Dβτmn,πτmn(a)](b)τmnf_{n}(b)=\|[D^{\tau^{n}_{m}}_{\beta},\pi_{\tau^{n}_{m}}(a)](b)\|_{\tau^{n}_{m}}. Note that fnf_{n} is continuous with respect to m\|\cdot\|_{{\mathcal{B}}_{m}} by finite dimensionality.

Define

Cn={bn:bτmn1},C_{n}=\{b\in{\mathcal{B}}_{n}:\|b\|_{\tau^{n}_{m}}\leqslant 1\},

which is compact with respect to m\|\cdot\|_{{\mathcal{B}}_{m}} by finite dimensionality.

Next, we verify that all the CnC_{n}’s are contained in one compact set. By finite dimensional, there exists a sharp νn>0\nu_{n}>0 such that

mνnτmn.\|\cdot\|_{{\mathcal{B}}_{m}}\leqslant\nu_{n}\cdot\|\cdot\|_{\tau^{n}_{m}}.

By [1, Proposition 3.10], we have that (νn)n(\nu_{n})_{n\in{\mathds{N}}} converges to ν\nu_{\infty}. Hence r=supn¯νn<r=\sup_{n\in\overline{{\mathds{N}}}}\nu_{n}<\infty. Now, let bCnb\in C_{n}, then τmn1\|\cdot\|_{\tau^{n}_{m}}\leqslant 1, and so

bmνnτmnνnr.\|b\|_{{\mathcal{B}}_{m}}\leqslant\nu_{n}\cdot\|\cdot\|_{\tau^{n}_{m}}\leqslant\nu_{n}\leqslant r.

Hence b{bm:bmr}b\in\{b\in{\mathcal{B}}_{m}:\|b\|_{{\mathcal{B}}_{m}}\leqslant r\}. Set C={bm:bmr}C^{\prime}=\{b\in{\mathcal{B}}_{m}:\|b\|_{{\mathcal{B}}_{m}}\leqslant r\}. We have that CC^{\prime} is compact by finite dimensionality and that CnCC_{n}\subseteq C^{\prime} for every nN¯n\in\overline{N}.

Next, by Proposition 2.1 and by a similar argument to [1, Proposition 3.6], we have that (fn)n(f_{n})_{n\in{\mathds{N}}} converges uniformly to ff_{\infty} on any compact subset of (m,m)({\mathcal{B}}_{m},\|\cdot\|_{{\mathcal{B}}_{m}}) including CC^{\prime}. Finally, we have that (Cn)n(C_{n})_{n\in{\mathds{N}}} converges to CC_{\infty} in the Hausdorff distance with respect to m\|\cdot\|_{{\mathcal{B}}_{m}} by weak* convergence by Proposition 2.7. Therefore by Lemma 2.6, we have that

(supbCnfn(b))n=(Lβτmn(a))n(\sup_{b\in C_{n}}f_{n}(b))_{n\in{\mathds{N}}}=(L^{\tau^{n}_{m}}_{\beta}(a))_{n\in{\mathds{N}}}

converges to supbCf(b)=Lβτm(a)\sup_{b\in C_{\infty}}f_{\infty}(b)=L^{\tau^{\infty}_{m}}_{\beta}(a) in the usual topology on .{\mathds{R}}.

3. Convergence of sequences of Effros-Shen algebras and UHF algebras

We will now provide our main convergence results. But first, we need notation for each of these applications. We begin with the Effros-Shen algebras which were first defined in [8].

Let θ\theta\in{\mathds{R}} be irrational. There exists a unique sequence of integers (rnθ)n(r^{\theta}_{n})_{n\in{\mathds{N}}} with rnθ>0r^{\theta}_{n}>0 for all n{0}n\in{\mathds{N}}\setminus\{0\} such that

θ=limnr0θ+1r1θ+1r2θ+1r3θ+1+1rnθ.\theta=\lim_{n\to\infty}r_{0}^{\theta}+\cfrac{1}{r^{\theta}_{1}+\cfrac{1}{r^{\theta}_{2}+\cfrac{1}{r^{\theta}_{3}+\cfrac{1}{\ddots+\cfrac{1}{r^{\theta}_{n}}}}}}.

When θ(0,1)\theta\in(0,1), we have that r0θ=0r^{\theta}_{0}=0. The sequence (rnθ)n0(r^{\theta}_{n})_{n\in{\mathds{N}}_{0}} is the continued fraction expansion of θ\theta [10].

For each nn\in{\mathds{N}}, define

p0θ=r0θ,p1θ=1 and q0θ=1,q1θ=r1θ,p_{0}^{\theta}=r_{0}^{\theta},\quad p_{1}^{\theta}=1\quad\text{ and }\quad q_{0}^{\theta}=1,\quad q_{1}^{\theta}=r^{\theta}_{1},

and set

pn+1θ=rn+1θpnθ+pn1θp_{n+1}^{\theta}=r^{\theta}_{n+1}p_{n}^{\theta}+p_{n-1}^{\theta}

and

qn+1θ=rn+1θqnθ+qn1θ.q_{n+1}^{\theta}=r^{\theta}_{n+1}q_{n}^{\theta}+q_{n-1}^{\theta}.

The sequence (pnθ/qnθ)n0\left(p_{n}^{\theta}/q_{n}^{\theta}\right)_{n\in\mathbb{N}_{0}} of convergents pnθ/qnθp^{\theta}_{n}/q^{\theta}_{n} converges to θ\theta. In fact, for each nn\in{\mathds{N}},

pnθqnθ=r0θ+1r1θ+1r2θ+1r3θ+1+1rnθ.\frac{p_{n}^{\theta}}{q_{n}^{\theta}}=r_{0}^{\theta}+\cfrac{1}{r^{\theta}_{1}+\cfrac{1}{r^{\theta}_{2}+\cfrac{1}{r^{\theta}_{3}+\cfrac{1}{\ddots+\cfrac{1}{r^{\theta}_{n}}}}}}.

We now define the terms for the inductive sequence that form the Effros-Shen algebras. Let θ,0={\mathcal{B}}_{\theta,0}=\mathbb{C} and, for each n0n\in\mathbb{N}_{0}, let

θ,n=Mqnθ()Mqn1θ(){\mathcal{B}}_{\theta,n}=M_{q_{n}^{\theta}}({\mathds{C}})\oplus M_{q_{n-1}^{\theta}}({\mathds{C}})

and for each nn\in{\mathds{N}}, set 𝒜θ,n=α(n)(θ,n){\mathcal{A}}_{\theta,n}=\alpha^{(n)}({\mathcal{B}}_{\theta,n}).

These form an inductive sequence with the maps

(3.1) αθ,n:abθ,ndiag(a,,a,b)aθ,n+1,\alpha_{\theta,n}:a\oplus b\in{\mathcal{B}}_{\theta,n}\mapsto\mathrm{diag}\left(a,\ldots,a,b\right)\oplus a\in{\mathcal{B}}_{\theta,n+1},

where there are rn+1θr^{\theta}_{n+1} copies of aa on the diagonal in the first summand of θ,n+1{\mathcal{B}}_{\theta,n+1}. This is a unital *-monomorphism by construction. For n=0n=0,

αθ,0:λθ,0diag(λ,,λ)λθ,1.\alpha_{\theta,0}:\lambda\in{\mathcal{B}}_{\theta,0}\mapsto\mathrm{diag}(\lambda,\ldots,\lambda)\oplus\lambda\ \in{\mathcal{B}}_{\theta,1}.

The Effros–Shen algebra associated to θ\theta is the inductive limit (see [20, Section 6.1])

𝒜θ=lim(θ,n,αθ,n)n{\mathcal{A}}_{\theta}=\underrightarrow{\lim}\ ({\mathcal{B}}_{\theta,n},\alpha_{\theta,n})_{n\in{\mathds{N}}}

by [8].

There exists a unique faithful tracial state τθ\tau^{\theta} on 𝒜θ{\mathcal{A}}_{\theta} such that for each n{0}n\in{\mathds{N}}\setminus\{0\}, τθ,n\tau_{\theta,n} (see Expression (2.4)) is defined for each (a,b)θ,n(a,b)\in{\mathcal{B}}_{\theta,n} by

τθ,n(a,b)=t(θ,n)1qnθTr(a)+(1t(θ,n))1qn1θTr(b),\tau_{\theta,n}(a,b)=t(\theta,n)\frac{1}{q_{n}^{\theta}}\mathrm{Tr}(a)+(1-t(\theta,n))\frac{1}{q_{n-1}^{\theta}}\mathrm{Tr}(b),

where

t(θ,n)=(1)n1qnθ(θqn1θpn1θ)(0,1)t(\theta,n)=(-1)^{n-1}q_{n}^{\theta}(\theta q_{n-1}^{\theta}-p_{n-1}^{\theta})\in(0,1)

(see [2, Lemma 5.5]).

For each nn\in{\mathds{N}}, define

(3.2) βnθ=1dim(𝒜θ,n)=1(qnθ)2+(qn1θ)2,\beta^{\theta}_{n}=\frac{1}{\dim({\mathcal{A}}_{\theta,n})}=\frac{1}{(q_{n}^{\theta})^{2}+(q_{n-1}^{\theta})^{2}},

and note that (βnθ)n(\beta^{\theta}_{n})_{n\in{\mathds{N}}} is summable by [10]. Finally, for each nn\in{\mathds{N}}, define

anτθ=cnτθβnθa^{\tau_{\theta}}_{n}=\frac{c^{\tau^{\theta}}_{n}}{\beta^{\theta}_{n}}

where cnτθc^{\tau^{\theta}}_{n} is given by Expression (2.1).

Theorem 3.1.

The map

θ(0,1)(𝒜θ,Lβθτθ)\theta\in(0,1)\setminus{\mathds{Q}}\longmapsto({\mathcal{A}}_{\theta},L^{\tau_{\theta}}_{\beta_{\theta}})

is continuous with respect to the quantum Gromov-Hausdorff propinquity of [15] where LβθτθL^{\tau_{\theta}}_{\beta_{\theta}} is given by Theorem 2.2.

Proof.

Note that for every θ(0,1)\theta\in(0,1)\setminus{\mathds{Q}} there exists a summable sequence of positive reals (βn)n(\beta_{n})_{n\in{\mathds{N}}} such that βnθβn\beta^{\theta}_{n}\leqslant\beta_{n} for every nn\in{\mathds{N}} (see the beginning of the proof of [2, Theorem 5.14]. Now, let (θn)n(\theta^{n})_{n\in{\mathds{N}}} be a sequence in (0,1)(0,1)\setminus{\mathds{Q}} that converges to some θ(0,1)\theta\in(0,1)\setminus{\mathds{Q}} with respect to the usual topology on {\mathds{R}}. Let ε>0\varepsilon>0. Choose N1N_{1}\in{\mathds{N}} such that k=nβn<ε/3\sum_{k=n}^{\infty}\beta_{n}<\varepsilon/3 for every nN1n\geqslant N_{1}. Now choose N2N_{2}\in{\mathds{N}} such that N2N1N_{2}\geqslant N_{1} and qkθn=qkθq^{\theta_{n}}_{k}=q^{\theta}_{k} for every nN1n\geqslant N_{1} and k{1,2,,N1}k\in\{1,2,\ldots,N_{1}\} which is possible by [2, Proposition 5.10]. Thus, for every nN2n\geqslant N_{2}, we have θn,k=θ,k{\mathcal{B}}_{\theta_{n},k}={\mathcal{B}}_{\theta,k} and αθn,k=αθ,k\alpha_{\theta_{n},k}=\alpha_{\theta,k} for every kN1k\leqslant N_{1}. Now by [2, Lemma 5.5], we have that (τθl,N1)lN2(\tau_{\theta_{l},N_{1}})_{l\geqslant N_{2}} converges to τθ,N1\tau_{\theta,N_{1}} in the weak* topology. Thus, by the same proof as [2, Lemma 5.13], we have that there exists N3N2N_{3}\geqslant N_{2} such that

Λ((N1,Lβnθτθn,N1),(N1,Lβθτθ,N1))<ε/3{\mathsf{\Lambda}}(({\mathcal{B}}_{N_{1}},L^{\tau_{\theta_{n},N_{1}}}_{\beta^{\theta}_{n}}),({\mathcal{B}}_{N_{1}},L^{\tau_{\theta,N_{1}}}_{\beta_{\theta}}))<\varepsilon/3

by Theorem 2.8. Let nN3n\geqslant N_{3}. By Theorem 2.4 and Theorem 2.5 and the triangle inequality, we have

Λ((𝒜θn,Lβθnτθn),(𝒜θ,Lβθτθ))\displaystyle{\mathsf{\Lambda}}(({\mathcal{A}}_{\theta_{n}},L^{\tau^{\theta_{n}}}_{\beta^{\theta_{n}}}),({\mathcal{A}}_{\theta},L^{\tau^{\theta}}_{\beta^{\theta}}))
Λ((𝒜θn,Lβθnτθn),(𝒜θn,N1,Lβθnτθn))\displaystyle\quad\leqslant{\mathsf{\Lambda}}(({\mathcal{A}}_{\theta_{n}},L^{\tau^{\theta_{n}}}_{\beta^{\theta_{n}}}),({\mathcal{A}}_{\theta_{n},N_{1}},L^{\tau^{\theta_{n}}}_{\beta^{\theta_{n}}}))
+Λ((𝒜θn,N1,Lβθnτθn),(𝒜θn,N1,Lβθτθ))\displaystyle\quad\quad+{\mathsf{\Lambda}}(({\mathcal{A}}_{\theta_{n},N_{1}},L^{\tau^{\theta_{n}}}_{\beta^{\theta_{n}}}),({\mathcal{A}}_{\theta_{n},N_{1}},L^{\tau^{\theta}}_{\beta^{\theta}}))
+Λ((𝒜θn,N1,Lβθτθ),(𝒜θ,Lβθτθ))\displaystyle\quad\quad\quad+{\mathsf{\Lambda}}(({\mathcal{A}}_{\theta_{n},N_{1}},L^{\tau^{\theta}}_{\beta^{\theta}}),({\mathcal{A}}_{\theta},L^{\tau^{\theta}}_{\beta^{\theta}}))
k=N1βkθn+Λ((𝒜θn,N1,Lβθnτθn),(𝒜θn,N1,Lβθτθ))+k=N1βkθn\displaystyle\leqslant\sum_{k=N_{1}}^{\infty}\beta^{\theta_{n}}_{k}+{\mathsf{\Lambda}}(({\mathcal{A}}_{\theta_{n},N_{1}},L^{\tau^{\theta_{n}}}_{\beta^{\theta_{n}}}),({\mathcal{A}}_{\theta_{n},N_{1}},L^{\tau^{\theta}}_{\beta^{\theta}}))+\sum_{k=N_{1}}^{\infty}\beta^{\theta_{n}}_{k}
k=N1βk+Λ((𝒜θn,N1,Lβθnτθn),(𝒜θn,N1,Lβθτθ))+k=N1βk\displaystyle\leqslant\sum_{k=N_{1}}^{\infty}\beta_{k}+{\mathsf{\Lambda}}(({\mathcal{A}}_{\theta_{n},N_{1}},L^{\tau^{\theta_{n}}}_{\beta^{\theta_{n}}}),({\mathcal{A}}_{\theta_{n},N_{1}},L^{\tau^{\theta}}_{\beta^{\theta}}))+\sum_{k=N_{1}}^{\infty}\beta_{k}
<ε/3+Λ((𝒜θn,N1,Lβθnτθn),(𝒜θn,N1,Lβθτθ))+ε/3\displaystyle<\varepsilon/3+{\mathsf{\Lambda}}(({\mathcal{A}}_{\theta_{n},N_{1}},L^{\tau^{\theta_{n}}}_{\beta^{\theta_{n}}}),({\mathcal{A}}_{\theta_{n},N_{1}},L^{\tau^{\theta}}_{\beta^{\theta}}))+\varepsilon/3
=2ε/3+Λ((N1,Lβnθτθn,N1),(N1,Lβθτθ,N1))\displaystyle=2\varepsilon/3+{\mathsf{\Lambda}}(({\mathcal{B}}_{N_{1}},L^{\tau_{\theta_{n},N_{1}}}_{\beta^{\theta}_{n}}),({\mathcal{B}}_{N_{1}},L^{\tau_{\theta,N_{1}}}_{\beta_{\theta}}))
<2ε/3+ε/3=ε\displaystyle<2\varepsilon/3+\varepsilon/3=\varepsilon

as desired. ∎

Next, we move to the UHF case.

Definition 3.2.

The Baire space 𝒩{\mathscr{N}} is the set ({0})({\mathds{N}}\setminus\{0\})^{\mathds{N}} endowed with the metric 𝖽\mathsf{d} defined, for any two (x(n))n(x(n))_{n\in{\mathds{N}}}, (y(n))n(y(n))_{n\in{\mathds{N}}} in 𝒩{\mathscr{N}}, by

d𝒩((x(n))n,(y(n))n)={0 if x(n)=y(n) for all n,2min{n:x(n)y(n)} otherwise.d_{\mathscr{N}}\left((x(n))_{n\in{\mathds{N}}},(y(n))_{n\in{\mathds{N}}}\right)=\begin{cases}0\ \ \ \ \text{ if $x(n)=y(n)$ for all $n\in{\mathds{N}}$},\\ \\ 2^{-\min\left\{n\in{\mathds{N}}:x(n)\not=y(n)\right\}}\ \ \ \text{ otherwise}\text{.}\end{cases}

Next, we define UHF algebras in a way that suits our needs. Given (β(n))n𝒩(\beta(n))_{n\in{\mathds{N}}}\in{\mathscr{N}}, let

β(n)={1 if n=0,j=0n1(β(j)+1) otherwise.\boxtimes\beta(n)=\begin{cases}1&\text{ if }n=0,\\ \prod_{j=0}^{n-1}(\beta(j)+1)&\text{ otherwise}.\end{cases}

For each nn\in{\mathds{N}}, define a unital *-monomorphism by

μβ,n:a𝔐β(n)()diag(a,a,,a)𝔐β(n+1)(),\mu_{\beta,n}:a\in{\mathfrak{M}}_{\boxtimes\beta(n)}({\mathds{C}})\longmapsto\mathrm{diag}(a,a,\ldots,a)\in{\mathfrak{M}}_{\boxtimes\beta(n+1)}({\mathds{C}}),

where there are β(n)+1\beta(n)+1 copies of aa in diag(a,a,,a)\mathrm{diag}(a,a,\ldots,a). Set 𝗎𝗁𝖿((β(n))n)=lim(𝔐β(n)(),μβ,n)n\mathsf{uhf}((\beta(n))_{n\in{\mathds{N}}})=\underrightarrow{\lim}\ ({\mathfrak{M}}_{\boxtimes\beta(n)}({\mathds{C}}),\mu_{\beta,n})_{n\in{\mathds{N}}}. The map

(β(n))n𝒩𝗎𝗁𝖿((β(n))n)(\beta(n))_{n\in{\mathds{N}}}\in{\mathscr{N}}\longmapsto\mathsf{uhf}((\beta(n))_{n\in{\mathds{N}}})

is a surjection onto the class of all UHF algebras up to *-isomorphism by [7, Chapter III.5].

For each nn\in{\mathds{N}}, let

γβ(n)=1dim(𝔐β(n)()),\gamma_{\beta}(n)=\frac{1}{\dim({\mathfrak{M}}_{\boxtimes\beta(n)}({\mathds{C}}))},

and let

ρβ\rho_{\beta}

be the unique faithful tracial state on uhf((β(n))n)uhf((\beta(n))_{n\in{\mathds{N}}}). We now state our result for continuity of UHF algebras with respect to the Baire space.

Theorem 3.3.

The map

β𝒩(𝗎𝗁𝖿(β),Lγβρβ)\beta\in{\mathscr{N}}\longmapsto(\mathsf{uhf}(\beta),L^{\rho_{\beta}}_{\gamma_{\beta}})

is continuous with respect to the quantum Gromov-Hausdorff propinquity of [15] where LγβρβL^{\rho_{\beta}}_{\gamma_{\beta}} is given by Theorem 2.2.

Proof.

This follows similarly as the proof of Theorem 3.1 since convergence in the Baire space is equivalent to convergence of irrationals by [2, Proposition 5.10]. ∎

References

  • [1] Konrad Aguilar, Stephan Ramon Garcia, Elena Kim, and Frédéric Latrémolière, The strongly Leibniz property and the Gromov-Hausdorff propinquity, J. Math. Anal. Appl. 529 (2024), no. 1, Paper No. 127572, 22. MR 4620742
  • [2] Konrad Aguilar and Frédéric Latrémolière, Quantum ultrametrics on AF algebras and the Gromov-Hausdorff propinquity, Studia Math. 231 (2015), no. 2, 149–193, arXiv: 1511.07114. MR 3465284
  • [3] Jacopo Bassi, Roberto Conti, Carla Farsi, and Frédéric Latrémolière, Isometry groups of inductive limits of metric spectral triples and gromov-hausdorff convergence, (2023), 38 pages, ArXiv: 2302.09117.
  • [4] Erik Christensen and Cristina Ivan, Spectral triples for AF CC^{*}-algebras and metrics on the Cantor set, J. Operator Theory 56 (2006), no. 1, 17–46. MR 2261610
  • [5] A. Connes, Compact metric spaces, Fredholm modules, and hyperfiniteness, Ergodic Theory Dynam. Systems 9 (1989), no. 2, 207–220. MR 1007407
  • [6] by same author, Noncommutative geometry, Academic Press, Inc., San Diego, CA, 1994. MR 1303779
  • [7] Kenneth R. Davidson, CC^{*}-algebras by example, Fields Institute Monographs, vol. 6, American Mathematical Society, Providence, RI, 1996. MR 1402012
  • [8] E. G. Effros and C. L. Shen, Approximately finite CC^{\ast}-algebras and continued fractions, Indiana Univ. Math. J. 29 (1980), no. 2, 191–204.
  • [9] Mikhael Gromov, Groups of polynomial growth and expanding maps, Inst. Hautes Études Sci. Publ. Math. (1981), no. 53, 53–73. MR 623534
  • [10] G. H. Hardy and E. M. Wright, An introduction to the theory of numbers, sixth ed., Oxford University Press, Oxford, 2008.
  • [11] D. Kerr, Matricial quantum Gromov-Hausdorff distance, J. Funct. Anal. 205 (2003), no. 1, 132–167, math.OA/0207282.
  • [12] F. Latrémolière, The modular Gromov-Hausdorff propinquity, Submitted (2016), 67 pages, ArXiv: 1608.04881.
  • [13] Frédéric Latrémolière, Convergence of fuzzy tori and quantum tori for the quantum Gromov-Hausdorff propinquity: an explicit approach, Münster J. Math. 8 (2015), no. 1, 57–98. MR 3549521
  • [14] by same author, The dual Gromov-Hausdorff propinquity, J. Math. Pures Appl. (9) 103 (2015), no. 2, 303–351. MR 3298361
  • [15] by same author, The quantum Gromov-Hausdorff propinquity, Trans. Amer. Math. Soc. 368 (2016), no. 1, 365–411. MR 3413867
  • [16] by same author, Quantum metric spaces and the Gromov-Hausdorff propinquity, Noncommutative geometry and optimal transport, Contemp. Math., vol. 676, Amer. Math. Soc., Providence, RI, 2016, pp. 47–133. MR 3578737
  • [17] by same author, Convergence of spectral triples on fuzzy tori to spectral triples on quantum tori, Comm. Math. Phys. 388 (2021), no. 2, 1049–1128. MR 4334254
  • [18] by same author, The Gromov-Hausdorff propinquity for metric spectral triples, Adv. Math. 404 (2022), no. part A, Paper No. 108393, 56. MR 4411527
  • [19] H. Li, CC^{\ast}-algebraic quantum Gromov-Hausdorff distance, (2003), ArXiv: math.OA/0312003.
  • [20] Gerard J. Murphy, CC^{*}-algebras and operator theory, Academic Press, Inc., Boston, MA, 1990. MR 1074574
  • [21] M. A. Rieffel, Metrics on states from actions of compact groups, Doc. Math. 3 (1998), 215–229. MR 1647515
  • [22] by same author, Gromov-Hausdorff distance for quantum metric spaces, vol. 168, 2004, Appendix 1 by Hanfeng Li, Gromov-Hausdorff distance for quantum metric spaces. Matrix algebras converge to the sphere for quantum Gromov-Hausdorff distance, pp. 1–65. MR 2055927
  • [23] by same author, Compact quantum metric spaces, Operator algebras, quantization, and noncommutative geometry, Contemporary Math, vol. 365, American Mathematical Society, 2005, ArXiv: 0308207, pp. 315–330.
  • [24] W. Wu, Non-commutative metric topology on matrix state spaces, Proc. Amer. Math. Soc. 134 (2006), no. 2, 443–453, ArXiv: math.OA/0410587.