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

Bunce-Deddens algebras as quantum Gromov-Hausorff distance limits of circle algebras

Konrad Aguilar [email protected] Department of Mathematics and Computer Science (IMADA)
University of Southern Denmark
Campusvej 55, DK-5230 Odense M, Denmark
Frédéric Latrémolière [email protected] http://www.math.du.edu/~frederic Department of Mathematics
University of Denver
Denver CO 80208
 and  Timothy Rainone [email protected] https://www.oxy.edu/academics/faculty/timothy-rainone Department of Mathematics
Occidental College
Los Angeles CA 90041
Abstract.

We show that Bunce-Deddens algebras, which are A𝕋{\mathds{T}}-algebras, are also limits of circle algebras for Rieffel’s quantum Gromov-Hausdorff distance, and moreover, form a continuous family indexed by the Baire space. To this end, we endow Bunce-Deddens algebras with a quantum metric structure, a step which requires that we reconcile the constructions of the Latrémolière’s Gromov-Hausdorff propinquity and Rieffel’s quantum Gromov-Hausdorff distance when working on order-unit quantum metric spaces. This work thus continues the study of the connection between inductive limits and metric limits.

Key words and phrases:
Noncommutative metric geometry, Gromov-Hausdorff convergence, Monge-Kantorovich distance, Quantum Metric Spaces, Lip-norms, Bunce-Deddens algebras, AT-algebras.
2000 Mathematics Subject Classification:
Primary: 46L89, 46L30, 58B34.
The first author gratefully acknowledges the financial support from the Independent Research Fund Denmark through the project ‘Classical and Quantum Distances’ (grant no. 9040-00107B)
The first and second author were partially supported by the grant H2020-MSCA-RISE-2015-691246-QUANTUM DY-NAMICS and the Polish Ministry of Science and Higher Education grant #3542/H2020/2016/2.
Some of this work was completed atThe Fields Institute for Research in Mathematical Sciences during theWorkshop on New Geometry of Quantum Dynamics August 12 - 16, 2019

1. Introduction

Noncommutative analogues of the Gromov-Hausdorff distance [31, 19, 16, 21, 18] allow for the discussion of limits for certain sequences unital C*-algebras endowed with a notion of a quantum metric, in a manner which generalizes the notion of convergence of compact metric spaces in the sense of the Edwards-Gromov-Hausdorff distance [9]. Of course, in C*-algebra theory, a very common notion of limit is given by inductive limits of inductive sequences, as used to great success in classification theory, among others. In particular, inductive limits of finite dimensional C*-algebras, called AF-algebras, can be seen as the beginning of the research on the classification of C*-algebras. Reconciling metric convergence and inductive limits for AF algebras has been the topic of a previous work from the first two authors [3], followed by several developments by the second author [2, 1]. A next, crucial chapter in the theory of classification, was the study of inductive limits of circle algebras, called A𝕋{\mathds{T}}-algebras. A well-known example of an A𝕋{\mathds{T}}-algebra is given by the Bunce-Deddens algebras [6], and the present work proposes to see how metric convergence can be reconciled with the notion of inductive limit for this particular family of A𝕋{\mathds{T}}-algebras. Moreover, we also prove that the function which maps an element of the Baire space — a sequence of natural numbers — to its associated Bunce-Deddens algebra is a continuous map for Rieffel’s quantum Gromov-Hausdorff distance.

The first noncommutative analogue of the Gromov-Hausdorff distance, motivated by questions from mathematical physics, was discovered by Rieffel [31]. It was constructed over the class of quantum metric order unit spaces, which Rieffel called quantum compact metric spaces — though as we shall briefly discuss, this last term’s meaning has evolved in time. The idea behind the definition of an quantum metric order unit space, inspired by an idea of Connes [7], is to generalize the structure of a space of Lipschitz maps, endowed with the Lipschitz seminorm. Rieffel especially noted that a key property of the Lipschitz seminorm on the space of real-valued functions over a compact metric space XX is that it induces by duality a distance on the space of Radon probability measures over XX which metrizes the weak* topology. This distance is of course the Monge-Kantorovich metric introduced by Kantorovich [11]. This property can be made sense of in the more general, noncommutative context.

Definition 1.1 ([4]).

A vector space with an order unit is a pair (V,1V)(V,1_{V}) of an ordered vector space (V,)(V,\leqslant) over {\mathds{R}} such that for all vVv\in V there exists λ>0\lambda>0 such that λ1Vvλ1V-\lambda 1_{V}\leqslant v\leqslant\lambda 1_{V}.

An order unit space 𝔄{\mathfrak{A}} is a vector space with an order unit 1𝔄1_{\mathfrak{A}} which is Archimedean, i.e. for all vVv\in V, if vλ1𝔄v\leqslant\lambda 1_{\mathfrak{A}} for all λ>0\lambda>0 then v0v\leqslant 0.

An order unit space comes equipped with a norm defined for all a𝔄a\in{\mathfrak{A}} by a𝔄=inf{λ>0λ1𝔄aλ1𝔄}\|a\|_{\mathfrak{A}}=\inf\{\lambda>0\mid-\lambda 1_{\mathfrak{A}}\leqslant a\leqslant\lambda 1_{\mathfrak{A}}\} and satisfying a𝔄1𝔄aa𝔄1𝔄-\|a\|_{\mathfrak{A}}1_{\mathfrak{A}}\leqslant a\leqslant\|a\|_{\mathfrak{A}}1_{\mathfrak{A}}.

Definition 1.2.

The state space 𝒮(𝔄){\mathscr{S}}({\mathfrak{A}}) of an order unit space 𝔄{\mathfrak{A}} is the set of all positive linear functionals from 𝔄{\mathfrak{A}} to {\mathds{C}} which maps the order unit of 𝔄{\mathfrak{A}} to 11.

The first occurrence of the term quantum compact metric space can be found in Connes’ work on spectral triples [7]. Rieffel proposed a definition for this term based on order unit spaces and seminorms which generalize Lipschitz seminorms on spaces of functions over compact metric spaces. The following definition encapsulates Rieffel’s notion as used in his construction of the quantum Gromov-Hausdorff distance.

Definition 1.3 ([27, 28, 31]).

A quantum metric order unit space (𝔄,𝖫)({\mathfrak{A}},{\mathsf{L}}) is an ordered pair of a norm-complete order unit space 𝔄{\mathfrak{A}} and a seminorm 𝖫{\mathsf{L}} defined on a norm-dense subspace of 𝔄{\mathfrak{A}} such that:

  1. (1)

    {a𝔄:𝖫(a)=0}=1𝔄\{a\in{\mathfrak{A}}:{\mathsf{L}}(a)=0\}={\mathds{R}}1_{\mathfrak{A}},

  2. (2)

    the Monge-Kantorovich metric, defined for any two φ,ψ𝒮(𝔄)\varphi,\psi\in{\mathscr{S}}({\mathfrak{A}}) by:

    𝗆𝗄𝖫(φ,ψ)=sup{|φ(a)ψ(a)|:a𝔄,𝖫(a)1}{\mathsf{mk}_{{\mathsf{L}}}}(\varphi,\psi)=\sup\left\{|\varphi(a)-\psi(a)|:a\in{\mathfrak{A}},{\mathsf{L}}(a)\leqslant 1\right\}\text{, }

    metrizes the weak* topology on 𝒮(𝔄){\mathscr{S}}({\mathfrak{A}}),

  3. (3)

    {a𝔄:𝖫(a)1}\{a\in{\mathfrak{A}}:{\mathsf{L}}(a)\leqslant 1\} is closed in 𝔄\|\cdot\|_{\mathfrak{A}}.

The seminorm 𝖫{\mathsf{L}} is called a Lip-norm on 𝔄{\mathfrak{A}}.

We denote the class of quantum metric order unit space by CQMSou\mathrm{CQMS}_{\mathrm{ou}}.

We note that the requirement that the unit ball of Lip-norms be closed is not included in [31]. However, as explained in [31], if a seminorm SS satisfies (1) and (2) but not (3) in Definition (1.3), then setting:

𝖫(a)=sup{|φ(a)ψ(a)|𝗆𝗄S(φ,ψ):φψ𝒮(𝔄)}{\mathsf{L}}(a)=\sup\left\{\frac{|\varphi(a)-\psi(a)|}{{\mathsf{mk}_{S}}(\varphi,\psi)}:\varphi\not=\psi\in{\mathscr{S}}({\mathfrak{A}})\right\}

allowing for \infty, gives a quantum metric order unit space (𝔄,𝖫)({\mathfrak{A}},{\mathsf{L}}) with 𝗆𝗄𝖫=𝗆𝗄S{\mathsf{mk}_{{\mathsf{L}}}}={\mathsf{mk}_{S}} and 𝖫{\mathsf{L}} is lower semicontinuous on 𝔄{\mathfrak{A}}. So Assumption (3) can always be made, and it simplifies the statement of many theorems.

Rieffel’s quantum Gromov-Hausdorff distance is thus a noncommutative analogue of the Gromov-Hausdorff distance for quantum metric order unit spaces. Its definition follows Edwards and Gromov’s ideas, though of course, the techniques needed to establish the properties of this new metric are quite different.

Definition 1.4.

Let (𝔄1,𝖫1)({\mathfrak{A}}_{1},{\mathsf{L}}_{1}) and (𝔄2,𝖫2)({\mathfrak{A}}_{2},{\mathsf{L}}_{2}) be two quantum metric order unit spaces. A Lip-norm 𝖫{\mathsf{L}} on 𝔄1𝔄2{\mathfrak{A}}_{1}\oplus{\mathfrak{A}}_{2} is admissible for 𝖫1{\mathsf{L}}_{1} and 𝖫2{\mathsf{L}}_{2} when:

j{1,2}a𝔄j𝖫j(a)=inf{𝖫(a,b):b𝔅}.\forall j\in\{1,2\}\quad\forall a\in{\mathfrak{A}}_{j}\quad{\mathsf{L}}_{j}(a)=\inf\left\{{\mathsf{L}}(a,b):b\in{\mathfrak{B}}\right\}\text{.}
Notation 1.5.

If (E,d)(E,d) is a metric space, the Hausdorff distance [10] induced by dd on the set of all closed subsets of EE is denoted by 𝖧𝖺𝗎𝗌d{\mathsf{Haus}_{d}}.

Definition 1.6.

The quantum Gromov-Hausdorff distance 𝖽𝗂𝗌𝗍q((𝔄1,𝖫1),(𝔄2,𝖫2)){\mathsf{dist}}_{q}(({\mathfrak{A}}_{1},{\mathsf{L}}_{1}),({\mathfrak{A}}_{2},{\mathsf{L}}_{2})) between two quantum metric order unit spaces (𝔄1,𝖫1)({\mathfrak{A}}_{1},{\mathsf{L}}_{1}) and (𝔄2,𝖫2)({\mathfrak{A}}_{2},{\mathsf{L}}_{2}) is:

inf{𝖧𝖺𝗎𝗌𝗆𝗄𝖫(𝒮(𝔄1),𝒮(𝔄2)):𝖫 is admissible for 𝖫1 and 𝖫2}.\inf\left\{{\mathsf{Haus}_{{\mathsf{mk}_{{\mathsf{L}}}}}}({\mathscr{S}}({\mathfrak{A}}_{1}),{\mathscr{S}}({\mathfrak{A}}_{2})):\text{${\mathsf{L}}$ is admissible for ${\mathsf{L}}_{1}$ and ${\mathsf{L}}_{2}$}\right\}\text{.}

Rieffel proved in [31] that the distance 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} is a complete pseudo-metric on the class of all quantum metric order unit spaces, which is zero between two quantum metric order unit spaces (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) if and only there exists a positive linear map π:𝔄𝔅\pi:{\mathfrak{A}}\rightarrow{\mathfrak{B}} such that 𝖫𝔅π=𝖫𝔄{\mathsf{L}}_{\mathfrak{B}}\circ\pi={\mathsf{L}}_{\mathfrak{A}}. Several examples of convergence for this metric were derived [14, 31, 29].

In time, it has become apparent that progress in noncommutative metric geometry requires a noncommutative analogue of the Gromov-Hausdorff distance for the class of quantum compact metric spaces defined, not on order unit spaces, but on actual C*-algebras, with the appropriate coincidence property. The Gromov-Hausdorff propinquity [16, 21] provides such an analogue.

Definition 1.7 ([18]).

A quantum compact metric space with the FF-Leibniz property (where F:[0,)4[0,)F:[0,\infty)^{4}\rightarrow[0,\infty) is a function which is increasing in the product order) is given by an ordered pair (𝔄,𝖫)({\mathfrak{A}},{\mathsf{L}}) where:

  1. (1)

    𝔄{\mathfrak{A}} is a unital C*-algebra,

  2. (2)

    𝖫{\mathsf{L}} is a seminorm defined on a dense domain of 𝔰𝔞(𝔄){\mathfrak{sa}\left({{\mathfrak{A}}}\right)}, where 𝔰𝔞(𝔄){\mathfrak{sa}\left({{\mathfrak{A}}}\right)} is the self-adjoint part of 𝔄{\mathfrak{A}},

  3. (3)

    (𝔰𝔞(𝔄),𝖫)({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}) is an quantum metric order unit space,

  4. (4)

    max{𝖫(ab+ba2),𝖫(abba2i)}F(a𝔄,b𝔄,𝖫(a),𝖫(b))\max\left\{{\mathsf{L}}\left(\frac{ab+ba}{2}\right),{\mathsf{L}}\left(\frac{ab-ba}{2i}\right)\right\}\leqslant F(\left\|{a}\right\|_{{\mathfrak{A}}},\left\|{b}\right\|_{{\mathfrak{A}}},{\mathsf{L}}(a),{\mathsf{L}}(b)) for all a,b𝔰𝔞(𝔄)a,b\in{\mathfrak{sa}\left({{\mathfrak{A}}}\right)}.

For any fixed function FF as above, the propinquity, denoted ΛF{\mathsf{\Lambda}^{\ast}_{F}}, is a complete metric on the class of quantum compact metric spaces with the FF-Leibniz property, with the property that the propinquity between two such quantum compact metric spaces (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) is null if and only if there exists a *-isomorphism π:𝔄𝔅\pi:{\mathfrak{A}}\rightarrow{\mathfrak{B}} such that 𝖫𝔅π=𝖫𝔄{\mathsf{L}}_{\mathfrak{B}}\circ\pi={\mathsf{L}}_{\mathfrak{A}}, and π\pi is called a full quantum isometry [18]. We note that the propinquity is a metric up to full quantum isometry on the class of all FF-Leibniz quantum compact metric spaces without fixed FF, which we denote Λ{\mathsf{\Lambda}^{\ast}}, but it is not complete in this case. The propinquity, when restricted to “classical” quantum metric spaces, is topologically equivalent to the Gromov-Hausdorff distance. New techniques are needed to prove the properties of the propinquity and utilize it since working with quantum compact metric spaces rather than quantum metric order unit spaces means working with a more rigid structure. On the other hand, the advantages of working with the propinquity become apparent as it allows for the discussion of convergence of modules [13] or convergence of group actions [12].

Many convergence results are known for the propinquity [16, 13, 3, 30]. In particular, [3], the two first authors initiated the study of convergence, in the metric sense, of sequences used to construct C*-algebras by taking inductive limits, by studying the metric properties of AF-algebras. This work was continued in subsequent papers [2]. As part of this particular line of investigation, the first author proved in [1] that the completeness of the propinquity makes it possible to define a quantum metric on the inductive limit of a sequence of C*-algebras endowed with quantum metric structures, as long as the connecting maps satisfy some natural properties.

The present work proposes to endow Bunce-Deddens algebras with quantum metrics using the same method as [1]. Thus, the idea is to introduce certain quantum metrics on circle algebras and a completeness-based argument to obtain a metric on the A𝕋{\mathds{T}}-algebra obtained as the inductive limit of these circle algebras.

There is however a difficulty in proceeding directly along these lines. Indeed, each quantum metric which we introduce on circle algebras does satisfy a form of Leibniz identity, but we can not prove that there is a uniform choice of such a Leibniz property for the entire inductive sequence for a given Bunce-Deddens algebra. This means that unfortunately, we work outside of any class where we know that the propinquity is a metric (it is of course a pseudo-metric).

Thus, in order to use the techniques of [1], we would like to understand how some of the relevant constructions in [19, 15] for the propinquity may remain valid without the Leibniz property assumptions. As we noted, there is no hope to keep the important coincidence property, but this is not directly used in [1]. However, the proof of completeness is central to the argument of [1].

We see in this paper that in fact, once we remove the constraints to work with C*-algebras and quasi-Leibniz Lip-norms, the construction of the propinquity, if mimicked, simply gives an alternate expression for Rieffel’s distance 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}. This is a very interesting fact, since it shows that indeed, the efforts placed in devising new techniques when working with the propinquity are exactly due to working with C*-algebras and Lip-norms with some Leibniz property. We stress that this does not mean that the restriction of 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} to quantum compact metric spaces is the propinquity — it is not as it still does not enjoy the appropriate coincidence property. What it means is that by allowing order unit spaces in the construction of the propinquity, we lose what makes the propinquity different from 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}. This observation is of independent interest.

Therefore, in this paper, we prove that Bunce-Deddens algebras are limits of their inductive sequences in a metric sense, for Rieffel’s distance 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}. There is one more point of subtlety which we must address here. While Rieffel proved that 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} is complete in [31], we need for our proof in this paper a different description of the limit of a Cauchy sequence for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}. The description we seek is essentially the one obtained in [16] for the propinquity. Thus, we spend some efforts carrying the proof of completeness in [16] to 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}. This turns out to be quite technical, but is carried out successfully in this paper. This is an example of how our new expression for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}, inspired by the propinquity, can lead to new observations about 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}.

Our paper is thus organized as follows. First, in the next section, we open with a general scheme to turn bi-Lipschitz morphisms between C*-algebras to quantum isometries by manipulating the quantum metrics on their codomain, under the assumption that the range of the morphisms is also the range of a conditional expectation on the codomain. This section is very general, and is concluded with a theorem about making Cauchy sequences for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} out of inductive limits of quantum compact metric spaces. We already observe that in general, this construction suffers from the problem that no quasi-Leibniz property emerges which is common to all the quantum metrics in the Cauchy sequence, forcing us to work with 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} rather than the propinquity. This second section contains the main idea of the construction of quantum metrics on circle algebras used in this paper.

We then apply our second section to the standard inductive sequences defining Bunce-Deddens algebras in the third and fourth section of this paper. We establish that all the needed ingredients required to apply the first section can be constructed for these circle algebras. We conclude with the fact that such an inductive sequence is naturally Cauchy for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}, and thus by completeness, must converge to some quantum metric order unit space. The question, of course, is whether this quantum metric order unit space is the self-adjoint part of a Bunce-Deddens algebra with some quantum metric.

To answer this question, we generalize [1] to 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} under appropriate hypothesis. In the fifth section, we establish a new expression for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} inspired, as discussed above, by the propinquity. We then see how the proof of completeness for the propinquity gives a new proof of completeness for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} which, importantly, gives a different expression for the limit of a Cauchy sequence. Of course, our expression for these limits are isomorphic as quantum metric order unit spaces to Rieffel’s, but our new expression makes it possible to relate the limit in the metric sense to the limit in the categorical sense. This matter is explained in the sixth section of the paper, where [1] is ported to our current framework. We then can answer our problem and prove that, indeed, Bunce-Deddens algebras are limits for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} of their standard inductive sequence. We use this result to also obtain a continuity result for the family of Bunce-Deddens algebras over the Baire space.

2. Distance from Conditional Expectations and bi-Lipschitz monomorphisms

We begin with a very simple observation: if a C*-subalgebra of a quantum compact metric space (𝔄,𝖫)({\mathfrak{A}},{\mathsf{L}}), containing the unit of 𝔄{\mathfrak{A}}, is the range of some conditional expectation on 𝔄{\mathfrak{A}} which is also contractive for 𝖫{\mathsf{L}}, then we can always modify 𝖫{\mathsf{L}} to make 𝔄{\mathfrak{A}} and 𝔅{\mathfrak{B}} arbitrarily close in the sense of the propinquity (though never at distance 0 unless 𝔄{\mathfrak{A}} and 𝔅{\mathfrak{B}} are *-isomorphic, of course).

During this section, we will keep track of the Leibniz conditions on our quantum metrics, precisely because in fact, it will make clear the difficulties we encountered with our construction about inductive limits. For this purpose, we set F2,0:(x,y,lx,ly)[0,)42(xly+ylx)F_{2,0}:(x,y,l_{x},l_{y})\in[0,\infty)^{4}\mapsto 2(xl_{y}+yl_{x}).

Lemma 2.1.

Let (𝔄,𝖫)({\mathfrak{A}},{\mathsf{L}}) be an FF-Leibniz quantum compact metric space and let 𝔅𝔄{\mathfrak{B}}\subseteq{\mathfrak{A}} be a C*-subalgebra of 𝔄{\mathfrak{A}} which contains the unit of 𝔅{\mathfrak{B}}. If there exists a conditional expectation 𝔼:𝔄𝔅\mathds{E}:{\mathfrak{A}}\twoheadrightarrow{\mathfrak{B}} of 𝔄{\mathfrak{A}} onto 𝔅{\mathfrak{B}} such that 𝖫𝔼𝖫{\mathsf{L}}\circ\mathds{E}\leqslant{\mathsf{L}}, then for all ε>0\varepsilon>0, if we set for all a𝔰𝔞(𝔄)a\in{\mathfrak{sa}\left({{\mathfrak{A}}}\right)}:

𝖫ε(a)=max{𝖫(a),1εa𝔼(a)𝔄}{\mathsf{L}}_{\varepsilon}(a)=\max\left\{{\mathsf{L}}(a),\frac{1}{\varepsilon}\left\|{a-\mathds{E}(a)}\right\|_{{\mathfrak{A}}}\right\}

then (𝔄,𝖫ε)({\mathfrak{A}},{\mathsf{L}}_{\varepsilon}) is a max{F,F2,0}\max\{F,F_{2,0}\}-Leibniz quantum compact metric space, 𝖫ε(b)=𝖫(b){\mathsf{L}}_{\varepsilon}(b)={\mathsf{L}}(b) for all b𝔰𝔞(𝔅)b\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}, and:

Λmax{F,F2,0}((𝔄,𝖫ε),(𝔅,𝖫))ε.{\mathsf{\Lambda}^{\ast}_{\max\{F,F_{2,0}\}}}(({\mathfrak{A}},{\mathsf{L}}_{\varepsilon}),({\mathfrak{B}},{\mathsf{L}}))\leqslant\varepsilon\text{.}
Proof.

The domain of 𝖫ε{\mathsf{L}}_{\varepsilon} is the domain of 𝖫{\mathsf{L}} since N:a𝔄1εa𝔼(a)𝔄N:a\in{\mathfrak{A}}\mapsto\frac{1}{\varepsilon}\left\|{a-\mathds{E}(a)}\right\|_{{\mathfrak{A}}} is continuous. Moreover, this also implies that 𝖫ε{\mathsf{L}}_{\varepsilon} is lower semi-continuous seminorm. Moreover, NN is (2,0)(2,0)-quasi-Leibniz by [3, Lemma 2.3], so 𝖫ε{\mathsf{L}}_{\varepsilon} is max{F,F2,0}\max\{F,F_{2,0}\}-Leibniz. Of course, if 𝖫ε(a)=0{\mathsf{L}}_{\varepsilon}(a)=0 then 𝖫(a)=0{\mathsf{L}}(a)=0 so a1𝔄a\in{\mathds{R}}1_{\mathfrak{A}}, and 𝖫ε(1𝔄)=0{\mathsf{L}}_{\varepsilon}(1_{\mathfrak{A}})=0.

Last, for any μ𝒮(𝔄)\mu\in{\mathscr{S}}({\mathfrak{A}}):

{a𝔰𝔞(𝔄):𝖫ε(a)1,μ(a)=0}{a𝔰𝔞(𝔄):𝖫(a)1,μ(a)=0}\left\{a\in{\mathfrak{sa}\left({{\mathfrak{A}}}\right)}:{\mathsf{L}}_{\varepsilon}(a)\leqslant 1,\mu(a)=0\right\}\subseteq\left\{a\in{\mathfrak{sa}\left({{\mathfrak{A}}}\right)}:{\mathsf{L}}(a)\leqslant 1,\mu(a)=0\right\}

and thus, as the set on the right hand side is totally bounded since 𝖫{\mathsf{L}} is an L-seminorm, so it the set on the left hand side. We have thus shown that (𝔄,𝖫ε)({\mathfrak{A}},{\mathsf{L}}_{\varepsilon}) is a max{F,F2,0}\max\{F,F_{2,0}\}-Leibniz quantum compact metric space using [25, Proposition 1.3].

By assumption, 𝔼(b)=b\mathds{E}(b)=b for all b𝔅b\in{\mathfrak{B}} so 𝖫ε(b)=𝖫(b){\mathsf{L}}_{\varepsilon}(b)={\mathsf{L}}(b) for all b𝔰𝔞(𝔅)b\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}.

Let π𝔄:(a,b)𝔄𝔅a𝔄\pi_{\mathfrak{A}}:(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}}\mapsto a\in{\mathfrak{A}} and π𝔅:(a,b)𝔄𝔅b𝔅\pi_{\mathfrak{B}}:(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}}\mapsto b\in{\mathfrak{B}}.

For all a𝔰𝔞(𝔄)a\in{\mathfrak{sa}\left({{\mathfrak{A}}}\right)} and b𝔰𝔞(𝔅)b\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}, we define:

𝖰(a,b)=max{𝖫ε(a),𝖫(b),1εba𝔄}.\mathsf{Q}(a,b)=\max\left\{{\mathsf{L}}_{\varepsilon}(a),{\mathsf{L}}(b),\frac{1}{\varepsilon}\left\|{b-a}\right\|_{{\mathfrak{A}}}\right\}\text{.}

If b𝔰𝔞(𝔅)b\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)} and 𝖫(b)=1{\mathsf{L}}(b)=1 then 𝖫ε(b)=1{\mathsf{L}}_{\varepsilon}(b)=1 and bb𝔄=0\left\|{b-b}\right\|_{{\mathfrak{A}}}=0. Thus 𝖰(b,b)=1\mathsf{Q}(b,b)=1. Since (a,b)𝖫(b)\mathds{Q}(a,b)\geqslant{\mathsf{L}}(b) for all a𝔰𝔞(𝔄)a\in{\mathfrak{sa}\left({{\mathfrak{A}}}\right)}, we conclude that π𝔅:𝔄𝔅𝔅\pi_{\mathfrak{B}}:{\mathfrak{A}}\oplus{\mathfrak{B}}\mapsto{\mathfrak{B}} is a quantum isometry.

If a𝔰𝔞(𝔄)a\in{\mathfrak{sa}\left({{\mathfrak{A}}}\right)} and 𝖫ε(a)=1{\mathsf{L}}_{\varepsilon}(a)=1 then 𝖫(𝔼(a))1{\mathsf{L}}(\mathds{E}(a))\leqslant 1, and moreover:

a𝔼(a)𝔄ε\left\|{a-\mathds{E}(a)}\right\|_{{\mathfrak{A}}}\leqslant\varepsilon

so 𝖰(a,𝔼(a))=1\mathsf{Q}(a,\mathds{E}(a))=1. Again, since 𝖰(a,b)𝖫ε(a)\mathsf{Q}(a,b)\geqslant{\mathsf{L}}_{\varepsilon}(a) for all b𝔰𝔞(𝔅)b\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}, we conclude that π𝔅\pi_{\mathfrak{B}} is a quantum isometry as well.

We thus gather that (𝔄𝔅,𝖰,π𝔄,π𝔅)({\mathfrak{A}}\oplus{\mathfrak{B}},\mathsf{Q},\pi_{\mathfrak{A}},\pi_{\mathfrak{B}}) is a tunnel from (𝔄,𝖫ε)({\mathfrak{A}},{\mathsf{L}}_{\varepsilon}) to (𝔅,𝖫)({\mathfrak{B}},{\mathsf{L}}) by [21, Definition 2.3]. It is of course max{F,F2,0}\max\{F,F_{2,0}\}-Leibniz. We now compute its extent.

Let φ𝒮(𝔅)\varphi\in{\mathscr{S}}({\mathfrak{B}}). Let ψ=φ𝔼𝒮(𝔄)\psi=\varphi\circ\mathds{E}\in{\mathscr{S}}({\mathfrak{A}}) and note that φ=φ𝔼\varphi=\varphi\circ\mathds{E}, so the restriction of ψ\psi to 𝔅{\mathfrak{B}} is φ\varphi — in fact for this proof, any extension ψ\psi of φ\varphi to a state of 𝔄{\mathfrak{A}} would work (but since we have our conditional expectation here, we need not invoke the Hahn-Banach theorem). Let (a,b)𝔰𝔞(𝔄𝔅)(a,b)\in{\mathfrak{sa}\left({{\mathfrak{A}}\oplus{\mathfrak{B}}}\right)} with 𝖰(a,b)1\mathsf{Q}(a,b)\leqslant 1. Then in particular, ab𝔄ε\left\|{a-b}\right\|_{{\mathfrak{A}}}\leqslant\varepsilon and thus:

|ψ(a)φ(b)|=|ψ(ab)|ε.\left|\psi(a)-\varphi(b)\right|=\left|\psi(a-b)\right|\leqslant\varepsilon\text{.}

Therefore 𝗆𝗄𝖰(φ,ψ)ε{\mathsf{mk}_{\mathsf{Q}}}(\varphi,\psi)\leqslant\varepsilon.

Let now φ𝒮(𝔄)\varphi\in{\mathscr{S}}({\mathfrak{A}}). Set ψ\psi be the restriction of φ\varphi to 𝔅{\mathfrak{B}}. Let (a,b)𝔰𝔞(𝔄𝔅)(a,b)\in{\mathfrak{sa}\left({{\mathfrak{A}}\oplus{\mathfrak{B}}}\right)} with 𝖰(a,b)1\mathsf{Q}(a,b)\leqslant 1. Then ab𝔄1\left\|{a-b}\right\|_{{\mathfrak{A}}}\leqslant 1 and thus:

|φ(b)ψ(a)|=|φ(ba)|ε.\left|\varphi(b)-\psi(a)\right|=\left|\varphi(b-a)\right|\leqslant\varepsilon\text{.}

We conclude 𝗆𝗄𝖰(φ,ψ)ε{\mathsf{mk}_{\mathsf{Q}}}(\varphi,\psi)\leqslant\varepsilon.

Consequently χ(τ)ε{\chi\left({\tau}\right)}\leqslant\varepsilon. This concludes our proof by [21, Definition 3.6]. ∎

We now turn to a result about bi-Lipschitz *-morphisms. There are several equivalent definitions of Lipschitz morphisms [28, 20], which we now recall.

Definition 2.2.

A Lipschitz morphism φ:(𝔄,𝖫𝔄)(𝔅,𝖫𝔅)\varphi:({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}})\rightarrow({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) between two quantum compact metric spaces (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) is a unital *-morphism from 𝔄{\mathfrak{A}} to 𝔅{\mathfrak{B}} such that:

φ(dom(𝖫𝔄))dom(𝖫𝔅).\varphi({\operatorname*{dom}\left({{\mathsf{L}}_{\mathfrak{A}}}\right)})\subseteq{\operatorname*{dom}\left({{\mathsf{L}}_{\mathfrak{B}}}\right)}\text{.}
Theorem 2.3 ([20]).

Let φ:𝔄𝔅\varphi:{\mathfrak{A}}\rightarrow{\mathfrak{B}} be a unital *-morphism between two unital C*-algebras 𝔄{\mathfrak{A}} and 𝔅{\mathfrak{B}}. If 𝖫𝔄{\mathsf{L}}_{\mathfrak{A}} and 𝖫𝔅{\mathsf{L}}_{\mathfrak{B}} are, respectively, Lip-norms on 𝔄{\mathfrak{A}} and 𝔅{\mathfrak{B}}, the the following assertions are equivalent:

  1. (1)

    φ\varphi is a Lipschitz morphism from (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) to (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}),

  2. (2)

    k0𝖫𝔅φk𝖫𝔄\exists k\geqslant 0\quad{\mathsf{L}}_{\mathfrak{B}}\circ\varphi\leqslant k{\mathsf{L}}_{\mathfrak{A}},

  3. (3)

    k0μ,ν𝒮(𝔅)𝗆𝗄𝖫𝔄(μφ,νφ)k𝗆𝗄𝖫𝔅(μ,ν)\exists k\geqslant 0\quad\forall\mu,\nu\in{\mathscr{S}}({\mathfrak{B}})\quad{\mathsf{mk}_{{\mathsf{L}}_{\mathfrak{A}}}}(\mu\circ\varphi,\nu\circ\varphi)\leqslant k{\mathsf{mk}_{{\mathsf{L}}_{\mathfrak{B}}}}(\mu,\nu).

If α:(𝔄,𝖫𝔄)(𝔅,𝖫𝔅)\alpha:({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}})\rightarrow({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) is a kk-Lipschitz morphism between two quantum compact metric spaces, then it naturally becomes a contractive morphism from (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) to (𝔅,1k𝖫𝔅)({\mathfrak{B}},\frac{1}{k}{\mathsf{L}}_{\mathfrak{B}}). Yet if α\alpha is actually injective and bi-Lipschitz, then it is usually not possible to adjust the quantum metrics to turn π\pi into a quantum isometry. However, under the hypothesis that we can find a conditional expectation 𝔼\mathds{E} of 𝔅{\mathfrak{B}} onto the range of α\alpha in 𝔅{\mathfrak{B}} such that 𝔼\mathds{E} is also a Lipschitz linear map, then it is indeed possible to modify 𝖫𝔅{\mathsf{L}}_{\mathfrak{B}} to turn α\alpha into a quantum isometry.

Lemma 2.4.

Let (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) be quantum compact metric spaces, with 𝖫𝔅{\mathsf{L}}_{\mathfrak{B}} an FF-Leibniz seminorm. If α:𝔄𝔅\alpha:{\mathfrak{A}}\rightarrow{\mathfrak{B}} is a unital *-monomorphism and 𝔼:𝔅α(𝔄)\mathds{E}:{\mathfrak{B}}\twoheadrightarrow\alpha({\mathfrak{A}}) is a conditional expectation such that for some mα,k𝔼>0m_{\alpha},k_{\mathds{E}}>0, the following conditions hold:

  • a𝔰𝔞(𝔄)mα𝖫𝔄(a)𝖫𝔅α(a)\forall a\in{\mathfrak{sa}\left({{\mathfrak{A}}}\right)}\quad m_{\alpha}{\mathsf{L}}_{\mathfrak{A}}(a)\leqslant{\mathsf{L}}_{\mathfrak{B}}\circ\alpha(a),

  • a𝔰𝔞(𝔅)𝖫𝔅(𝔼(a))k𝔼𝖫𝔅(a)\forall a\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}\quad{\mathsf{L}}_{\mathfrak{B}}(\mathds{E}(a))\leqslant k_{\mathds{E}}{\mathsf{L}}_{\mathfrak{B}}(a),

then if we set:

b𝔰𝔞(𝔅)𝖫𝔅(b)=max{𝖫𝔅(b),𝖫𝔄(α1𝔼(a))}\forall b\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}\quad{\mathsf{L}}^{\prime}_{\mathfrak{B}}(b)=\max\left\{{\mathsf{L}}_{\mathfrak{B}}(b),{\mathsf{L}}_{\mathfrak{A}}(\alpha^{-1}\circ\mathds{E}(a))\right\}

where α1\alpha^{-1} is the inverse of the *-isomorphism α:𝔄α(𝔄)\alpha:{\mathfrak{A}}\rightarrow\alpha({\mathfrak{A}}), then (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}^{\prime}_{\mathfrak{A}}) is a max{1,k𝔼mα}F\max\left\{1,\frac{k_{\mathds{E}}}{m_{\alpha}}\right\}F-Leibniz quantum compact metric space and:

aα(𝔄)𝖫𝔄(a)𝖫𝔅(α(a)),\forall a\in\alpha({\mathfrak{A}})\quad{\mathsf{L}}_{\mathfrak{A}}(a)\leqslant{\mathsf{L}}^{\prime}_{\mathfrak{B}}(\alpha(a))\text{,}

and dom(𝖫𝔅)=dom(𝖫𝔅){\operatorname*{dom}\left({{\mathsf{L}}^{\prime}_{\mathfrak{B}}}\right)}={\operatorname*{dom}\left({{\mathsf{L}}_{\mathfrak{B}}}\right)}.

If, moreover, there exists kα>0k_{\alpha}>0 such that 𝖫𝔅αkα𝖫𝔄{\mathsf{L}}_{\mathfrak{B}}\circ\alpha\leqslant k_{\alpha}{\mathsf{L}}_{\mathfrak{A}} then:

a𝔰𝔞(𝔄)𝖫𝔄(a)𝖫𝔅(α(a))max{1,kα}𝖫𝔄(a)\forall a\in{\mathfrak{sa}\left({{\mathfrak{A}}}\right)}\quad{\mathsf{L}}_{\mathfrak{A}}(a)\leqslant{\mathsf{L}}^{\prime}_{\mathfrak{B}}(\alpha(a))\leqslant\max\{1,k_{\alpha}\}{\mathsf{L}}_{\mathfrak{A}}(a)

In particular, if kα=1k_{\alpha}=1 then α:(𝔄,𝖫𝔄)(α(𝔄),𝖫𝔅)\alpha:({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}})\rightarrow(\alpha({\mathfrak{A}}),{\mathsf{L}}^{\prime}_{\mathfrak{B}}) is a full quantum isometry.

Proof.

Let bdom(𝖫𝔅)b\in{\operatorname*{dom}\left({{\mathsf{L}}_{\mathfrak{B}}}\right)}. By assumption, 𝖫𝔅(𝔼(b))k𝔼𝖫𝔅(b)<{\mathsf{L}}_{\mathfrak{B}}(\mathds{E}(b))\leqslant k_{\mathds{E}}{\mathsf{L}}_{\mathfrak{B}}(b)<\infty and therefore 𝖫𝔄(α1(𝔼(b)))1mα𝖫𝔅(𝔼(b))k𝔼mα𝖫𝔅(b)<{\mathsf{L}}_{\mathfrak{A}}(\alpha^{-1}(\mathds{E}(b)))\leqslant\frac{1}{m_{\alpha}}{\mathsf{L}}_{\mathfrak{B}}(\mathds{E}(b))\leqslant\frac{k_{\mathds{E}}}{m_{\alpha}}{\mathsf{L}}_{\mathfrak{B}}(b)<\infty. So bdom(𝖫𝔅)b\in{\operatorname*{dom}\left({{\mathsf{L}}^{\prime}_{\mathfrak{B}}}\right)}. Therefore dom(𝖫𝔅)dom(𝖫𝔅){\operatorname*{dom}\left({{\mathsf{L}}_{\mathfrak{B}}}\right)}\subseteq{\operatorname*{dom}\left({{\mathsf{L}}^{\prime}_{\mathfrak{B}}}\right)} and since dom(𝖫𝔅){\operatorname*{dom}\left({{\mathsf{L}}_{\mathfrak{B}}}\right)} is dense in 𝔰𝔞(𝔅){\mathfrak{sa}\left({{\mathfrak{B}}}\right)}, so is dom(𝖫𝔅){\operatorname*{dom}\left({{\mathsf{L}}^{\prime}_{\mathfrak{B}}}\right)}. Moreover, if 𝖫𝔅(a)={\mathsf{L}}_{\mathfrak{B}}(a)=\infty then 𝖫𝔅(a)={\mathsf{L}}^{\prime}_{\mathfrak{B}}(a)=\infty, so dom(𝖫𝔅)dom(𝖫𝔅){\operatorname*{dom}\left({{\mathsf{L}}^{\prime}_{\mathfrak{B}}}\right)}\subseteq{\operatorname*{dom}\left({{\mathsf{L}}_{\mathfrak{B}}}\right)}. Thus dom(𝖫𝔅)=dom(𝖫𝔅){\operatorname*{dom}\left({{\mathsf{L}}_{\mathfrak{B}}}\right)}={\operatorname*{dom}\left({{\mathsf{L}}^{\prime}_{\mathfrak{B}}}\right)}.

If 𝖫𝔅(b)=0{\mathsf{L}}^{\prime}_{\mathfrak{B}}(b)=0 then 𝖫𝔅(b)=0{\mathsf{L}}_{\mathfrak{B}}(b)=0 so b1𝔅b\in{\mathds{R}}1_{\mathfrak{B}}. Of course, 𝖫𝔅(1𝔅)=0{\mathsf{L}}^{\prime}_{\mathfrak{B}}(1_{\mathfrak{B}})=0 since 𝔼\mathds{E} and α\alpha are unital.

We now check the Leibniz property of 𝖫𝔅{\mathsf{L}}^{\prime}_{\mathfrak{B}}. For all a,b𝔰𝔞(𝔅)a,b\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}, we compute:

𝖫𝔄(α1(𝔼(ab)))\displaystyle{\mathsf{L}}_{\mathfrak{A}}(\alpha^{-1}(\mathds{E}(ab))) 1mα𝖫𝔅(𝔼(ab))\displaystyle\leqslant\frac{1}{m_{\alpha}}{\mathsf{L}}_{\mathfrak{B}}(\mathds{E}(ab))
k𝔼mα𝖫𝔅(ab)\displaystyle\leqslant\frac{k_{\mathds{E}}}{m_{\alpha}}{\mathsf{L}}_{\mathfrak{B}}(ab)
k𝔼mαF(𝖫𝔅(a),𝖫𝔅(b),a𝔅,b𝔅).\displaystyle\leqslant\frac{k_{\mathds{E}}}{m_{\alpha}}F\left({\mathsf{L}}_{\mathfrak{B}}(a),{\mathsf{L}}_{\mathfrak{B}}(b),\left\|{a}\right\|_{{\mathfrak{B}}},\left\|{b}\right\|_{{\mathfrak{B}}}\right)\text{.}

Therefore, 𝖫𝔅{\mathsf{L}}^{\prime}_{\mathfrak{B}} is max{1,k𝔼mα}F\max\left\{1,\frac{k_{\mathds{E}}}{m_{\alpha}}\right\}F-Leibniz.

As the supremum of two lower semi-continuous seminorms, 𝖫𝔅{\mathsf{L}}^{\prime}_{\mathfrak{B}} is a lower semi-continuous seminorm as well. Moreover, if μ𝒮(𝔅)\mu\in{\mathscr{S}}({\mathfrak{B}}), since:

{b𝔰𝔞(𝔅):𝖫𝔅(b)1,μ(b)=0}{b𝔰𝔞(𝔅):𝖫𝔅(b)1,μ(b)=0}\left\{b\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}:{\mathsf{L}}^{\prime}_{\mathfrak{B}}(b)\leqslant 1,\mu(b)=0\right\}\\ \subseteq\left\{b\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}:{\mathsf{L}}_{\mathfrak{B}}(b)\leqslant 1,\mu(b)=0\right\}

and since 𝖫𝔅{\mathsf{L}}_{\mathfrak{B}} is an Lip-norm, we conclude that the set on the right hand side, and therefore the set on the left hand side, is totally bounded.

Therefore 𝖫𝔅{\mathsf{L}}^{\prime}_{\mathfrak{B}} is a max{1,k𝔼mα}F\max\left\{1,\frac{k_{\mathds{E}}}{m_{\alpha}}\right\}F-quasi-Leibniz Lip-norm as a claimed.

By construction of 𝖫𝔅{\mathsf{L}}^{\prime}_{\mathfrak{B}}:

𝖫𝔄(a)=𝖫𝔄(α1(α(a)))=𝖫𝔄(α1(𝔼(α(a))))𝖫𝔅(α(a)).{\mathsf{L}}_{\mathfrak{A}}(a)={\mathsf{L}}_{\mathfrak{A}}(\alpha^{-1}(\alpha(a)))={\mathsf{L}}_{\mathfrak{A}}(\alpha^{-1}(\mathds{E}(\alpha(a))))\leqslant{\mathsf{L}}^{\prime}_{\mathfrak{B}}(\alpha(a))\text{.}

Last, assume that there exists kα>0k_{\alpha}>0 such that 𝖫𝔅αkα𝖫𝔄{\mathsf{L}}_{\mathfrak{B}}\circ\alpha\leqslant k_{\alpha}{\mathsf{L}}_{\mathfrak{A}}. We will note that, of course, if a𝔄a\in{\mathfrak{A}} then α1(𝔼(α(a)))=a\alpha^{-1}(\mathds{E}(\alpha(a)))=a, and thus:

𝖫𝔅(α(a))=max{𝖫𝔅(α(a)),𝖫𝔄(α1𝔼(α(a)))}max{kα,1}𝖫𝔄(a).{\mathsf{L}}^{\prime}_{\mathfrak{B}}(\alpha(a))=\max\{{\mathsf{L}}_{\mathfrak{B}}(\alpha(a)),{\mathsf{L}}_{\mathfrak{A}}(\alpha^{-1}\mathds{E}(\alpha(a)))\}\leqslant\max\{k_{\alpha},1\}{\mathsf{L}}_{\mathfrak{A}}(a)\text{.}

If kα=1k_{\alpha}=1 then 𝖫𝔄(a)=𝖫𝔅(α(a)){\mathsf{L}}_{\mathfrak{A}}(a)={\mathsf{L}}^{\prime}_{\mathfrak{B}}(\alpha(a)) for all aα(𝔄)a\in\alpha({\mathfrak{A}}). Thus α\alpha is a full quantum isometry form (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) onto (α(𝔄),𝖫𝔅)(\alpha({\mathfrak{A}}),{\mathsf{L}}^{\prime}_{\mathfrak{B}}). ∎

Remark 2.5.

If we drop the assumption that 𝖫𝔅{\mathsf{L}}_{\mathfrak{B}} is FF-Leibniz, then we still can conclude that 𝖫𝔅{\mathsf{L}}^{\prime}_{\mathfrak{B}} is a lower semi-continuous Lip-norm such that α1\alpha^{-1} is 11-Lipschitz.

We now bring our two previous observations in one theorem which will be key to our construction.

Theorem 2.6.

Let (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) be quantum compact metric spaces, with 𝖫𝔅{\mathsf{L}}_{\mathfrak{B}} a FF-Leibniz seminorm. If α:𝔄𝔅\alpha:{\mathfrak{A}}\rightarrow{\mathfrak{B}} is a unital *-monomorphism and 𝔼:𝔅α(𝔄)\mathds{E}:{\mathfrak{B}}\twoheadrightarrow\alpha({\mathfrak{A}}) is a conditional expectation such that for some mα>0m_{\alpha}>0:

  • a𝔰𝔞(𝔄)mα𝖫𝔄(a)𝖫𝔅α(a)𝖫(a)\forall a\in{\mathfrak{sa}\left({{\mathfrak{A}}}\right)}\quad m_{\alpha}{\mathsf{L}}_{\mathfrak{A}}(a)\leqslant{\mathsf{L}}_{\mathfrak{B}}\circ\alpha(a)\leqslant{\mathsf{L}}(a),

  • a𝔰𝔞(𝔅)𝖫𝔅(𝔼(a))𝖫𝔅(a)\forall a\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}\quad{\mathsf{L}}_{\mathfrak{B}}(\mathds{E}(a))\leqslant{\mathsf{L}}_{\mathfrak{B}}(a),

and if ε>0\varepsilon>0, then setting:

b𝔰𝔞(𝔅)𝖫𝔅ε(b)=max{𝖫𝔅(b),1εb𝔼(b)𝔅,𝖫𝔄(α1𝔼(a))}\forall b\in{\mathfrak{sa}\left({{\mathfrak{B}}}\right)}\quad{\mathsf{L}}^{\varepsilon}_{\mathfrak{B}}(b)=\max\left\{{\mathsf{L}}_{\mathfrak{B}}(b),\frac{1}{\varepsilon}\left\|{b-\mathds{E}(b)}\right\|_{{\mathfrak{B}}},{\mathsf{L}}_{\mathfrak{A}}(\alpha^{-1}\circ\mathds{E}(a))\right\}

then

ΛG((𝔄,𝖫𝔄),(𝔅,𝖫𝔅))ε,{\mathsf{\Lambda}^{\ast}_{G}}(({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}),({\mathfrak{B}},{\mathsf{L}}^{\prime}_{\mathfrak{B}}))\leqslant\varepsilon\text{,}

where G=max{1,1mα}FG=\max\{1,\frac{1}{m_{\alpha}}\}\cdot F.

Proof.

By Lemma (2.4), if we define 𝖫𝔅=max{𝖫𝔅,𝖫𝔄α1𝔼}{\mathsf{L}}^{\prime}_{\mathfrak{B}}=\max\{{\mathsf{L}}_{\mathfrak{B}},{\mathsf{L}}_{\mathfrak{A}}\circ\alpha^{-1}\circ\mathds{E}\} then α\alpha is a full quantum isometry from (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) to (α(𝔄),𝖫𝔅)(\alpha({\mathfrak{A}}),{\mathsf{L}}^{\prime}_{\mathfrak{B}}). So:

ΛG((𝔄,𝖫𝔄),(α(𝔄),𝖫𝔅))=0.{\mathsf{\Lambda}^{\ast}_{G}}(({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}),(\alpha({\mathfrak{A}}),{\mathsf{L}}^{\prime}_{\mathfrak{B}}))=0\text{.}

The seminorm 𝖫𝔅ε{\mathsf{L}}^{\varepsilon}_{\mathfrak{B}} is then obtained by applying Lemma (2.1) to 𝖫𝔅{\mathsf{L}}^{\prime}_{\mathfrak{B}}. In particular, 𝖫𝔅ε(b)=𝖫𝔅(b){\mathsf{L}}^{\varepsilon}_{\mathfrak{B}}(b)={\mathsf{L}}^{\prime}_{\mathfrak{B}}(b) for all b𝔰𝔞(α(𝔄))b\in{\mathfrak{sa}\left({\alpha({\mathfrak{A}})}\right)}, and:

ΛG((α(𝔄),𝖫𝔅),(𝔅,𝖫𝔅ε))ε.{\mathsf{\Lambda}^{\ast}_{G}}((\alpha({\mathfrak{A}}),{\mathsf{L}}^{\prime}_{\mathfrak{B}}),({\mathfrak{B}},{\mathsf{L}}^{\varepsilon}_{\mathfrak{B}}))\leqslant\varepsilon\text{.}

Thus, by the triangle inequality, we conclude:

ΛG((𝔄,𝖫𝔄),(𝔅,𝖫𝔅ε))ε{\mathsf{\Lambda}^{\ast}_{G}}(({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}),({\mathfrak{B}},{\mathsf{L}}^{\varepsilon}_{\mathfrak{B}}))\leqslant\varepsilon

as desired. ∎

We record that the tunnel constructed in Theorem (2.6) is given as:

(2.1) 𝖰(a,b)=max{𝖫𝔄(a),𝖫𝔅ε(b),1εbα(a)𝔅}\mathsf{Q}(a,b)=\max\left\{{\mathsf{L}}_{\mathfrak{A}}(a),{\mathsf{L}}^{\varepsilon}_{\mathfrak{B}}(b),\frac{1}{\varepsilon}\left\|{b-\alpha(a)}\right\|_{{\mathfrak{B}}}\right\}

for all (a,b)𝔰𝔞(𝔄𝔅).(a,b)\in{\mathfrak{sa}\left({{\mathfrak{A}}\oplus{\mathfrak{B}}}\right)}.

We have now worked out how, given a bi-Lipschitz morphism between two quantum compact metric spaces, it is possible to change the quantum metric on the codomain of this morphism to turn it into a full quantum isometry onto its range and to make its range and its codomain arbitrarily close in the quantum propinquity, at the cost of relaxing the Leibniz inequality.

We now apply this construction repeatedly to an inductive sequence of quantum compact metric spaces whose connecting maps are all bi-Lipschitz morphisms. The problem which arises is that unfortunately, the Leibniz condition of the seminorms we construct typically worsen at each stage. We will address this matter after we prove that we can indeed make a Cauchy sequence for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} out of any inductive sequence of quantum compact metric spaces where the connecting maps are bi-Lipschitz.

Theorem 2.7.

Let:

𝔄0α0𝔄1α1𝔄2α2𝔄3α3{\mathfrak{A}}_{0}\xrightarrow{\alpha_{0}}{\mathfrak{A}}_{1}\xrightarrow{\alpha_{1}}{\mathfrak{A}}_{2}\xrightarrow{\alpha_{2}}{\mathfrak{A}}_{3}\xrightarrow{\alpha_{3}}\ldots

be an inductive sequence of unital C*-algebras, where the connecting maps are unital *-monomorphisms, such that:

  • for each nn\in{\mathds{N}}, we are given an FnF_{n}-Leibniz Lip-norm 𝖫n{\mathsf{L}}_{n} on 𝔰𝔞(𝔄n){\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},

  • for each n{0}n\in{\mathds{N}}\setminus\{0\}, there exists a conditional expectation 𝔼n\mathds{E}_{n} from 𝔄n{\mathfrak{A}}_{n} onto αn1(𝔄n1)\alpha_{n-1}({\mathfrak{A}}_{n-1}) such that 𝖫n𝔼n𝖫n{\mathsf{L}}_{n}\circ\mathds{E}_{n}\leqslant{\mathsf{L}}_{n},

  • for each n{0}n\in{\mathds{N}}\setminus\{0\}, there exists cn,dn>0c_{n},d_{n}>0 such that:

    cn𝖫n1𝖫nαn1dn𝖫n1c_{n}{\mathsf{L}}_{n-1}\leqslant{\mathsf{L}}_{n}\circ\alpha_{n-1}\leqslant d_{n}{\mathsf{L}}_{n-1}

then, setting 𝖲0=𝖫0\mathsf{S}_{0}={\mathsf{L}}_{0} and for all n{0}n\in{\mathds{N}}\setminus\{0\}:

a𝔰𝔞(𝔄n)𝖲n(a)=max{ϰn𝖫n(a),𝖲n1αn11𝔼n(a),12na𝔼n(a)𝔄n}\forall a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}\quad\mathsf{S}_{n}(a)=\max\left\{\varkappa_{n}{\mathsf{L}}_{n}(a),\mathsf{S}_{n-1}\circ\alpha_{n-1}^{-1}\circ\mathds{E}_{n}(a),\frac{1}{2^{n}}\left\|{a-\mathds{E}_{n}(a)}\right\|_{{\mathfrak{A}}_{n}}\right\}

where

ϰn=j=0n1dj\varkappa_{n}=\prod_{j=0}^{n}\frac{1}{d_{j}}

then:

Λ((𝔄n,𝖲n),(𝔄n+1,𝖲n+1))12n+1.{\mathsf{\Lambda}^{\ast}}(({\mathfrak{A}}_{n},\mathsf{S}_{n}),({\mathfrak{A}}_{n+1},\mathsf{S}_{n+1}))\leqslant\frac{1}{2^{n+1}}\text{.}

Consequently, there exists a quantum metric order unit space (O,𝖫)(O,{\mathsf{L}}) such that:

limn𝖽𝗂𝗌𝗍q((𝔰𝔞(𝔄n),𝖲n),(O,𝖫))=0.\lim_{n\rightarrow\infty}{\mathsf{dist}}_{q}(({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},\mathsf{S}_{n}),(O,{\mathsf{L}}))=0\text{.}
Proof.

This proof is by induction. Since c0𝖫0𝖫1α0d0𝖫0c_{0}{\mathsf{L}}_{0}\leqslant{\mathsf{L}}_{1}\circ\alpha_{0}\leqslant d_{0}{\mathsf{L}}_{0}, the map α0\alpha_{0} is a contraction from (𝔄0,𝖫0)({\mathfrak{A}}_{0},{\mathsf{L}}_{0}) to (𝔄1,1d0𝖫1)({\mathfrak{A}}_{1},\frac{1}{d_{0}}{\mathsf{L}}_{1}), and c0d0𝖫0𝖫1α0\frac{c_{0}}{d_{0}}{\mathsf{L}}_{0}\leqslant{\mathsf{L}}_{1}\circ\alpha_{0}. Thus by Theorem (2.6), if we set:

a𝔰𝔞(𝔄1)𝖲1(a)=max{1d0𝖫1(a),𝖫0α01𝔼0(a),a𝔼0(a)𝔄1}\forall a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{1}}\right)}\quad\mathsf{S}_{1}(a)=\max\left\{\frac{1}{d_{0}}{\mathsf{L}}_{1}(a),{\mathsf{L}}_{0}\circ\alpha_{0}^{-1}\circ\mathds{E}_{0}(a),\left\|{a-\mathds{E}_{0}(a)}\right\|_{{\mathfrak{A}}_{1}}\right\}

then (𝔄1,𝖲1)({\mathfrak{A}}_{1},\mathsf{S}_{1}) is an FF-Leibniz quantum compact metric space for some FF, and Λ((𝔄0,𝖫0),(𝔄1,𝖲1))1{\mathsf{\Lambda}^{\ast}}(({\mathfrak{A}}_{0},{\mathsf{L}}_{0}),({\mathfrak{A}}_{1},\mathsf{S}_{1}))\leqslant 1.

Assume that, for some nn\in{\mathds{N}}, we have now shown that Λ((𝔄k,𝕊k),(𝔄k+1,𝖲k+1))12k{\mathsf{\Lambda}^{\ast}}(({\mathfrak{A}}_{k},\mathds{S}_{k}),({\mathfrak{A}}_{k+1},\mathsf{S}_{k+1}))\leqslant\frac{1}{2^{k}} and dom(𝖲k)=dom(𝖫k){\operatorname*{dom}\left({\mathsf{S}_{k}}\right)}={\operatorname*{dom}\left({{\mathsf{L}}_{k}}\right)} for all k{1,,n}k\in\{1,\ldots,n\}.

By assumption

𝖫n+1αndn+1𝖫ndn+1ϰn𝖲n=1ϰn+1𝖲n.{\mathsf{L}}_{n+1}\circ\alpha_{n}\leqslant d_{n+1}{\mathsf{L}}_{n}\leqslant\frac{d_{n+1}}{\varkappa_{n}}\mathsf{S}_{n}=\frac{1}{\varkappa_{n+1}}\mathsf{S}_{n}\text{.}

Let a𝔰𝔞(αn(𝔄n))a\in{\mathfrak{sa}\left({\alpha_{n}({\mathfrak{A}}_{n})}\right)} and ε>0\varepsilon>0. There exists aεdom(𝖫n+1)a_{\varepsilon}\in{\operatorname*{dom}\left({{\mathsf{L}}_{n+1}}\right)} such that aaε𝔄n+1<ε\left\|{a-a_{\varepsilon}}\right\|_{{\mathfrak{A}}_{n+1}}<\varepsilon. By assumption, 𝖫n+1𝔼n+1(aε)𝖫n+1(aε)<{\mathsf{L}}_{n+1}\circ\mathds{E}_{n+1}(a_{\varepsilon})\leqslant{\mathsf{L}}_{n+1}(a_{\varepsilon})<\infty, Moreover, 𝔼n+1(aε)αn(𝔄n+1)\mathds{E}_{n+1}(a_{\varepsilon})\in\alpha_{n}({\mathfrak{A}}_{n+1}) and

𝔼n+1(aε)a𝔄n+1\displaystyle\left\|{\mathds{E}_{n+1}(a_{\varepsilon})-a}\right\|_{{\mathfrak{A}}_{n+1}} =𝔼n+1(aε)𝔼n+1(a)𝔄n+1\displaystyle=\left\|{\mathds{E}_{n+1}(a_{\varepsilon})-\mathds{E}_{n+1}(a)}\right\|_{{\mathfrak{A}}_{n+1}}
aεa𝔄n+1<ε.\displaystyle\leqslant\left\|{a_{\varepsilon}-a}\right\|_{{\mathfrak{A}}_{n+1}}<\varepsilon\text{.}

Thus dom(𝖫n+1)𝔰𝔞(αn(𝔄n)){\operatorname*{dom}\left({{\mathsf{L}}_{n+1}}\right)}\cap{\mathfrak{sa}\left({\alpha_{n}({\mathfrak{A}}_{n})}\right)} is dense in 𝔰𝔞(αn(𝔄n)){\mathfrak{sa}\left({\alpha_{n}({\mathfrak{A}}_{n})}\right)}. Again by our assumption, if a𝔰𝔞(αn(𝔄n))dom(𝖫n+1)a\in{\mathfrak{sa}\left({\alpha_{n}({\mathfrak{A}}_{n})}\right)}\cap{\operatorname*{dom}\left({{\mathsf{L}}_{n+1}}\right)}, then 𝖫n(αn1(a))1cn𝖫n(a)<{\mathsf{L}}_{n}(\alpha_{n}^{-1}(a))\leqslant\frac{1}{c_{n}}{\mathsf{L}}_{n}(a)<\infty. Thus by our induction hypothesis, we conclude αn1(a)dom(Sn)\alpha_{n}^{-1}(a)\in{\operatorname*{dom}\left({S_{n}}\right)}. Now αn1\alpha_{n}^{-1} is a unital *-morphism from (αn(𝔄n),𝖫n+1)(\alpha_{n}({\mathfrak{A}}_{n}),{\mathsf{L}}_{n+1}) onto (𝔄n,𝖲n)({\mathfrak{A}}_{n},\mathsf{S}_{n}) which maps dom(𝖫n+1){\operatorname*{dom}\left({{\mathsf{L}}_{n+1}}\right)} to dom(𝖲n){\operatorname*{dom}\left({\mathsf{S}_{n}}\right)}. By Theorem (2.3), we conclude that there exists m>0m>0 such that m𝖲n𝖫n+1αnm\mathsf{S}_{n}\leqslant{\mathsf{L}}_{n+1}\circ\alpha_{n}. We will compute an estimate for mm in the next lemma but its actual value is not very important for us.

Thus, by Theorem (2.6), if we set:

a𝔰𝔞(𝔄n+1)𝖲n+1(a)=max{ϰn+1𝖫n+1(a),𝖲nαn1𝔼n(a),12n+1a𝔼n+1(a)𝔄n+1}\forall a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n+1}}\right)}\quad\mathsf{S}_{n+1}(a)=\\ \max\left\{\varkappa_{n+1}{\mathsf{L}}_{n+1}(a),\mathsf{S}_{n}\circ\alpha_{n}^{-1}\circ\mathds{E}_{n}(a),\frac{1}{2^{n+1}}\left\|{a-\mathds{E}_{n+1}(a)}\right\|_{{\mathfrak{A}}_{n+1}}\right\}

then 𝖲n+1\mathsf{S}_{n+1} is an FF-Leibniz Lip-norm for some FF and:

Λ((𝔄n+1,𝖲n+1),(𝔄n,𝖲n))12n+1,{\mathsf{\Lambda}^{\ast}}(({\mathfrak{A}}_{n+1},\mathsf{S}_{n+1}),({\mathfrak{A}}_{n},\mathsf{S}_{n}))\leqslant\frac{1}{2^{n+1}}\text{,}

as claimed. Moreover dom(𝖲n+1)=dom(𝖫n+1){\operatorname*{dom}\left({\mathsf{S}_{n+1}}\right)}={\operatorname*{dom}\left({{\mathsf{L}}_{n+1}}\right)}. Our induction hypothesis holds for all nn\in{\mathds{N}}.

The conclusion of the theorem then follows from the observation that 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} is complete and is dominated by Λ{\mathsf{\Lambda}^{\ast}} by [16, Theorem 5.5]. ∎

We note that the above theorem highlights the issue we have with the lack of a uniform Leibniz rule for all nn\in{\mathds{N}}. Indeed, we see that we can calculate distances between any term in our inductive sequence using propinquity since propinquity is a distance on the class of all quantum compact metric spaces equipped with any FF-Leibniz property as discussed after Definition 1.7. However, we are unable to find a uniform bound on the FF-Leibniz properties over all nn\in{\mathds{N}} (even in the explicit setting of the Bunce-Deddens algebras) and propinquity is only complete over classes of quantum compact metric spaces with uniform FF-Leibniz property. But, this is where distq\mathrm{dist}_{q} has an advantage. It is complete on the class of all quantum metric order unit spaces. Although distq\mathrm{dist}_{q} doesn’t have the optimal coincidence property, it still has this one crucial advantage, and this is why the above theorem ends with distq\mathrm{dist}_{q} rather than Λ{\mathsf{\Lambda}^{\ast}}.

Of course, we want to relate the limit in Theorem (2.7) with the inductive limit of the given sequence. This is achieved by proving two observations:

  • the limit of the sequence is described in the proof of the completeness of 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q},

  • the tunnels used in the proof of Theorem (2.6), which are given by Expression (2.1), are actually at once related to the inductive limit and to the metric limit computations.

Before we move in this direction, we however begin with our core example for Theorem(2.7): the Bunce-Deddens algebras.

3. The Bunce-Deddens C*-algebras

We denote the C*-algebra of n×nn\times n matrices over {\mathds{C}} by 𝔐n(){\mathfrak{M}}_{n}({\mathds{C}}).

Definition 3.1 ([23]).

Let 𝒩=({0,1}){0}{\mathscr{N}}=({\mathds{N}}\setminus\{0,1\})^{{\mathds{N}}\setminus\{0\}}. For each x,y𝒩x,y\in{\mathscr{N}} set

𝖽𝒩(x,y)={0 if x=y,2min{m:x(m)y(m)} otherwise.\mathsf{d}_{{\mathscr{N}}}(x,y)=\begin{cases}0\text{ if $x=y$,}\\ 2^{-\min\{m\in{\mathds{N}}:x(m)\neq y(m)\}}\text{ otherwise.}\end{cases}

The metric space (𝒩,𝖽𝒩)({\mathscr{N}},\mathsf{d}_{{\mathscr{N}}}) is called the Baire space.

Notation 3.2.

If σ𝒩\sigma\in{\mathscr{N}} then, for m{0}m\in{\mathds{N}}\setminus\{0\}, we set:

σm=j=1mσm;\boxtimes\sigma_{m}=\prod_{j=1}^{m}\sigma_{m}\text{;}

and we set σ0=1\boxtimes\sigma_{0}=1, and we denote σ=(σm)m\boxtimes\sigma=(\boxtimes\sigma_{m})_{m\in{\mathds{N}}}.

Notation 3.3.

For each tt\in{\mathds{R}}, let z(t)=exp(2πit)z(t)=\exp(2\pi it). Let m{0}m\in{\mathds{N}}\setminus\{0\}.

  • For all j,k{1,,m}j,k\in\{1,\ldots,m\} and tt\in{\mathds{R}}, we define

    zj,km(t)=1mz((mj)(t+k1)/m).z^{m}_{j,k}(t)=\frac{1}{\sqrt{m}}z((m-j)(t+k-1)/m)\text{.}

    The function zj,kmz^{m}_{j,k} is a continuous, mm-periodic {\mathds{C}}-valued function over {\mathds{R}}.

  • We define the 𝔐m(){\mathfrak{M}}_{m}({\mathds{C}})-valued continuous function:

    Um=(zj,km)j,k{1,,m}=(z1,1mz1,mmzm,1mzm,mm).U_{m}=(z^{m}_{j,k})_{j,k\in\{1,\ldots,m\}}=\begin{pmatrix}z^{m}_{1,1}&\cdots&z^{m}_{1,m}\\ \vdots&&\vdots\\ z^{m}_{m,1}&\cdots&z^{m}_{m,m}\end{pmatrix}\text{.}

In addition, we set U0=1U_{0}=1. The function UmU_{m} is a continuous mm-periodic 𝔐m(){\mathfrak{M}}_{m}({\mathds{C}})-valued function. Moreover, by [8, Chapter V.3], the map UmU_{m} is unitary.

Lemma 3.4.

For all m{0}m\in{\mathds{N}}\setminus\{0\}, the map UmU_{m} is a unitary and is mm-periodic and if Vm=(00011000010000010)𝔐m()V_{m}=\begin{pmatrix}0&0&\cdots&0&1\\ 1&0&\cdots&0&0\\ 0&1&\cdots&0&0\\ \vdots&\vdots&\ddots&\vdots&0\\ 0&0&\cdots&1&0\end{pmatrix}\in{\mathfrak{M}}_{m}({\mathds{C}}):

tUm(t+1)=Um(t)Vm.\forall t\in{\mathds{R}}\quad U_{m}(t+1)=U_{m}(t)V_{m}\text{.}
Proof.

This is an immediate computation. ∎

Notation 3.5.

Let σ=(σm)m𝒩\sigma=(\sigma_{m})_{m\in{\mathds{N}}}\in{\mathscr{N}}. For m{0}m\in{\mathds{N}}\setminus\{0\}, the C*-algebra of 𝔐σm(){\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}})-valued, continuous, 11-periodic functions over {\mathds{R}} is denoted by 𝔓(σ,m){{\mathfrak{CP}}\left({\sigma},{m}\right)}.

We then define:

Uσ,m=Uσmidσm1=(z1,1σmidσm1z1,σmσmidσm1zσm,1σmidσm1zσm,σmσmidσm1)U_{\sigma,m}=U_{\sigma_{m}}\otimes\mathrm{id}_{\boxtimes\sigma_{m-1}}=\begin{pmatrix}z_{1,1}^{\sigma_{m}}\cdot\mathrm{id}_{\boxtimes\sigma_{m-1}}&&z_{1,\sigma_{m}}^{\sigma_{m}}\cdot\mathrm{id}_{\boxtimes\sigma_{m-1}}\\ \vdots&&\vdots\\ z_{\sigma_{m},1}^{\sigma_{m}}\cdot\mathrm{id}_{\boxtimes\sigma_{m-1}}&&z_{\sigma_{m},\sigma_{m}}^{\sigma_{m}}\cdot\mathrm{id}_{\boxtimes\sigma_{m-1}}\end{pmatrix}

and note that Uσ,mU_{\sigma,m} is a 𝔐σm(){\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}})-valued continuous function over {\mathds{R}}.

Lemma 3.6.

The map Uσ,mU_{\sigma,m} is a unitary, σm\sigma_{m}-periodic function such that if Wσ,m=Vσmidσm1=(000idσm1idσm10000idσm100000idσm10),W_{\sigma,m}=V_{\sigma_{m}}\otimes\mathrm{id}_{\boxtimes\sigma_{m-1}}=\begin{pmatrix}0&0&\cdots&0&\mathrm{id}_{\boxtimes\sigma_{m-1}}\\ \mathrm{id}_{\boxtimes\sigma_{m-1}}&0&\cdots&0&0\\ 0&\mathrm{id}_{\boxtimes\sigma_{m-1}}&\cdots&0&0\\ \vdots&\vdots&\ddots&\vdots&0\\ 0&0&\cdots&\mathrm{id}_{\boxtimes\sigma_{m-1}}&0\end{pmatrix}, then:

tUσ,m(t+1)=Uσ,m(t)Wσ,m.\forall t\in{\mathds{R}}\quad U_{\sigma,m}(t+1)=U_{\sigma,m}(t)W_{\sigma,m}\text{.}
Proof.

This is an immediate computation. ∎

Notation 3.7.

Let σ𝒩\sigma\in{\mathscr{N}}. For m,a𝔓(σ,m),tm\in{\mathds{N}},a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)},t\in{\mathds{R}}, we define:

ασ,m(a)(t)=Uσ,m+1(t)(a(tσm+1)a(t+1σm+1)a(t+σm+11σm+1))Uσ,m+1(t),\begin{split}&\alpha_{\sigma,m}(a)(t)\\ &=U_{\sigma,m+1}(t)\begin{pmatrix}a\left(\frac{t}{\sigma_{m+1}}\right)&&&\\ &a\left(\frac{t+1}{\sigma_{m+1}}\right)&&\\ &&\ddots&\\ &&&a\left(\frac{t+\sigma_{m+1}-1}{\sigma_{m+1}}\right)\end{pmatrix}U_{\sigma,m+1}^{\ast}(t),\end{split}

and note ασ,m(a)Cb(,𝔐σm+1()).\alpha_{\sigma,m}(a)\in C_{b}({\mathds{R}},{\mathfrak{M}}_{\boxtimes\sigma_{m+1}}({\mathds{C}}))\text{.}

The following lemma is presented in [8, Section V.3], but we provide more details for the proof here since our notation differs some from this reference.

Lemma 3.8.

Let σ𝒩\sigma\in{\mathscr{N}} and m{0}m\in{\mathds{N}}\setminus\{0\}. If a𝔓(σ,m1)a\in{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}, then ασ,m1(a)𝔓(σ,m)\alpha_{\sigma,m-1}(a)\in{{\mathfrak{CP}}\left({\sigma},{m}\right)}.

The map ασ,m1\alpha_{\sigma,m-1} thus defined is a unital *-monomorphism from 𝔓(σ,m1){{\mathfrak{CP}}\left({\sigma},{m-1}\right)} to 𝔓(σ,m){{\mathfrak{CP}}\left({\sigma},{m}\right)}.

Proof.

We use the notations of Lemma (3.6). If a=(aj,k)1j,kσm𝔐σm()a=(a_{j,k})_{1\leqslant j,k\leqslant\sigma_{m}}\in{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}}), with a1,1a_{1,1}, …, aσm,σma_{\sigma_{m},\sigma_{m}} elements of 𝔐σm1(){\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}}), then Wσ,maWσ,mW_{\sigma,m}aW_{\sigma,m}^{\ast} is the matrix (aπ(j),π(k))1j,kσm(a_{\pi(j),\pi(k)})_{1\leqslant j,k\leqslant\sigma_{m}} where π\pi is the permutation (σm 1 2σm1)(\sigma_{m}\ 1\ 2\ \ldots\ \sigma_{m}-1).

By Lemma (3.6), we also have:

tUσ,m(t+1)=Uσ,m(t)Wσ,m.\forall t\in{\mathds{R}}\quad U_{\sigma,m}(t+1)=U_{\sigma,m}(t)W_{\sigma,m}\text{.}

Let a𝔓(σ,m1)a\in{{\mathfrak{CP}}\left({\sigma},{m-1}\right)} and tt\in{\mathds{R}}. Since aa is 11-periodic, we note that

a(t+σm1+1σm)=a(t+σmσm)=a(tσm+1)=a(tσm).a\left(\frac{t+\sigma_{m}-1+1}{\sigma_{m}}\right)=a\left(\frac{t+\sigma_{m}}{\sigma_{m}}\right)=a\left(\frac{t}{\sigma_{m}}+1\right)=a\left(\frac{t}{\sigma_{m}}\right).

We then compute:

ασ,m1(a)(t+1)\displaystyle\alpha_{\sigma,m-1}(a)(t+1)
=Uσ,m(t+1)(a(t+1σm)a(t+2σm)a(t+σm1σm)a(t+σm1+1σm))\displaystyle=U_{\sigma,m}(t+1)\begin{pmatrix}a\left(\frac{t+1}{\sigma_{m}}\right)&&&&\\ &a\left(\frac{t+2}{\sigma_{m}}\right)&&&\\ &&\ddots&&\\ &&&a\left(\frac{t+\sigma_{m}-1}{\sigma_{m}}\right)&\\ &&&&a\left(\frac{t+\sigma_{m}-1+1}{\sigma_{m}}\right)\end{pmatrix}
Uσ,m(t+1)\displaystyle\quad\quad\cdot U_{\sigma,m}^{\ast}(t+1)
=Uσ,m(t+1)(a(t+1σm)a(t+2σm)a(t+σm1σm)a(tσm))\displaystyle=U_{\sigma,m}(t+1)\begin{pmatrix}a\left(\frac{t+1}{\sigma_{m}}\right)&&&&\\ &a\left(\frac{t+2}{\sigma_{m}}\right)&&&\\ &&\ddots&&\\ &&&a\left(\frac{t+\sigma_{m}-1}{\sigma_{m}}\right)&\\ &&&&a\left(\frac{t}{\sigma_{m}}\right)\end{pmatrix}
Uσ,m(t+1)\displaystyle\quad\quad\cdot U_{\sigma,m}^{\ast}(t+1)
=Uσ,m(t)Wσ,m(a(t+1σm)a(t+2σm)a(t+σm1σm)a(tσm))\displaystyle=U_{\sigma,m}(t)W_{\sigma,m}\begin{pmatrix}a\left(\frac{t+1}{\sigma_{m}}\right)&&&&\\ &a\left(\frac{t+2}{\sigma_{m}}\right)&&&\\ &&\ddots&&\\ &&&a\left(\frac{t+\sigma_{m}-1}{\sigma_{m}}\right)&\\ &&&&a\left(\frac{t}{\sigma_{m}}\right)\end{pmatrix}
Wσ,mUσ,m(t)\displaystyle\quad\quad\cdot W_{\sigma,m}^{\ast}U_{\sigma,m}^{\ast}(t)
=Uσ,m(t)(a(tσm)a(t+1σm)a(t+σm1σm))Uσ,m(t)\displaystyle=U_{\sigma,m}(t)\begin{pmatrix}a\left(\frac{t}{\sigma_{m}}\right)&&&\\ &a\left(\frac{t+1}{\sigma_{m}}\right)&&\\ &&\ddots&\\ &&&a\left(\frac{t+\sigma_{m}-1}{\sigma_{m}}\right)\end{pmatrix}U^{\ast}_{\sigma,m}(t)
=ασ,m1(a)(t).\displaystyle=\alpha_{\sigma,m-1}(a)(t)\text{.}

Thus ασ,m1(a)\alpha_{\sigma,m-1}(a) is 11-periodic. It is of course a continuous function over {\mathds{R}} valued in 𝔐σm(){\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}}), so ασ,m1(a)𝔓(σ,m)\alpha_{\sigma,m-1}(a)\in{{\mathfrak{CP}}\left({\sigma},{m}\right)}. Since Uσ,mU_{\sigma,m} is a unitary, it is immediate that ασ,m1\alpha_{\sigma,m-1} is a unital *-monomorphism. ∎

Definition 3.9 ([8, Section V.3]).

The Bunce-Deddens algebra 𝔅𝔇(σ){{\mathfrak{BD}}\left({\sigma}\right)} is the C*-algebra inductive limit [24, Section 6.1] of the sequence:

𝔅𝔇(σ)=lim(𝔓(σ,m),ασ,m)m.{{\mathfrak{BD}}\left({\sigma}\right)}=\underrightarrow{\lim}\ \left({{\mathfrak{CP}}\left({\sigma},{m}\right)},\alpha_{\sigma,m}\right)_{m\in{\mathds{N}}}\text{.}
Remark 3.10.

The Bunce-Deddens algebra 𝔅𝔇(σ){{\mathfrak{BD}}\left({\sigma}\right)} is denoted by 𝔅(σ){\mathfrak{B}}(\boxtimes\sigma) in [8, Section V.3].

Notation 3.11.

Let σ𝒩\sigma\in{\mathscr{N}}. For each mm\in{\mathds{N}}, we let

ασ(m):𝔓(σ,m)𝔅𝔇(σ)\alpha^{(m)}_{\sigma}:{{\mathfrak{CP}}\left({\sigma},{m}\right)}\longrightarrow{{\mathfrak{BD}}\left({\sigma}\right)}

denote the canonical unital *-monomorphism such that ασ(m+1)ασ,m=ασ(m)\alpha^{(m+1)}_{\sigma}\circ\alpha_{\sigma,m}=\alpha^{(m)}_{\sigma} given by [24, Section 6.1].

The Bunce-Deddens algebras have a unique faithful tracial state.

Notation 3.12.

Let σ𝒩\sigma\in{\mathscr{N}}. We denote the unique faithful tracial state on 𝔅𝔇(σ){{\mathfrak{BD}}\left({\sigma}\right)} by τσ\tau_{\sigma} [8, Theorem V.3.6]. For each m{0}m\in{\mathds{N}}\setminus\{0\}, we denote

τσ,m=τσασ(m):𝔓(σ,m),\tau_{\sigma,m}=\tau_{\sigma}\circ\alpha^{(m)}_{\sigma}:{{\mathfrak{CP}}\left({\sigma},{m}\right)}\longrightarrow{\mathds{C}}\text{,}

which is a faithful tracial state on 𝔓(σ,m){{\mathfrak{CP}}\left({\sigma},{m}\right)}. Also, note that:

τσ,m+1ασ,m=τσ,m\tau_{\sigma,m+1}\circ\alpha_{\sigma,m}=\tau_{\sigma,m}

by [8, Theorem V.3.6] and its proof.

At the core of our construction of a quantum metric on the Bunce-Deddens algebras 𝔅𝔇(σ){{\mathfrak{BD}}\left({\sigma}\right)} lies a conditional expectation from a circle algebra to the image by a connecting morphism of a previous circle algebra in the inductive sequence defining 𝔅𝔇(σ){{\mathfrak{BD}}\left({\sigma}\right)}. We now construct this conditional expectation.

Notation 3.13.

If M𝔐σm()M\in{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}}), and if we write M=(N1,1N1,σmNσm,1Nσm,σm)M=\begin{pmatrix}N_{1,1}&\cdots&N_{1,\sigma_{m}}\\ \vdots&&\vdots\\ N_{\sigma_{m},1}&\cdots&N_{\sigma_{m},\sigma_{m}}\end{pmatrix} with N1,1N_{1,1}, …, Nσm,σmN_{\sigma_{m},\sigma_{m}} all in 𝔐σm1(){\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}}), then we define:

Dσ,m(M)=(N1,1N2,2Nσm,σm).D_{\sigma,m}(M)=\begin{pmatrix}N_{1,1}&&&\\ &N_{2,2}&&\\ &&\ddots&\\ &&&N_{\sigma_{m},\sigma_{m}}\end{pmatrix}\text{.}

Modeled on [3], the map Dσ,mD_{\sigma,m} is a conditional expectation from 𝔐σ(m)(){\mathfrak{M}}_{\boxtimes\sigma(m)}({\mathds{C}}) to the C*-subalgebra of block-diagonal matrices with blocks all square matrices of order σm1\boxtimes\sigma_{m-1}.

Lemma 3.14.

Let σ𝒩\sigma\in{\mathscr{N}} and m{0}m\in{\mathds{N}}\setminus\{0\}. If for all a𝔓(σ,m)a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)}, we define:

𝔼σ,m(a):tUσ,m(t)[Dσ,m(Uσ,m(t)a(t)Uσ,m(t))]Uσ,m(t),{\mathds{E}_{\sigma,m}\left({a}\right)}:t\in{\mathds{R}}\mapsto U_{\sigma,m}(t)\left[D_{\sigma,m}\left(U_{\sigma,m}(t)^{\ast}a(t)U_{\sigma,m}(t)\right)\right]U_{\sigma,m}(t)^{\ast}\text{,}

then 𝔼σ,m:a𝔓(σ,m)𝔼σ,m(a)ασ,m1(𝔓(σ,m1))\mathds{E}_{\sigma,m}:a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)}\mapsto{\mathds{E}_{\sigma,m}\left({a}\right)}\in\alpha_{\sigma,m-1}({{\mathfrak{CP}}\left({\sigma},{m-1}\right)}) is a conditional expectation onto ασ,m1(𝔓(σ,m1))\alpha_{\sigma,m-1}({{\mathfrak{CP}}\left({\sigma},{m-1}\right)}) such that τσ,m𝔼σ,m=τσ,m\tau_{\sigma,m}\circ\mathds{E}_{\sigma,m}=\tau_{\sigma,m}.

Proof.

We use the notations of Lemma (3.6) and use the computations of Lemma (3.8).

If a=(aj,k)1j,kσm𝔐σm()a=(a_{j,k})_{1\leqslant j,k\leqslant\sigma_{m}}\in{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}}), with a1,1a_{1,1}, …, aσm,σma_{\sigma_{m},\sigma_{m}} elements of 𝔐σm1(){\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}}), then Wσ,maWσ,mW_{\sigma,m}^{\ast}aW_{\sigma,m} is the matrix (aπ(j),π(k))1j,kσm(a_{\pi(j),\pi(k)})_{1\leqslant j,k\leqslant\sigma_{m}} where π\pi is the permutation (2 3m 1)(2\,3\,\ldots\,m\,1). Thus in particular,

Dσ,m(Wσ,maWσ,m)=(a2,2a3,3aσm,σma1,1)=Wσ,mDσ,m(a)Wσ,m.D_{\sigma,m}(W_{\sigma,m}^{\ast}aW_{\sigma,m})=\begin{pmatrix}a_{2,2}&&&&\\ &a_{3,3}&&&\\ &&\ddots&&\\ &&&a_{\sigma_{m},\sigma_{m}}&\\ &&&&a_{1,1}\end{pmatrix}=W_{\sigma,m}^{\ast}D_{\sigma,m}(a)W_{\sigma,m}.

Hence Dσ,mD_{\sigma,m} commutes with AdWσ,m\mathrm{Ad}_{W_{\sigma,m}} (and similarly with AdWσ,m\mathrm{Ad}_{W_{\sigma,m}^{\ast}}). Thus

Wσ,mDσ,m(Wσ,maWσ,m)Wσ,m=(a1,1a2,2aσm,σm)=Dσ,m(a).W_{\sigma,m}D_{\sigma,m}(W_{\sigma,m}^{\ast}aW_{\sigma,m})W_{\sigma,m}^{\ast}=\begin{pmatrix}a_{1,1}&&&\\ &a_{2,2}\\ &&\ddots&\\ &&&a_{\sigma_{m},\sigma_{m}}\end{pmatrix}=D_{\sigma,m}(a)\text{.}

By Lemma (3.6), we also have:

tUσ,m(t+1)=Uσ,m(t)Wσ,m.\forall t\in{\mathds{R}}\quad U_{\sigma,m}(t+1)=U_{\sigma,m}(t)W_{\sigma,m}\text{.}

Let now a𝔓(σ,m)a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)} — note that aa is 11-periodic. We then compute for any tt\in{\mathds{R}}:

Uσ,m(t+1)[Dσ,m(Uσ,m(t+1)a(t+1)Uσ,m(t+1))]Uσ,m(t+1)=Uσ,m(t)Wσ,m[Dσ,m(Wσ,mUσ,m(t)a(t)Uσ,m(t)Wσ,m)]Wσ,mUσ,m(t)=Uσ,m(t)[Dσ,m(Uσ,m(t)a(t)Uσ,m(t))]Uσ,m(t).U_{\sigma,m}(t+1)\left[D_{\sigma,m}(U_{\sigma,m}^{\ast}(t+1)\cdot a(t+1)\cdot U_{\sigma,m}(t+1)^{\ast})\right]U_{\sigma,m}^{\ast}(t+1)\\ \begin{aligned} &=U_{\sigma,m}(t)W_{\sigma,m}\left[D_{\sigma,m}(W_{\sigma,m}^{\ast}U_{\sigma,m}^{\ast}(t)\cdot a(t)\cdot U_{\sigma,m}(t)W_{\sigma,m})\right]W_{\sigma,m}^{\ast}U_{\sigma,m}^{\ast}(t)\\ &=U_{\sigma,m}(t)\left[D_{\sigma,m}(U_{\sigma,m}^{\ast}(t)\cdot a(t)\cdot U_{\sigma,m}(t))\right]U_{\sigma,m}^{\ast}(t)\text{.}\end{aligned}

Therefore, 𝔼σ,m(a)\mathds{E}_{\sigma,m}(a) is 11-periodic, and obviously continuous, so it is an element of 𝔓(σ,m){{\mathfrak{CP}}\left({\sigma},{m}\right)}.

Now, we wish to find f𝔓(σ,m1)f\in{{\mathfrak{CP}}\left({\sigma},{m-1}\right)} such that ασ,m1(f)=𝔼σ,m(a)\alpha_{\sigma,m-1}(f)=\mathds{E}_{\sigma,m}(a). In particular, we wish to find f𝔓(σ,m1)f\in{{\mathfrak{CP}}\left({\sigma},{m-1}\right)} such that:

t(f(tσm)f(t+1σm)f(t+σm1σm))=Dσ,m(Uσ,m(t)a(t)Uσ,m(t)).\forall t\in{\mathds{R}}\quad\begin{pmatrix}f\left(\frac{t}{\sigma_{m}}\right)&&&\\ &f\left(\frac{t+1}{\sigma_{m}}\right)&&\\ &&\ddots&\\ &&&f\left(\frac{t+\sigma_{m}-1}{\sigma_{m}}\right)\end{pmatrix}=D_{\sigma,m}(U_{\sigma,m}^{\ast}(t)\cdot a(t)\cdot U_{\sigma,m}(t))\text{.}

To this end, if M𝔐σm()M\in{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}}), and if we write M=(N1,1N1,σmNσm,1Nσm,σm)M=\begin{pmatrix}N_{1,1}&\cdots&N_{1,\sigma_{m}}\\ \vdots&&\vdots\\ N_{\sigma_{m},1}&\cdots&N_{\sigma_{m},\sigma_{m}}\end{pmatrix} with N1,1N_{1,1}, …, Nσm,σmN_{\sigma_{m},\sigma_{m}} all in 𝔐σm1(){\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}}), then for j{1,,σm}j\in\{1,\ldots,\sigma_{m}\}, we define:

Fσ,m,j(M)=Nj,j,F_{\sigma,m,j}(M)=N_{j,j},

Modeled on [3], the map Fσ,m,jF_{\sigma,m,j} is a unital completely positive contraction from 𝔐σ(m)(){\mathfrak{M}}_{\boxtimes\sigma(m)}({\mathds{C}}) onto 𝔐σ(m1)(){\mathfrak{M}}_{\boxtimes\sigma(m-1)}({\mathds{C}}). Also, a similar calculation to the one involving Dσ,mD_{\sigma,m} shows that Fσ,m,j(Wσ,mMWσ,m)=Fσ,m,j+1(M)F_{\sigma,m,j}(W_{\sigma,m}^{\ast}MW_{\sigma,m})=F_{\sigma,m,j+1}(M) for j{1,,σm1}j\in\{1,\ldots,\sigma_{m}-1\} and Fσ,m,σm(Wσ,mMWσ,m)=Fσ,m,1(M)F_{\sigma,m,\sigma_{m}}(W_{\sigma,m}^{\ast}MW_{\sigma,m})=F_{\sigma,m,1}(M).

Next, fix j{0,,σm1}j\in\{0,\ldots,\sigma_{m}-1\}. For all t[jσm,j+1σm]t\in\left[\frac{j}{\sigma_{m}},\frac{j+1}{\sigma_{m}}\right], set

fj(t)=Fσ,m,j+1(Dσ,m(Uσ,m(tσmj)a(tσmj)Uσ,m(tσmj))).f_{j}(t)=F_{\sigma,m,j+1}(D_{\sigma,m}(U_{\sigma,m}^{\ast}(t\sigma_{m}-j)\cdot a(t\sigma_{m}-j)\cdot U_{\sigma,m}(t\sigma_{m}-j))).

Since Fσ,m,j+1F_{\sigma,m,j+1} is continuous, then so is fjf_{j} on [jσm,j+1σm]\left[\frac{j}{\sigma_{m}},\frac{j+1}{\sigma_{m}}\right].

Now, let j{0,,σm1}j\in\{0,\ldots,\sigma_{m}-1\}. We have since aa is 11-periodic

fj(j+1σm)=Fσ,m,j+1(Dσ,m(Uσ,m(1)a(1)Uσ,m(1)))=Fσ,m,j+1(Dσ,m(Wσ,mUσ,m(0)a(0)Uσ,m(0)Wσ,m))=Fσ,m,j+1(Wσ,mDσ,m(Uσ,m(0)a(0)Uσ,m(0))Wσ,m)=Fσ,m,j+2(Dσ,m(Uσ,m(0)a(0)Uσ,m(0)))=fj+1(j+1σm).\begin{split}f_{j}\left(\frac{j+1}{\sigma_{m}}\right)&=F_{\sigma,m,j+1}(D_{\sigma,m}(U_{\sigma,m}^{\ast}(1)\cdot a(1)\cdot U_{\sigma,m}(1)))\\ &=F_{\sigma,m,j+1}(D_{\sigma,m}(W_{\sigma,m}^{\ast}U_{\sigma,m}^{\ast}(0)\cdot a(0)\cdot U_{\sigma,m}(0)W_{\sigma,m}))\\ &=F_{\sigma,m,j+1}(W_{\sigma,m}^{\ast}D_{\sigma,m}(U_{\sigma,m}^{\ast}(0)\cdot a(0)\cdot U_{\sigma,m}(0))W_{\sigma,m})\\ &=F_{\sigma,m,j+2}(D_{\sigma,m}(U_{\sigma,m}^{\ast}(0)\cdot a(0)\cdot U_{\sigma,m}(0)))=f_{j+1}\left(\frac{j+1}{\sigma_{m}}\right).\end{split}

A similar computation shows that fσm(1)=f0(0)f_{\sigma_{m}}(1)=f_{0}(0). Thus, the map defined for all t[0,1]t\in[0,1] by

f(t)={fj(t) if t[jσm,j+1σm]j{0,,σm1},f(t)=\begin{cases}f_{j}(t)&\text{ if }t\in\left[\frac{j}{\sigma_{m}},\frac{j+1}{\sigma_{m}}\right]\ \land\ j\in\{0,\ldots,\sigma_{m}-1\},\end{cases}

is well-defined and continuous and f(0)=f(1)f(0)=f(1). Thus, ff extends uniquely to an element in 𝔓(σ,m1){{\mathfrak{CP}}\left({\sigma},{m-1}\right)}, which we will still denote by ff. And, by construction, we have that ασ,m1(f)=𝔼σ,m(a)\alpha_{\sigma,m-1}(f)=\mathds{E}_{\sigma,m}(a).

Next, it remains to show that if bασ,m1(𝔓(σ,m1))b\in\alpha_{\sigma,m-1}({{\mathfrak{CP}}\left({\sigma},{m-1}\right)}), then 𝔼σ,m(b)=b\mathds{E}_{\sigma,m}(b)=b. Let bασ,m1(𝔓(σ,m1))b\in\alpha_{\sigma,m-1}({{\mathfrak{CP}}\left({\sigma},{m-1}\right)}). Thus, there exists c𝔓(σ,m1)c\in{{\mathfrak{CP}}\left({\sigma},{m-1}\right)} such that b=ασ,m1(c)b=\alpha_{\sigma,m-1}(c).

Hence, for all tt\in{\mathds{R}}, we have

𝔼σ,m(b)(t)=Uσ,m(t)[Dσ,m(Uσ,m(t)ασ,m1(c)(t)Uσ,m(t))]Uσ,m(t)=Uσ,m(t)[Dσ,m(c(tσm)c(t+1σm)c(t+σm1σm))]Uσ,m(t)=Uσ,m(t)(c(tσm)c(t+1σm)c(t+σm1σm))Uσ,m(t)=ασ,m1(c)(t)=b(t).\begin{split}&\mathds{E}_{\sigma,m}(b)(t)\\ &=U_{\sigma,m}(t)\left[D_{\sigma,m}\left(U_{\sigma,m}(t)^{\ast}\alpha_{\sigma,m-1}(c)(t)U_{\sigma,m}(t)\right)\right]U_{\sigma,m}(t)^{\ast}\\ &=U_{\sigma,m}(t)\left[D_{\sigma,m}\begin{pmatrix}c\left(\frac{t}{\sigma_{m}}\right)&&&\\ &c\left(\frac{t+1}{\sigma_{m}}\right)&&\\ &&\ddots&\\ &&&c\left(\frac{t+\sigma_{m}-1}{\sigma_{m}}\right)\end{pmatrix}\right]U_{\sigma,m}(t)^{\ast}\\ &=U_{\sigma,m}(t)\begin{pmatrix}c\left(\frac{t}{\sigma_{m}}\right)&&&\\ &c\left(\frac{t+1}{\sigma_{m}}\right)&&\\ &&\ddots&\\ &&&c\left(\frac{t+\sigma_{m}-1}{\sigma_{m}}\right)\end{pmatrix}U_{\sigma,m}(t)^{\ast}\\ &=\alpha_{\sigma,m-1}(c)(t)=b(t).\end{split}

Now, 𝔼σ,m\mathds{E}_{\sigma,m} is positive and contractive by construction, as the composition of *-isomorphisms and a conditional expectation. So 𝔼σ,m\mathds{E}_{\sigma,m} is a conditional expectation onto ασ,m1(𝔓(σ,m1))\alpha_{\sigma,m-1}({{\mathfrak{CP}}\left({\sigma},{m-1}\right)}) by [5, Tomiyama’s Theorem 1.5.10 and Theorem 3.5.3].

Finally, following [8, Theorem V.3.6], we have since Dσ,mD_{\sigma,m} preserves the trace, Tr\mathrm{Tr}:

τσ,m(𝔼σ,m(a))=01σm1Tr(𝔼σ,m(a)(t))dt=01σm1Tr(Uσ,m(t)[Dσ,m(Uσ,m(t)a(t)Uσ,m(t))]Uσ,m(t))dt=01σm1Tr(Dσ,m(Uσ,m(t)a(t)Uσ,m(t)))dt=01σm1Tr(Uσ,m(t)a(t)Uσ,m(t))dt=01σm1Tr(a(t))dt=τσ,m(a),\begin{split}\tau_{\sigma,m}(\mathds{E}_{\sigma,m}(a))&=\int_{0}^{1}\boxtimes\sigma_{m}^{-1}\mathrm{Tr}(\mathds{E}_{\sigma,m}(a)(t))\ dt\\ &=\int_{0}^{1}\boxtimes\sigma_{m}^{-1}\mathrm{Tr}(U_{\sigma,m}(t)\left[D_{\sigma,m}\left(U_{\sigma,m}(t)^{\ast}a(t)U_{\sigma,m}(t)\right)\right]U_{\sigma,m}(t)^{\ast})\ dt\\ &=\int_{0}^{1}\boxtimes\sigma_{m}^{-1}\mathrm{Tr}(D_{\sigma,m}\left(U_{\sigma,m}(t)^{\ast}a(t)U_{\sigma,m}(t)\right))\ dt\\ &=\int_{0}^{1}\boxtimes\sigma_{m}^{-1}\mathrm{Tr}\left(U_{\sigma,m}(t)^{\ast}a(t)U_{\sigma,m}(t)\right)\ dt\\ &=\int_{0}^{1}\boxtimes\sigma_{m}^{-1}\mathrm{Tr}\left(a(t)\right)\ dt=\tau_{\sigma,m}(a),\end{split}

which completes the proof. ∎

4. The metric geometry of the class of the Bunce-Deddens Algebras

In this section, we construct our Lip-norms on circle algebras that are meant to be suitable with both the inductive limit structure and the conditional expectations presented in the previous section. This will then allow us to utilize Theorem 2.7 to get one step closer to building Lip-norms on the Bunce-Deddens algebras. We begin with some classical structure.

Notation 4.1.

Let (X,𝖽)(X,\mathsf{d}) be a locally compact metric space, and let n{0}n\in{\mathds{N}}\setminus\{0\}. Let 𝔅{\mathfrak{B}} be a unital C*-subalgebra of the C*-algebra Cb(X,𝔐n())C_{b}(X,{\mathfrak{M}}_{n}({\mathds{C}})) of bounded 𝔐n(){\mathfrak{M}}_{n}({\mathds{C}})-valued continuous functions over XX, such that the unit 1𝔅1_{\mathfrak{B}} is the unit of Cb(X,𝔐n())C_{b}(X,{\mathfrak{M}}_{n}({\mathds{C}})) — the constant function equal to the identity in 𝔐n(){\mathfrak{M}}_{n}({\mathds{C}}).

For all a𝔅a\in{\mathfrak{B}}, we define:

l𝖽n(a)=supx,yX,xya(x)a(y)𝔐n()𝖽(x,y).l_{\mathsf{d}}^{n}(a)=\sup_{x,y\in X,x\neq y}\frac{\|a(x)-a(y)\|_{{\mathfrak{M}}_{n}({\mathds{C}})}}{\mathsf{d}(x,y)}\text{.}

Last, if XX is a normed vector space with norm X\left\|{\cdot}\right\|_{X}, then we write lXnl_{\left\|{\cdot}\right\|_{X}}^{n} for ldnl_{d}^{n} with dd the metric induced by X\left\|{\cdot}\right\|_{X}.

We make two simple but important remarks:

  • a𝔅ldn(a)=ldn(a)\forall a\in{\mathfrak{B}}\quad l_{d}^{n}(a^{\ast})=l_{d}^{n}(a),

  • a,b𝔅ldn(ab)a𝔅ldn(b)+ldn(a)b𝔅\forall a,b\in{\mathfrak{B}}\quad l_{d}^{n}(ab)\leqslant\left\|{a}\right\|_{{\mathfrak{B}}}l_{d}^{n}(b)+l_{d}^{n}(a)\left\|{b}\right\|_{{\mathfrak{B}}}.

Lemma 4.2.

Let m{0}m\in{\mathds{N}}\setminus\{0\}. We use Notation (3.3).

  • For all j,k{1,,m}j,k\in\{1,\ldots,m\}, we estimate:

    l||(zj,km)mjm3/2.l_{|\cdot|}(z^{m}_{j,k})\leqslant\frac{m-j}{m^{3/2}}\text{.}
  • We have:

    l||m(Um)2m2+3m+16m.l_{|\cdot|}^{m}(U_{m})\leqslant\sqrt{\frac{2m^{2}+3m+1}{6m}}.
Proof.

Fix m{0}m\in{\mathds{N}}\setminus\{0\} and j,k{1,,n}j,k\in\{1,\ldots,n\}. Of course l||(z)1l_{|\cdot|}(z)\leqslant 1. Hence, if r,tr,t\in{\mathds{R}}, then:

|zj,km(t)zj,km(r)|\displaystyle\left|z^{m}_{j,k}(t)-z^{m}_{j,k}(r)\right|
=|1mz((mj)(t+k1)m)1mz((mj)(r+k1)m)|\displaystyle=\left|\frac{1}{\sqrt{m}}z\left(\frac{(m-j)(t+k-1)}{m}\right)-\frac{1}{\sqrt{m}}z\left(\frac{(m-j)(r+k-1)}{m}\right)\right|
1m|(mj)(t+k1)m(mj)(r+k1)m|\displaystyle\leqslant\frac{1}{\sqrt{m}}\left|\frac{(m-j)(t+k-1)}{m}-\frac{(m-j)(r+k-1)}{m}\right|
1m(mjm)|tr|.\displaystyle\leqslant\frac{1}{\sqrt{m}}\left(\frac{m-j}{m}\right)|t-r|.

Thus l||(zj,km)mjm3/2l_{|\cdot|}(z^{m}_{j,k})\leqslant\frac{m-j}{m^{3/2}}.

Let t,st,s\in{\mathds{R}}. Let ξ=(ξ1,,ξm)\xi=(\xi_{1},\ldots,\xi_{m}) with ξ2=j=1m|ξj|21\left\|{\xi}\right\|_{2}=\sqrt{\sum_{j=1}^{m}|\xi_{j}|^{2}}\leqslant 1. We compute:

(Um(t)Um(s))ξ=(k=1m(zj,k(t)zj,k(s))ξk)j{1,,m}.(U_{m}(t)-U_{m}(s))\xi=\left(\sum_{k=1}^{m}(z_{j,k}(t)-z_{j,k}(s))\xi_{k}\right)_{j\in\{1,\ldots,m\}}\text{.}

Now for all j{1,,m}j\in\{1,\ldots,m\}:

|k=1m(zj,k(t)zj,k(s))ξk|2\displaystyle\left|\sum_{k=1}^{m}(z_{j,k}(t)-z_{j,k}(s))\xi_{k}\right|^{2} (k=1m|zj,k(t)zj,k(s)|2)(k=1m|ξk|2)\displaystyle\leqslant\left(\sum_{k=1}^{m}\left|z_{j,k}(t)-z_{j,k}(s)\right|^{2}\right)\left(\sum_{k=1}^{m}|\xi_{k}|^{2}\right)
(k=1m(mjmm|ts|)2)1\displaystyle\leqslant\left(\sum_{k=1}^{m}\left(\frac{m-j}{m\sqrt{m}}|t-s|\right)^{2}\right)\cdot 1
(mj)2m2|ts|2.\displaystyle\leqslant\frac{(m-j)^{2}}{m^{2}}|t-s|^{2}\text{.}

Now

(Um(t)Um(s))ξ2\displaystyle\left\|{(U_{m}(t)-U_{m}(s))\xi}\right\|_{2} j=1m(mj)2m2|ts|2\displaystyle\leqslant\sqrt{\sum_{j=1}^{m}\frac{(m-j)^{2}}{m^{2}}|t-s|^{2}}
|ts|1m2+4m2++1\displaystyle\leqslant|t-s|\sqrt{\frac{1}{m^{2}}+\frac{4}{m^{2}}+\ldots+1}
=|ts|(2m+1)(m+1)m6m2\displaystyle=|t-s|\sqrt{\frac{(2m+1)(m+1)m}{6m^{2}}}
=|ts|2m2+3m+16m.\displaystyle=|t-s|\sqrt{\frac{2m^{2}+3m+1}{6m}}\text{.}

Thus supξ21(Um(t)Um(s))ξ22m2+3m+16m|ts|\sup_{\left\|{\xi}\right\|_{2}\leqslant 1}\left\|{(U_{m}(t)-U_{m}(s))\xi}\right\|_{2}\leqslant\sqrt{\frac{2m^{2}+3m+1}{6m}}|t-s|. Hence

Um(s)Um(t)𝔐m()2m2+3m+16m|ts|.\left\|{U_{m}(s)-U_{m}(t)}\right\|_{{\mathfrak{M}}_{m}({\mathds{C}})}\leqslant\sqrt{\frac{2m^{2}+3m+1}{6m}}|t-s|.

This concludes our proof. ∎

We first define a natural (2,0)(2,0)-Leibniz Lip-norm on the circle algebras.

Definition 4.3.

Let σ𝒩\sigma\in{\mathscr{N}} and m{0}m\in{\mathds{N}}\setminus\{0\}. We define a𝔰𝔞(𝔓(σ,m))\forall a\in{\mathfrak{sa}\left({{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\right)}:

𝖫σ,m(a)=max{l||σm(Uσ,maUσ,m),aτm,σ(a)1𝔓(σ,m)𝔓(σ,m)}.{\mathsf{L}}_{\sigma,m}(a)=\max\left\{l_{|\cdot|}^{\boxtimes\sigma_{m}}(U_{\sigma,m}^{\ast}aU_{\sigma,m}),\left\|{a-\tau_{m,\sigma}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\right\}\text{.}

The main motivation for our choice of Lip-norm is that it is well-adapted to the conditional expectation. Before we prove that Definition (4.3) actually gives Lip-norms on circle algebras, we prove the following key result.

Lemma 4.4.

Let σ𝒩\sigma\in{\mathscr{N}} and let m{0}m\in{\mathds{N}}\setminus\{0\}. If a𝔓(σ,m)a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)} then:

𝖫σ,m(𝔼σ,m(a))𝖫σ,m(a).{\mathsf{L}}_{\sigma,m}({\mathds{E}_{\sigma,m}\left({a}\right)})\leqslant{\mathsf{L}}_{\sigma,m}(a)\text{.}
Proof.

Let n=σmn=\boxtimes\sigma_{m} and a𝔓(σ,m)a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)}. To ease notations, we just write ll for the seminorm l||σml_{|\cdot|}^{\boxtimes\sigma_{m}}.

Since |Dσ,m|𝔐n()1{\left|\mkern-1.5mu\left|\mkern-1.5mu\left|{D_{\sigma,m}}\right|\mkern-1.5mu\right|\mkern-1.5mu\right|_{{\mathfrak{M}}_{n}({\mathds{C}})}}\leqslant 1, we compute:

l(Uσ,m𝔼σ,m(a)Uσ,m)=l(Dσ,m(Uσ,maUσ,m))=supx,yxyDσ,m(Uσ,m(x)aUσ,m(x))Dσ,m(Uσ,m(y)aUσ,m(y))𝔐n()|xy|=supx,yxyDσ,m(Uσ,m(x)aUσ,m(x)Uσ,m(y)aUσ,m(y))𝔐n()|xy|supx,yxyUσ,m(x)aUσ,m(x)Uσ,m(y)aUσ,m(y)𝔐n()|xy|=l(Uσ,maUσ,m).l\left(U_{\sigma,m}^{\ast}{\mathds{E}_{\sigma,m}\left({a}\right)}U_{\sigma,m}\right)\\ \begin{aligned} &=l\left(D_{\sigma,m}\left(U_{\sigma,m}^{\ast}aU_{\sigma,m}\right)\right)\\ &=\sup_{\begin{subarray}{c}x,y\in{\mathds{R}}\\ x\not=y\end{subarray}}\frac{\left\|{D_{\sigma,m}\left(U_{\sigma,m}(x)^{\ast}aU_{\sigma,m}(x)\right)-D_{\sigma,m}\left(U_{\sigma,m}^{\ast}(y)aU_{\sigma,m}(y)\right)}\right\|_{{\mathfrak{M}}_{n}({\mathds{C}})}}{|x-y|}\\ &=\sup_{\begin{subarray}{c}x,y\in{\mathds{R}}\\ x\not=y\end{subarray}}\frac{\left\|{D_{\sigma,m}\left(U_{\sigma,m}^{\ast}(x)aU_{\sigma,m}(x)-U_{\sigma,m}^{\ast}(y)aU_{\sigma,m}(y)\right)}\right\|_{{\mathfrak{M}}_{n}({\mathds{C}})}}{|x-y|}\\ &\leqslant\sup_{\begin{subarray}{c}x,y\in{\mathds{R}}\\ x\not=y\end{subarray}}\frac{\left\|{U_{\sigma,m}^{\ast}(x)aU_{\sigma,m}(x)-U_{\sigma,m}^{\ast}(y)aU_{\sigma,m}(y)}\right\|_{{\mathfrak{M}}_{n}({\mathds{C}})}}{|x-y|}\\ &=l(U_{\sigma,m}^{\ast}aU_{\sigma,m})\text{.}\end{aligned}

Furthermore, by Lemma 3.14:

𝔼σ,m(a)τσ,m(𝔼σ,m(a))1𝔓(σ,m)𝔓(σ,m)\displaystyle\left\|{{\mathds{E}_{\sigma,m}\left({a}\right)}-\tau_{\sigma,m}({\mathds{E}_{\sigma,m}\left({a}\right)})1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}
=𝔼σ,m(aτσ,m(a)1𝔓(σ,m))𝔓(σ,m)\displaystyle=\left\|{{\mathds{E}_{\sigma,m}\left({a-\tau_{\sigma,m}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}}\right)}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}
aτσ,m(a)1𝔓(σ,m)𝔓(σ,m).\displaystyle\leqslant\left\|{a-\tau_{\sigma,m}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\text{.}

Therefore:

𝖫σ,m(Eσ,m(a))\displaystyle{\mathsf{L}}_{\sigma,m}(E_{\sigma,m}(a))
=max{l(Uσ,m𝔼σ,m(a)Uσ,m),𝔼σ,m(a)τσ,m(𝔼σ,m(a))1𝔓(σ,m)𝔓(σ,m)}\displaystyle=\max\big{\{}l(U_{\sigma,m}^{\ast}{\mathds{E}_{\sigma,m}\left({a}\right)}U_{\sigma,m}),\left\|{{\mathds{E}_{\sigma,m}\left({a}\right)}-\tau_{\sigma,m}({\mathds{E}_{\sigma,m}\left({a}\right)})1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\big{\}}
max{l(Uσ,maUσ,m),aτσ,m(a)1𝔓(σ,m)𝔓(σ,m)}\displaystyle\leqslant\max\big{\{}l(U_{\sigma,m}^{\ast}aU_{\sigma,m}),\left\|{a-\tau_{\sigma,m}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\big{\}}
=𝖫σ,m(a).\displaystyle={\mathsf{L}}_{\sigma,m}(a)\text{.}

This concludes our result. ∎

Theorem 4.5.

If σ𝒩\sigma\in{\mathscr{N}} and m{0}m\in{\mathds{N}}\setminus\{0\}, then (𝔓(σ,m),𝖫σ,m)\left({{\mathfrak{CP}}\left({\sigma},{m}\right)},{\mathsf{L}}_{\sigma,m}\right) is a (2,0)(2,0)-quantum compact metric space.

Proof.

As the maximum of two lower semi-continuous seminorms, 𝖫σ,m{\mathsf{L}}_{\sigma,m} is a lower semi-continuous seminorm (allowing for the value \infty).

By [3, Lemma 2.3], the seminorm a𝔓(σ,m)aτσ,m(a)1𝔓(σ,m)𝔓(σ,m)a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)}\mapsto\left\|{a-\tau_{\sigma,m}(a)1_{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}} is (2,0)-quasi-Leibniz. As a𝔓(σ,m)l||σm(Uσ,maUσ,m)a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)}\mapsto l_{|\cdot|}^{\boxtimes\sigma_{m}}(U_{\sigma,m}^{\ast}aU_{\sigma,m}) is of course Leibniz, we conclude that 𝖫σ,m{\mathsf{L}}_{\sigma,m} is (2,0)(2,0)-quasi-Leibniz.

If 𝖫σ,m(a)=0{\mathsf{L}}_{\sigma,m}(a)=0 then l||σm(Uσ,maUσ,m)=0l_{|\cdot|}^{\boxtimes\sigma_{m}}(U_{\sigma,m}^{\ast}aU_{\sigma,m})=0, so there exists T𝔐σm()T\in{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}}) such that a=Uσ,mTUσ,ma=U_{\sigma,m}TU_{\sigma,m}^{\ast}. On the other hand:

0\displaystyle 0 =aτσ,m(a)1𝔓(σ,m)𝔓(σ,m)\displaystyle=\left\|{a-\tau_{\sigma,m}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}
=Uσ,m(aτσ,m(a)1𝔓(σ,m))Uσ,m𝔓(σ,m) since Uσ,m is unitary,\displaystyle=\left\|{U_{\sigma,m}^{\ast}(a-\tau_{\sigma,m}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}})U_{\sigma,m}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\text{ since $U_{\sigma,m}$ is unitary,}
=Uσ,maUσ,nτσ,m(a)1𝔓(σ,m)𝔓(σ,m)\displaystyle=\left\|{U_{\sigma,m}^{\ast}aU_{\sigma,n}-\tau_{\sigma,m}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}

so T=τσ,m(a)idσmT=\tau_{\sigma,m}(a)\mathrm{id}_{\boxtimes\sigma_{m}}, and therefore a=τσ,m(a)1𝔓(σ,m)1𝔓(σ,m)a=\tau_{\sigma,m}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\in{\mathds{C}}1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}. Of course, 𝖫σ,m(1𝔓(σ,m))=0{\mathsf{L}}_{\sigma,m}(1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}})=0.

If f𝔓(σ,0)f\in{{\mathfrak{CP}}\left({\sigma},{0}\right)} with l||1(f)<l_{|\cdot|}^{1}(f)<\infty and if T𝔓(σ,m)T\in{{\mathfrak{CP}}\left({\sigma},{m}\right)} then:

(4.1) l||σm(fT)l||1(f)T𝔓(σ,m)+fCb()l||σm(T)l_{|\cdot|}^{\boxtimes\sigma_{m}}(f\otimes T)\leqslant l_{|\cdot|}^{1}(f)\left\|{T}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}+\left\|{f}\right\|_{C_{b}({\mathds{R}})}l_{|\cdot|}^{\boxtimes\sigma_{m}}(T)

using the standard *-isomorphism between 𝔓(σ,0)𝔐σm(){{\mathfrak{CP}}\left({\sigma},{0}\right)}\otimes{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}}) and 𝔓(σ,m){{\mathfrak{CP}}\left({\sigma},{m}\right)} given on elementary tensors by fT𝔓(σ,0)𝔐σm()(tf(t)T𝔐σm())𝔓(σ,m)f\otimes T\in{{\mathfrak{CP}}\left({\sigma},{0}\right)}\otimes{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}})\mapsto(t\in{\mathds{R}}\mapsto f(t)T\in{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}}))\in{{\mathfrak{CP}}\left({\sigma},{m}\right)} [24, Theorem 6.4.17].

Using the same *-isomorphism, if a𝔓(σ,m)a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)} and ε>0\varepsilon>0, then there exist f1,,fk𝔓(σ,0)f_{1},\ldots,f_{k}\in{{\mathfrak{CP}}\left({\sigma},{0}\right)} and T1,,Tk𝔐σm()T_{1},\ldots,T_{k}\in{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}}) such that:

aj=1kfjTj𝔓(σ,m)<ε2.\left\|{a-\sum_{j=1}^{k}f_{j}\otimes T_{j}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}{}<\frac{\varepsilon}{2}.

Let K=kmax{TjMσm():j{1,,k}}K=k\max\left\{\left\|{T_{j}}\right\|_{M_{\boxtimes\sigma_{m}}({\mathds{C}})}:j\in\{1,\ldots,k\}\right\}.

As Lipschitz functions are dense in 𝔓(σ,0){{\mathfrak{CP}}\left({\sigma},{0}\right)}, there exists g1,,gk𝔓(σ,0)g_{1},\ldots,g_{k}\in{{\mathfrak{CP}}\left({\sigma},{0}\right)} such that fjgj𝔓(σ,0)<ε2K\left\|{f_{j}-g_{j}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{0}\right)}}<\frac{\varepsilon}{2K} while l||1(gj)<l_{|\cdot|}^{1}(g_{j})<\infty for all j{1,,k}j\in\{1,\ldots,k\}. Therefore:

aj=1kgjTj𝔓(σ,m)\displaystyle\left\|{a-\sum_{j=1}^{k}g_{j}\otimes T_{j}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}} aj=1kfjTj𝔓(σ,m)\displaystyle\leqslant\left\|{a-\sum_{j=1}^{k}f_{j}\otimes T_{j}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}
+j=1kfjTjj=1kgjTj𝔓(σ,m)\displaystyle\quad\quad+\left\|{\sum_{j=1}^{k}f_{j}\otimes T_{j}-\sum_{j=1}^{k}g_{j}\otimes T_{j}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}
ε2+j=1kfjgj𝔓(σ,0)Kε.\displaystyle\leqslant\frac{\varepsilon}{2}+\sum_{j=1}^{k}\left\|{f_{j}-g_{j}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{0}\right)}}K\leqslant\varepsilon\text{.}

Last, using the Leibniz property of l||σml_{|\cdot|}^{\boxtimes\sigma_{m}} and the fact that l||σm(Uσ,m)<l_{|\cdot|}^{\boxtimes\sigma_{m}}(U_{\sigma,m})<\infty, we conclude that l||σm(Uσ,mj=1k(gjTj)Uσ,m)<l_{|\cdot|}^{\boxtimes\sigma_{m}}(U_{\sigma,m}^{*}\sum_{j=1}^{k}(g_{j}\otimes T_{j})U_{\sigma,m})<\infty by Expression (4.1). This concludes the proof that the domain of 𝖫σ,m{\mathsf{L}}_{\sigma,m} is dense in 𝔓(σ,m){{\mathfrak{CP}}\left({\sigma},{m}\right)} since the other seminorm in its definition are actually continuous on 𝔓(σ,m){{\mathfrak{CP}}\left({\sigma},{m}\right)}.

Last, let (an)n(a_{n})_{n\in{\mathds{N}}} be a sequence in 𝔓(σ,m){{\mathfrak{CP}}\left({\sigma},{m}\right)} such that 𝖫σ,m(an)1{\mathsf{L}}_{\sigma,m}(a_{n})\leqslant 1 and τσ,m(an)=0\tau_{\sigma,m}(a_{n})=0. for all nn\in{\mathds{N}}. Since l||σm(an)1l_{|\cdot|}^{\boxtimes\sigma_{m}}(a_{n})\leqslant 1 for all nn\in{\mathds{N}}, the set {Uσ,manUσ,m:n}\{U_{\sigma,m}^{\ast}a_{n}U_{\sigma,m}:n\in{\mathds{N}}\} is equicontinuous, and thus {an:n}\{a_{n}:n\in{\mathds{N}}\} is equicontinuous as Uσ,mU_{\sigma,m} is unitary. Moreover, we also have:

nan𝔓(σ,m)anτσ,m(an)1𝔓(σ,m)𝔓(σ,m)1.\forall n\in{\mathds{N}}\quad\left\|{a_{n}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\leqslant\left\|{a_{n}-\tau_{\sigma,m}(a_{n})1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\leqslant 1\text{.}

So {an:n}\{a_{n}:n\in{\mathds{N}}\} is an equicontinuous set of continuous 11-periodic functions on {\mathds{R}}, all valued in the closed unit disk, which is compact. By Arzéla-Ascoli theorem, we thus conclude that {an:n}\{a_{n}:n\in{\mathds{N}}\} is totally bounded for the norm (note: apply Arzéla-Ascoli theorem to the restriction of these functions to the compact [0,1][0,1] and then conclude using periodicity). Therefore, the sequence (an)n(a_{n})_{n\in{\mathds{N}}} admits a Cauchy subsequence (ar(n))n(a_{r(n)})_{n\in{\mathds{N}}}. As 𝔓(σ,m){{\mathfrak{CP}}\left({\sigma},{m}\right)} is complete, (ar(n))n(a_{r(n)})_{n\in{\mathds{N}}} converges to some a𝔓(σ,m)a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)}. As 𝖫σ,m{\mathsf{L}}_{\sigma,m} is lower semi-continuous, we get 𝖫σ,m(a)1{\mathsf{L}}_{\sigma,m}(a)\leqslant 1.

Thus {a𝔓(σ,m):𝖫σ,m(a)1,τσ,m(a)=0}\{a\in{{\mathfrak{CP}}\left({\sigma},{m}\right)}:{\mathsf{L}}_{\sigma,m}(a)\leqslant 1,\tau_{\sigma,m}(a)=0\} is compact. By [25, Proposition 1.3], we thus can conclude that (𝔓(σ,m),𝖫σ,m)\left({{\mathfrak{CP}}\left({\sigma},{m}\right)},{\mathsf{L}}_{\sigma,m}\right) is a quantum compact metric space. ∎

Now, we study the metric properties of the connecting maps defining the Bunce-Deddens algebras.

Lemma 4.6.

Let σ𝒩\sigma\in{\mathscr{N}} and m{0}m\in{\mathds{N}}\setminus\{0\}. Let km=max{1,1+2l||σm1(Uσ,m1)σm}k_{m}=\max\left\{1,\frac{1+2l_{|\cdot|}^{\boxtimes\sigma_{m-1}}(U_{\sigma,m-1})}{\sigma_{m}}\right\}. If a𝔓(σ,m1)a\in{{\mathfrak{CP}}\left({\sigma},{m-1}\right)} then:

1σm2km𝖫σ,m1(a)𝖫σ,m(ασ,m1(a))km𝖫σ,m1(a).\frac{1}{\sigma_{m}^{2}k_{m}}{\mathsf{L}}_{\sigma,m-1}(a)\leqslant{\mathsf{L}}_{\sigma,m}(\alpha_{\sigma,m-1}(a))\leqslant k_{m}{\mathsf{L}}_{\sigma,m-1}(a)\text{.}
Proof.

We denote l||σml_{|\cdot|}^{\boxtimes\sigma_{m}} by ll and l||σm1l_{|\cdot|}^{\boxtimes\sigma_{m-1}} by l1l_{-1} in this proof.

For a𝔓(σ,m1)a\in{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}, set:

θ(a):t(a(tσm)a(t+1σm)a(t+σm1σm))𝔐σm().\theta(a):t\in{\mathds{R}}\mapsto\begin{pmatrix}a\left(\frac{t}{\sigma_{m}}\right)&&&\\ &a\left(\frac{t+1}{\sigma_{m}}\right)&&\\ &&\ddots&\\ &&&a\left(\frac{t+\sigma_{m}-1}{\sigma_{m}}\right)\end{pmatrix}\in{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}})\text{.}

We compute:

l(θ(a))\displaystyle l(\theta(a)) =supx,yxy{θ(a)(x)θ(a)(y)𝔐σm()|xy|}\displaystyle=\sup_{\begin{subarray}{c}x,y\in{\mathds{R}}\\ x\not=y\end{subarray}}\left\{\frac{\left\|{\theta(a)(x)-\theta(a)(y)}\right\|_{{\mathfrak{M}}_{\boxtimes\sigma_{m}}({\mathds{C}})}}{|x-y|}\right\}
=supx,yxy{maxj{0,,σm1}a(x+jσm)a(y+jσm)𝔐σm1()|xy|}\displaystyle=\sup_{\begin{subarray}{c}x,y\in{\mathds{R}}\\ x\not=y\end{subarray}}\left\{\max_{j\in\{0,\ldots,\sigma_{m}-1\}}\frac{\left\|{a\left(\frac{x+j}{\sigma_{m}}\right)-a\left(\frac{y+j}{\sigma_{m}}\right)}\right\|_{{\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}})}}{|x-y|}\right\}
=1σmsupx,yxy{maxj{0,,σm1}a(x+jσm)a(y+jσm)𝔐σm1()1σm|xy|}\displaystyle=\frac{1}{\sigma_{m}}\cdot\sup_{\begin{subarray}{c}x,y\in{\mathds{R}}\\ x\not=y\end{subarray}}\left\{\max_{j\in\{0,\ldots,\sigma_{m}-1\}}\frac{\left\|{a\left(\frac{x+j}{\sigma_{m}}\right)-a\left(\frac{y+j}{\sigma_{m}}\right)}\right\|_{{\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}})}}{\frac{1}{\sigma_{m}}|x-y|}\right\}
=1σmsupx,yxy{maxj{0,,σm1}a(x+jσm)a(y+jσm)𝔐σm1()|x+jσmy+jσm|}\displaystyle=\frac{1}{\sigma_{m}}\cdot\sup_{\begin{subarray}{c}x,y\in{\mathds{R}}\\ x\not=y\end{subarray}}\left\{\max_{j\in\{0,\ldots,\sigma_{m}-1\}}\frac{\left\|{a\left(\frac{x+j}{\sigma_{m}}\right)-a\left(\frac{y+j}{\sigma_{m}}\right)}\right\|_{{\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}})}}{\left|\frac{x+j}{\sigma_{m}}-\frac{y+j}{\sigma_{m}}\right|}\right\}
=1σml1(a).\displaystyle=\frac{1}{\sigma_{m}}l_{-1}(a)\text{.}

Therefore:

l(Uσ,mασ,m1(a)Un,m)\displaystyle l(U_{\sigma,m}^{\ast}\alpha_{\sigma,m-1}(a)U_{n,m})
=l(θ(a))=1σml1(a)=1σml1(Uσ,m1Uσ,m1aUσ,m1Uσ,m1)\displaystyle=l(\theta(a))=\frac{1}{\sigma_{m}}l_{-1}(a)=\frac{1}{\sigma_{m}}l_{-1}(U_{\sigma,m-1}U_{\sigma,m-1}^{\ast}aU_{\sigma,m-1}U_{\sigma,m-1}^{\ast})
1σm(l1(Uσ,m1aUσ,m1)+2l1(Uσ,m1)Uσ,m1aUσ,m1Cb(,𝔐σm1()))\displaystyle\leqslant\frac{1}{\sigma_{m}}\left(l_{-1}(U_{\sigma,m-1}^{\ast}aU_{\sigma,m-1})+2l_{-1}(U_{\sigma,m-1})\left\|{U_{\sigma,m-1}^{\ast}aU_{\sigma,m-1}}\right\|_{C_{b}({\mathds{R}},{\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}}))}\right)
1σm(𝖫σ,m1(a)+2l1(Uσ,m1)Uσ,m1aUσ,m1Cb(,𝔐σm1())).\displaystyle\leqslant\frac{1}{\sigma_{m}}\left({\mathsf{L}}_{\sigma,m-1}(a)+2l_{-1}(U_{\sigma,m-1})\left\|{U_{\sigma,m-1}^{\ast}aU_{\sigma,m-1}}\right\|_{C_{b}({\mathds{R}},{\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}}))}\right)\text{.}

Now,

Uσ,m1(aτσ,m1(a)1𝔓(σ,m1))Uσ,m1Cb(,𝔐σm1())aτσ,m1(a)1𝔓(σ,m1)𝔓(σ,m1)𝖫σ,m1(a).\begin{split}&\left\|{U_{\sigma,m-1}^{\ast}(a-\tau_{\sigma,m-1}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}})U_{\sigma,m-1}}\right\|_{C_{b}({\mathds{R}},{\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}}))}\\ &\leqslant\left\|{a-\tau_{\sigma,m-1}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}}\\ &\leqslant{\mathsf{L}}_{\sigma,m-1}(a).\end{split}

Thus since the seminorms ll and 𝖫σ,m1{\mathsf{L}}_{\sigma,m-1} vanish on scalars, we have

l(Uσ,mασ,m1(a)Un,m)=l(Uσ,mασ,m1(aτσ,m1(a)1𝔓(σ,m1))Un,m)1σm(𝖫σ,m1(aτσ,m1(a)1𝔓(σ,m1))+2l1(Uσ,m1)Uσ,m1(aτσ,m1(a)1𝔓(σ,m1))Uσ,m1Cb(,𝔐σm1()))1σm(𝖫σ,m1(a)+2l1(Uσ,m1)𝖫σ,m1(a))=1+2l1(Uσ,m1)σm𝖫σ,m1(a).\begin{split}&l(U_{\sigma,m}^{\ast}\alpha_{\sigma,m-1}(a)U_{n,m})\\ &=l(U_{\sigma,m}^{\ast}\alpha_{\sigma,m-1}(a-\tau_{\sigma,m-1}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}})U_{n,m})\\ &\leqslant\frac{1}{\sigma_{m}}\Big{(}{\mathsf{L}}_{\sigma,m-1}(a-\tau_{\sigma,m-1}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}})\\ &\quad\quad+2l_{-1}(U_{\sigma,m-1})\left\|{U_{\sigma,m-1}^{\ast}(a-\tau_{\sigma,m-1}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}})U_{\sigma,m-1}}\right\|_{C_{b}({\mathds{R}},{\mathfrak{M}}_{\boxtimes\sigma_{m-1}}({\mathds{C}}))}\Big{)}\\ &\leqslant\frac{1}{\sigma_{m}}\left({\mathsf{L}}_{\sigma,m-1}(a)+2l_{-1}(U_{\sigma,m-1}){\mathsf{L}}_{\sigma,m-1}(a)\right)\\ &=\frac{1+2l_{-1}(U_{\sigma,m-1})}{\sigma_{m}}{\mathsf{L}}_{\sigma,m-1}(a).\end{split}

Next, we note that

ασ,m1(a)τσ,m(ασ,m1(a))1𝔓(σ,m)𝔓(σ,m)=aτσ,m1(a)1𝔓(σ,m1)𝔓(σ,m1)𝖫σ,m1(a).\begin{split}&\left\|{\alpha_{\sigma,m-1}(a)-\tau_{\sigma,m}(\alpha_{\sigma,m-1}(a))1_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\\ &=\left\|{a-\tau_{\sigma,m-1}(a)1_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}}\leqslant{\mathsf{L}}_{\sigma,m-1}(a).\end{split}

Since α(c)τσ,n(ασ,m(c))𝔓(σ,n)=cτσ,m1(c)𝔓(σ,m1)\left\|{\alpha(c)-\tau_{\sigma,n}(\alpha_{\sigma,m}(c))}\right\|_{{{\mathfrak{CP}}\left({\sigma},{n}\right)}}=\left\|{c-\tau_{\sigma,m-1}(c)}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}} by construction, we conclude:

𝖫σ,m(ασ,m(a))km𝖫σ,m1(a).{\mathsf{L}}_{\sigma,m}(\alpha_{\sigma,m}(a))\leqslant k_{m}{\mathsf{L}}_{\sigma,m-1}(a)\text{.}

Let c𝔓(σ,m1)c\in{{\mathfrak{CP}}\left({\sigma},{m-1}\right)} such that 𝖫σ,m(ασ,m1(c))=1{\mathsf{L}}_{\sigma,m}(\alpha_{\sigma,m-1}(c))=1. By the above computation, we see that l1(c)σml_{-1}(c)\leqslant\sigma_{m} and cτσ,m1(c)1𝔓(σ,m1)𝔓(σ,m1)1\left\|{c-\tau_{\sigma,m-1}(c)1_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}}\leqslant 1. So:

l1(Uσ,m1cUσ,m1)2l1(Uσ,m1)c𝔓(σ,m1)+l1(c).\begin{split}l_{-1}(U_{\sigma,m-1}^{\ast}cU_{\sigma,m-1})\leqslant 2l_{-1}(U_{\sigma,m-1})\|c\|_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}}+l_{-1}(c).\end{split}

Thus, since l1l_{-1} vanishes on scalars, we have

l1(Uσ,m1cUσ,m1)=l1(Uσ,m1(cτσ,m1(c)1𝔓(σ,m1))Uσ,m1)2l1(Uσ,m1)cτσ,m1(c)1𝔓(σ,m1)𝔓(σ,m1)+l1(c)2l1(Uσ,m1)+l1(c)2l1(Uσ,m1)σm+σm=(1+2l1(Uσ,m1))σm.\begin{split}&l_{-1}(U_{\sigma,m-1}^{\ast}cU_{\sigma,m-1})\\ &=l_{-1}(U_{\sigma,m-1}^{\ast}(c-\tau_{\sigma,m-1}(c)1_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}})U_{\sigma,m-1})\\ &\leqslant 2l_{-1}(U_{\sigma,m-1})\|c-\tau_{\sigma,m-1}(c)1_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}}\|_{{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}}+l_{-1}(c)\\ &\leqslant 2l_{-1}(U_{\sigma,m-1})+l_{-1}(c)\\ &\leqslant 2l_{-1}(U_{\sigma,m-1})\sigma_{m}+\sigma_{m}=(1+2l_{-1}(U_{\sigma,m-1}))\sigma_{m}.\end{split}

As above, we conclude:

c𝔓(σ,m1)𝖫σ,m1(c)(1+2l1(Uσ,m1))σm𝖫σ,m(ασ,m1(c)).\forall c\in{{\mathfrak{CP}}\left({\sigma},{m-1}\right)}\quad{\mathsf{L}}_{\sigma,m-1}(c)\leqslant(1+2l_{-1}(U_{\sigma,m-1}))\sigma_{m}{\mathsf{L}}_{\sigma,m}(\alpha_{\sigma,m-1}(c))\text{.}

This concludes our proof. ∎

Theorem 4.7.

If σ𝒩\sigma\in{\mathscr{N}}, and if for all mm\in{\mathds{N}} we set 𝖲σ,0=𝖫σ,0\mathsf{S}_{\sigma,0}={\mathsf{L}}_{\sigma,0} on 𝔰𝔞(𝔓(σ,0)){\mathfrak{sa}\left({{{\mathfrak{CP}}\left({\sigma},{0}\right)}}\right)} and for all m{0}m\in{\mathds{N}}\setminus\{0\}:

a𝔰𝔞(𝔓(σ,m))𝖲σ,m(a)=max{ϰm𝖫σ,m(a),𝖲σ,m1ασ,m11𝔼σ,m(a),12ma𝔼σ,m(a)𝔓(σ,m)}\forall a\in{\mathfrak{sa}\left({{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\right)}\quad\mathsf{S}_{\sigma,m}(a)=\\ \max\left\{\varkappa_{m}{\mathsf{L}}_{\sigma,m}(a),\mathsf{S}_{\sigma,m-1}\circ\alpha_{\sigma,m-1}^{-1}\circ{\mathds{E}_{\sigma,m}\left({a}\right)},\frac{1}{2^{m}}\left\|{a-{\mathds{E}_{\sigma,m}\left({a}\right)}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\right\}

where for all m{0}m\in{\mathds{N}}\setminus\{0\}, we have km=1+2l||σm1(Uσ,m1)σmk_{m}=\frac{1+2l_{|\cdot|}^{\boxtimes\sigma_{m-1}}(U_{\sigma,m-1})}{\sigma_{m}}, and:

nϰn={1 if n{0,1},ϰn1kn otherwise\forall n\in{\mathds{N}}\quad\varkappa_{n}=\begin{cases}1\text{ if $n\in\{0,1\}$,}\\ \frac{\varkappa_{n-1}}{k_{n}}\text{ otherwise}\end{cases}

then (𝔰𝔞(𝔓(σ,m)),𝖲σ,m)({\mathfrak{sa}\left({{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\right)},\mathsf{S}_{\sigma,m}) is a quantum metric order unit space and there exists a quantum metric order unit space (O(σ),𝖲)(O(\sigma),\mathsf{S}) such that:

limm𝖽𝗂𝗌𝗍q(O(σ),𝖲),(𝔰𝔞(𝔓(σ,m)),𝖲σ,m))=0.\lim_{m\rightarrow\infty}{\mathsf{dist}}_{q}(O(\sigma),\mathsf{S}),({\mathfrak{sa}\left({{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\right)},\mathsf{S}_{\sigma,m}))=0\text{.}
Proof.

We apply Theorem (2.7) to the sequence ((𝔓(σ,m),𝖫σ,m),ασ,m)m(({{\mathfrak{CP}}\left({\sigma},{m}\right)},{\mathsf{L}}_{\sigma,m}),\alpha_{\sigma,m})_{m\in{\mathds{N}}} and the conditional expectations 𝔼σ,m\mathds{E}_{\sigma,m}. ∎

Of course, we now want to show that O(σ)O(\sigma), as defined in Theorem (4.7), is 𝔰𝔞(𝔅𝔇(σ)){\mathfrak{sa}\left({{{\mathfrak{BD}}\left({\sigma}\right)}}\right)}. This requires us to generalize techniques from [1] to our current setting.

5. The Propinquity for order-unit-based quantum metric spaces and Rieffel’s quantum Gromov-Hausdorff distance

The proof of completeness of the propinquity [15] and the construction of the inductive limit of an inductive sequence of C*-algebras share some obvious patterns, which were first exploited in [1]. In order to extend the techniques in [1] to the setting of Rieffel’s distance, we first derive a new expression for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} motivated by the construction of the propinquity which we now follow, but in this much more relaxed framework of quantum metric order unit spaces.

Definition 5.1.

A quantum order-unit isometry π:(𝔄,𝖫𝔄)(𝔅,𝖫𝔅)\pi:({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}})\rightarrow({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) between two quantum metric order unit spaces (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) is a positive linear map π:𝔄𝔅\pi:{\mathfrak{A}}\rightarrow{\mathfrak{B}} which maps the order unit of 𝔄{\mathfrak{A}} to the unity order of 𝔅{\mathfrak{B}}, such that:

b𝔅𝖫𝔅(b)=inf{𝖫𝔄(a):π(a)=b}.\forall b\in{\mathfrak{B}}\quad{\mathsf{L}}_{\mathfrak{B}}(b)=\inf\left\{{\mathsf{L}}_{\mathfrak{A}}(a):\pi(a)=b\right\}\text{.}
Definition 5.2.

If (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) are two quantum metric order unit spaces, then an order unit tunnel τ=(𝔇,𝖫𝔇,π𝔄.π𝔅)\tau=({\mathfrak{D}},{\mathsf{L}}_{\mathfrak{D}},\pi_{\mathfrak{A}}.\pi_{\mathfrak{B}}) is an ordered quadruple such that (𝔇,𝖫𝔇)({\mathfrak{D}},{\mathsf{L}}_{\mathfrak{D}}) is an quantum metric order unit space, while π𝔄\pi_{\mathfrak{A}} and π𝔅\pi_{\mathfrak{B}} are quantum order unit isometries from (𝔇,𝖫𝔇)({\mathfrak{D}},{\mathsf{L}}_{\mathfrak{D}}) onto, respectively, (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}).

Definition 5.3.

The extent χ(τ){\chi\left({\tau}\right)} of an order unit tunnel τ\tau from (𝔄1,𝖫1)({\mathfrak{A}}_{1},{\mathsf{L}}_{1}) to (𝔄2,𝖫2)({\mathfrak{A}}_{2},{\mathsf{L}}_{2}) is:

maxj{1,2}𝖧𝖺𝗎𝗌𝗆𝗄𝖫𝔇(𝒮(𝔇),{φπj:φ𝒮(𝔄j)}).\max_{j\in\{1,2\}}{\mathsf{Haus}_{{\mathsf{mk}_{{\mathsf{L}}_{\mathfrak{D}}}}}}\left({\mathscr{S}}({\mathfrak{D}}),\left\{\varphi\circ\pi_{j}:\varphi\in{\mathscr{S}}({\mathfrak{A}}_{j})\right\}\right)\text{.}

We remark that if 𝖫{\mathsf{L}} is an admissible Lip-norm for (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) then we can form the tunnel:

(𝔄𝔅,𝖫,(a,b)𝔄𝔅a,(a,b)𝔄𝔅b).\left({\mathfrak{A}}\oplus{\mathfrak{B}},{\mathsf{L}},(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}}\mapsto a,(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}}\mapsto b\right)\text{.}

The following lemma reconciles the extent of this tunnel with Rieffel’s computation of 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}.

Lemma 5.4.

If (𝔄1,𝖫1)({\mathfrak{A}}_{1},{\mathsf{L}}_{1}) and (𝔄2,𝖫2)({\mathfrak{A}}_{2},{\mathsf{L}}_{2}) are two quantum metric order unit spaces, and if τ=(𝔄1𝔄2,𝖫,π1,π2)\tau=({\mathfrak{A}}_{1}\oplus{\mathfrak{A}}_{2},{\mathsf{L}},\pi_{1},\pi_{2}) is a tunnel from (𝔄1,𝖫1)({\mathfrak{A}}_{1},{\mathsf{L}}_{1}) to (𝔄2,𝖫2)({\mathfrak{A}}_{2},{\mathsf{L}}_{2}) with πj:(a1,a2)𝔄1𝔄2aj\pi_{j}:(a_{1},a_{2})\in{\mathfrak{A}}_{1}\oplus{\mathfrak{A}}_{2}\mapsto a_{j} for both j{1,2}j\in\{1,2\}, then:

χ(τ)=𝖧𝖺𝗎𝗌𝗆𝗄𝖫((𝔄1,𝖫1),(𝔄2,𝖫2)).{\chi\left({\tau}\right)}={\mathsf{Haus}_{{\mathsf{mk}_{{\mathsf{L}}}}}}(({\mathfrak{A}}_{1},{\mathsf{L}}_{1}),({\mathfrak{A}}_{2},{\mathsf{L}}_{2}))\text{.}
Proof.

Write λ=𝖧𝖺𝗎𝗌𝗆𝗄𝖫((𝔄1,𝖫1),(𝔄2,𝖫2))\lambda={\mathsf{Haus}_{{\mathsf{mk}_{{\mathsf{L}}}}}}(({\mathfrak{A}}_{1},{\mathsf{L}}_{1}),({\mathfrak{A}}_{2},{\mathsf{L}}_{2})).

Let φ𝒮(𝔄1𝔄2)\varphi\in{\mathscr{S}}({\mathfrak{A}}_{1}\oplus{\mathfrak{A}}_{2}). There exists t1[0,1]t_{1}\in[0,1], φ1𝒮(𝔄1)\varphi_{1}\in{\mathscr{S}}({\mathfrak{A}}_{1}), and φ2𝒮(𝔄2)\varphi_{2}\in{\mathscr{S}}({\mathfrak{A}}_{2}) such that φ=t1φ1+(1t1)φ2\varphi=t_{1}\varphi_{1}+(1-t_{1})\varphi_{2}. Now, there exists ψ1𝒮(𝔄1)\psi_{1}\in{\mathscr{S}}({\mathfrak{A}}_{1}) such that 𝗆𝗄𝖫(ψ1,φ2)λ{\mathsf{mk}_{{\mathsf{L}}}}(\psi_{1},\varphi_{2})\leqslant\lambda by definition of λ\lambda. Set θ=t1φ1+(1t1)ψ1\theta=t_{1}\varphi_{1}+(1-t_{1})\psi_{1}. We then compute:

𝗆𝗄𝖫(φ,θ)\displaystyle{\mathsf{mk}_{{\mathsf{L}}}}(\varphi,\theta) =sup{|φ(a,b)θ(a)|:𝖫(a,b)1}\displaystyle=\sup\left\{|\varphi(a,b)-\theta(a)|:{\mathsf{L}}(a,b)\leqslant 1\right\}
(1t1)sup{|φ2(b)ψ1(a)|:𝖫(a,b)1}\displaystyle\leqslant(1-t_{1})\sup\left\{|\varphi_{2}(b)-\psi_{1}(a)|:{\mathsf{L}}(a,b)\leqslant 1\right\}
𝗆𝗄𝖫(ψ1,φ2)λ.\displaystyle\leqslant{\mathsf{mk}_{{\mathsf{L}}}}(\psi_{1},\varphi_{2})\leqslant\lambda\text{.}

By symmetry in 𝔄1{\mathfrak{A}}_{1} and 𝔄2{\mathfrak{A}}_{2}, we conclude that χ(τ)λ{\chi\left({\tau}\right)}\leqslant\lambda.

On the other hand, let φ𝒮(𝔄1)\varphi\in{\mathscr{S}}({\mathfrak{A}}_{1}). Of course, with the usual identification, φ𝒮(𝔄1𝔄2)\varphi\in{\mathscr{S}}({\mathfrak{A}}_{1}\oplus{\mathfrak{A}}_{2}) (with φ(a1,a2)=φ(a1)\varphi(a_{1},a_{2})=\varphi(a_{1}) for all (a1,a2)𝔄1𝔄2(a_{1},a_{2})\in{\mathfrak{A}}_{1}\oplus{\mathfrak{A}}_{2}). By definition of χ(τ){\chi\left({\tau}\right)}, there exists ψ𝒮(𝔄2)\psi\in{\mathscr{S}}({\mathfrak{A}}_{2}) such that 𝗆𝗄𝖫(φ,ψ)χ(τ){\mathsf{mk}_{{\mathsf{L}}}}(\varphi,\psi)\leqslant{\chi\left({\tau}\right)}. As φ\varphi is arbitrary and by symmetry in 𝔄1{\mathfrak{A}}_{1} and 𝔄2{\mathfrak{A}}_{2}, we conclude λχ(τ)\lambda\leqslant{\chi\left({\tau}\right)}. ∎

Thus, we obtain a new expression for Rieffel’s distance in the spirit of the propinquity.

Theorem 5.5.

If (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) are two quantum metric order unit spaces, then:

𝖽𝗂𝗌𝗍q((𝔄,𝖫𝔄),(𝔅,𝖫𝔅))=inf{χ(τ): τ is an order unit tunnel from (𝔄,𝖫𝔄) to (𝔅,𝖫𝔅)}.{\mathsf{dist}}_{q}(({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}),({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}))=\inf\{{\chi\left({\tau}\right)}:\\ \text{ $\tau$ is an order unit tunnel from $({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}})$ to $({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}})$}\}\text{.}
Proof.

Let τ=(𝔇,𝖫,π𝔄,π𝔅)\tau=({\mathfrak{D}},{\mathsf{L}},\pi_{\mathfrak{A}},\pi_{\mathfrak{B}}) be an order unit tunnel. Let ε>0\varepsilon>0. For all d1,d2𝔇d_{1},d_{2}\in{\mathfrak{D}}, we define:

𝖫(d1,d2)=max{𝖫(d1),𝖫(d2),1εd1d2𝔇}.{\mathsf{L}}^{\prime}(d_{1},d_{2})=\max\left\{{\mathsf{L}}(d_{1}),{\mathsf{L}}(d_{2}),\frac{1}{\varepsilon}\left\|{d_{1}-d_{2}}\right\|_{{\mathfrak{D}}}\right\}\text{.}

If 𝖫(d1,d2)=0{\mathsf{L}}^{\prime}(d_{1},d_{2})=0 for some d1,d2𝔇d_{1},d_{2}\in{\mathfrak{D}}, then d1=d2d_{1}=d_{2}, 𝖫(d1)=𝖫(d2)=0{\mathsf{L}}(d_{1})={\mathsf{L}}(d_{2})=0 and thus d1=d21𝔇d_{1}=d_{2}\in{\mathds{R}}1_{\mathfrak{D}}.

Let φ𝒮(𝔇)\varphi\in{\mathscr{S}}({\mathfrak{D}}). By construction:

(5.1) {(d1,d2)𝔇𝔇:𝖫(d1,d2)1,φ(d1)=0}{d𝔇:𝖫(d)1,φ(d)=0}×{d𝔇:𝖫(d)1,|φ(d)|ε}.\left\{(d_{1},d_{2})\in{\mathfrak{D}}\oplus{\mathfrak{D}}:{\mathsf{L}}^{\prime}(d_{1},d_{2})\leqslant 1,\varphi(d_{1})=0\right\}\\ \subseteq\left\{d\in{\mathfrak{D}}:{\mathsf{L}}(d)\leqslant 1,\varphi(d)=0\right\}\times\left\{d\in{\mathfrak{D}}:{\mathsf{L}}(d)\leqslant 1,|\varphi(d)|\leqslant\varepsilon\right\}\text{.}

Both factors in the Cartesian product on the right hand-side of Expression (5.1) are compact since 𝖫{\mathsf{L}} is a Lip-norm, so the left hand side is a subset of a compact set in 𝔇𝔇{\mathfrak{D}}\oplus{\mathfrak{D}}. Since 𝖫{\mathsf{L}}^{\prime}, as the maximum of lower semi-continuous functions, is lower-semicontinuous (and since φ\varphi is continuous), the set:

{(d1,d2)𝔇𝔇:𝖫(d1,d2)1,φ(d1)=0}\left\{(d_{1},d_{2})\in{\mathfrak{D}}\oplus{\mathfrak{D}}:{\mathsf{L}}^{\prime}(d_{1},d_{2})\leqslant 1,\varphi(d_{1})=0\right\}

is closed, and thus it is compact as well.

We thus have shown that 𝖫{\mathsf{L}}^{\prime} is a Lip-norm using [25, Proposition 1.3].

For all (a,b)𝔄𝔅(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}}, we set:

𝖫′′(a,b)=inf{𝖫(d,d):π𝔄(d)=a,π𝔅(d)=b}.{\mathsf{L}}^{\prime\prime}(a,b)=\inf\left\{{\mathsf{L}}^{\prime}(d,d^{\prime}):\pi_{\mathfrak{A}}(d)=a,\pi_{\mathfrak{B}}(d^{\prime})=b\right\}\text{.}

By [31], the seminorm 𝖫′′{\mathsf{L}}^{\prime\prime} — the quotient of 𝖫{\mathsf{L}}^{\prime} for the map (d,d)𝔇𝔇(π𝔄(d),π𝔅(d))𝔄𝔅(d,d^{\prime})\in{\mathfrak{D}}\oplus{\mathfrak{D}}\mapsto(\pi_{\mathfrak{A}}(d),\pi_{\mathfrak{B}}(d^{\prime}))\in{\mathfrak{A}}\oplus{\mathfrak{B}} — is a Lip-norm on 𝔄𝔅{\mathfrak{A}}\oplus{\mathfrak{B}}.

Now, let a𝔄a\in{\mathfrak{A}} with 𝖫𝔄(a)1{\mathsf{L}}_{\mathfrak{A}}(a)\leqslant 1. Since τ\tau is a tunnel, there exists d𝔇d\in{\mathfrak{D}} with 𝖫(d)1{\mathsf{L}}(d)\leqslant 1 and π𝔄(d)=a\pi_{\mathfrak{A}}(d)=a. Let b=π𝔅(d)b=\pi_{\mathfrak{B}}(d). As π𝔅\pi_{\mathfrak{B}} is 1-Lipschitz, we have 𝖫𝔅(b)1{\mathsf{L}}_{\mathfrak{B}}(b)\leqslant 1.

Thus the canonical surjection (a,b)𝔄𝔅a𝔄(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}}\mapsto a\in{\mathfrak{A}} is a quantum isometry from (𝔄𝔅,𝖫′′)({\mathfrak{A}}\oplus{\mathfrak{B}},{\mathsf{L}}^{\prime\prime}) onto (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}). Similarly, (a,b)𝔄𝔅b𝔅(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}}\mapsto b\in{\mathfrak{B}} is also a quantum isometry.

Let now φ𝒮(𝔄)\varphi\in{\mathscr{S}}({\mathfrak{A}}). As τ\tau is a tunnel, there exists ψ𝒮(𝔅)\psi\in{\mathscr{S}}({\mathfrak{B}}) such that 𝗆𝗄𝖫(φ,ψ)χ(τ){\mathsf{mk}_{{\mathsf{L}}}}(\varphi,\psi)\leqslant{\chi\left({\tau}\right)}. Let (a,b)𝔄𝔅(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}} with 𝖫′′(a,b)1{\mathsf{L}}^{\prime\prime}(a,b)\leqslant 1. By definition of 𝖫′′{\mathsf{L}}^{\prime\prime}, there exists d1,d2𝔇d_{1},d_{2}\in{\mathfrak{D}} such that π𝔄(d1)=a\pi_{\mathfrak{A}}(d_{1})=a and π𝔅(d2)=b\pi_{\mathfrak{B}}(d_{2})=b, with 𝖫(d1,d2)1{\mathsf{L}}^{\prime}(d_{1},d_{2})\leqslant 1. We then estimate:

|φ(a)ψ(b)|\displaystyle|\varphi(a)-\psi(b)| |φ(π𝔄(d1))ψ(π𝔅(d2))|\displaystyle\leqslant|\varphi(\pi_{\mathfrak{A}}(d_{1}))-\psi(\pi_{\mathfrak{B}}(d_{2}))|
|φ(π𝔄(d1))ψ(π𝔅(d1))|+|ψ(π𝔅(d1))ψ(π𝔅(d2))|\displaystyle\leqslant|\varphi(\pi_{\mathfrak{A}}(d_{1}))-\psi(\pi_{\mathfrak{B}}(d_{1}))|+|\psi(\pi_{\mathfrak{B}}(d_{1}))-\psi(\pi_{\mathfrak{B}}(d_{2}))|
𝗆𝗄𝖫(φπ𝔄,ψπ𝔅)+|ψ(π𝔅(d1))ψ(π𝔅(d2))|\displaystyle\leqslant{\mathsf{mk}_{{\mathsf{L}}}}(\varphi\circ\pi_{\mathfrak{A}},\psi\circ\pi_{\mathfrak{B}})+|\psi(\pi_{\mathfrak{B}}(d_{1}))-\psi(\pi_{\mathfrak{B}}(d_{2}))|
χ(τ)+d1d2𝔇\displaystyle\leqslant{\chi\left({\tau}\right)}+\left\|{d_{1}-d_{2}}\right\|_{{\mathfrak{D}}}
χ(τ)+ε.\displaystyle\leqslant{\chi\left({\tau}\right)}+\varepsilon\text{.}

Consequently, 𝗆𝗄𝖫′′(φ,ψ)χ(τ)+ε{\mathsf{mk}_{{\mathsf{L}}^{\prime\prime}}}(\varphi,\psi)\leqslant{\chi\left({\tau}\right)}+\varepsilon. By symmetry, we conclude:

𝖧𝖺𝗎𝗌𝗆𝗄𝖫′′(𝒮(𝔄),𝒮(𝔅))χ(τ)+ε.{\mathsf{Haus}_{{\mathsf{mk}_{{\mathsf{L}}^{\prime\prime}}}}}({\mathscr{S}}({\mathfrak{A}}),{\mathscr{S}}({\mathfrak{B}}))\leqslant{\chi\left({\tau}\right)}+\varepsilon\text{.}

Thus, by [31], we conclude:

𝖽𝗂𝗌𝗍q((𝔄,𝖫𝔄),(𝔅,𝖫𝔅))χ(τ)+ε{\mathsf{dist}}_{q}(({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}),({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}))\leqslant{\chi\left({\tau}\right)}+\varepsilon

and since ε>0\varepsilon>0 is arbitrary, we conclude 𝖽𝗂𝗌𝗍q((𝔄,𝖫𝔄),(𝔅,𝖫𝔅))χ(τ){\mathsf{dist}}_{q}(({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}),({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}))\leqslant{\chi\left({\tau}\right)}.

Since τ\tau is an arbitrary tunnel between (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}), we conclude:

𝖽𝗂𝗌𝗍q((𝔄,𝖫𝔄),(𝔅,𝖫𝔅))inf{χ(τ): τ is an order unit tunnel from (𝔄,𝖫𝔄) to (𝔅,𝖫𝔅)}.{\mathsf{dist}}_{q}(({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}),({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}))\leqslant\inf\{{\chi\left({\tau}\right)}:\\ \text{ $\tau$ is an order unit tunnel from $({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}})$ to $({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}})$}\}\text{.}

On the other hand, if 𝖫{\mathsf{L}}^{\prime} is an admissible Lip-norm on 𝔄𝔅{\mathfrak{A}}\oplus{\mathfrak{B}} then (𝔄𝔅,𝖫,π𝔄,π𝔅)({\mathfrak{A}}\oplus{\mathfrak{B}},{\mathsf{L}}^{\prime},\pi_{\mathfrak{A}},\pi_{\mathfrak{B}}), with π𝔄:(a,b)a\pi_{\mathfrak{A}}:(a,b)\mapsto a and π𝔅:(a,b)b\pi_{\mathfrak{B}}:(a,b)\mapsto b, is a tunnel from (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) to (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}). By Lemma (5.4), we then have: 𝖧𝖺𝗎𝗌𝗆𝗄𝖫(𝒮(𝔄),𝒮(𝔅))=χ(τ){\mathsf{Haus}_{{\mathsf{mk}_{{\mathsf{L}}^{\prime}}}}}({\mathscr{S}}({\mathfrak{A}}),{\mathscr{S}}({\mathfrak{B}}))={\chi\left({\tau}\right)}.

This completes our theorem. ∎

Remarkably if π:(𝔄,𝖫𝔄)(𝔅,𝖫𝔅)\pi:({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}})\rightarrow({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}) is a quantum isometry between two quantum metric order unit spaces (𝔄,𝖫𝔄)({\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}) and (𝔅,𝖫𝔅)({\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}), it need not be a quotient map, in the following sense:

Definition 5.6.

A surjection π:𝔄𝔅\pi:{\mathfrak{A}}\twoheadrightarrow{\mathfrak{B}} between two normed vector spaces 𝔄{\mathfrak{A}} and 𝔅{\mathfrak{B}} is a quotient map when:

b𝔅b𝔅=inf{a𝔄:π(a)=b}.\forall b\in{\mathfrak{B}}\quad\left\|{b}\right\|_{{\mathfrak{B}}}=\inf\left\{\left\|{a}\right\|_{{\mathfrak{A}}}:\pi(a)=b\right\}\text{.}

Thus, it may be natural to require that quantum isometries are also quotient maps. In [31], this matter is noted but seems inconsequential. In particular, we note that if 𝔄1{\mathfrak{A}}_{1} and 𝔄2{\mathfrak{A}}_{2} are two order unit spaces, then the maps (a1,a2)𝔄1𝔄2aj(a_{1},a_{2})\in{\mathfrak{A}}_{1}\oplus{\mathfrak{A}}_{2}\mapsto a_{j}, for j=1,2j=1,2, are in fact quotient maps, and also that an order-isomorphism between order unit maps is automatically a quotient map. Thus, the quantum isometries which play any role in [31], including in the definition of 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}, are all already quotient maps. This issue also does not arise in [16] and subsequent work since *-epimorphisms are always quotient maps as well. However, in general, Definition (5.2) allows for tunnels constructed out of maps which may not be quotient maps. But it is immediate that, if we restrict ourselves to tunnels constructed with quantum isometries which are also quotient maps, then Theorem (5.5) still holds, as the only point of note is that Lemma (5.4) involves quantum isometries which are quotient maps (the rest of the argument follows unchanged).

Theorem (5.5) suggests other techniques used in the theory of the Gromov-Hausdorff propinquity may be applied to Rieffel’s distance. We will indeed follow this idea and provide an alternate proof of completeness for Rieffel’s distance along the lines of the propinquity’s proof of completeness. The reason for doing so is that the construction in the propinquity’s proof are well-behaved with respect to C*-algebras, which will be helpful in our current context. Indeed, the limit is not just defined as an order unit space of continuous affine functions over some compact convex sets, but as a quotient of an order unit space.

However, it is important to stress that Theorem (5.5) does not state that the propinquity is the restriction of 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} to C*-algebra-based quantum compact metric spaces. In fact, the proof of Theorem (5.5) involves taking a quotient of a Lip-norm, which would create difficulties when working with quasi-Leibniz seminorms, which is part of the basic framework of the propinquity. In fact, 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} and the propinquity have different coincidence properties and are different metrics, even when restricted to C*-algebras with (quasi)-Leibniz Lip-norms. Instead, Theorem (5.5) states that the efforts put in deriving new techniques for the propinquity are indeed worthwhile, since removing the constraints of working only with quasi-Leibniz Lip-norms on C*-algebras (including those involved in the definition of tunnels!) simply lead us back to Rieffel’s distance.

In any case, in order to use the results of [1], we turn to the interesting exercise to adapt the proof of completeness of the propinquity from [16] to 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}. This comes with some interesting subtleties. We proceed our result with a standard definition and a well-known description of order unit spaces.

Notation 5.7.

If ZAZ\subseteq A is a closed convex subset of a topological {\mathds{R}}-vector space AA, then the vector space of all the continuous affine functions from ZZ to {\mathds{R}} is denoted by 𝒜(Z)\mathcal{AF}(Z), where φ:Z\varphi:Z\rightarrow{\mathds{R}} is affine when for all a,bZa,b\in Z and t[0,1]t\in[0,1], we have φ(ta+(1t)b)=tφ(a)+(1t)φ(b)\varphi(ta+(1-t)b)=t\varphi(a)+(1-t)\varphi(b).

We recall a classical result due to Kadison that provides a functional representation of complete order unit spaces.

Theorem 5.8 ([4, Theorem II.1.8]).

If 𝔄{\mathfrak{A}} is a complete order unit space, and if for all a𝔄a\in{\mathfrak{A}} we set a^:φ𝒮(𝔄)φ(a)\widehat{a}:\varphi\in{\mathscr{S}}({\mathfrak{A}})\mapsto\varphi(a), then the map a𝔄𝒜(𝒮(𝔄))a\in{\mathfrak{A}}\mapsto\mathcal{AF}({\mathscr{S}}({\mathfrak{A}})) is an order isomorphism.

Theorem 5.9.

Let (𝔄n,𝖫n)n({\mathfrak{A}}_{n},{\mathsf{L}}_{n})_{n\in{\mathds{N}}} be a sequence of quantum metric order unit spaces such that for all nn\in{\mathds{N}}, there exists an order unit tunnel τn=(𝔇n,𝖫n,πn,ρn)\tau_{n}=({\mathfrak{D}}_{n},{\mathsf{L}}^{n},\pi_{n},\rho_{n}) from (𝔄n,𝖫n)({\mathfrak{A}}_{n},{\mathsf{L}}_{n}) to (𝔄n+1,𝖫n+1)({\mathfrak{A}}_{n+1},{\mathsf{L}}_{n+1}) with n=0χ(τn)<\sum_{n=0}^{\infty}{\chi\left({\tau_{n}}\right)}<\infty.

Let:

𝔅={(dn)nn𝔇n:supndn𝔇n<}{\mathfrak{B}}=\left\{(d_{n})_{n\in{\mathds{N}}}\in\prod_{n\in{\mathds{N}}}{\mathfrak{D}}_{n}:\sup_{n\in{\mathds{N}}}\left\|{d_{n}}\right\|_{{\mathfrak{D}}_{n}}<\infty\right\}

endowed with (dn)n𝔅=supndn𝔇n\left\|{(d_{n})_{n\in{\mathds{N}}}}\right\|_{{\mathfrak{B}}}=\sup_{n\in{\mathds{N}}}\left\|{d_{n}}\right\|_{{\mathfrak{D}}_{n}} for all (dn)n𝔅(d_{n})_{n\in{\mathds{N}}}\in{\mathfrak{B}}.

Now, let:

𝔎={(dn)n𝔅:nρn(dn)=πn+1(dn+1)}{\mathfrak{K}}=\left\{(d_{n})_{n\in{\mathds{N}}}\in{\mathfrak{B}}:\forall n\in{\mathds{N}}\quad\rho_{n}(d_{n})=\pi_{n+1}(d_{n+1})\right\}

and

𝔏={(dn)n𝔎:supn𝖫n(dn)<}.{\mathfrak{L}}=\{(d_{n})_{n\in{\mathds{N}}}\in{\mathfrak{K}}:\sup_{n\in{\mathds{N}}}{\mathsf{L}}^{n}(d_{n})<\infty\}.

Let 𝔈{\mathfrak{E}} be the closure of 𝔏{\mathfrak{L}} for 𝔅\left\|{\cdot}\right\|_{{\mathfrak{B}}}.

Let 𝔍={(dn)n𝔈:limndn𝔇n=0}{\mathfrak{J}}=\left\{(d_{n})_{n\in{\mathds{N}}}\in{\mathfrak{E}}:\lim_{n\rightarrow\infty}\left\|{d_{n}}\right\|_{{\mathfrak{D}}_{n}}=0\right\}.

The space 𝔈{\mathfrak{E}} is a complete order unit space, and 𝔍{\mathfrak{J}} is an order ideal of 𝔈{\mathfrak{E}}. There exists a weak* compact convex subset ZZ of 𝒮(𝔈){\mathscr{S}}({\mathfrak{E}}) such that 𝔍=Z{\mathfrak{J}}=Z^{\perp} and limn𝖧𝖺𝗎𝗌𝗐(Z,𝒮(𝔇n))=0\lim_{n\rightarrow\infty}{\mathsf{Haus}_{\mathsf{w}}}(Z,{\mathscr{S}}({\mathfrak{D}}_{n}))=0 where 𝗐\mathsf{w} is any metric which induces the weak* topology on the state space 𝒮(𝔈){\mathscr{S}}({\mathfrak{E}}) of 𝔈{\mathfrak{E}}.

If 𝔉=𝔈/𝔍{\mathfrak{F}}={\raisebox{1.99997pt}{${\mathfrak{E}}$}\left/\raisebox{-1.99997pt}{${\mathfrak{J}}$}\right.}, then 𝔉{\mathfrak{F}}, endowed with the quotient order, is order-isomorphic to the order unit space of 𝒜(Z)\mathcal{AF}(Z).

If, moreover, ρn\rho_{n} is a quotient map for all nn\in{\mathds{N}}, then the norm induced by the quotient order on 𝔉{\mathfrak{F}} and the quotient norm are equal.

Remark 5.10.

We emphasize that 𝔉{\mathfrak{F}} has two possible orders: its quotient order and the order from its structure as a space of affine functions. This already endows 𝔉{\mathfrak{F}} with two potentially distinct seminorms. Moreover, 𝔉{\mathfrak{F}} has a norm from being the quotient of a normed vector space by a closed subspace. A priori, this norm is different from the two other order seminorms. Our result reconciles these structures under appropriate hypothesis.

Proof.

It is easy to check that 𝔅{\mathfrak{B}} and 𝔈{\mathfrak{E}} are complete order unit spaces with order unit 1𝔈=(1n)n1_{{\mathfrak{E}}}=(1_{n})_{n\in{\mathds{N}}} (note that 𝔈{\mathfrak{E}} is closed in 𝔅{\mathfrak{B}} by continuity if the maps πn\pi_{n} and ρn\rho_{n} for all nn\in{\mathds{N}}).

Now, 𝔍{\mathfrak{J}} is a closed subspace of 𝔈{\mathfrak{E}}, so 𝔉=𝔈/𝔍{\mathfrak{F}}={\raisebox{1.99997pt}{${\mathfrak{E}}$}\left/\raisebox{-1.99997pt}{${\mathfrak{J}}$}\right.} is a normed vector space with norm 𝔉\left\|{\cdot}\right\|_{{\mathfrak{F}}}. Let q:𝔈𝔉q:{\mathfrak{E}}\twoheadrightarrow{\mathfrak{F}} be the canonical surjection. Moreover, if 0(dn)n(jn)n0\leqslant(d_{n})_{n\in{\mathds{N}}}\leqslant(j_{n})_{n\in{\mathds{N}}} with (jn)n𝔍(j_{n})_{n\in{\mathds{N}}}\in{\mathfrak{J}}, then for all nn\in{\mathds{N}}, we have 0dnjn0\leqslant d_{n}\leqslant j_{n} and thus dn𝔇njn𝔇n\left\|{d_{n}}\right\|_{{\mathfrak{D}}_{n}}\leqslant\left\|{j_{n}}\right\|_{{\mathfrak{D}}_{n}}, from which we readily conclude that (jn)n𝔍(j_{n})_{n\in{\mathds{N}}}\in{\mathfrak{J}}. Thus 𝔍{\mathfrak{J}} is a order ideal in 𝔈{\mathfrak{E}}. Consequently, 𝔉{\mathfrak{F}} is also an order vector space (with the quotient order) with an order unit by [26] — though this order unit is not necessarily Archimedean. Nonetheless, qq is a positive linear map which maps the order unit of 𝔈{\mathfrak{E}} to the order unit of 𝔉{\mathfrak{F}}. We denote the quotient order on 𝔉{\mathfrak{F}} simply as \leqslant, and its order unit as e=q(1𝔈)e=q(1_{{\mathfrak{E}}}). There is also a seminorm 𝔉,\left\|{\cdot}\right\|_{{\mathfrak{F}},\leqslant} induced on 𝔉{\mathfrak{F}} by the ordered vector space with an order unit structure on 𝔉{\mathfrak{F}}.

Any φ𝒮(𝔇n)\varphi\in{\mathscr{S}}({\mathfrak{D}}_{n}) for any nn\in{\mathds{N}} defines a state of 𝔈{\mathfrak{E}} by setting (dk)k𝔈φ(dn)(d_{k})_{k\in{\mathds{N}}}\in{\mathfrak{E}}\mapsto\varphi(d_{n}), and we will henceforth identify 𝒮(𝔇n){\mathscr{S}}({\mathfrak{D}}_{n}) with its image for this map.

Now, by [16], the sequence (𝒮(𝔇n))n({\mathscr{S}}({\mathfrak{D}}_{n}))_{n\in{\mathds{N}}} converges, for the Hausdorff distance induced by any metric 𝗐\mathsf{w} for the weak* topology on 𝒮(𝔈){\mathscr{S}}({\mathfrak{E}}) (which exists since 𝔈{\mathfrak{E}} is separable), to a weak* compact and convex set Z𝒮(𝔈)Z\subseteq{\mathscr{S}}({\mathfrak{E}}), and moreover:

𝔍={d𝔈:φZφ(d)=0}=Z.{\mathfrak{J}}=\left\{d\in{\mathfrak{E}}:\forall\varphi\in Z\quad\varphi(d)=0\right\}=Z^{\perp}\text{.}

We note in passing that the metric used in [16] to define ZZ is the Monge-Kantorovich metric induced by a Lip-norm on 𝔈{\mathfrak{E}}, but actually, the topologies induced on 𝒮(𝔈){\mathscr{S}}({\mathfrak{E}}) by the Hausdorff distance for any metrization 𝗐\mathsf{w} of the weak* topology all agree with the Vietoris topology on the weak* closed subsets of 𝒮(𝔈){\mathscr{S}}({\mathfrak{E}}), so the exact metric involved is not important for this part of the proof.

If φ𝒮(𝔉)\varphi\in{\mathscr{S}}({\mathfrak{F}}), then φq𝒮(𝔈)\varphi\circ q\in{\mathscr{S}}({\mathfrak{E}}) and by construction, φq(𝔍)={0}\varphi\circ q({\mathfrak{J}})=\{0\}, thus φqZ\varphi\circ q\in Z. Conversely, if φZ\varphi\in Z then φ\varphi induces a state on 𝔉{\mathfrak{F}} since Z=𝔍Z^{\perp}={\mathfrak{J}}. These two maps are inverse to each other and allow us to identify ZZ with 𝒮(𝔉){\mathscr{S}}({\mathfrak{F}}).

For any f𝔉f\in{\mathfrak{F}}, and for any φZ\varphi\in Z, we set f^(φ)=φ(f)\widehat{f}(\varphi)=\varphi(f). The map f𝔉f^f\in{\mathfrak{F}}\mapsto\widehat{f} is linear and maps ee to the order unit of 𝒜(Z)\mathcal{AF}(Z), i.e. the constant function 11. This map is injective: if f^=0\widehat{f}=0, then, for any d𝔈d\in{\mathfrak{E}} with d+𝔍=fd+{\mathfrak{J}}=f, we have dZ=𝔍d\in Z^{\perp}={\mathfrak{J}} and thus f=0f=0. Let F={f^:f𝔉}F=\left\{\widehat{f}:f\in{\mathfrak{F}}\right\}.

We now prove that this injective linear map is positive.

Let f𝔉f\in{\mathfrak{F}} such that f0f\geqslant 0. There exists d𝔈d\in{\mathfrak{E}} and b𝔍b\in{\mathfrak{J}} such that dbd\geqslant b and d+𝔍=fd+{\mathfrak{J}}=f. Let φZ\varphi\in Z. Then φ(d)φ(b)=0\varphi(d)\geqslant\varphi(b)=0 and thus, for all φZ\varphi\in Z, we have φ(f)0\varphi(f)\geqslant 0. So f^0\widehat{f}\geqslant 0.

We now check that the inverse map of f𝔉f^f\in{\mathfrak{F}}\mapsto\widehat{f} is also a positive map from FF onto 𝔉{\mathfrak{F}}.

Let f𝔉f\in{\mathfrak{F}} such that f^0\widehat{f}\geqslant 0, i.e. for all φZ\varphi\in Z, we have φ(f)0\varphi(f)\geqslant 0. Let d𝔈d\in{\mathfrak{E}} such that d+𝔍=fd+{\mathfrak{J}}=f. Now, suppose that for some ε>0\varepsilon>0, for all NN\in{\mathds{N}}, there exists nNn\geqslant N such that dnε1nd_{n}\leqslant-\varepsilon 1_{n}.

By induction, there exists a strictly increasing function θ:\theta:{\mathds{N}}\rightarrow{\mathds{N}} and, for each nn\in{\mathds{N}}, there exists φn𝒮(𝔇θ(n))\varphi_{n}\in{\mathscr{S}}({\mathfrak{D}}_{\theta(n)}) such that φn(dθ(n))ε\varphi_{n}(d_{\theta(n)})\leqslant-\varepsilon. By compactness of 𝒮(𝔈){\mathscr{S}}({\mathfrak{E}}), there exists a weak* limit ψ𝒮(𝔈)\psi\in{\mathscr{S}}({\mathfrak{E}}) for a subsequence of (φn)n(\varphi_{n})_{n\in{\mathds{N}}}. Now, by construction, ψZ\psi\in Z since ZZ is the Hausdorff limit of 𝒮(𝔇θ(n)){\mathscr{S}}({\mathfrak{D}}_{\theta(n)}). Now, ψ(d)=limnφn(d)ε\psi(d)=\lim_{n\rightarrow\infty}\varphi_{n}(d)\leqslant-\varepsilon. By assumption, ψ(d)=ψ(f)0\psi(d)=\psi(f)\geqslant 0. This is a contradiction. Hence, for all ε>0\varepsilon>0, there exists NN\in{\mathds{N}} such that if nNn\geqslant N then dnε1nd_{n}\geqslant-\varepsilon 1_{n}. Consequently, by an easy induction, we can find a sequence (εn)n(\varepsilon_{n})_{n\in{\mathds{N}}} converging to 0 and such that for all nn\in{\mathds{N}}, we have dnεn1nd_{n}\geqslant-\varepsilon_{n}1_{n}. Set b=(εn1n)nb=(\varepsilon_{n}1_{n})_{n\in{\mathds{N}}}: by construction, bn𝔍b_{n}\in{\mathfrak{J}}, and d+b0d+b\geqslant 0. Therefore f0f\geqslant 0.

Consequently, 𝔉{\mathfrak{F}} is order-isomorphic to FF. In turn, this proves that 𝔉{\mathfrak{F}} is an order unit space, with state space ZZ, and as 𝔉{\mathfrak{F}} is complete, the map f𝔉f^𝒜(Z)f\in{\mathfrak{F}}\mapsto\widehat{f}\in\mathcal{AF}(Z) is onto as well. The quotient order seminorm 𝔉,\left\|{\cdot}\right\|_{{\mathfrak{F}},\leqslant} on 𝔉{\mathfrak{F}} is given by supφZ|φ()|\sup_{\varphi\in Z}|\varphi(\cdot)| — and it is a norm.

We now turn to the relationship between the order norm 𝔉,\left\|{\cdot}\right\|_{{\mathfrak{F}},\leqslant} and the quotient norm 𝔉\left\|{\cdot}\right\|_{{\mathfrak{F}}}.

First, let f𝔉f\in{\mathfrak{F}}. Let ε>0\varepsilon>0. There exists d𝔈d\in{\mathfrak{E}} such that d+𝔍=fd+{\mathfrak{J}}=f and f𝔉d𝔈f𝔉+ε\left\|{f}\right\|_{{\mathfrak{F}}}\leqslant\left\|{d}\right\|_{{\mathfrak{E}}}\leqslant\left\|{f}\right\|_{{\mathfrak{F}}}+\varepsilon. In particular, d𝔈1𝔈dd𝔈1𝔈-\left\|{d}\right\|_{{\mathfrak{E}}}1_{{\mathfrak{E}}}\leqslant d\leqslant\left\|{d}\right\|_{{\mathfrak{E}}}1_{{\mathfrak{E}}} and thus d𝔈efd𝔈e-\left\|{d}\right\|_{{\mathfrak{E}}}e\leqslant f\leqslant\left\|{d}\right\|_{{\mathfrak{E}}}e. We conclude that f𝔉,f𝔉+ε\left\|{f}\right\|_{{\mathfrak{F}},\leqslant}\leqslant\left\|{f}\right\|_{{\mathfrak{F}}}+\varepsilon, and thus 𝔉,𝔉\left\|{\cdot}\right\|_{{\mathfrak{F}},\leqslant}\leqslant\left\|{\cdot}\right\|_{{\mathfrak{F}}} as ε>0\varepsilon>0 is arbitrary.

In general, we do not have much more to say about these two norms. However, if, for all nn\in{\mathds{N}}, the map ρn\rho_{n} is a quotient map — as is the case, for instance, when working with C*-algebras — then more can be concluded. Assume henceforth that ρn\rho_{n} is a quotient map for all nn\in{\mathds{N}}.

Let f𝔉f\in{\mathfrak{F}}. Let ε>0\varepsilon>0. Let d𝔈d\in{\mathfrak{{\mathfrak{E}}}} such that d+𝔍=fd+{\mathfrak{J}}=f and d𝔈f𝔉+ε\left\|{d}\right\|_{{\mathfrak{E}}}\leqslant\left\|{f}\right\|_{{\mathfrak{F}}}+\varepsilon. Note that f𝔉d𝔈\left\|{f}\right\|_{{\mathfrak{F}}}\leqslant\left\|{d}\right\|_{{\mathfrak{E}}}. We now make an observation.

Let NN\in{\mathds{N}}. Since πN(dN)𝔄NdN𝔇N\left\|{\pi_{N}(d_{N})}\right\|_{{\mathfrak{A}}_{N}}\leqslant\left\|{d_{N}}\right\|_{{\mathfrak{D}}_{N}}, and since ρN\rho_{N} is a quotient map, there exists dN1𝔇N1d^{\prime}_{N-1}\in{\mathfrak{D}}_{N-1} such that dN1𝔇N1πN(dN)𝔄N+εNdN𝔇N+εN\left\|{d^{\prime}_{N-1}}\right\|_{{\mathfrak{D}}_{N-1}}\leqslant\left\|{\pi_{N}(d_{N})}\right\|_{{\mathfrak{A}}_{N}}+\frac{\varepsilon}{N}\leqslant\left\|{d_{N}}\right\|_{{\mathfrak{D}}_{N}}+\frac{\varepsilon}{N} and ρN1(dN1)=πN(dN)\rho_{N-1}(d^{\prime}_{N-1})=\pi_{N}(d_{N}). Now, repeating this process, a simple induction show that we can find d0𝔇0,,dN1𝔇N1d_{0}^{\prime}\in{\mathfrak{D}}_{0},\ldots,d_{N-1}^{\prime}\in{\mathfrak{D}}_{N-1} such that dj𝔇jdN𝔇N+jεNdN𝔇N+ε\left\|{d^{\prime}_{j}}\right\|_{{\mathfrak{D}}_{j}}\leqslant\left\|{d_{N}}\right\|_{{\mathfrak{D}}_{N}}+\frac{j\varepsilon}{N}\leqslant\left\|{d_{N}}\right\|_{{\mathfrak{D}}_{N}}+\varepsilon and d=(d0,,dN1,dN,dN+1,)𝔈d^{\prime}=(d^{\prime}_{0},\ldots,d^{\prime}_{N-1},d_{N},d_{N+1},\ldots)\in{\mathfrak{E}}. By construction, dd𝔍d-d^{\prime}\in{\mathfrak{J}}, d+𝔍=fd^{\prime}+{\mathfrak{J}}=f and, in particular, f𝔉d𝔈\left\|{f}\right\|_{{\mathfrak{F}}}\leqslant\left\|{d^{\prime}}\right\|_{{\mathfrak{E}}}.

We note in passing that, by construction:

d𝔈\displaystyle\left\|{d^{\prime}}\right\|_{{\mathfrak{E}}} =supndn𝔇n\displaystyle=\sup_{n\in{\mathds{N}}}\left\|{d^{\prime}_{n}}\right\|_{{\mathfrak{D}}_{n}}
=max{maxj=0Ndj𝔇j,supj>Ndj𝔇j}\displaystyle=\max\left\{\max_{j=0}^{N}\left\|{d^{\prime}_{j}}\right\|_{{\mathfrak{D}}_{j}},\sup_{j>N}\left\|{d_{j}}\right\|_{{\mathfrak{D}}_{j}}\right\}
max{d𝔈+ϵ,d𝔈}\displaystyle\leqslant\max\left\{\left\|{d}\right\|_{{\mathfrak{E}}}+\epsilon,\left\|{d}\right\|_{{\mathfrak{E}}}\right\}
d𝔈+ϵf𝔉+2ϵ\displaystyle\leqslant\left\|{d}\right\|_{{\mathfrak{E}}}+\epsilon\leqslant\left\|{f}\right\|_{{\mathfrak{F}}}+2\epsilon

so we actually have f𝔉d𝔈f𝔉+2ε\left\|{f}\right\|_{{\mathfrak{F}}}\leqslant\left\|{d^{\prime}}\right\|_{{\mathfrak{E}}}\leqslant\left\|{f}\right\|_{{\mathfrak{F}}}+2\varepsilon — though, for our proof, only the lower bound on the norm of dd^{\prime} matters.

Thus, starting from f𝔉d𝔈\left\|{f}\right\|_{{\mathfrak{F}}}\leqslant\left\|{d^{\prime}}\right\|_{{\mathfrak{E}}}, and by definition of 𝔈\left\|{\cdot}\right\|_{{\mathfrak{E}}}, there exists nn\in{\mathds{N}} such that f𝔉εdn𝔈f𝔉+2ε\left\|{f}\right\|_{{\mathfrak{F}}}-\varepsilon\leqslant\left\|{d^{\prime}_{n}}\right\|_{{\mathfrak{E}}}\leqslant\left\|{f}\right\|_{{\mathfrak{F}}}+2\varepsilon. If nNn\leqslant N then dn𝔈dN𝔇N+ϵ\left\|{d^{\prime}_{n}}\right\|_{{\mathfrak{E}}}\leqslant\left\|{d_{N}}\right\|_{{\mathfrak{D}}_{N}}+\epsilon, and thus dN𝔇𝔑f𝔉2ε\left\|{d_{N}}\right\|_{{\mathfrak{D_{N}}}}\geqslant\left\|{f}\right\|_{{\mathfrak{F}}}-2\varepsilon. Thus, we have shown that for all NN\in{\mathds{N}}, there exists nNn\geqslant N such that:

f𝔉2εdn𝔇n+f𝔉+ε.\left\|{f}\right\|_{{\mathfrak{F}}}-2\varepsilon\leqslant\left\|{d_{n}}\right\|_{{\mathfrak{D}}_{n}}+\leqslant\left\|{f}\right\|_{{\mathfrak{F}}}+\varepsilon\text{.}

Thus we can find a sequence of states φn\varphi_{n} and a strictly increasing function g:g:{\mathds{N}}\rightarrow{\mathds{N}} such that φn𝒮(𝔇g(n))\varphi_{n}\in{\mathscr{S}}({\mathfrak{D}}_{g(n)}) for all nn\in{\mathds{N}}, and |φn(dg(n))|[f𝔈2ϵ,f𝔈+ε]\left|\varphi_{n}(d_{g(n)})\right|\in[\left\|{f}\right\|_{{\mathfrak{E}}}-2\epsilon,\left\|{f}\right\|_{{\mathfrak{E}}}+\varepsilon]. By compactness, the sequence (φn)n(\varphi_{n})_{n\in{\mathds{N}}} admits a weak* convergent subsequence, whose limit is denoted by ψ\psi. By definition of ZZ, we then have ψZ\psi\in Z, and by construction, |ψ(f)|[f𝔈2ε,f𝔈+ε]|\psi(f)|\in[\left\|{f}\right\|_{{\mathfrak{E}}}-2\varepsilon,\left\|{f}\right\|_{{\mathfrak{E}}}+\varepsilon]. We conclude that:

f𝔈|ψ(f)|+2εsupφZ|φ(f)|+2ε=f𝔉,+2ε,\left\|{f}\right\|_{{\mathfrak{E}}}\leqslant|\psi(f)|+2\varepsilon\leqslant\sup_{\varphi\in Z}|\varphi(f)|+2\varepsilon=\left\|{f}\right\|_{{\mathfrak{F}},\leqslant}+2\varepsilon\text{,}

where the last equality follows from our proof that the order norm is indeed the order unit norm obtained from identifying 𝔉{\mathfrak{F}} with 𝒜(Z)\mathcal{AF}(Z).

As ε>0\varepsilon>0 is arbitrary, we conclude that f𝔉f𝔉,\left\|{f}\right\|_{{\mathfrak{F}}}\leqslant\left\|{f}\right\|_{{\mathfrak{F}},\leqslant}.

We thus have shown 𝔉=𝔉,\left\|{\cdot}\right\|_{{\mathfrak{F}}}=\left\|{\cdot}\right\|_{{\mathfrak{F}},\leqslant}. ∎

Corollary 5.11.

Using the hypothesis of Theorem (5.9), and setting for all f𝔉f\in{\mathfrak{F}}:

S(f)=inf{supn𝖫n(dn):q((dn)n)=f}S(f)=\inf\left\{\sup_{n\in{\mathds{N}}}{\mathsf{L}}^{n}(d_{n}):q((d_{n})_{n\in{\mathds{N}}})=f\right\}

then (𝔉,S)({\mathfrak{F}},S), when 𝔉{\mathfrak{F}} is endowed with the order norm 𝔉,\left\|{\cdot}\right\|_{{\mathfrak{F}},\leqslant}, is a quantum metric order unit space and:

limn𝖽𝗂𝗌𝗍q((𝔄n,𝖫n),(𝔉,S))=0.\lim_{n\rightarrow\infty}{\mathsf{dist}}_{q}(({\mathfrak{A}}_{n},{\mathsf{L}}_{n}),({\mathfrak{F}},S))=0\text{.}

If for all nn\in{\mathds{N}}, the surjections ρn\rho_{n} are quotient maps, then 𝔉=𝔉,\left\|{\cdot}\right\|_{{\mathfrak{F}}}=\left\|{\cdot}\right\|_{{\mathfrak{F}},\leqslant} and thus we can identify 𝔉{\mathfrak{F}} with 𝒜(Z)\mathcal{AF}(Z) with no ambiguity.

Proof.

This follows from Theorem (5.9) and appropriate choices of techniques in [16]. ∎

Any Cauchy sequence for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} admits a subsequence which meets the hypothesis of Theorem (5.9), and thus the limits of Cauchy sequences for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q} is described by Theorem (5.9).

6. Inductive limits of Order Unit spaces and compact quantum metric spaces

A method for placing a quantum metric on an inductive limit of C*-algebras was introduced in [1]. This method did not assume any quantum metric structure on the inductive limit, but rather assumed quantum metric structure on each term of the inductive sequence with some compatibility conditions between the quantum metrics of each consecutive term of the inductive sequence. However, this method relied heavily on quasi-Leibniz Lip-norms and the C*-algebra structure, which works well when one has such structure. In our main example of this paper, we have seen that we are not in this position in that our Lip-norms on our terms of our inductive sequence, while quasi-Leibniz, are not all satisfying some common Leibniz property. Therefore, our next goal is to translate the methods of [1] to the setting of order unit spaces with Lip-norms. The key to this lies in Theorem 5.9. A result such as Theorem 5.9 was automatically given by the C*-algebraic structure in [1]. First, we recall some definitions and known results from [19].

Definition 6.1 ([19, Definition 3.6, Lemma 3.4]).

Let 𝔄{\mathfrak{A}}, 𝔅{\mathfrak{B}} be two unital C*-algebras. A bridge γ\gamma from 𝔄{\mathfrak{A}} to 𝔅{\mathfrak{B}} is a 4-tuple γ=(𝔇,ω,π𝔄,π𝔅)\gamma=({\mathfrak{D}},\omega,\pi_{\mathfrak{A}},\pi_{\mathfrak{B}}) such that

  1. (1)

    𝔇{\mathfrak{D}} is a unital C*-algebra and ω𝔇\omega\in{\mathfrak{D}},

  2. (2)

    the set 𝒮1(ω)={ψ𝒮(𝔇):d𝔇,ψ(d)=ψ(ωd)=ψ(dω)}{\mathscr{S}}_{1}(\omega)=\{\psi\in{\mathscr{S}}({\mathfrak{D}}):\forall d\in{\mathfrak{D}},\psi(d)=\psi(\omega d)=\psi(d\omega)\} is non-empty, in which case ω\omega is called the pivot, and

  3. (3)

    π𝔄:𝔄𝔇\pi_{\mathfrak{A}}:{\mathfrak{A}}\rightarrow{\mathfrak{D}} and π𝔅:𝔅𝔇\pi_{\mathfrak{B}}:{\mathfrak{B}}\rightarrow{\mathfrak{D}} are unital *-monomorphisms.

The next lemma produces a characterization of lengths of the types of bridges that appear in this article, which follows immediately from definition. But, first, we introduce a definition for the types of bridges that appear in this article.

Definition 6.2.

Let 𝔄{\mathfrak{A}} be a unital C*-algebra, and let 𝔅𝔄{\mathfrak{B}}\subseteq{\mathfrak{A}} be a unital C*-subalgebra of 𝔄{\mathfrak{A}}. We call the 4-tuple (𝔄,1𝔄,ι,id𝔄)({\mathfrak{A}},1_{\mathfrak{A}},\iota,\mathrm{id}_{\mathfrak{A}}) the evident bridge from 𝔅{\mathfrak{B}} to 𝔄{\mathfrak{A}}, where ι:𝔅𝔄\iota:{\mathfrak{B}}\rightarrow{\mathfrak{A}} is the inclusion mapping and id𝔄:𝔄𝔄\mathrm{id}_{\mathfrak{A}}:{\mathfrak{A}}\rightarrow{\mathfrak{A}} is the identity map.

Lemma 6.3 ([19, Definition 3.17]).

Let 𝔄,𝔅{\mathfrak{A}},{\mathfrak{B}} be two unital C*-algebras and let (𝔰𝔞(𝔄),𝖫𝔄),(𝔰𝔞(𝔅),𝖫𝔅)({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}_{\mathfrak{A}}),({\mathfrak{sa}\left({{\mathfrak{B}}}\right)},{\mathsf{L}}_{\mathfrak{B}}) be two quantum metric order unit spaces. If a bridge γ\gamma from 𝔄{\mathfrak{A}} to 𝔅{\mathfrak{B}} is of the form γ=(𝔇,1𝔇,π𝔄,π𝔅)\gamma=({\mathfrak{D}},1_{\mathfrak{D}},\pi_{\mathfrak{A}},\pi_{\mathfrak{B}}), then the length of the bridge is

λ(γ|𝖫𝔄,𝖫𝔅)=max{supa𝔄,𝖫𝔄(a)1{infb𝔅,𝖫𝔅(b)1{π𝔄(a)π𝔅(b)𝔇}},supb𝔅,𝖫𝔅(b)1{infa𝔄,𝖫𝔄(a)1{π𝔄(a)π𝔅(b)𝔇}}}\begin{split}&\lambda(\gamma|{\mathsf{L}}_{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{B}})=\\ &\max\left\{\begin{array}[]{c}\sup_{a\in{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}(a)\leqslant 1}\left\{\inf_{b\in{\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}(b)\leqslant 1}\left\{\|\pi_{\mathfrak{A}}(a)-\pi_{\mathfrak{B}}(b)\|_{\mathfrak{D}}\right\}\right\},\\ \ \ \ \ \sup_{b\in{\mathfrak{B}},{\mathsf{L}}_{\mathfrak{B}}(b)\leqslant 1}\left\{\inf_{a\in{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{A}}(a)\leqslant 1}\left\{\|\pi_{\mathfrak{A}}(a)-\pi_{\mathfrak{B}}(b)\|_{\mathfrak{D}}\right\}\right\}\end{array}\right\}\end{split}

In particular, this holds for evident bridges.

Next, we see how lengths of bridges can be used to estimate lengths of certain tunnels. We note that the length of any bridge between two compact quantum metric ou-spaces is finite (see the discussion preceding [19, Definition 3.14]).

Theorem 6.4 ([17, Theorem 3.48]).

Let 𝔄,𝔅{\mathfrak{A}},{\mathfrak{B}} be two unital C*-algebras and let (𝔰𝔞(𝔄),𝖫𝔄),(𝔰𝔞(𝔅),𝖫𝔅)({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}_{\mathfrak{A}}),({\mathfrak{sa}\left({{\mathfrak{B}}}\right)},{\mathsf{L}}_{\mathfrak{B}}) be two quantum metric order unit spaces. Let γ=(𝔇,ω,π𝔄,π𝔅)\gamma=({\mathfrak{D}},\omega,\pi_{\mathfrak{A}},\pi_{\mathfrak{B}}) be a bridge from 𝔄{\mathfrak{A}} to 𝔅{\mathfrak{B}}. Fix any r>λ(γ|𝖫𝔄,𝖫𝔅)r>\lambda(\gamma|{\mathsf{L}}_{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{B}}), where λ(γ|𝖫𝔄,𝖫𝔅)\lambda(\gamma|{\mathsf{L}}_{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{B}}) is the length of the bridge γ\gamma.

If we define for all (a,b)𝔄𝔅(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}}

𝖫γ|𝖫𝔄,𝖫𝔅r(a,b)=max{𝖫𝔄(a),𝖫𝔅(b),π𝔄(a)ωωπ𝔅(b)𝔇r}{\mathsf{L}}^{r}_{\gamma|{\mathsf{L}}_{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{B}}}(a,b)=\max\left\{{\mathsf{L}}_{\mathfrak{A}}(a),{\mathsf{L}}_{\mathfrak{B}}(b),\frac{\|\pi_{\mathfrak{A}}(a)\omega-\omega\pi_{\mathfrak{B}}(b)\|_{\mathfrak{D}}}{r}\right\}

and we let p𝔄:(a,b)𝔄𝔅a𝔄p_{\mathfrak{A}}:(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}}\to a\in{\mathfrak{A}} and p𝔅:(a,b)𝔄𝔅b𝔅p_{\mathfrak{B}}:(a,b)\in{\mathfrak{A}}\oplus{\mathfrak{B}}\to b\in{\mathfrak{B}} denote the canonical surjections, then τ=(𝔄𝔅,𝖫γ|𝖫𝔄,𝖫𝔅r,p𝔄,p𝔅)\tau=({\mathfrak{A}}\oplus{\mathfrak{B}},{\mathsf{L}}^{r}_{\gamma|{\mathsf{L}}_{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{B}}},p_{\mathfrak{A}},p_{\mathfrak{B}}) is an order unit tunnel from (𝔰𝔞(𝔄),𝖫𝔄)({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}_{\mathfrak{A}}) to (𝔰𝔞(𝔅),𝖫𝔅)({\mathfrak{sa}\left({{\mathfrak{B}}}\right)},{\mathsf{L}}_{\mathfrak{B}}) with length λ(τ)r\lambda(\tau)\leqslant r, and

𝖽𝗂𝗌𝗍q((𝔰𝔞(𝔄),𝖫𝔄),(𝔰𝔞(𝔅),𝖫𝔅))2r.{\mathsf{dist}}_{q}\left(\left({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}_{\mathfrak{A}}\right),\left({\mathfrak{sa}\left({{\mathfrak{B}}}\right)},{\mathsf{L}}_{\mathfrak{B}}\right)\right)\leqslant 2r.
Proof.

This theorem follows from the methods in [17, Theorem 3.48] and our Theorem (5.5). ∎

This allows us to define:

Definition 6.5.

Let 𝔄,𝔅{\mathfrak{A}},{\mathfrak{B}} be two unital C*-algebras and let (𝔰𝔞(𝔄),𝖫𝔄),(𝔰𝔞(𝔅),𝖫𝔅)({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}_{\mathfrak{A}}),({\mathfrak{sa}\left({{\mathfrak{B}}}\right)},{\mathsf{L}}_{\mathfrak{B}}) be two quantum metric order unit spaces. Let γ=(𝔇,ω,π𝔄,π𝔅)\gamma=({\mathfrak{D}},\omega,\pi_{\mathfrak{A}},\pi_{\mathfrak{B}}) be a bridge from 𝔄{\mathfrak{A}} to 𝔅{\mathfrak{B}}. We call the order unit tunnel (𝔄𝔅,𝖫γ|𝖫𝔄,𝖫𝔅r,p𝔄,p𝔅)({\mathfrak{A}}\oplus{\mathfrak{B}},{\mathsf{L}}^{r}_{\gamma|{\mathsf{L}}_{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{B}}},p_{\mathfrak{A}},p_{\mathfrak{B}}) from (𝔰𝔞(𝔄),𝖫𝔄)({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}_{\mathfrak{A}}) to (𝔰𝔞(𝔅),𝖫𝔅)({\mathfrak{sa}\left({{\mathfrak{B}}}\right)},{\mathsf{L}}_{\mathfrak{B}}) of Theorem 6.4 the (r,γ|𝖫𝔄,𝖫𝔅)(r,\gamma|{\mathsf{L}}_{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{B}})-evident tunnel associated to the bridge γ\gamma, Lip-norms 𝖫𝔄,𝖫𝔅{\mathsf{L}}_{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{B}}, and r>λ(γ|𝖫𝔄,𝖫𝔅).r>\lambda(\gamma|{\mathsf{L}}_{\mathfrak{A}},{\mathsf{L}}_{\mathfrak{B}}).

Hypothesis 6.6.

Let 𝔄=n𝔄n¯𝔄{\mathfrak{A}}=\overline{\cup_{n\in{\mathds{N}}}{\mathfrak{A}}_{n}}^{\|\cdot\|_{\mathfrak{A}}} be a unital C*-algebra such that for each nn\in{\mathds{N}}, we have that 𝔄n{\mathfrak{A}}_{n} is a unital C*-subalgebra of 𝔄{\mathfrak{A}} and 𝔄n𝔄n+1{\mathfrak{A}}_{n}\subseteq{\mathfrak{A}}_{n+1}. For each nn\in{\mathds{N}}, let 𝖫n{\mathsf{L}}_{n} be a Lip-norm on 𝔰𝔞(𝔄n){\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}, so that (𝔰𝔞(𝔄n),𝖫n)({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},{\mathsf{L}}_{n}) is a quantum metric order unit space. Assume for all a𝔰𝔞(𝔄n)a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)} that 𝖫n+1(a)𝖫n(a).{\mathsf{L}}_{n+1}(a)\leqslant{\mathsf{L}}_{n}(a). Let (β(n))n(\beta(n))_{n\in{\mathds{N}}} be a sequence of positive real numbers such that n=0β(n)<.\sum_{n=0}^{\infty}\beta(n)<\infty. Let γn=(𝔄n+1,1𝔄,ιn,id𝔄n+1)\gamma_{n}=({\mathfrak{A}}_{n+1},1_{\mathfrak{A}},\iota_{n},\mathrm{id}_{{\mathfrak{A}}_{n+1}}) be the evident bridge from 𝔄n{\mathfrak{A}}_{n} to 𝔄n+1{\mathfrak{A}}_{n+1}, and assume λ(γn|𝖫n,𝖫n+1)β(n)\lambda(\gamma_{n}|{\mathsf{L}}_{n},{\mathsf{L}}_{n+1})\leqslant\beta(n). Denote the associated (2β(n),γn|𝖫n,𝖫n+1)(2\beta(n),\gamma_{n}|{\mathsf{L}}_{n},{\mathsf{L}}_{n+1})-evident tunnel by τn\tau_{n} an denote 𝔄n𝔄n+1=𝔇n{\mathfrak{A}}_{n}\oplus{\mathfrak{A}}_{n+1}={\mathfrak{D}}_{n} and its Lip-norm by 𝖫n{\mathsf{L}}^{n}.

We now begin listing some results that are more or less immediate from [1] since these results are not affected by the lack of C*-algebraic structure.

Proposition 6.7.

Given Hypothesis 6.6, it holds that for each nn\in{\mathds{N}}, we have

𝖽𝗂𝗌𝗍q((𝔰𝔞(𝔄n),𝖫n),(𝔰𝔞(𝔄n+1),𝖫n+1))4β(n){\mathsf{dist}}_{q}(({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},{\mathsf{L}}_{n}),({\mathfrak{sa}\left({{\mathfrak{A}}_{n+1}}\right)},{\mathsf{L}}_{n+1}))\leqslant 4\beta(n)

and therefore ((𝔰𝔞(𝔄n),𝖫n))n(({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},{\mathsf{L}}_{n}))_{n\in{\mathds{N}}} is Cauchy with respect to 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}, and we denote its limit given by Theorem 5.9 and Corollary 5.11 by (𝔉𝔄,S𝔄)({\mathfrak{F}}_{\mathfrak{A}},S_{\mathfrak{A}}).

Proof.

This is the same proof as [1, Proposition 2.6]. ∎

Now, we give a more explicit description of what the limit quantum metric order unit space (𝔉𝔄,S𝔄)({\mathfrak{F}}_{\mathfrak{A}},S_{\mathfrak{A}}) in Proposition (6.7) looks like under Hypothesis 6.6.

Proposition 6.8.

Assume Hypothesis 6.6. Using notation from Proposition 6.8 and Theorem 5.9, we have that

  1. (1)
    𝔅\displaystyle{\mathfrak{B}} ={((ann,an+1n))n𝔰𝔞(n(𝔄n𝔄n+1)):\displaystyle=\Bigg{\{}((a^{n}_{n},a_{n+1}^{n}))_{n\in{\mathds{N}}}\in{\mathfrak{sa}\left({\prod_{n\in{\mathds{N}}}({\mathfrak{A}}_{n}\oplus{\mathfrak{A}}_{n+1})}\right)}:
    supn{max{ann𝔄,an+1n𝔄<}}}\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\sup_{n\in{\mathds{N}}}\{\max\{\|a_{n}^{n}\|_{{\mathfrak{A}}},\|a_{n+1}^{n}\|_{{\mathfrak{A}}}<\infty\}\}\Bigg{\}}
  2. (2)

    𝔎={((ann,an+1n))n𝔅:n,an+1n=an+1n+1}{\mathfrak{K}}=\{((a^{n}_{n},a_{n+1}^{n}))_{n\in{\mathds{N}}}\in{\mathfrak{B}}:\forall n\in{\mathds{N}},a^{n}_{n+1}=a_{n+1}^{n+1}\},

  3. (3)

    and for all d=((ann,an+1n))n𝔅d=((a^{n}_{n},a_{n+1}^{n}))_{n\in{\mathds{N}}}\in{\mathfrak{B}}, if we define

    S0(d)=sup{𝖫n((ann,an+1n)):n},S_{0}(d)=\sup\{{\mathsf{L}}^{n}((a^{n}_{n},a_{n+1}^{n})):n\in{\mathds{N}}\},

    then for all d=((ann,an+1n))n𝔎d=((a^{n}_{n},a_{n+1}^{n}))_{n\in{\mathds{N}}}\in{\mathfrak{K}}, it holds that

    S0(d)=supn{max{𝖫n(ann),annan+1n+1𝔄2β(n)}},S_{0}(d)=\sup_{n\in{\mathds{N}}}\left\{\max\left\{{\mathsf{L}}_{n}\left(a_{n}^{n}\right),\frac{\left\|a^{n}_{n}-a^{n+1}_{n+1}\right\|_{\mathfrak{A}}}{2\beta(n)}\right\}\right\},
  4. (4)

    and

    S𝔄(f)=inf{S0(d):d𝔈,q(d)=f}S_{\mathfrak{A}}(f)=\inf\{S_{0}(d):d\in{\mathfrak{E}},q(d)=f\}

    for all f𝔉𝔄f\in{\mathfrak{F}}_{\mathfrak{A}},

  5. (5)

    and for each nn\in{\mathds{N}}, it holds that

    𝖽𝗂𝗌𝗍q((𝔰𝔞(𝔄n),𝖫n),(𝔉𝔄,S𝔄))4j=nβ(j).{\mathsf{dist}}_{q}(({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},{\mathsf{L}}_{n}),({\mathfrak{F}}_{\mathfrak{A}},S_{\mathfrak{A}}))\leqslant 4\sum_{j=n}^{\infty}\beta(j).
Proof.

This is the same proof as [1, Proposition 2.10] and the fact that our tunnels are built using the canonical surjections associated to 𝔄n𝔄n+1{\mathfrak{A}}_{n}\oplus{\mathfrak{A}}_{n+1} for all nn\in{\mathds{N}}, which are quotient maps. ∎

Now, we want to show that 𝔰𝔞(𝔄){\mathfrak{sa}\left({{\mathfrak{A}}}\right)} is order isomorphic to 𝔉𝔄{\mathfrak{F}}_{\mathfrak{A}}. This particular part came much more easily in [1] since 𝔉𝔄{\mathfrak{F}}_{\mathfrak{A}} is a C*-algebra there and injective *-homorphisms between C*-algebras are automatically *-isomorphisms onto their image. In our current setting, it is not as simple to provide order isomorphisms. Hence, we have to do more work.

Definition 6.9.

Assuming Hypothesis 6.6 and using notation from Proposition 6.8, define ψ0:𝔰𝔞(𝔄0)𝔅\psi_{0}:{\mathfrak{sa}\left({{\mathfrak{A}}_{0}}\right)}\rightarrow{\mathfrak{B}} by

ψ0(a0)=((a0,a0))n,\psi_{0}(a_{0})=((a_{0},a_{0}))_{n\in{\mathds{N}}},

and for n{0}n\in{\mathds{N}}\setminus\{0\}, define ψn:𝔰𝔞(𝔄n)𝔅\psi_{n}:{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}\rightarrow{\mathfrak{B}} by

ψn(an)=((0,0),,(0,0),(0,an),(an,an),(an,an),),\psi_{n}(a_{n})=((0,0),\ldots,(0,0),(0,a_{n}),(a_{n},a_{n}),(a_{n},a_{n}),\ldots),

where (0,an)𝔇n1=𝔰𝔞(𝔄n1)𝔰𝔞(𝔄n)(0,a_{n})\in{\mathfrak{D}}_{n-1}={\mathfrak{sa}\left({{\mathfrak{A}}_{n-1}}\right)}\oplus{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}.

Lemma 6.10.

Assuming Hypothesis 6.6 and using notation from Proposition 6.8, the map ψn:𝔰𝔞(𝔄n)𝔅\psi_{n}:{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}\rightarrow{\mathfrak{B}} is an order isometry (not necessarily unital) such that ψn(𝔰𝔞(𝔄n))𝔈\psi_{n}({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)})\subseteq{\mathfrak{E}} for all nn\in{\mathds{N}}.

Proof.

Fix nn\in{\mathds{N}}. We have that ψn\psi_{n} is an isometry by construction. Since the order on 𝔅{\mathfrak{B}} is just the coordinate order induced by each 𝔰𝔞(𝔄n){\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}, we have that ψn\psi_{n} is a positive map and thus an order isometry. We will now show that ψn({a𝔰𝔞(𝔄n):𝖫n(a)<})𝔏\psi_{n}\left(\{a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}:{\mathsf{L}}_{n}(a)<\infty\}\right)\subseteq{\mathfrak{L}} where 𝔏{\mathfrak{L}} was defined in Theorem 5.9. Let a𝔰𝔞(𝔄n)a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)} such that 𝖫n(a)<{\mathsf{L}}_{n}(a)<\infty. If a1𝔄a\in{\mathds{R}}1_{\mathfrak{A}}, then S0(ψn(a))=0<S_{0}(\psi_{n}(a))=0<\infty. So, assume a1𝔄a\not\in{\mathds{R}}1_{\mathfrak{A}}. By construction, ψn(𝔰𝔞(𝔄n))𝔎\psi_{n}({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)})\subseteq{\mathfrak{K}} by Proposition 6.8, and we thus have S0(ψn(a))=max{𝖫n(a),a𝔄/(2β(n1))}<S_{0}(\psi_{n}(a))=\max\{{\mathsf{L}}_{n}(a),\|a\|_{\mathfrak{A}}/(2\beta(n-1))\}<\infty. Therefore, ψn({a𝔰𝔞(𝔄n):𝖫n(a)<})𝔏\psi_{n}\left(\{a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}:{\mathsf{L}}_{n}(a)<\infty\}\right)\subseteq{\mathfrak{L}}, and thus

ψn(𝔰𝔞(𝔄n))=ψn({a𝔰𝔞(𝔄n):𝖫n(a)<}¯𝔄)ψn({a𝔰𝔞(𝔄n):𝖫n(a)<})¯𝔈𝔏¯𝔈=𝔈.\begin{split}\psi_{n}({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)})&=\psi_{n}\left(\overline{\{a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}:{\mathsf{L}}_{n}(a)<\infty\}}^{\|\cdot\|_{\mathfrak{A}}}\right)\\ &\subseteq\overline{\psi_{n}\left(\{a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}:{\mathsf{L}}_{n}(a)<\infty\}\right)}^{\|\cdot\|_{\mathfrak{E}}}\subseteq\overline{{\mathfrak{L}}}^{\|\cdot\|_{\mathfrak{E}}}={\mathfrak{E}}.\end{split}

by continuity. ∎

This lemma allows us to define:

Definition 6.11.

Assuming Hypothesis 6.6, for each nn\in{\mathds{N}}, by Lemma 6.10 we may define ψ(n):𝔰𝔞(𝔄n)𝔉𝔄\psi^{(n)}:{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}\rightarrow{\mathfrak{F}}_{{\mathfrak{A}}} by ψ(n):=qψn\psi^{(n)}:=q\circ\psi_{n} where q:𝔈𝔉𝔄q:{\mathfrak{E}}\rightarrow{\mathfrak{F}}_{{\mathfrak{A}}} is the quotient map.

Lemma 6.12.

Assuming Hypothesis 6.6, the map ψ(n):𝔰𝔞(𝔄n)𝔉𝔄\psi^{(n)}:{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}\rightarrow{\mathfrak{F}}_{{\mathfrak{A}}} of Definition 6.11 is an order unit isomorphism onto its image for each nn\in{\mathds{N}}. Furthermore ψ(n+1)\psi^{(n+1)} restricted to 𝔰𝔞(𝔄n){\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)} is ψ(n)\psi^{(n)} for all nn\in{\mathds{N}}.

Proof.

Fix nn\in{\mathds{N}}. The map ψ(n)\psi^{(n)} is linear by construction. For unital, we note that ψn(1𝔄)1𝔅𝔍\psi_{n}(1_{\mathfrak{A}})-1_{\mathfrak{B}}\in{\mathfrak{J}}, and thus ψ(n)(1𝔄)=1𝔉𝔄.\psi^{(n)}(1_{\mathfrak{A}})=1_{{\mathfrak{F}}_{\mathfrak{A}}}. For injectivity, assume a𝔰𝔞(𝔄n)a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)} and ψ(n)(a)=0\psi^{(n)}(a)=0. Thus ψn(a)𝔍\psi_{n}(a)\in{\mathfrak{J}}. Hence 0=limna𝔄=a𝔄0=\lim_{n\to\infty}\|a\|_{\mathfrak{A}}=\|a\|_{\mathfrak{A}} which implies a=0a=0. Next, let a𝔰𝔞(𝔄n)𝔰𝔞(𝔄n+1)a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}\subseteq{\mathfrak{sa}\left({{\mathfrak{A}}_{n+1}}\right)}. Then, again we have ψn(a)ψn+1(a)𝔍\psi_{n}(a)-\psi_{n+1}(a)\in{\mathfrak{J}}, and thus ψ(n)(a)=ψ(n+1)(a).\psi^{(n)}(a)=\psi^{(n+1)}(a).

Now, we will show that ψ(n)\psi^{(n)} and its inverse defined on ψ(n)(𝔰𝔞(𝔄n))\psi^{(n)}({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}) are positive. Note by Theorem 5.9, we use the quotient order on 𝔉𝔄{\mathfrak{F}}_{\mathfrak{A}}. We begin with showing that ψ(n)\psi^{(n)} is positive. Let a𝔰𝔞(𝔄n)a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)} such that a0a\geqslant 0. By Lemma 6.10, we have that ψn(a)0\psi_{n}(a)\geqslant 0 with respect to the order on 𝔈{\mathfrak{E}}, which is the same order on 𝔅{\mathfrak{B}}. Now, consider d=((0,0))n𝔈d=((0,0))_{n\in{\mathds{N}}}\in{\mathfrak{E}}. We have that d𝔍d\in{\mathfrak{J}} and ψn(a)+d=ψn(a)0\psi_{n}(a)+d=\psi_{n}(a)\geqslant 0. Thus ψ(n)(a)0\psi^{(n)}(a)\geqslant 0 with respect to the quotient order on 𝔉𝔄{\mathfrak{F}}_{\mathfrak{A}}. Hence ψ(n)\psi^{(n)} is positive.

Next, we show that the inverse of ψ(n)\psi^{(n)} on ψ(n)(𝔰𝔞(𝔄n))\psi^{(n)}({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}) is positive. Let a𝔰𝔞(𝔄n)a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)} and assume that ψ(n)(a)0\psi^{(n)}(a)\geqslant 0 in the quotient order on 𝔉𝔄{\mathfrak{F}}_{\mathfrak{A}}. Thus, there exists d𝔍d\in{\mathfrak{J}} such that ψn(a)+d0\psi_{n}(a)+d\geqslant 0. Now d=((ann,an+1n))nn(𝔰𝔞(𝔄n)𝔰𝔞(𝔄n+1))d=((a_{n}^{n},a^{n}_{n+1}))_{n\in{\mathds{N}}}\in\prod_{n\in{\mathds{N}}}({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)}\oplus{\mathfrak{sa}\left({{\mathfrak{A}}_{n+1}}\right)}) such that limnmax{ann𝔄,an+1n𝔄}=0.\lim_{n\to\infty}\max\{\|a_{n}^{n}\|_{{\mathfrak{A}}},\|a^{n}_{n+1}\|_{{\mathfrak{A}}}\}=0. Now, we have that a+amm0a+a^{m}_{m}\geqslant 0 for all mn+1m\geqslant n+1 since ψn(a)+d0\psi_{n}(a)+d\geqslant 0. Hence for all mn+1m\geqslant n+1, we have

a+amm0\displaystyle a+a^{m}_{m}\geqslant 0 aammamm𝔄1𝔄.\displaystyle\implies a\geqslant-a_{m}^{m}\geqslant-\|a_{m}^{m}\|_{\mathfrak{A}}1_{\mathfrak{A}}.

Hence, as limnann𝔄=0\lim_{n\to\infty}\|a^{n}_{n}\|_{\mathfrak{A}}=0, we have that ar1𝔄a\geqslant-r1_{\mathfrak{A}} for all r,r>0r\in{\mathds{R}},r>0. Since the given order on 𝔰𝔞(𝔄n){\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)} is Archimedean, we have that a0.a\geqslant 0.

Thus ψ(n)\psi^{(n)} is an order unit isomorphism onto its image. ∎

Finally, we are ready to build an order unit isomorphism from 𝔰𝔞(𝔄){\mathfrak{sa}\left({{\mathfrak{A}}}\right)} onto 𝔉𝔄{\mathfrak{F}}_{\mathfrak{A}}. This follows the same process of [1, Theorem 2.15] up to some crucial details concerning order unit spaces.

Theorem 6.13.

Assume Hypothesis 6.6. Using notation from Proposition 6.7, the following hold:

  1. (1)

    there exists an order unit isomorphism ψ:𝔰𝔞(𝔄)𝔉𝔄\psi:{\mathfrak{sa}\left({{\mathfrak{A}}}\right)}\rightarrow{\mathfrak{F}}_{\mathfrak{A}} such that the restriction of ψ\psi to 𝔰𝔞(𝔄n){\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)} is ψ(n)\psi^{(n)} for all nn\in{\mathds{N}} of Definition 6.11, and

  2. (2)

    if we define 𝖫𝔄β:=S𝔄ψ{\mathsf{L}}^{\beta}_{\mathfrak{A}}:=S_{\mathfrak{A}}\circ\psi, then (𝔰𝔞(𝔄),𝖫𝔄β)({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}^{\beta}_{\mathfrak{A}}) is a quantum metric order unit space such that ndom(𝖫n)dom(𝖫𝔄β)\cup_{n\in{\mathds{N}}}{\operatorname*{dom}\left({{\mathsf{L}}_{n}}\right)}\subseteq{\operatorname*{dom}\left({{\mathsf{L}}^{\beta}_{\mathfrak{A}}}\right)} with for each nn\in{\mathds{N}}

    𝖽𝗂𝗌𝗍q((𝔰𝔞(𝔄n),𝖫n),(𝔰𝔞(𝔄),𝖫𝔄β))4j=nβ(j){\mathsf{dist}}_{q}\left(\left({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},{\mathsf{L}}_{n}\right),\left({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}^{\beta}_{\mathfrak{A}}\right)\right)\leqslant 4\sum_{j=n}^{\infty}\beta(j)

    and therefore limn𝖽𝗂𝗌𝗍q((𝔰𝔞(𝔄n),𝖫n),(𝔰𝔞(𝔄),𝖫𝔄β))=0.\lim_{n\to\infty}{\mathsf{dist}}_{q}\left(\left({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},{\mathsf{L}}_{n}\right),\left({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}^{\beta}_{\mathfrak{A}}\right)\right)=0.

Proof.

The fact that there exists an order unit isomorphism ψ:𝔰𝔞(𝔄)𝔉𝔄\psi:{\mathfrak{sa}\left({{\mathfrak{A}}}\right)}\rightarrow{\mathfrak{F}}_{\mathfrak{A}} such that the restriction of ψ\psi to 𝔰𝔞(𝔄n){\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)} is ψ(n)\psi^{(n)} for all nn\in{\mathds{N}} follows from [22, Remark 3.6 (ii)] and Lemma 6.12.

Next, we show ψ(𝔰𝔞(𝔄))=𝔉𝔄.\psi({\mathfrak{sa}\left({{\mathfrak{A}}}\right)})={\mathfrak{F}}_{\mathfrak{A}}. Let a+𝔍𝔉𝔄a+{\mathfrak{J}}\in{\mathfrak{F}}_{\mathfrak{A}}. Let ε>0\varepsilon>0. There exists b=(bn)n=((bnn,bn+1n))n𝔏b=(b_{n})_{n\in{\mathds{N}}}=((b_{n}^{n},b^{n}_{n+1}))_{n\in{\mathds{N}}}\in{\mathfrak{L}} such that a+𝔍b+𝔍𝔉𝔄<ε/2\|a+{\mathfrak{J}}-b+{\mathfrak{J}}\|_{{\mathfrak{F}}_{\mathfrak{A}}}<\varepsilon/2 by density. Hence, there exists r,r>0r\in{\mathds{R}},r>0 such that S0(b)rS_{0}(b)\leqslant r. Also, there exists N,N>1N\in{\mathds{N}},N>1 such that 2rj=Nβ(j)<ε/42r\cdot\sum_{j=N}^{\infty}\beta(j)<\varepsilon/4.

Define c=((cnn,cn+1n))n𝔅c=((c_{n}^{n},c^{n}_{n+1}))_{n\in{\mathds{N}}}\in{\mathfrak{B}} in the following way:

(cnn,cn+1n)={(0,0):0nN2(0,bNN1):n=N1(bnn,bn+1n):nN.(c_{n}^{n},c^{n}_{n+1})=\begin{cases}(0,0)&:0\leqslant n\leqslant N-2\\ (0,b^{N-1}_{N})&:n=N-1\\ (b_{n}^{n},b^{n}_{n+1})&:n\geqslant N.\end{cases}

Therefore c𝔏c\in{\mathfrak{L}} and cb𝔍c-b\in{\mathfrak{J}}, which implies that c+𝔍=b+𝔍𝔉𝔄c+{\mathfrak{J}}=b+{\mathfrak{J}}\in{\mathfrak{F}}_{\mathfrak{A}}.

Now, consider ψN(bNN1)\psi_{N}(b^{N-1}_{N}) and recall that bNN1=bNNb^{N-1}_{N}=b^{N}_{N} and that bn+1n=bn+1n+1b^{n}_{n+1}=b^{n+1}_{n+1} for all nn\in{\mathds{N}} by Proposition 6.8. Therefore

ψN(bNN1)c𝔈=supkbNNbN+k+1N+k+1𝔄.\left\|\psi_{N}(b^{N-1}_{N})-c\right\|_{\mathfrak{E}}=\sup_{k\in{\mathds{N}}}\left\|b^{N}_{N}-b^{N+k+1}_{N+k+1}\right\|_{\mathfrak{A}}.

Since S0(b)r<S_{0}(b)\leqslant r<\infty, we have that bnnbn+1n+1𝔄r2β(n)\|b_{n}^{n}-b^{n+1}_{n+1}\|_{\mathfrak{A}}\leqslant r\cdot 2\beta(n) for all nn\in{\mathds{N}} by Proposition 6.8. Hence for all kk\in{\mathds{N}}

bNNbN+k+1N+k+1𝔄2rj=NN+kβ(j)2rj=Nβ(j)<ε/4.\left\|b^{N}_{N}-b^{N+k+1}_{N+k+1}\right\|_{\mathfrak{A}}\leqslant 2r\cdot\sum_{j=N}^{N+k}\beta(j)\leqslant 2r\cdot\sum_{j=N}^{\infty}\beta(j)<\varepsilon/4.

Thus ψN(bNN1)c𝔈ε/4\|\psi_{N}(b^{N-1}_{N})-c\|_{\mathfrak{E}}\leqslant\varepsilon/4. Therefore, since ψ(bNN1)=ψ(N)(bNN1))\psi(b^{N-1}_{N})=\psi^{(N)}(b^{N-1}_{N})) as bNN1𝔰𝔞(𝔄N)b^{N-1}_{N}\in{\mathfrak{sa}\left({{\mathfrak{A}}_{N}}\right)}, we gather

a+𝔍ψ(bNN1)𝔉𝔄a+𝔍b+𝔍𝔉𝔄+ψ(bNN1)b+𝔍𝔉𝔄<ε/2+ψ(bNN1)c+𝔍𝔉𝔄ε/2+ψN(bNN1)c𝔈ε/2+ε/4<ε.\begin{split}\|a+{\mathfrak{J}}-\psi(b^{N-1}_{N})\|_{{\mathfrak{F}}_{\mathfrak{A}}}&\leqslant\|a+{\mathfrak{J}}-b+{\mathfrak{J}}\|_{{\mathfrak{F}}_{\mathfrak{A}}}+\|\psi(b^{N-1}_{N})-b+{\mathfrak{J}}\|_{{\mathfrak{F}}_{\mathfrak{A}}}\\ &<\varepsilon/2+\|\psi(b^{N-1}_{N})-c+{\mathfrak{J}}\|_{{\mathfrak{F}}_{\mathfrak{A}}}\\ &\leqslant\varepsilon/2+\|\psi_{N}(b^{N-1}_{N})-c\|_{\mathfrak{E}}\leqslant\varepsilon/2+\varepsilon/4<\varepsilon.\end{split}

by definition of quotient norm. In particular, the set ψ(𝔰𝔞(𝔄))\psi({\mathfrak{sa}\left({{\mathfrak{A}}}\right)}) is dense in 𝔉𝔄{{\mathfrak{F}}_{\mathfrak{A}}} with respect to the quotient norm. However, as our tunnels are built from quotient maps since they are built using the canonical surjections (a,b)𝔄n𝔄n+1a(a,b)\in{\mathfrak{A}}_{n}\oplus{\mathfrak{A}}_{n+1}\mapsto a and (a,b)𝔄n𝔄n+1b(a,b)\in{\mathfrak{A}}_{n}\oplus{\mathfrak{A}}_{n+1}\mapsto b (see the discussion after Definition 5.6), we have that the order norm and quotient norm equal by Theorem 5.9. As ψ\psi is an isometry with respect to the order norms since it is an order unit isomorphism between (Archimedean) order unit spaces (see [4, Corollary II.1.4]) and 𝔰𝔞(𝔄){\mathfrak{sa}\left({{\mathfrak{A}}}\right)} is complete, it must be the case that ψ(𝔰𝔞(𝔄))=𝔉𝔄\psi({\mathfrak{sa}\left({{\mathfrak{A}}}\right)})={{\mathfrak{F}}_{\mathfrak{A}}}.

Now, assume that andom(𝖫n)a\in\cup_{n\in{\mathds{N}}}{\operatorname*{dom}\left({{\mathsf{L}}_{n}}\right)}, then there exists N,N>1N\in{\mathds{N}},N>1 such that a𝔰𝔞(𝔄N)a\in{\mathfrak{sa}\left({{\mathfrak{A}}_{N}}\right)} and 𝖫N(a)<{\mathsf{L}}_{N}(a)<\infty. Thus by Proposition 6.8

𝖫𝔄ψ(a)=S𝔄ψ(a)=S𝔄ψ(N)(a)S0(ψN(a))=max{𝖫N(a),a𝔄/(2β(N1))}<.\begin{split}{\mathsf{L}}^{\psi}_{\mathfrak{A}}(a)&=S_{\mathfrak{A}}\circ\psi(a)=S_{\mathfrak{A}}\circ\psi^{(N)}(a)\leqslant S_{0}(\psi_{N}(a))\\ &=\max\{{\mathsf{L}}_{N}(a),\|a\|_{\mathfrak{A}}/(2\beta(N-1))\}<\infty.\end{split}

The remaining follows from Proposition 6.8 and Corollary 5.11. ∎

As a corollary, we prove the previous result for the following description of inductive limits.

Corollary 6.14.

Let 𝔄=lim(𝔄n,αn)n{\mathfrak{A}}=\underrightarrow{\lim}\ ({\mathfrak{A}}_{n},\alpha_{n})_{n\in{\mathds{N}}} be an inductive limit of C*-algebras (see [24, Section 6.1]), where for each nn\in{\mathds{N}}, 𝔄n{\mathfrak{A}}_{n} is a unital C*-algebra and αn\alpha_{n} is a unital *-monomorphism. For each nn\in{\mathds{N}}, let (𝔰𝔞(𝔄n),𝖫n)({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},{\mathsf{L}}_{n}) be quantum metric order unit space. Let (ψ(n))n(\psi(n))_{n\in{\mathds{N}}} be a summable sequence of positive real numbers.

If for each nn\in{\mathds{N}}, the bridge δn=(𝔄n,𝔄n+1,1𝔄n+1,αn,id𝔄n+1)\delta_{n}=({\mathfrak{A}}_{n},{\mathfrak{A}}_{n+1},1_{{\mathfrak{A}}_{n+1}},\alpha_{n},\mathrm{id}_{{\mathfrak{A}}_{n+1}}) has length λ(δn|𝖫n,𝖫n+1)ψ(n),\lambda(\delta_{n}|{\mathsf{L}}_{n},{\mathsf{L}}_{n+1})\leqslant\psi(n), then there exists a Lip-norm 𝖫𝔄ψ{\mathsf{L}}^{\psi}_{\mathfrak{A}} on 𝔰𝔞(𝔄){\mathfrak{sa}\left({{\mathfrak{A}}}\right)} such that for each nn\in{\mathds{N}},

𝖽𝗂𝗌𝗍q((𝔰𝔞(𝔄n),𝖫n),(𝔰𝔞(𝔄),𝖫𝔄ψ))8j=nψ(j),{\mathsf{dist}}_{q}(({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},{\mathsf{L}}_{n}),({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}^{\psi}_{\mathfrak{A}}))\leqslant 8\sum_{j=n}^{\infty}\psi(j),

and thus

limn𝖽𝗂𝗌𝗍q((𝔰𝔞(𝔄n),𝖫n),(𝔰𝔞(𝔄),𝖫𝔄ψ))=0.\lim_{n\to\infty}{\mathsf{dist}}_{q}(({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},{\mathsf{L}}_{n}),({\mathfrak{sa}\left({{\mathfrak{A}}}\right)},{\mathsf{L}}^{\psi}_{\mathfrak{A}}))=0.
Proof.

By [24, Section 6.1], for each nn\in{\mathds{N}}, let α(n):𝔄n𝔄\alpha^{(n)}:{\mathfrak{A}}_{n}\rightarrow{\mathfrak{A}} be the canonical unital *-monomorphism associated to αn\alpha_{n}, where α(n)(𝔄n)\alpha^{(n)}({\mathfrak{A}}_{n}) is a unital C*-subalgebra of 𝔄{\mathfrak{A}} such that α(n)(𝔄n)α(n+1)(𝔄n+1)\alpha^{(n)}({\mathfrak{A}}_{n})\subseteq\alpha^{(n+1)}({\mathfrak{A}}_{n+1}) and 𝔄=nα(n)(𝔄n)¯{\mathfrak{A}}=\overline{\cup_{n\in{\mathds{N}}}\alpha^{(n)}({\mathfrak{A}}_{n})}.

Next, for each nn\in{\mathds{N}}, we have (𝔰𝔞(α(n)(𝔄n)),𝖫n(α(n))1)({\mathfrak{sa}\left({\alpha^{(n)}({\mathfrak{A}}_{n})}\right)},{\mathsf{L}}_{n}\circ(\alpha^{(n)})^{-1}) is a quantum metric order unit space such that

(6.1) distq((𝔰𝔞(α(n)(𝔄n)),𝖫n(α(n))1),(𝔰𝔞(𝔄n),𝖫n))=0.\mathrm{dist}_{q}(({\mathfrak{sa}\left({\alpha^{(n)}({\mathfrak{A}}_{n})}\right)},{\mathsf{L}}_{n}\circ(\alpha^{(n)})^{-1}),({\mathfrak{sa}\left({{\mathfrak{A}}_{n}}\right)},{\mathsf{L}}_{n}))=0.

Now, consider the bridge γn=(α(n)(𝔄n),α(n+1)(𝔄n+1),1𝔄,ιn,idα(n+1)(𝔄n+1))\gamma_{n}=(\alpha^{(n)}({\mathfrak{A}}_{n}),\alpha^{(n+1)}({\mathfrak{A}}_{n+1}),1_{\mathfrak{A}},\iota_{n},\mathrm{id}_{\alpha^{(n+1)}({\mathfrak{A}}_{n+1})}). We will show that its length λ(γn|𝖫n(α(n)))1,𝖫n+1(α(n+1))1)2ψ(n)\lambda(\gamma_{n}|{\mathsf{L}}_{n}\circ(\alpha^{(n)}))^{-1},{\mathsf{L}}_{n+1}\circ(\alpha^{(n+1)})^{-1})\leqslant 2\psi(n).

Let aα(n)(𝔄n)a\in\alpha^{(n)}({\mathfrak{A}}_{n}) such that 𝖫n(α(n))1(a)1{\mathsf{L}}_{n}\circ(\alpha^{(n)})^{-1}(a)\leqslant 1. Set a=α(n)(a)a=\alpha^{(n)}(a^{\prime}) for some a𝔄na^{\prime}\in{\mathfrak{A}}_{n}. Then, we have that 𝖫n(a)1{\mathsf{L}}_{n}(a^{\prime})\leqslant 1. Thus, by Lemma 6.3, there exists b𝔄n+1b^{\prime}\in{\mathfrak{A}}_{n+1} such that 𝖫𝔄n+1(b)1{\mathsf{L}}_{{\mathfrak{A}}_{n+1}}(b^{\prime})\leqslant 1 and αn(a)b𝔄n+12ψ(n).\|\alpha_{n}(a^{\prime})-b^{\prime}\|_{{\mathfrak{A}}_{n+1}}\leqslant 2\psi(n). Now, set b=α(n+1)(b)b=\alpha^{(n+1)}(b^{\prime}) and note 𝖫n+1(α(n+1))1(b)1.{\mathsf{L}}_{n+1}\circ(\alpha^{(n+1)})^{-1}(b)\leqslant 1. Next, by [24, Section 6.1], we have

ab𝔄=α(n)(a)α(n+1)(b)𝔄=α(n+1)(αn(a))α(n+1)(b)𝔄=αn(a)b𝔄n+12ψ(n).\begin{split}\|a-b\|_{\mathfrak{A}}&=\|\alpha^{(n)}(a^{\prime})-\alpha^{(n+1)}(b^{\prime})\|_{\mathfrak{A}}=\|\alpha^{(n+1)}(\alpha_{n}(a^{\prime}))-\alpha^{(n+1)}(b^{\prime})\|_{\mathfrak{A}}\\ &=\|\alpha_{n}(a^{\prime})-b^{\prime}\|_{{\mathfrak{A}}_{n+1}}\leqslant 2\psi(n).\end{split}

The argument is symmetric if one begins with the space (α(n+1)(𝔄n+1),𝖫n+1(α(n+1))1)(\alpha^{(n+1)}({\mathfrak{A}}_{n+1}),{\mathsf{L}}_{n+1}\circ(\alpha^{(n+1)})^{-1}). Thus, by Lemma 6.3, it holds that λ(γn|𝖫n(α(n)))1,𝖫n+1(α(n+1))1)2ψ(n)\lambda(\gamma_{n}|{\mathsf{L}}_{n}\circ(\alpha^{(n)}))^{-1},{\mathsf{L}}_{n+1}\circ(\alpha^{(n+1)})^{-1})\leqslant 2\psi(n). Therefore, the proof is complete by Theorem 6.13 and Expression (6.1), and we denote 𝖫𝔄2ψ{\mathsf{L}}^{2\psi}_{\mathfrak{A}} by 𝖫𝔄ψ{\mathsf{L}}^{\psi}_{\mathfrak{A}}. ∎

Now, we may provide quantum metrics on inductive limits built from quantum metrics on the spaces of the inductive sequence without the requirement of any quasi-Leibniz rule. Of course, this comes at the loss of capturing the multiplicative structure of the C*-algebra, but this opens up many more possibilities for convergence results in Rieffel’s quantum distance 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}.


We thus are now able to state a main result for this paper.

Theorem 6.15.

If σ𝒩\sigma\in{\mathscr{N}}, and if for all mm\in{\mathds{N}} we set 𝖲σ,0=𝖫σ,0\mathsf{S}_{\sigma,0}={\mathsf{L}}_{\sigma,0} on 𝔰𝔞(𝔓(σ,0)){\mathfrak{sa}\left({{{\mathfrak{CP}}\left({\sigma},{0}\right)}}\right)} and for all m{0}m\in{\mathds{N}}\setminus\{0\}:

a𝔰𝔞(𝔓(σ,m))𝖲σ,m(a)=max{ϰm𝖫σ,m(a),𝖲σ,m1ασ,m1𝔼σ,m(a),12ma𝔼σ,m(a)𝔓(σ,m)}\forall a\in{\mathfrak{sa}\left({{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\right)}\quad\mathsf{S}_{\sigma,m}(a)=\\ \max\left\{\varkappa_{m}{\mathsf{L}}_{\sigma,m}(a),\mathsf{S}_{\sigma,m-1}\circ\alpha_{\sigma,m}^{-1}\circ{\mathds{E}_{\sigma,m}\left({a}\right)},\frac{1}{2^{m}}\left\|{a-{\mathds{E}_{\sigma,m}\left({a}\right)}}\right\|_{{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\right\}

where for all m{0}m\in{\mathds{N}}\setminus\{0\}, we have km=1+2l||σm1(Uσ,m1)σmk_{m}=\frac{1+2l_{|\cdot|}^{\boxtimes\sigma_{m-1}}(U_{\sigma,m-1})}{\sigma_{m}}, and:

nϰn={1 if n{0,1},ϰn1kn otherwise\forall n\in{\mathds{N}}\quad\varkappa_{n}=\begin{cases}1\text{ if $n\in\{0,1\}$,}\\ \frac{\varkappa_{n-1}}{k_{n}}\text{ otherwise}\end{cases}

then (𝔓(σ,m),𝖲σ,m)({{\mathfrak{CP}}\left({\sigma},{m}\right)},\mathsf{S}_{\sigma,m}) is a quantum metric order unit space and there exists a Lip-norm 𝖲σ\mathsf{S}_{\sigma} on the Bunce-Deddens algebra 𝔅𝔇(σ){{\mathfrak{BD}}\left({\sigma}\right)} such that:

limm𝖽𝗂𝗌𝗍q((𝔰𝔞(𝔅𝔇(σ)),𝖲σ),(𝔰𝔞(𝔓(σ,m)),𝖲σ,m))=0.\lim_{m\rightarrow\infty}{\mathsf{dist}}_{q}(({\mathfrak{sa}\left({{{\mathfrak{BD}}\left({\sigma}\right)}}\right)},\mathsf{S}_{\sigma}),({\mathfrak{sa}\left({{{\mathfrak{CP}}\left({\sigma},{m}\right)}}\right)},\mathsf{S}_{\sigma,m}))=0\text{.}
Proof.

We apply Corollary (6.14) to the conclusions of Theorem (4.7) and the proofs of Lemma 2.1 and Theorem 2.6, where the length of the bridges in Corollary (6.14) are calculated. ∎

We conclude with a consequence of our construction: the map which sends an element of the Baire space to its Bunce-Deddens algebra is continuous for 𝖽𝗂𝗌𝗍q{\mathsf{dist}}_{q}.

Theorem 6.16.

Using the notations of Theorem (6.15), the map

β(𝒩,𝖽𝒩)(𝔰𝔞(𝔅𝔇(β)),𝖫𝔅𝔇(β)ψβ)(CQMSou,𝖽𝗂𝗌𝗍q).\beta\in(\mathcal{N},\mathsf{d}_{\mathcal{N}})\longmapsto({\mathfrak{sa}\left({{\mathfrak{B}}{\mathfrak{D}}(\beta)}\right)},{\mathsf{L}}^{\psi_{\beta}}_{{\mathfrak{B}}{\mathfrak{D}}(\beta)})\in(\mathrm{CQMS}_{\mathrm{ou}},{\mathsf{dist}}_{q}).

is Lipschitz with Lipschitz constant at most 3232.

Proof.

Let β,η𝒩\beta,\eta\in\mathcal{N} such that βη\beta\neq\eta. Set n=min{k:β(k)η(k)}n=\min\{k\in{\mathds{N}}:\beta(k)\neq\eta(k)\} (so d(η,β)=2nd(\eta,\beta)=2^{-n}). Hence β(k)=η(k)\boxtimes\beta(k)=\boxtimes\eta(k) for all k{0,,n}k\in\{0,\ldots,n\}. By induction, we note that (𝔓(β,m),𝖲β,m)=(𝔓(η,m),𝖲η,m)({{\mathfrak{CP}}\left({\beta},{m}\right)},\mathsf{S}_{\beta,m})=({{\mathfrak{CP}}\left({\eta},{m}\right)},\mathsf{S}_{\eta,m}) for all mnm\leqslant n.

Thus, by the triangle inequality and Theorem 6.15, it holds that

𝖽𝗂𝗌𝗍q((𝔰𝔞(𝔅𝔇(β)),𝖫𝔅𝔇(β)ψβ),(𝔰𝔞(𝔅𝔇(η)),𝖫𝔅𝔇(η)ψη))8j=n2j+0+8j=n2j=821n+821n=322n=32𝖽𝒩(β,η),\begin{split}{\mathsf{dist}}_{q}(({\mathfrak{sa}\left({{\mathfrak{B}}{\mathfrak{D}}(\beta)}\right)},{\mathsf{L}}^{\psi_{\beta}}_{{\mathfrak{B}}{\mathfrak{D}}(\beta)}),({\mathfrak{sa}\left({{\mathfrak{B}}{\mathfrak{D}}(\eta)}\right)},{\mathsf{L}}^{\psi_{\eta}}_{{\mathfrak{B}}{\mathfrak{D}}(\eta)}))&\leqslant 8\sum_{j=n}^{\infty}2^{-j}+0+8\sum_{j=n}^{\infty}2^{-j}\\ &=8\cdot 2^{1-n}+8\cdot 2^{1-n}\\ &=32\cdot 2^{-n}\\ &=32\cdot\mathsf{d}_{\mathcal{N}}(\beta,\eta),\end{split}

which completes the proof. ∎

References

  • [1] K. Aguilar, Inductive limits of C*-algebras and compact quantum metric spaces, 24 pages, accepted (2020), to appear in Journal of the Australian Mathematical Society, ArXiv: 1807.10424.
  • [2] K. Aguilar, Fell topologies for AF-algebras and the quantum propinquity, J. Operator Theory 82 (2019), no. 2, 469–514, ArXiv:1608.07016. MR 4015960
  • [3] K. Aguilar and F. Latrémolière, Quantum ultrametrics on AF algebras and the Gromov-Hausdorff propinquity, Studia Mathematica 231 (2015), no. 2, 149 –193, ArXiv: 1511.07114.
  • [4] E. M. Alfsen, Compact convex sets and boundary integrals, Ergebnisse Math., vol. 57, Springer-Verlag, 1971.
  • [5] N. P. Brown and N. Ozawa, C*-algebras and finite-dimensional approximations, Graudate Studies in Mathematics, vol. 88, American Mathematical Society, 2008.
  • [6] John W. Bunce and James A. Deddens, A family of simple CC^{\ast}-algebras related to weighted shift operators, J. Functional Analysis 19 (1975), 13–24. MR 0365157
  • [7] A. Connes, Compact metric spaces, Fredholm modules and hyperfiniteness, Ergodic Theory and Dynamical Systems 9 (1989), no. 2, 207–220.
  • [8] K. R. Davidson, C*–algebras by example, Fields Institute Monographs, American Mathematical Society, 1996.
  • [9] M. Gromov, Metric structures for Riemannian and non-Riemannian spaces, Progress in Mathematics, Birkhäuser, 1999.
  • [10] F. Hausdorff, Grundzüge der Mengenlehre, Verlag Von Veit und Comp., 1914.
  • [11] L. V. Kantorovich, On one effective method of solving certain classes of extremal problems, Dokl. Akad. Nauk. USSR 28 (1940), 212–215.
  • [12] F. Latrémolière, The covariant Gromov-Hausdorff propinquity, Submitted (2018), 29 pages, ArXiv: 1805.11229.
  • [13] by same author, The modular Gromov-Hausdorff propinquity, Submitted (2016), 67 pages, ArXiv: 1608.04881.
  • [14] by same author, Approximation of the quantum tori by finite quantum tori for the quantum Gromov-Hausdorff distance, Journal of Funct. Anal. 223 (2005), 365–395, math.OA/0310214.
  • [15] by same author, Convergence of fuzzy tori and quantum tori for the quantum Gromov-Hausdorff propinquity: an explicit approach, Münster J. Math. 8 (2015), no. 1, ArXiv: math/1312.0069.
  • [16] by same author, The dual Gromov–Hausdorff Propinquity, Journal de Mathématiques Pures et Appliquées 103 (2015), no. 2, 303–351, ArXiv: 1311.0104.
  • [17] 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, ArXiv: 1506.04341, pp. 47–133.
  • [18] by same author, A compactness theorem for the dual Gromov-Hausdorff propinquity, Indiana Univ. Math. J. 66 (2017), no. 5, 1707–1753. MR 3718439
  • [19] by same author, The quantum Gromov-Hausdorff Propinquity, Trans. Amer. Math. Soc. (electronically published on May 22, 2015), : 49 Pages, http://dx.doi.org/10.1090/tran/6334, to appear in print, ArXiv: 1302.4058.
  • [20] Frédéric Latrémolière, Equivalence of quantum metrics with a common domain, J. Math. Anal. Appl. 443 (2016), no. 2, 1179–1195, ArXiv: 1604.00755.
  • [21] by same author, The triangle inequality and the dual Gromov-Hausdorff propinquity, Indiana Univ. Math. J. 66 (2017), no. 1, 297–313, ArXiv: 1404.6633.
  • [22] Linda Mawhinney and Ivan G. Todorov, Inductive limits in the operator system and related categories, Dissertationes Math. 536 (2018), 1–57, arXiv:1705.04663. MR 3898342
  • [23] A. W. Miller, Descriptive set theory and forcing: how to prove theorems about borel sets the hard way, Springer-Verlag, 1995.
  • [24] G. J. Murphy, C{C^{\ast}}-algebras and operator theory, Academic Press, San Diego, 1990.
  • [25] N. Ozawa and M. A. Rieffel, Hyperbolic group CC^{\ast}-algebras and free products CC^{\ast}-algebras as compact quantum metric spaces, Canad. J. Math. 57 (2005), 1056–1079, ArXiv: math/0302310.
  • [26] Vern I. Paulsen and Mark Tomforde, Vector spaces with an order unit, Indiana Univ. Math. J. 58 (2009), no. 3, 1319–1359. MR 2542089
  • [27] M. A. Rieffel, Metrics on states from actions of compact groups, Documenta Mathematica 3 (1998), 215–229, math.OA/9807084.
  • [28] by same author, Metrics on state spaces, Documenta Math. 4 (1999), 559–600, math.OA/9906151.
  • [29] by same author, Matrix algebras converge to the sphere for quantum Gromov–Hausdorff distance, Mem. Amer. Math. Soc. 168 (2004), no. 796, 67–91, math.OA/0108005.
  • [30] by same author, Matricial bridges for “matrix algebras converge to the sphere”, Operator algebras and their applications, Contemp. Math., vol. 671, Amer. Math. Soc., Providence, RI, 2016, ArXiv: 1502.00329, pp. 209–233.
  • [31] by same author, Gromov-Hausdorff distance for quantum metric spaces, Mem. Amer. Math. Soc. 168 (March 2004), no. 796, math.OA/0011063.