Constructing equivalence bimodules between noncommutative solenoids: a two-pronged approach
Abstract.
We revisit and generalize the application of a method introduced by Latrémolière and Packer for constructing finitely generated projective modules over the noncommutative solenoid C*-algebras. By realizing them as direct limits of rotation algebras, the method constructs directed systems of equivalence bimodules between rotation algebras that satisfy the necessary compatibility conditions to build Morita equivalence bimodules between the direct limit C*-algebras. In the irrational case, we use a fixed projection in a matrix algebra over the rotation algebra satisfying a key condition to build an equivalence bimodule at each stage following a construction of Rieffel. From this, our main result shows that two irrational noncommutative solenoids are Morita equivalent if and only if such a projection exists. We also make additional observations about the Heisenberg bimodules construction studied by the aforementioned two authors and connect the two constructions.
Key words and phrases:
C*-algebras; direct limit; noncommutative solenoids; projective modules; -adic analysis2020 Mathematics Subject Classification:
Primary: 46M40, 46L08; Secondary 46L80, 19K141. Introduction
In this paper, we continue the study of the noncommutative solenoids introduced by Latrémolière and Packer [8]. For a fixed prime , they are defined as the twisted group C*-algebras on , the additive group of the ring of integers adjoining the multiplicative inverse of , with the discrete topology. These C*-algebras are one of the first examples of twisted group C*-algebras associated to non-compactly generated abelian groups with in-depth analysis. A focus of Latrémolière and Packer’s earlier work was on explicit constructions of finitely generated projective modules over the noncommutative solenoids [6, 7]. Since then, the same two authors have also studied these C*-algebras in the context of noncommutative metric geometry [5]. More recently, N. Brownlowe, M. Hawkins, and A. Sims introduced a class of Toeplitz extensions to these C*-algebras and investigated their associated Kubo-Martin-Schwinger (KMS) states [1]. Another reason to study these C*-algebras is that they are -stable and fit into the Elliott classification program, but the fact that their -groups are infinitely generated makes the study highly non-trivial. Yet another point of interest is that these C*-algebras can also be expressed as groupoid C*-algebras, which have recently proven to be of interest in the research program of R. Deeley, I. Putnam, and K. Strung [2, 3], which is also centered on classification.
This paper is organized as follows. In the preliminaries section, we recall the definition of noncommutative solenoids as twisted group C*-algebras. More importantly, we focus on the alternative description that realizes as a direct limit of rotation algebras. Specifically, fix prime and let be a sequence of real numbers from such that for all , for some . Then the sequence of rotation algebras converges to , where the connecting maps send the generators and of to the th powers of the corresponding generators of . Moreover, we can naturally associate to each a -adic integer that is given by . We will exploit this connection to -adic analysis throughout our study. Since for any , changing any entry of by an integer does not affect the resulting solenoid. For this reason, we introduce an additive group of real number sequences satisfying and define a unique noncommutative solenoid from as a direct limit of rotation algebras the same way as before. When is an irrational sequence, is a direct limit of irrational rotation algebras, which have been studied extensively. Since much is known about the structure of their projective modules, we primarily concern ourselves with the irrational noncommutative solenoids.
In Section 3, we recall a construction of Latrémolière and Packer for forming projective modules over noncommutative solenoids. The construction is originally due ot Rieffel [14] and commonly referred to as the Heisenberg bimodules. In summary, one can embed into the self-dual group as a discrete cocompact subgroup, where is the field of -adic numbers. The embedding can be done in such a way so that the twisted group C*-algebra associated to the image in (with the Heisenberg multiplier) is isomorphic to and the one associated to the annihilator of the image (with the conjugated Heisenberg multiplier) is isomorphic to for some . In this case, , suitably completed, is an equivalence bimodule between and . Our contribution here is to simply give an explicit formula for , using basic facts about the -adic numbers. This formula will be useful when we relate our second construction to the Heisenberg bimodules in the irrational case.
The main concern of this paper is the open problem of classifying the irrational noncommutative solenoids up to (strong) Morita equivalence. Since these C*-algebras are unital, and are Morita equivalent if and only if one is a full corner of the algebra of matrices over the other for a suitable . That is, there exists a projection such that . It is sufficient to assume that for some , as and the connecting maps are all unital embeddings. For the same reason, one can regard as a projection in for all . In Section 4, using such a projection and a formulation of Rieffel, we establish a Morita equivalence bimodule between and , where is the appropriate fractional linear transformation of for each . In [7], Latrémolière and Packer introduced a notion of directed systems of equivalence bimodules adapted to directed systems of unital C*-algebras that gives rise to an equivalence bimodule between the direct limit C*-algebras. In their definition, the bimodules embeddings must satisfy the necessary compatibility conditions in relation to the connecting maps in the direct limit C*-algebras. We define explicit embeddings of into that give us a directed system of equivalence bimodules between two sequences of irrational rotation algebras, each of which converges to a noncommutative solenoid. Much of Section 4 is devoted to checking the bimodule embeddings are compatible with the connecting maps required to form a noncommutative solenoid. Additionally, motivated by finding all such projections, we give a key condition on projections that is both necessary and sufficient to establish equivalence bimodules between solenoids using the directed systems construction.
In Section 5, we address the question of Morita equivalence for irrational noncommutative solenoids. Specifically, we conclude that and are Morita equivalent if and only if there exists a projection in a matrix algebra over that satisfies our key condition.
Notation 1.1.
Throughout this paper, the natural numbers include zero. Unless otherwise specified, we always follow the convention that if .
2. Review of the Noncommutative Solenoids
In this section, we recall the necessary definitions concerning the noncommutative solenoids.
Fix prime , and let
be the additive subgroup of consisting of rational numbers whose denominators are nonnegative integral powers of . Naturally, we insist that all elements of are written in their unique reduced form, that is, the exponent of in the denominator is minimal. We endow with the discrete topology.
Definition 2.1.
Fix prime . A noncommutative solenoid is a twisted group C*-algebra of the form
where is a multiplier of the group . To easy notation, we will denote by the discrete group .
Recall that a multiplier on a locally compact group is a Borel function , where is the unit circle in , satisfying
-
(1)
;
-
(2)
,
for all , with being the identity of . In other words, a multiplier is a normalized -cocycle for the trivial group action of on . The set of multipliers is denoted by and two multipliers , are cohomologous, or equivalent, if there is a Borel function such that for all . The set of multipliers cohomologous to the multiplier that is identically form a normal subgroup of , denoted by . The quotient group is denoted by . If two multipliers and are cohomologous, then the twisted group C*-algebras and are -isomorphic. It is then necessary to classify the multipliers up to their cohomology classes. This was first addressed in [8] and we record it here, as it is crucial to our computations in Section 3.
Theorem 2.2.
[8, Theorem 2.3] Fix prime , and the set
is a group under pointwise addition modulo one. There exists a group isomorphism such that if and , and if is a multiplier of equivalence class , then is cohomologous to the multiplier:
It follows immediately that and are cohomologous if and only if . Notice that changing any number of entries of by integer values does not change the multiplier, hence the noncommutative solenoid it defines. However, changing any entry of by a nonzero integer takes out of the group , so it is convenient to define the following group:
with pointwise addition. The map given by defines a surjective group homomorphism. Given any , the corresponding noncommutative solenoid can also be defined by any element in the fiber of under . Frequently, a noncommutative solenoid will be defined by some , and we make no distinction between and for any .
Remark 2.3.
It is easy to see that for any , determines for all . To uniquely determine an element in , it is sufficient to know infinitely many entries (for example, for all ). This is not the case for , as many different elements of could agree at infinitely many entries. However, it is straightforward to check that if and , both in and agree at infinitely many entries, then and .
Recall that for any , the rotation algebra is the universal C*-algebra generated by two unitaries and satisfying the relation . The following alternative characterization of the noncommutative solenoids as direct limits of rotation algebras will be crucial to us in Section 4. Specifically, it allows us to build equivalence bimodules over irrational noncommutative solenoids from equivalence bimodules over irrational rotation algebras.
Theorem 2.4.
[8, Theorem 3.7] Let , and for each , let be the unique homomorphism given by
The noncommutative solenoid is the direct limit .
Since the rotation algebras and are isomorphic for any , we can again replace with any satisfying without changing the direct limit defined from . When there is no confusion, we will take this alternative characterization as definition for the noncommutative solenoid.
Remark 2.5.
It is shown in [8, Proposition 3.3] that can also be written as , where is the -solenoid group and the action of on is given by
Hence it is a a groupoid C*-algebra corresponding to the associated transformation C*-algebra.
In addition to characterizing each embedding by the generators, it will be necessary to describe it in greater detail, especially when each is identified as the crossed product C*-algebra . For details of the following, see [17, page 68].
Since is the completion of the -algebra with the appropriate convolution product and involution, it suffices for our purposes to extend the embedding to the dense subset of spanned by
where
Lemma 2.6.
With notation as above, for any and ,
Proof.
In this realization, the generators and are given by and , respectively, where and for all . Following the proof of [17, Proposition 2.56], we denote by the embedding of into given by , then for any . Since spans a dense subalgebra of , and for any ,
it follows that for any . We compute that for any ,
∎
Extending by linearity, we have that for any finite sum ,
3. Forming Heisenberg Bimodules over Noncommutative Solenoids
3.1. The field of -adic numbers
We first take the standard algebraic approach to define the field of -adic numbers and refer to Chapter 1 of [15] for a more extensive exposition. For a fixed prime , recall that a -adic integer is a formal series with integral coefficients satisfying . Under the usual addition and multiplication, the ring of -adic integers form an integral domain with additive identity , and multiplicative identity . If , then there exists a unique natural number such that and for all . We call the order of and denote it by , with the usual convention that .
Lemma 3.1.
The group of invertible -adic integers consists exactly of the -adic integers of order . That is, if and only if .
Proof.
This is well known. ∎
The field of -adic numbers is then defined to be the field of fractions of . Each nonzero -adic number can be uniquely written as with possibly negative, , and for all . The fractional part of a -adic number is given by
Naturally, if and only if is a -adic integer. Extending the definition of order for -adic integers, the order of is given by . For each , if , then it is easy to deduce that .
Alternatively, taking an analytic approach, is the completion of with respect to the -adic absolute value : for with and both indivisible by (then is necessarily unique), . We note that the inclusion of into is an injective ring homomorphism with image exactly the set of -adic numbers with coefficients that are eventually periodic. Moreover, -adic arithmetic extends the ordinary arithmetic of the rationals.
Remark 3.2.
Unless otherwise specified, we identify a rational number with its image in the -adic numbers. This also means that when convenient, we identify elements of with the corresponding -adic numbers.
The following lemmas will be useful.
Lemma 3.3.
For , with ,
where for .
Proof.
This is trivial if , in which case and . For , observe that
and
It is now visible that ∎
Lemma 3.4.
Let be a -adic integer with and inverse . Then for all nonnegative integer ,
Proof.
This is equivalent to
for all nonnegative integer , which holds for any invertible -adic integer with inverse . ∎
3.2. Heisenberg bimodules of Rieffel
In this section, we only assume that is a sequence in with . Using a construction of Rieffel [14] known as the Heisenberg bimodules, explicit --equivalence bimodules are constructed in [6], where is isomorphic to a noncommutative solenoid. We will summarize the construction when applied to noncommutative solenoids and give a formula for such that in Theorem 3.6 at the end of the section.
As before, we denote by the discrete group . Consider the group . Since both and are self-dual, is self-dual. Specifically, It is shown in [6] that every character of is given by
for some , where
are characters of and , respectively. Again, is the fractional part of the -adic number .
Following Rieffel [14], the Heisenberg multiplier is defined to be
The symmetrized version of , denoted by , is the following multiplier on :
with for .
Now, for each pair of and , let be the embedding of as a lattice (discrete cocompact subgroup) into given by
We denote the image of in by . For each such embedding, the so-called annihilator of is defined as
Since is a lattice in , so is [14, Lemma 3.1].
Applying the result in [14, Theorem 2.15] to , and , we have that , suitably completed, has the structure of a --equivalence bimodule and implements the Morita equivalence between and . We now state the main theorem of this section.
Theorem 3.5.
Let be prime and let such that with and . Then is isomorphic to the noncommutative solenoid , where and . Since , we write (uniquely) . Via the Heisenberg equivalence bimodule of Rieffel, is Morita equivalent to a noncommutative solenoid , where and
Proof.
The statement of this theorem gives an explicit formula for the in [6, Theorem 5.6], with the additional assumption that . We note that in the proof of the aforementioned theorem, one only needs to assume . This means that is isomorphic even when is taken from . We simply need to show that is isomorphic to to conclude that the two noncommutative solenoids and are Morita equivalent due to Rieffel’s method.
It is shown in [6, Lemma 5.5] that
For convenience, we choose to substitute with and write
Let denote the embedding of into , so that the image of is . We then have
Note that for a -adic number of order ,
It is clear that if , then . If , then . It follows that for , either or . In either case, we have
By Theorem 2.2, the multiplier on is given by:
By the Lemma 3.3, , so we have
This shows that , for all , , , and in , as desired. Therefore, is isomorphic to . It follows that is Morita equivalent to . ∎
Thus, one sees that given any in satisfying and invertible in the -adic integers, we can find a (satisfying the same conditions), such that and are Morita equivalent. It is then not unexpected that if we started with , the formula in this theorem recovers exactly.
We conclude this section with the following slightly more general result, in which we only assume the -adic integer associated to is nonzero.
Theorem 3.6.
Let be prime and let such that with and . Then is isomorphic to the noncommutative solenoid , where and . Since , we write (uniquely) . Via the Heisenberg equivalence bimodule of Rieffel, is Morita equivalent to a noncommutative solenoid , where is given by
4. Directed systems of equivalence bimodules
4.1. Directed system of equivalence bimodules.
In [7], a notion of directed system of equivalence bimodules was introduced. We recall it here.
Definition 4.1.
[[7], Definition 3.1] Let
and
be two directed systems of unital C*-algebras, whose *-morphisms are all unital maps. A sequence is a directed system of equivalence bimodule adapted to the sequence and when is an --equivalence bimodule, whose - and -valued inner products are denoted respectively by and , for all , and such that the sequence
is a directed sequence of modules satisfying
(4.1) |
and
(4.2) |
with analogous but symmetric equalities holding for the viewed as left Hilbert -modules:
(4.3) |
and
(4.4) |
As stated in [7, Theorem 3.4], Definition 4.1 provides the necessary structure to construct an equivalence bimodule that implements the Morita equivalence of the direct limit C*-algebras.
In this section, we present a construction for producing families of irrational noncommutative solenoids that are Morita equivalent to a given one.
In short, first fix an irrational noncommutative solenoid and a projection from an matrix algebra over , both satisfying certain conditions. We can then form a noncommutative solenoid from the directed sequence of irrational rotation algebras with connecting maps given by Theorem 2.4. We will show that , with the appropriate connecting maps, is a directed system of equivalence bimodules adapted to the sequences and . This allows us to state this section’s main result as Theorem 5.3, which generalizes the results in Section 4 of [7].
For the connecting maps between the approximating bimodules, it requires algebraically intensive steps to check that they satisfy Equations 4.1 through 4.4. Much of the current section is devoted to this.
We begin by establishing the equivalence bimodules at each stage.
Notation 4.2.
Let be an irrational sequence and for all . Let be an irrational noncommutative solenoid realized as the direct limit of the sequence of irrational rotation algebras . We denote the unique tracial state on by . We use the same notation for the standard unnormalized trace on , . Note that is then the unique tracial state on .
Let be a nontrivial projection in with that satisfies the following condition:
Condition 4.3.
.
We follow the convention that for any positive integer . For details of the existence and construction of such a projection, see [11]. When is considered as a projection in for any , it is not hard to see that , but , so the trace of there is
We let
so that in .
The follow lemma states Condition 4.3 in a form that will be useful to us later.
Lemma 4.4.
If , then for all .
Proof.
First, it is clear that implies , as any common divisor of and divides both and .
Fix . We show that if , then . It then follows that implies for all .
Assume , then either divides , or divides for some divisor of that is strictly greater than . In the former case, for some integer ,
so . This makes a common divisor of and . In the latter case, , making a common divisor of and . This shows that implies for all . ∎
4.2. Review of Rieffel’s standard bimodule between irrational rotation algebras
Notation 4.6.
Throughout this section, we fix an irrational number and let and be a pair of integers that generate with and . Set . Moreover, we let and be any two integers such that , and let .
For any , let be the standard --equivalence bimodule defined by Rieffel in [13]. We start with a lemma that provides an alternative presentation of this bimodule.
Lemma 4.7.
[4, Lemma 6] With notation as above, let be a projection in , , with unnormalized trace , then is isomorphic to as a right -module.
Since is a nontrivial projection in the simple C*-algebra , is full and is a --equivalence bimodule. It follows from this observation and the lemma above that as C*-algebras.
We now recall Rieffel’s construction of the standard --equivalence bimodule for the case , which is sufficient for our purposes. The rest of this section is based on [13, Theorem 1.1] and its proof. It is an application of Phil Green’s Symmetric Imprimitivity Theorem that is proved in [12, Situation 10], or see [10].
Theorem 4.8.
[13, Theorem 1.1] With notation as above, let (If is a negative integer, we still define to be ) and consider the following subgroups of :
Let act on (the right cosets of ) by right translation, and let act on (the left cosets of ) by left translation. Then the transformation group C*-algebras and are isomorphic to and , respectively. Furthermore, , suitably completed and structured, provides an --equivalence bimodule.
Both isomorphisms rely on identifying the quotient group with the circle group and realizing the corresponding group action of (both and are isomorphic to ) gives rise to a rotation algebra. In particular, we employ the following identification of and with . The map given by
is a group homomorphism with kernel exactly . We can then identify with . Similarly, the map given by
is a group homomorphism with kernel exactly , so we can identify with .
Remark 4.9.
We use an inverse identification of with than what is presented in [13]. Over there, the identification is derived from the map given by . One of reasons for choosing this inverse identification is to later have , instead of , where and are the usual generating unitaries of .
In the following two lemmas, we record the formulas for the generators of , the left -action on , and the -valued inner product, all under the present identification of with . To ease notation, we just write to represent . These formulas are derived originally in [10] and applied to (matrix algebras over) rotation algebras in [13]. In the proof of [7, Proposition 4.2], they are given for the case where , and , for any .
For , we denote by the class of in . In the second lemma, the left -action is define on the dense subspace of , and is identified with via the map .
Lemma 4.10.
The generators and of are given by
and
We compute that
and
It follows that . After writing as , a few steps of algebra will show that , so we have .
Lemma 4.11.
For any , , and ,
(4.5) |
and
(4.6) |
In particular, it follows that the generators and act on in the following way:
(4.7) |
and
(4.8) |
The right -module structure is given by the following lemmas.
Lemma 4.12.
The generators and of are given by
and
Lemma 4.13.
For any , , and ,
(4.9) |
and
(4.10) |
In particular, it follows that the generators and act on in the following way:
(4.11) |
and
(4.12) |
4.3. Explicit construction of equivalence bimodules at each stage.
Proposition 4.14.
We follow the notation established in Notation 4.2 . Let be a projection in with . Assume satisfies Condition 4.3. For each , let and be a pair of integers satisfying and set
In particular, we can uniquely choose and so that is between and . Then for each , , suitably completed, has the structure of a --equivalence bimodule.
Moreover, we can form a noncommutative solenoid by taking the direct limit of irrational rotation algebras with the appropriate connecting maps.
Proof.
Since each connecting map is a unital inclusion, we can view as a projection in for all . By Lemma 4.4, and together generate . The --equivalence bimodule structure is then the one from Theorem 4.8.
Since and are coprime, we can find integers and such that . Moreover, if and are a pair of such integers, as are and for any integer . If , then
It follows that there is a unique pair of and putting is between and .
We show that the family of irrational rotation algebras form a noncommutative solenoid by taking the direct limit with the prescribed connecting maps. To this end, we simply need to check that for all .
To ease notation, we denote , , , , , , , and by , , , , , , , and , respectively. By definition of , and for some of . We have
We show that is an integer. With a few steps of algebra, one checks that
Note that , so , and arrive at the following simplification:
Since , it must be the case that for some integer . Thus, . Note that if both and are chosen to be between and , then . ∎
Remark 4.15.
In the above proposition, the sequence determines a unique element of by Remark 2.3. By the same remark, all other choices for the ’s and ’s give rise to sequences in that lead to the same noncommutative solenoid .
Let denote the --equivalence bimodule given in the above proposition. When working with , results are almost always formulated in terms for the following dense subset. For and , let be the function in given by
One can regard as the indicator function on . Then the set
spans a dense subalgebra of .
We now define the key embedding of in for each that gives us the directed systems of equivalence bimodules.
Definition 4.16.
We define on functions of the form and extend by linearity (Since is finite, extending by linearity is especially simple):
(4.13) |
for and . That is, evaluates to when for and , and zero otherwise.
It remains to show that satisfies Definition 4.1. Specifically, we need to check that Equations 4.1 - 4.4 are satisfied.
Notation 4.17.
Lemma 4.18.
For each , preserves left module action. That is,
(4.14) |
for all and .
Proof.
We check that respects left module action by showing that for any and ,
Since , we have (see Notation 4.6), and we simply denote this number by .
For any , it follows from Equation 4.7 that
On the other hand,
On the other hand, using the fact that , we have
which is nonzero for for any and . Notice that the evaluation is independent of the choice of representative for , so we let and
Moreover, since for some ,
with integer . This shows that
and allows us to the conclude that . ∎
Lemma 4.19.
For each , preserves right module action. That is,
(4.15) |
for all and .
Proof.
We check that respects left module action by showing that for any and ,
Since , we have (see Notation 4.6), and we simply denote this number by .
For any , it follows from Equation 4.11 that
On the other hand,
Since ,
For , goes through the same equivalence classes of as , so
and .
On the other hand, since , we have
which is nonzero for for any and . Choose (the computations are independent of this choice), we have that for ,
The last equality follow from . In conclusion, . ∎
We now prove the following major lemma of this paper.
Lemma 4.20.
For each , preserves left inner product. That is,
(4.16) |
for all .
We first give a formula for in the following lemma.
Lemma 4.21.
Fix and . Let and for any ,
and otherwise.
Proof.
We begin by computing for and . For any , it follow from Equation 4.5 that
First, the summand associated to is only nonzero when . Since , if and only if , so
Since is only nonzero when , is only nonzero when for , in which case
and otherwise. The formula for then follows from the embedding of into characterized in Lemma 2.6. ∎
Proof of Lemma 4.20.
We now compute . For a fixed pair of and in and , we have
and
For and , we compute
We focus on the inner product involving and and compute
when for , and the inner product evaluates to zero otherwise.
Since and are both integers between and , must be for some . Specifically, there are pairs of , with , such that , and pairs of , with , such that . Together, there are exactly nonzero inner products contributing to the value of for . Therefore, for each fixed , there are nonzero inner products, each with value
Therefore,
Since is relatively prime to , as it is relatively prime to and goes through the modular classes of exactly once as goes from to . Therefore,
so
Since and are fixed, goes through the integers exactly once as runs through , so the last equality follows.
Lastly, as goes from to , and goes over all integers, the sum will go over the integers exactly once, so we have that, for ,
and otherwise, which is exactly in Lemma 4.21. We conclude that
∎
The main result of this section is the following.
Theorem 4.22.
Fix prime and let be an irrational sequence with for all . Let be a projection in , , that satisfies Condition 4.3. Then is a directed system of equivalence bimodules adapted to the sequence and .
Proof.
We showed in Proposition 4.14 that is an --equivalence bimodule for each . By Lemma 4.18 and 4.20, each embedding preserves left module action and left inner product, that is,
for and , , and in . Moreover, Lemma 4.19 shows that each embedding respects the right modules action.
It only remains to show that for all . First, observe that for any , we have
so
or
By Lemma 4.20, the left side of the equality is equal to , which is equal to . Therefore, for any fixed ,
for all . This is sufficient for us to conclude that
for all as follows. We have shown in Lemma 4.20 that . One checks that . Therefore, we can regard as a --equivalence bimodule. By [9, Proposition 3.8], the map from to the compact operators on given by is an isomorphism of C*-algebras. Let and , we have on all of , which implies that , as we desired to show. ∎
Corollary 4.23.
Let and be given as in Theorem 4.22. Then and are Morita equivalent.
Proof.
This follows directly from [7, Theorem 3.4] by taking the direct limit of . ∎
4.4. Relating back to the Heisenberg bimodules.
We now state how the two constructions of equivalence bimodules between noncommutative solenoids are related. Fix a noncommutative solenoid with and , for all . Assume also that and , so the associated -adic integer has inverse .
For such a fixed , the Morita equivalent solenoid that arises from the Heisenberg bimodules construction (see Theorem 3.5) is given by
Following Remark 4.5 and Notation 4.2, the projection with satisfies Condition 4.3, so by Proposition 4.14, the noncommutative solenoid is determined by
with
By Lemma 3.4, is an integer and one checks that for all . Observe that, since ,
which are exactly the even entries of calculated from the Heisenberg bimodules construction. By Remark 2.3, since the two ’s agree at infinitely many entries, they must determine the same noncommutative solenoid.
We summarize the above discussion as the following result.
Proposition 4.24.
Fix a noncommutative solenoid with and , for all . Assume also that and . The Morita equivalent solenoid formed from the directed system of equivalence bimodule construction, using projection with , is the same as the one from the Heisenberg bimodule construction.
5. Morita equivalence of irrational noncommutative solenoids
We are now ready to prove our main result, which addresses the Morita equivalence problem for irrational noncommutative solenoids. We start with the following observation.
Lemma 5.1.
Let be an irrational noncommutative solenoid, then it does not contain a unital C*-subalgebra isomorphic to for any strictly greater than .
Proof.
In [6], Latrémolière and Packer showed that an irrational noncommutative solenoid is simple and has a unique tracial state, which we denote by . Moreover, the range of this trace on is given by . The restriction of to any unital C*-subalgebra B of is again a tracial state on . For a contradiction, suppose that for , then the restriction of must be the unique tracial state on . Since is a projection with , the range on must contain the fraction and we arrive at a contradiction. ∎
It follows from the lemma below that every projection in is unitarily equivalent to a projection for some .
Lemma 5.2.
Let be the direct limit of the direct sequence of C*-algebras and suppose that is the natural map for each . If is a projection in , then there is a and a projection such that and are unitarily equivalent.
Proof.
Theorem 5.3.
Let , be prime numbers. Let and . Then the following statements are equivalent.
-
(1)
and are Morita equivalent.
-
(2)
and there exists a projection for some that satisfies Condition 4.3, and .
Remark 5.4.
It is clear that truncating the first terms of does not change the resulting noncommutative solenoid. For this reason, we say that a projection satisfies Condition 4.3 to mean: and . Essentially, our results in the previous section apply to any projection in for any by truncating the first terms of starting from zero. Without loss of generality, we can always assume is even.
Proof of Theorem 5.3.
implying follows from the directed system of equivalence bimodules construction, as in Theorem 4.22 and its Corollary, with projection and truncated sequence .
We show that implies as follows. In [8, Theorem 3.7], the -theory of noncommutative solenoids is computed: in particular, for ,
It is clear that for and primes, if and only if . Since Morita equivalent C*-algebras have isomorphic groups, we must have .
Since both and are unital, for a suitable and a projection [11, Proposition 2.1]. Recall that is the direct limit of the sequence of increasing irrational rotation algebras associated to the ’s. By Lemma 5.2, we can then assume that for a large enough . It is evident that
It remains to show that must satisfy Condition 4.3. Assume, for the sake of contradiction, that does not satisfy the Condition. Then by Lemma 4.4, there exists such that , with . As discussed in the paragraph following Lemma 4.7, is then isomorphic to an matrix algebra over some irrational rotation algebra, say . Moreover, it is easy to check that is a common divisor of and for all . So the direct limit C*-algebra contains as a subalgebra. This is a contradiction to Lemma 5.1. ∎
We end this section with an example where Condition 4.3 is not satisfied. In particular, we see that the direct limit of can be a direct limit C*-algebra of matrix algebras that are strictly increasing in size.
Example 5.5.
For , consider the irrational noncommutative solenoid given by
Let be the projection in with trace . As a projection in . By Lemma 4.7, . Therefore, by Lemma 5.1, cannot be a noncommutative solenoid. However, it would be interesting to further investigate the structure of the C*-algebras , and how they are related to the noncommutative solenoids.
Acknowledgments
This work makes up part of the author’s Ph.D. thesis for the University of Colorado Boulder. The author would like to thank his thesis advisor Dr. Judith Packer. He is beyond grateful for her continued support and guidance. This work would not have come to fruition without her.
References
- [1] Nathan Brownlowe, Mitchell Hawkins, and Aidan Sims. The Toeplitz noncommutative solenoid and its Kubo-Martin-Schwinger states. Ergodic Theory Dynam. Systems, 39(1):105–131, 2019. Corrigendum, 39(9):2592, 2019.
- [2] Robin J. Deeley, Ian F. Putnam, and Karen R. Strung. Constructing minimal homeomorphisms on point-like spaces and a dynamical presentation of the Jiang-Su algebra. J. Reine Angew. Math., 742:241–261, 2018.
- [3] Robin J. Deeley, Ian F. Putnam, and Karen R. Strung. Constructions in minimal amenable dynamics and applications to the classification of -algebras, 2019.
- [4] Kazunori Kodaka. Endomorphisms of certain irrational rotation -algebras. Illinois J. Math., 36(4):643–658, 1992.
- [5] Frédéric Latrémolière and Judith Packer. Noncommutative solenoids and the Gromov-Hausdorff propinquity. Proc. Amer. Math. Soc., 145(5):2043–2057, 2017.
- [6] Frédéric Latrémolière and Judith A. Packer. Noncommutative solenoids and their projective modules. In Commutative and noncommutative harmonic analysis and applications, volume 603 of Contemp. Math., pages 35–53. Amer. Math. Soc., Providence, RI, 2013.
- [7] Frédéric Latrémolière and Judith A. Packer. Explicit construction of equivalence bimodules between noncommutative solenoids. In Trends in harmonic analysis and its applications, volume 650 of Contemp. Math., pages 111–140. Amer. Math. Soc., Providence, RI, 2015.
- [8] Frédéric Latrémolière and Judith A. Packer. Noncommutative solenoids. New York J. Math., 24A:155–191, 2018.
- [9] Iain Raeburn and Dana P. Williams. Morita equivalence and continuous-trace -algebras, volume 60 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 1998.
- [10] Marc A. Rieffel. Strong Morita equivalence of certain transformation group -algebras. Math. Ann., 222(1):7–22, 1976.
- [11] Marc A. Rieffel. -algebras associated with irrational rotations. Pacific J. Math., 93(2):415–429, 1981.
- [12] Marc A. Rieffel. Applications of strong Morita equivalence to transformation group -algebras. In Operator algebras and applications, Part I (Kingston, Ont., 1980), volume 38 of Proc. Sympos. Pure Math., pages 299–310. Amer. Math. Soc., Providence, R.I., 1982.
- [13] Marc A. Rieffel. The cancellation theorem for projective modules over irrational rotation -algebras. Proc. London Math. Soc. (3), 47(2):285–302, 1983.
- [14] Marc A. Rieffel. Projective modules over higher-dimensional noncommutative tori. Canad. J. Math., 40(2):257–338, 1988.
- [15] Alain M. Robert. A course in -adic analysis, volume 198 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2000.
- [16] N. E. Wegge-Olsen. -theory and -algebras. Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, 1993. A friendly approach.
- [17] Dana P. Williams. Crossed products of -algebras, volume 134 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2007.