Bunce-Deddens algebras as quantum Gromov-Hausorff distance limits of circle algebras
Abstract.
We show that Bunce-Deddens algebras, which are A-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, 58B34.1. 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 A-algebras. A well-known example of an A-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 A-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 is that it induces by duality a distance on the space of Radon probability measures over 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 of an ordered vector space over such that for all there exists such that .
An order unit space is a vector space with an order unit which is Archimedean, i.e. for all , if for all then .
An order unit space comes equipped with a norm defined for all by and satisfying .
Definition 1.2.
The state space of an order unit space is the set of all positive linear functionals from to which maps the order unit of to .
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 is an ordered pair of a norm-complete order unit space and a seminorm defined on a norm-dense subspace of such that:
-
(1)
,
-
(2)
the Monge-Kantorovich metric, defined for any two by:
metrizes the weak* topology on ,
-
(3)
is closed in .
The seminorm is called a Lip-norm on .
We denote the class of quantum metric order unit space by .
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 satisfies (1) and (2) but not (3) in Definition (1.3), then setting:
allowing for , gives a quantum metric order unit space with and is lower semicontinuous on . 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 and be two quantum metric order unit spaces. A Lip-norm on is admissible for and when:
Notation 1.5.
If is a metric space, the Hausdorff distance [10] induced by on the set of all closed subsets of is denoted by .
Definition 1.6.
The quantum Gromov-Hausdorff distance between two quantum metric order unit spaces and is:
Rieffel proved in [31] that the distance 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 and if and only there exists a positive linear map such that . 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 -Leibniz property (where is a function which is increasing in the product order) is given by an ordered pair where:
-
(1)
is a unital C*-algebra,
-
(2)
is a seminorm defined on a dense domain of , where is the self-adjoint part of ,
-
(3)
is an quantum metric order unit space,
-
(4)
for all .
For any fixed function as above, the propinquity, denoted , is a complete metric on the class of quantum compact metric spaces with the -Leibniz property, with the property that the propinquity between two such quantum compact metric spaces and is null if and only if there exists a *-isomorphism such that , and 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 -Leibniz quantum compact metric spaces without fixed , which we denote , 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 A-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 . 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 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 . 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 . There is one more point of subtlety which we must address here. While Rieffel proved that is complete in [31], we need for our proof in this paper a different description of the limit of a Cauchy sequence for . 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 . 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 , inspired by the propinquity, can lead to new observations about .
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 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 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 , 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 under appropriate hypothesis. In the fifth section, we establish a new expression for 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 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 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 , containing the unit of , is the range of some conditional expectation on which is also contractive for , then we can always modify to make and arbitrarily close in the sense of the propinquity (though never at distance unless and 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 .
Lemma 2.1.
Let be an -Leibniz quantum compact metric space and let be a C*-subalgebra of which contains the unit of . If there exists a conditional expectation of onto such that , then for all , if we set for all :
then is a -Leibniz quantum compact metric space, for all , and:
Proof.
The domain of is the domain of since is continuous. Moreover, this also implies that is lower semi-continuous seminorm. Moreover, is -quasi-Leibniz by [3, Lemma 2.3], so is -Leibniz. Of course, if then so , and .
Last, for any :
and thus, as the set on the right hand side is totally bounded since is an L-seminorm, so it the set on the left hand side. We have thus shown that is a -Leibniz quantum compact metric space using [25, Proposition 1.3].
By assumption, for all so for all .
Let and .
For all and , we define:
If and then and . Thus . Since for all , we conclude that is a quantum isometry.
If and then , and moreover:
so . Again, since for all , we conclude that is a quantum isometry as well.
We thus gather that is a tunnel from to by [21, Definition 2.3]. It is of course -Leibniz. We now compute its extent.
Let . Let and note that , so the restriction of to is — in fact for this proof, any extension of to a state of would work (but since we have our conditional expectation here, we need not invoke the Hahn-Banach theorem). Let with . Then in particular, and thus:
Therefore .
Let now . Set be the restriction of to . Let with . Then and thus:
We conclude .
Consequently . 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 between two quantum compact metric spaces and is a unital *-morphism from to such that:
Theorem 2.3 ([20]).
Let be a unital *-morphism between two unital C*-algebras and . If and are, respectively, Lip-norms on and , the the following assertions are equivalent:
-
(1)
is a Lipschitz morphism from to ,
-
(2)
,
-
(3)
.
If is a -Lipschitz morphism between two quantum compact metric spaces, then it naturally becomes a contractive morphism from to . Yet if is actually injective and bi-Lipschitz, then it is usually not possible to adjust the quantum metrics to turn into a quantum isometry. However, under the hypothesis that we can find a conditional expectation of onto the range of in such that is also a Lipschitz linear map, then it is indeed possible to modify to turn into a quantum isometry.
Lemma 2.4.
Let and be quantum compact metric spaces, with an -Leibniz seminorm. If is a unital *-monomorphism and is a conditional expectation such that for some , the following conditions hold:
-
•
,
-
•
,
then if we set:
where is the inverse of the *-isomorphism , then is a -Leibniz quantum compact metric space and:
and .
If, moreover, there exists such that then:
In particular, if then is a full quantum isometry.
Proof.
Let . By assumption, and therefore . So . Therefore and since is dense in , so is . Moreover, if then , so . Thus .
If then so . Of course, since and are unital.
We now check the Leibniz property of . For all , we compute:
Therefore, is -Leibniz.
As the supremum of two lower semi-continuous seminorms, is a lower semi-continuous seminorm as well. Moreover, if , since:
and since 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 is a -quasi-Leibniz Lip-norm as a claimed.
By construction of :
Last, assume that there exists such that . We will note that, of course, if then , and thus:
If then for all . Thus is a full quantum isometry form onto . ∎
Remark 2.5.
If we drop the assumption that is -Leibniz, then we still can conclude that is a lower semi-continuous Lip-norm such that is -Lipschitz.
We now bring our two previous observations in one theorem which will be key to our construction.
Theorem 2.6.
Let and be quantum compact metric spaces, with a -Leibniz seminorm. If is a unital *-monomorphism and is a conditional expectation such that for some :
-
•
,
-
•
,
and if , then setting:
then
where .
Proof.
By Lemma (2.4), if we define then is a full quantum isometry from to . So:
The seminorm is then obtained by applying Lemma (2.1) to . In particular, for all , and:
Thus, by the triangle inequality, we conclude:
as desired. ∎
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 out of any inductive sequence of quantum compact metric spaces where the connecting maps are bi-Lipschitz.
Theorem 2.7.
Let:
be an inductive sequence of unital C*-algebras, where the connecting maps are unital *-monomorphisms, such that:
-
•
for each , we are given an -Leibniz Lip-norm on ,
-
•
for each , there exists a conditional expectation from onto such that ,
-
•
for each , there exists such that:
then, setting and for all :
where
then:
Consequently, there exists a quantum metric order unit space such that:
Proof.
This proof is by induction. Since , the map is a contraction from to , and . Thus by Theorem (2.6), if we set:
then is an -Leibniz quantum compact metric space for some , and .
Assume that, for some , we have now shown that and for all .
By assumption
Let and . There exists such that . By assumption, , Moreover, and
Thus is dense in . Again by our assumption, if , then . Thus by our induction hypothesis, we conclude . Now is a unital *-morphism from onto which maps to . By Theorem (2.3), we conclude that there exists such that . We will compute an estimate for in the next lemma but its actual value is not very important for us.
Thus, by Theorem (2.6), if we set:
then is an -Leibniz Lip-norm for some and:
as claimed. Moreover . Our induction hypothesis holds for all .
The conclusion of the theorem then follows from the observation that is complete and is dominated by 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 . 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 -Leibniz property as discussed after Definition 1.7. However, we are unable to find a uniform bound on the -Leibniz properties over all (even in the explicit setting of the Bunce-Deddens algebras) and propinquity is only complete over classes of quantum compact metric spaces with uniform -Leibniz property. But, this is where has an advantage. It is complete on the class of all quantum metric order unit spaces. Although doesn’t have the optimal coincidence property, it still has this one crucial advantage, and this is why the above theorem ends with rather than .
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 ,
- •
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 matrices over by .
Notation 3.2.
If then, for , we set:
and we set , and we denote .
Notation 3.3.
For each , let . Let .
-
•
For all and , we define
The function is a continuous, -periodic -valued function over .
-
•
We define the -valued continuous function:
In addition, we set . The function is a continuous -periodic -valued function. Moreover, by [8, Chapter V.3], the map is unitary.
Lemma 3.4.
For all , the map is a unitary and is -periodic and if :
Proof.
This is an immediate computation. ∎
Notation 3.5.
Let . For , the C*-algebra of -valued, continuous, -periodic functions over is denoted by .
We then define:
and note that is a -valued continuous function over .
Lemma 3.6.
The map is a unitary, -periodic function such that if then:
Proof.
This is an immediate computation. ∎
Notation 3.7.
Let . For , we define:
and note
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 and . If , then .
The map thus defined is a unital *-monomorphism from to .
Proof.
We use the notations of Lemma (3.6). If , with , …, elements of , then is the matrix where is the permutation .
By Lemma (3.6), we also have:
Let and . Since is -periodic, we note that
We then compute:
Thus is -periodic. It is of course a continuous function over valued in , so . Since is a unitary, it is immediate that is a unital *-monomorphism. ∎
Definition 3.9 ([8, Section V.3]).
The Bunce-Deddens algebra is the C*-algebra inductive limit [24, Section 6.1] of the sequence:
Remark 3.10.
The Bunce-Deddens algebra is denoted by in [8, Section V.3].
Notation 3.11.
Let . For each , we let
denote the canonical unital *-monomorphism such that given by [24, Section 6.1].
The Bunce-Deddens algebras have a unique faithful tracial state.
Notation 3.12.
At the core of our construction of a quantum metric on the Bunce-Deddens algebras 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 . We now construct this conditional expectation.
Notation 3.13.
If , and if we write with , …, all in , then we define:
Modeled on [3], the map is a conditional expectation from to the C*-subalgebra of block-diagonal matrices with blocks all square matrices of order .
Lemma 3.14.
Let and . If for all , we define:
then is a conditional expectation onto such that .
Proof.
If , with , …, elements of , then is the matrix where is the permutation . Thus in particular,
Hence commutes with (and similarly with ). Thus
By Lemma (3.6), we also have:
Let now — note that is -periodic. We then compute for any :
Therefore, is -periodic, and obviously continuous, so it is an element of .
Now, we wish to find such that . In particular, we wish to find such that:
To this end, if , and if we write with , …, all in , then for , we define:
Modeled on [3], the map is a unital completely positive contraction from onto . Also, a similar calculation to the one involving shows that for and .
Next, fix . For all , set
Since is continuous, then so is on .
Now, let . We have since is -periodic
A similar computation shows that . Thus, the map defined for all by
is well-defined and continuous and . Thus, extends uniquely to an element in , which we will still denote by . And, by construction, we have that .
Next, it remains to show that if , then . Let . Thus, there exists such that .
Hence, for all , we have
Now, is positive and contractive by construction, as the composition of *-isomorphisms and a conditional expectation. So is a conditional expectation onto by [5, Tomiyama’s Theorem 1.5.10 and Theorem 3.5.3].
Finally, following [8, Theorem V.3.6], we have since preserves the trace, :
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 be a locally compact metric space, and let . Let be a unital C*-subalgebra of the C*-algebra of bounded -valued continuous functions over , such that the unit is the unit of — the constant function equal to the identity in .
For all , we define:
Last, if is a normed vector space with norm , then we write for with the metric induced by .
We make two simple but important remarks:
-
•
,
-
•
.
Lemma 4.2.
Let . We use Notation (3.3).
-
•
For all , we estimate:
-
•
We have:
Proof.
Fix and . Of course . Hence, if , then:
Thus .
Let . Let with . We compute:
Now for all :
Now
Thus . Hence
This concludes our proof. ∎
We first define a natural -Leibniz Lip-norm on the circle algebras.
Definition 4.3.
Let and . We define :
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 and let . If then:
Proof.
Let and . To ease notations, we just write for the seminorm .
Since , we compute:
Furthermore, by Lemma 3.14:
Therefore:
This concludes our result. ∎
Theorem 4.5.
If and , then is a -quantum compact metric space.
Proof.
As the maximum of two lower semi-continuous seminorms, is a lower semi-continuous seminorm (allowing for the value ).
By [3, Lemma 2.3], the seminorm is (2,0)-quasi-Leibniz. As is of course Leibniz, we conclude that is -quasi-Leibniz.
If then , so there exists such that . On the other hand:
so , and therefore . Of course, .
If with and if then:
(4.1) |
using the standard *-isomorphism between and given on elementary tensors by [24, Theorem 6.4.17].
Using the same *-isomorphism, if and , then there exist and such that:
Let .
As Lipschitz functions are dense in , there exists such that while for all . Therefore:
Last, using the Leibniz property of and the fact that , we conclude that by Expression (4.1). This concludes the proof that the domain of is dense in since the other seminorm in its definition are actually continuous on .
Last, let be a sequence in such that and . for all . Since for all , the set is equicontinuous, and thus is equicontinuous as is unitary. Moreover, we also have:
So is an equicontinuous set of continuous -periodic functions on , all valued in the closed unit disk, which is compact. By Arzéla-Ascoli theorem, we thus conclude that is totally bounded for the norm (note: apply Arzéla-Ascoli theorem to the restriction of these functions to the compact and then conclude using periodicity). Therefore, the sequence admits a Cauchy subsequence . As is complete, converges to some . As is lower semi-continuous, we get .
Thus is compact. By [25, Proposition 1.3], we thus can conclude that 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 and . Let . If then:
Proof.
We denote by and by in this proof.
For , set:
We compute:
Therefore:
Now,
Thus since the seminorms and vanish on scalars, we have
Next, we note that
Since by construction, we conclude:
Let such that . By the above computation, we see that and . So:
Thus, since vanishes on scalars, we have
As above, we conclude:
This concludes our proof. ∎
Theorem 4.7.
If , and if for all we set on and for all :
where for all , we have , and:
then is a quantum metric order unit space and there exists a quantum metric order unit space such that:
Proof.
We apply Theorem (2.7) to the sequence and the conditional expectations . ∎
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 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 between two quantum metric order unit spaces and is a positive linear map which maps the order unit of to the unity order of , such that:
Definition 5.2.
If and are two quantum metric order unit spaces, then an order unit tunnel is an ordered quadruple such that is an quantum metric order unit space, while and are quantum order unit isometries from onto, respectively, and .
Definition 5.3.
The extent of an order unit tunnel from to is:
We remark that if is an admissible Lip-norm for and then we can form the tunnel:
The following lemma reconciles the extent of this tunnel with Rieffel’s computation of .
Lemma 5.4.
If and are two quantum metric order unit spaces, and if is a tunnel from to with for both , then:
Proof.
Write .
Let . There exists , , and such that . Now, there exists such that by definition of . Set . We then compute:
By symmetry in and , we conclude that .
On the other hand, let . Of course, with the usual identification, (with for all ). By definition of , there exists such that . As is arbitrary and by symmetry in and , we conclude . ∎
Thus, we obtain a new expression for Rieffel’s distance in the spirit of the propinquity.
Theorem 5.5.
If and are two quantum metric order unit spaces, then:
Proof.
Let be an order unit tunnel. Let . For all , we define:
If for some , then , and thus .
Let . By construction:
(5.1) |
Both factors in the Cartesian product on the right hand-side of Expression (5.1) are compact since is a Lip-norm, so the left hand side is a subset of a compact set in . Since , as the maximum of lower semi-continuous functions, is lower-semicontinuous (and since is continuous), the set:
is closed, and thus it is compact as well.
We thus have shown that is a Lip-norm using [25, Proposition 1.3].
For all , we set:
By [31], the seminorm — the quotient of for the map — is a Lip-norm on .
Now, let with . Since is a tunnel, there exists with and . Let . As is 1-Lipschitz, we have .
Thus the canonical surjection is a quantum isometry from onto . Similarly, is also a quantum isometry.
Let now . As is a tunnel, there exists such that . Let with . By definition of , there exists such that and , with . We then estimate:
Consequently, . By symmetry, we conclude:
Since is an arbitrary tunnel between and , we conclude:
On the other hand, if is an admissible Lip-norm on then , with and , is a tunnel from to . By Lemma (5.4), we then have: .
This completes our theorem. ∎
Remarkably if is a quantum isometry between two quantum metric order unit spaces and , it need not be a quotient map, in the following sense:
Definition 5.6.
A surjection between two normed vector spaces and is a quotient map when:
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 and are two order unit spaces, then the maps , for , 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 , 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 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, 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 . 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 is a closed convex subset of a topological -vector space , then the vector space of all the continuous affine functions from to is denoted by , where is affine when for all and , we have .
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 is a complete order unit space, and if for all we set , then the map is an order isomorphism.
Theorem 5.9.
Let be a sequence of quantum metric order unit spaces such that for all , there exists an order unit tunnel from to with .
Let:
endowed with for all .
Now, let:
and
Let be the closure of for .
Let .
The space is a complete order unit space, and is an order ideal of . There exists a weak* compact convex subset of such that and where is any metric which induces the weak* topology on the state space of .
If , then , endowed with the quotient order, is order-isomorphic to the order unit space of .
If, moreover, is a quotient map for all , then the norm induced by the quotient order on and the quotient norm are equal.
Remark 5.10.
We emphasize that has two possible orders: its quotient order and the order from its structure as a space of affine functions. This already endows with two potentially distinct seminorms. Moreover, 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 and are complete order unit spaces with order unit (note that is closed in by continuity if the maps and for all ).
Now, is a closed subspace of , so is a normed vector space with norm . Let be the canonical surjection. Moreover, if with , then for all , we have and thus , from which we readily conclude that . Thus is a order ideal in . Consequently, 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, is a positive linear map which maps the order unit of to the order unit of . We denote the quotient order on simply as , and its order unit as . There is also a seminorm induced on by the ordered vector space with an order unit structure on .
Any for any defines a state of by setting , and we will henceforth identify with its image for this map.
Now, by [16], the sequence converges, for the Hausdorff distance induced by any metric for the weak* topology on (which exists since is separable), to a weak* compact and convex set , and moreover:
We note in passing that the metric used in [16] to define is the Monge-Kantorovich metric induced by a Lip-norm on , but actually, the topologies induced on by the Hausdorff distance for any metrization of the weak* topology all agree with the Vietoris topology on the weak* closed subsets of , so the exact metric involved is not important for this part of the proof.
If , then and by construction, , thus . Conversely, if then induces a state on since . These two maps are inverse to each other and allow us to identify with .
For any , and for any , we set . The map is linear and maps to the order unit of , i.e. the constant function . This map is injective: if , then, for any with , we have and thus . Let .
We now prove that this injective linear map is positive.
Let such that . There exists and such that and . Let . Then and thus, for all , we have . So .
We now check that the inverse map of is also a positive map from onto .
Let such that , i.e. for all , we have . Let such that . Now, suppose that for some , for all , there exists such that .
By induction, there exists a strictly increasing function and, for each , there exists such that . By compactness of , there exists a weak* limit for a subsequence of . Now, by construction, since is the Hausdorff limit of . Now, . By assumption, . This is a contradiction. Hence, for all , there exists such that if then . Consequently, by an easy induction, we can find a sequence converging to and such that for all , we have . Set : by construction, , and . Therefore .
Consequently, is order-isomorphic to . In turn, this proves that is an order unit space, with state space , and as is complete, the map is onto as well. The quotient order seminorm on is given by — and it is a norm.
We now turn to the relationship between the order norm and the quotient norm .
First, let . Let . There exists such that and . In particular, and thus . We conclude that , and thus as is arbitrary.
In general, we do not have much more to say about these two norms. However, if, for all , the map is a quotient map — as is the case, for instance, when working with C*-algebras — then more can be concluded. Assume henceforth that is a quotient map for all .
Let . Let . Let such that and . Note that . We now make an observation.
Let . Since , and since is a quotient map, there exists such that and . Now, repeating this process, a simple induction show that we can find such that and . By construction, , and, in particular, .
We note in passing that, by construction:
so we actually have — though, for our proof, only the lower bound on the norm of matters.
Thus, starting from , and by definition of , there exists such that . If then , and thus . Thus, we have shown that for all , there exists such that:
Thus we can find a sequence of states and a strictly increasing function such that for all , and . By compactness, the sequence admits a weak* convergent subsequence, whose limit is denoted by . By definition of , we then have , and by construction, . We conclude that:
where the last equality follows from our proof that the order norm is indeed the order unit norm obtained from identifying with .
As is arbitrary, we conclude that .
We thus have shown . ∎
Corollary 5.11.
Using the hypothesis of Theorem (5.9), and setting for all :
then , when is endowed with the order norm , is a quantum metric order unit space and:
If for all , the surjections are quotient maps, then and thus we can identify with with no ambiguity.
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 , be two unital C*-algebras. A bridge from to is a 4-tuple such that
-
(1)
is a unital C*-algebra and ,
-
(2)
the set is non-empty, in which case is called the pivot, and
-
(3)
and 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 be a unital C*-algebra, and let be a unital C*-subalgebra of . We call the 4-tuple the evident bridge from to , where is the inclusion mapping and is the identity map.
Lemma 6.3 ([19, Definition 3.17]).
Let be two unital C*-algebras and let be two quantum metric order unit spaces. If a bridge from to is of the form , then the length of the bridge is
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 be two unital C*-algebras and let be two quantum metric order unit spaces. Let be a bridge from to . Fix any , where is the length of the bridge .
If we define for all
and we let and denote the canonical surjections, then is an order unit tunnel from to with length , and
This allows us to define:
Definition 6.5.
Let be two unital C*-algebras and let be two quantum metric order unit spaces. Let be a bridge from to . We call the order unit tunnel from to of Theorem 6.4 the -evident tunnel associated to the bridge , Lip-norms , and
Hypothesis 6.6.
Let be a unital C*-algebra such that for each , we have that is a unital C*-subalgebra of and . For each , let be a Lip-norm on , so that is a quantum metric order unit space. Assume for all that Let be a sequence of positive real numbers such that Let be the evident bridge from to , and assume . Denote the associated -evident tunnel by an denote and its Lip-norm by .
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.
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 in Proposition (6.7) looks like under Hypothesis 6.6.
Proposition 6.8.
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 for all , which are quotient maps. ∎
Now, we want to show that is order isomorphic to . This particular part came much more easily in [1] since 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.
Lemma 6.10.
Proof.
Fix . We have that is an isometry by construction. Since the order on is just the coordinate order induced by each , we have that is a positive map and thus an order isometry. We will now show that where was defined in Theorem 5.9. Let such that . If , then . So, assume . By construction, by Proposition 6.8, and we thus have . Therefore, , and thus
by continuity. ∎
This lemma allows us to define:
Definition 6.11.
Lemma 6.12.
Proof.
Fix . The map is linear by construction. For unital, we note that , and thus For injectivity, assume and . Thus . Hence which implies . Next, let . Then, again we have , and thus
Now, we will show that and its inverse defined on are positive. Note by Theorem 5.9, we use the quotient order on . We begin with showing that is positive. Let such that . By Lemma 6.10, we have that with respect to the order on , which is the same order on . Now, consider . We have that and . Thus with respect to the quotient order on . Hence is positive.
Next, we show that the inverse of on is positive. Let and assume that in the quotient order on . Thus, there exists such that . Now such that Now, we have that for all since . Hence for all , we have
Hence, as , we have that for all . Since the given order on is Archimedean, we have that
Thus is an order unit isomorphism onto its image. ∎
Finally, we are ready to build an order unit isomorphism from onto . This follows the same process of [1, Theorem 2.15] up to some crucial details concerning order unit spaces.
Theorem 6.13.
Proof.
The fact that there exists an order unit isomorphism such that the restriction of to is for all follows from [22, Remark 3.6 (ii)] and Lemma 6.12.
Next, we show Let . Let . There exists such that by density. Hence, there exists such that . Also, there exists such that .
Define in the following way:
Therefore and , which implies that .
Now, consider and recall that and that for all by Proposition 6.8. Therefore
Since , we have that for all by Proposition 6.8. Hence for all
Thus . Therefore, since as , we gather
by definition of quotient norm. In particular, the set is dense in with respect to the quotient norm. However, as our tunnels are built from quotient maps since they are built using the canonical surjections and (see the discussion after Definition 5.6), we have that the order norm and quotient norm equal by Theorem 5.9. As 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 is complete, it must be the case that .
As a corollary, we prove the previous result for the following description of inductive limits.
Corollary 6.14.
Let be an inductive limit of C*-algebras (see [24, Section 6.1]), where for each , is a unital C*-algebra and is a unital *-monomorphism. For each , let be quantum metric order unit space. Let be a summable sequence of positive real numbers.
If for each , the bridge has length then there exists a Lip-norm on such that for each ,
and thus
Proof.
By [24, Section 6.1], for each , let be the canonical unital *-monomorphism associated to , where is a unital C*-subalgebra of such that and .
Next, for each , we have is a quantum metric order unit space such that
(6.1) |
Now, consider the bridge . We will show that its length .
Let such that . Set for some . Then, we have that . Thus, by Lemma 6.3, there exists such that and Now, set and note Next, by [24, Section 6.1], we have
The argument is symmetric if one begins with the space . Thus, by Lemma 6.3, it holds that . Therefore, the proof is complete by Theorem 6.13 and Expression (6.1), and we denote by . ∎
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 .
We thus are now able to state a main result for this paper.
Theorem 6.15.
If , and if for all we set on and for all :
where for all , we have , and:
then is a quantum metric order unit space and there exists a Lip-norm on the Bunce-Deddens algebra such that:
Proof.
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 .
Theorem 6.16.
Proof.
Let such that . Set (so ). Hence for all . By induction, we note that for all .
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 -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, -algebras and operator theory, Academic Press, San Diego, 1990.
- [25] N. Ozawa and M. A. Rieffel, Hyperbolic group -algebras and free products -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.