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

Constructing equivalence bimodules between noncommutative solenoids: a two-pronged approach

Shen Lu Department of Mathematics, University of Colorado Boulder, Campus Box 395, Boulder, CO 80309-0395, USA shen.lu@colorado.edu
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; pp-adic analysis
2020 Mathematics Subject Classification:
Primary: 46M40, 46L08; Secondary 46L80, 19K14

1. Introduction

In this paper, we continue the study of the noncommutative solenoids introduced by Latrémolière and Packer [8]. For a fixed prime pp, they are defined as the twisted group C*-algebras on [1p]\mathbb{Z}\left[\frac{1}{p}\right], the additive group of the ring of integers adjoining the multiplicative inverse of pp, 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 𝒵\mathcal{Z}-stable and fit into the Elliott classification program, but the fact that their KK-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 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} as twisted group C*-algebras. More importantly, we focus on the alternative description that realizes 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} as a direct limit of rotation algebras. Specifically, fix prime pp and let α=(αn)n\alpha=\left(\alpha_{n}\right)_{n\in\mathbb{N}} be a sequence of real numbers from [0,1)[0,1) such that for all nn\in\mathbb{N}, pαn+1=αn+xnp\alpha_{n+1}=\alpha_{n}+x_{n} for some xn{0,,p1}x_{n}\in\{0,\dotsc,p-1\}. Then the sequence of rotation algebras Aα0Aα2Aα4A_{\alpha_{0}}\rightarrow A_{\alpha_{2}}\rightarrow A_{\alpha_{4}}\rightarrow\cdots converges to 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}}, where the connecting maps send the generators Uα2nU_{\alpha_{2n}} and Vβ2nV_{\beta_{2n}} of Aα2nA_{\alpha_{2n}} to the ppth powers of the corresponding generators of Aα2n+2A_{\alpha_{2n+2}}. Moreover, we can naturally associate to each α\alpha a pp-adic integer that is given by xα=j=0xjpjx_{\alpha}=\sum_{j=0}^{\infty}x_{j}p^{j}. We will exploit this connection to pp-adic analysis throughout our study. Since AθAθ+nA_{\theta}\cong A_{\theta+n} for any nn\in\mathbb{Z}, changing any entry of α\alpha by an integer does not affect the resulting solenoid. For this reason, we introduce an additive group Ωp\Omega_{p} of real number sequences satisfying pαn+1αnmodp\alpha_{n+1}\equiv\alpha_{n}\mod\mathbb{Z} and define a unique noncommutative solenoid from αΩp\alpha\in\Omega_{p} as a direct limit of rotation algebras the same way as before. When α\alpha is an irrational sequence, 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} 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 [1/p]×[1/p]\mathbb{Z}\left[1/p\right]\times\mathbb{Z}\left[1/p\right] into the self-dual group M=×pM=\mathbb{R}\times\mathbb{Q}_{p} as a discrete cocompact subgroup, where p\mathbb{Q}_{p} is the field of pp-adic numbers. The embedding can be done in such a way so that the twisted group C*-algebra associated to the image in MM (with the Heisenberg multiplier) is isomorphic to 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} and the one associated to the annihilator of the image (with the conjugated Heisenberg multiplier) is isomorphic to 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} for some βΩp\beta\in\Omega_{p}. In this case, Cc(M)C_{c}\left(M\right), suitably completed, is an equivalence bimodule between 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} and 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}}. Our contribution here is to simply give an explicit formula for β\beta, using basic facts about the pp-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, 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} and 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} are Morita equivalent if and only if one is a full corner of the algebra of k×kk\times k matrices over the other for a suitable kk. That is, there exists a projection PMk(𝒜α𝒮)P\in M_{k}\left(\mathscr{A}_{\alpha}^{\mathscr{S}}\right) such that 𝒜β𝒮PMk(𝒜α𝒮)P\mathscr{A}_{\beta}^{\mathscr{S}}\cong PM_{k}\left(\mathscr{A}_{\alpha}^{\mathscr{S}}\right)P. It is sufficient to assume that PMk(Aα2N)P\in M_{k}\left(A_{\alpha_{2N}}\right) for some NN\in\mathbb{N}, as 𝒜α𝒮=limAα2n\mathscr{A}_{\alpha}^{\mathscr{S}}=\varinjlim A_{\alpha_{2n}} and the connecting maps are all unital embeddings. For the same reason, one can regard PP as a projection in Mk(Aα2n)M_{k}\left(A_{\alpha_{2n}}\right) for all nNn\geq N. In Section 4, using such a projection and a formulation of Rieffel, we establish a Morita equivalence bimodule X2nX_{2n} between Aα2nA_{\alpha_{2n}} and Aβ2nA_{\beta_{2n}}, where β2n\beta_{2n} is the appropriate fractional linear transformation of α2n\alpha_{2n} for each nn. 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 X2nX_{2n} into X2n+2X_{2n+2} 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 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} and 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} are Morita equivalent if and only if there exists a projection in a matrix algebra over 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} that satisfies our key condition.

Notation 1.1.

Throughout this paper, the natural numbers \mathbb{N} include zero. Unless otherwise specified, we always follow the convention that j=in=0\sum_{j=i}^{n}\cdots=0 if n<in<i.

2. Review of the Noncommutative Solenoids

In this section, we recall the necessary definitions concerning the noncommutative solenoids.

Fix prime pp, and let

[1p]:={jpk:j,k}\mathbb{Z}\left[\frac{1}{p}\right]:=\left\{\frac{j}{p^{k}}\in\mathbb{Q}\>:j\in\mathbb{Z},k\in\mathbb{N}\right\}

be the additive subgroup of \mathbb{Q} consisting of rational numbers whose denominators are nonnegative integral powers of pp. Naturally, we insist that all elements of [1/p]\mathbb{Z}\left[1/p\right] are written in their unique reduced form, that is, the exponent of pp in the denominator is minimal. We endow [1/p]\mathbb{Z}\left[1/p\right] with the discrete topology.

Definition 2.1.

Fix prime pp. A noncommutative solenoid is a twisted group C*-algebra of the form

C([1p]×[1p],σ),C^{\ast}\left(\mathbb{Z}\left[\frac{1}{p}\right]\times\mathbb{Z}\left[\frac{1}{p}\right],\sigma\right),

where σ\sigma is a multiplier of the group [1/p]×[1/p]\mathbb{Z}\left[1/p\right]\times\mathbb{Z}\left[1/p\right]. To easy notation, we will denote by Γ\Gamma the discrete group [1/p]×[1/p]\mathbb{Z}\left[1/p\right]\times\mathbb{Z}\left[1/p\right].

Recall that a multiplier σ\sigma on a locally compact group GG is a Borel function G×G𝕋G\times G\rightarrow\mathbb{T}, where 𝕋\mathbb{T} is the unit circle in \mathbb{C}, satisfying

  1. (1)

    σ(r,s)σ(r+s,t)=σ(r,s+t)σ(s,t)\sigma(r,s)\sigma(r+s,t)=\sigma(r,s+t)\sigma(s,t);

  2. (2)

    σ(s,e)=σ(e,s)=1\sigma(s,e)=\sigma(e,s)=1,

for all r,s,tGr,s,t\in G, with ee being the identity of GG. In other words, a multiplier is a normalized 22-cocycle for the trivial group action of GG on 𝕋\mathbb{T}. The set of multipliers is denoted by Z2(G,𝕋)Z^{2}(G,\mathbb{T}) and two multipliers σ\sigma, τ\tau are cohomologous, or equivalent, if there is a Borel function ρ:G𝕋\rho:G\rightarrow\mathbb{T} such that σ(s,t)=ρ(s)ρ(t)ρ(s+t)1τ(s,t)\sigma(s,t)=\rho(s)\rho(t)\rho(s+t)^{-1}\tau(s,t) for all s,tGs,t\in G. The set of multipliers cohomologous to the multiplier that is identically 11 form a normal subgroup of Z2(G,𝕋)Z^{2}(G,\mathbb{T}), denoted by B2(G,𝕋)B^{2}(G,\mathbb{T}). The quotient group is denoted by H2(G,𝕋)H^{2}(G,\mathbb{T}). If two multipliers σ\sigma and τ\tau are cohomologous, then the twisted group C*-algebras C(G,σ)C^{\ast}(G,\sigma) and C(G,τ)C^{\ast}(G,\tau) are \ast-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 pp, and the set

Ξp:={(αn)n:α0[0,1) and n,xn{0,,p1} such that pαn+1=αn+xn}\Xi_{p}:=\{(\alpha_{n})_{n\in\mathbb{N}}\>:\alpha_{0}\in[0,1)\text{ and }\forall n\in\mathbb{N},\exists x_{n}\in\{0,\dotsc,p-1\}\text{ such that }p\alpha_{n+1}=\alpha_{n}+x_{n}\}

is a group under pointwise addition modulo one. There exists a group isomorphism ρ:H2(Γ,𝕋)Ξp\rho:H^{2}(\Gamma,\mathbb{T})\rightarrow\Xi_{p} such that if σH2(Γ,𝕋)\sigma\in H^{2}(\Gamma,\mathbb{T}) and α=ρ(σ)\alpha=\rho\left(\sigma\right), and if ff is a multiplier of equivalence class σ\sigma, then ff is cohomologous to the multiplier:

Ψα:{Γ×Γ𝕋((j1pk1,j2pk2),(j3pk3,j4pk4))exp(2πiα(k1+k4)j1j4).\Psi_{\alpha}:\begin{cases}\qquad\qquad\quad\Gamma\times\Gamma&\rightarrow\mathbb{T}\\ \left(\left(\dfrac{j_{1}}{p^{k_{1}}},\dfrac{j_{2}}{p^{k_{2}}}\right),\left(\dfrac{j_{3}}{p^{k_{3}}},\dfrac{j_{4}}{p^{k_{4}}}\right)\right)&\mapsto\exp\left(2\pi i\alpha_{(k_{1}+k_{4})}j_{1}j_{4}\right).\end{cases}

It follows immediately that Ψα\Psi_{\alpha} and Ψβ\Psi_{\beta} are cohomologous if and only if α=βΞp\alpha=\beta\in\Xi_{p}. Notice that changing any number of entries of αΞp\alpha\in\Xi_{p} by integer values does not change the multiplier, hence the noncommutative solenoid it defines. However, changing any entry of α\alpha by a nonzero integer takes α\alpha out of the group Ξp\Xi_{p}, so it is convenient to define the following group:

Ωp:={(αn)n:n,xn such that pαn+1=αn+xn},\Omega_{p}:=\{(\alpha_{n})_{n\in\mathbb{N}}\in\mathbb{R}^{\mathbb{N}}\>:\forall n\in\mathbb{N},\exists x_{n}\in\mathbb{Z}\text{ such that }p\alpha_{n+1}=\alpha_{n}+x_{n}\},

with pointwise addition. The map h:ΩpΞph:\Omega_{p}\rightarrow\Xi_{p} given by h((αn)n)=(αnmod)nh\left((\alpha_{n})_{n\in\mathbb{N}}\right)=\left(\alpha_{n}\mod\mathbb{Z}\right)_{n\in\mathbb{N}} defines a surjective group homomorphism. Given any αΞp\alpha\in\Xi_{p}, the corresponding noncommutative solenoid 𝒜α𝒮=C(Γ,Ψα)\mathscr{A}^{\mathscr{S}}_{\alpha}=C^{*}(\Gamma,\Psi_{\alpha}) can also be defined by any element in the fiber of α\alpha under hh. Frequently, a noncommutative solenoid will be defined by some αΩp\alpha\in\Omega_{p}, and we make no distinction between C(Γ,Ψα)C^{\ast}(\Gamma,\Psi_{\alpha}) and C(Γ,Ψh(α))C^{\ast}(\Gamma,\Psi_{h\left(\alpha\right)}) for any αΩp\alpha\in\Omega_{p}.

Remark 2.3.

It is easy to see that for any αΞp\alpha\in\Xi_{p}, αN\alpha_{N} determines αn\alpha_{n} for all nNn\leq N. To uniquely determine an element α=(αn)n\alpha=(\alpha_{n})_{n\in\mathbb{N}} in Ξp\Xi_{p}, it is sufficient to know infinitely many entries (for example, α2n\alpha_{2n} for all nn\in\mathbb{N}). This is not the case for Ωp\Omega_{p}, as many different elements of Ωp\Omega_{p} could agree at infinitely many entries. However, it is straightforward to check that if α\alpha and α~\widetilde{\alpha}, both in Ωp\Omega_{p} and agree at infinitely many entries, then h(α)=h(α~)h(\alpha)=h(\widetilde{\alpha}) and C(Γ,Ψα)C(Γ,Ψα~)C^{\ast}(\Gamma,\Psi_{\alpha})\cong C^{\ast}(\Gamma,\Psi_{\widetilde{\alpha}}).

Recall that for any θ\theta\in\mathbb{R}, the rotation algebra AθA_{\theta} is the universal C*-algebra generated by two unitaries UθU_{\theta} and VθV_{\theta} satisfying the relation UθVθ=e2πiθVθUθU_{\theta}V_{\theta}=e^{2\pi i\theta}V_{\theta}U_{\theta}. 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 α=(αn)n=0Ξp\alpha=\left(\alpha_{n}\right)_{n=0}^{\infty}\in\Xi_{p}, and for each nn\in\mathbb{N}, let φn:Aα2nAα2n+2\varphi_{n}:A_{\alpha_{2n}}\rightarrow A_{\alpha_{2n+2}} be the unique homomorphism given by

φn(Uα2n)=Uα2n+2pandφn(Vα2n)=Vα2n+2p.\varphi_{n}(U_{\alpha_{2n}})=U_{\alpha_{2n+2}}^{p}\quad\text{and}\quad\varphi_{n}(V_{\alpha_{2n}})=V_{\alpha_{2n+2}}^{p}.

The noncommutative solenoid 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} is the direct limit lim(Aα2n,φn)\varinjlim(A_{\alpha_{2n}},\varphi_{n}).

Since the rotation algebras AθA_{\theta} and Aθ+nA_{\theta+n} are isomorphic for any nn\in\mathbb{Z}, we can again replace αΞp\alpha\in\Xi_{p} with any α~Ωp\widetilde{\alpha}\in\Omega_{p} satisfying h(α~)=αh\left(\widetilde{\alpha}\right)=\alpha without changing the direct limit defined from α\alpha. 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 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} can also be written as C(𝒮p)θα[1/p]C(\mathscr{S}_{p})\rtimes_{{\theta^{\alpha}}}\mathbb{Z}[1/p], where 𝒮p\mathscr{S}_{p} is the pp-solenoid group and the action of [1/p]\mathbb{Z}\left[1/p\right] on 𝒮p\mathscr{S}_{p} is given by

θjpkα((zn)n)=(exp(2πiαk+nj)zn)n.\theta^{\alpha}_{\frac{j}{p^{k}}}\left((z_{n})_{n\in\mathbb{N}}\right)=\left(\exp\left(2\pi i\alpha_{k+n}j\right)z_{n}\right)_{n\in\mathbb{N}}.

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 Aα2nA_{\alpha_{2n}} is identified as the crossed product C*-algebra C(𝕋)α2nC\left(\mathbb{T}\right)\rtimes_{\alpha_{2n}}\mathbb{Z}. For details of the following, see [17, page 68].

Since Aα2nA_{\alpha_{2n}} is the completion of the \ast-algebra Cc(×𝕋)C_{c}\left(\mathbb{Z}\times\mathbb{T}\right) with the appropriate convolution product and involution, it suffices for our purposes to extend the embedding φn:Aα2nAα2n+2\varphi_{n}:A_{\alpha_{2n}}\rightarrow A_{\alpha_{2n+2}} to the dense subset of Cc(×𝕋)C_{c}\left(\mathbb{Z}\times\mathbb{T}\right) spanned by

{fδj:fC(𝕋) and j},\{f\delta_{j}\>:f\in C\left(\mathbb{T}\right)\text{ and }j\in\mathbb{Z}\},

where

fδj(t,k)={f(t)if k=j,0otherwise.f\delta_{j}(t,k)=\begin{cases}f(t)&\text{if }k=j,\\ 0&\text{otherwise}.\end{cases}
Lemma 2.6.

With notation as above, for any fC(𝕋)f\in C\left(\mathbb{T}\right) and jj\in\mathbb{Z},

φn(fδj)(t,k)={f(pt) if k=jp0 otherwise.\varphi_{n}\left(f\delta_{j}\right)\left(t,k\right)=\begin{cases}f(pt)&\text{ if }k=jp\\ 0&\text{ otherwise.}\end{cases}
Proof.

In this realization, the generators Uα2nU_{\alpha_{2n}} and Vα2nV_{\alpha_{2n}} are given by 1C(𝕋)δ11_{C(\mathbb{T})}\delta_{1} and ι𝕋δ0\iota_{\mathbb{T}}\delta_{0}, respectively, where 1C(𝕋)(t)=11_{C(\mathbb{T})}(t)=1 and ι𝕋(t)=e2πit\iota_{\mathbb{T}}(t)=e^{2\pi it} for all tt\in\mathbb{R}. Following the proof of [17, Proposition 2.56], we denote by iC(𝕋)i_{C(\mathbb{T})} the embedding of C(𝕋)C(\mathbb{T}) into Cc(𝕋×)C_{c}\left(\mathbb{T}\times\mathbb{Z}\right) given by ffδ0f\mapsto f\delta_{0}, then fδj=iC(𝕋)(f)Uα2njf\delta_{j}=i_{C(\mathbb{T})}(f)\ast U_{\alpha_{2n}}^{j} for any jj\in\mathbb{Z}. Since {Vα2nm:m}\{V_{\alpha_{2n}}^{m}:m\in\mathbb{Z}\} spans a dense subalgebra of iC(𝕋)(C(𝕋))i_{C(\mathbb{T})}\left(C(\mathbb{T})\right), and for any (t,k)×(t,k)\in\mathbb{R}\times\mathbb{Z},

φn(Vα2nm)(t,k)=Vα2n+2mp(t,k)=Vα2n+2m(pt,k),\displaystyle\varphi_{n}\left({V_{\alpha_{2n}}^{m}}\right)(t,k)=V_{\alpha_{2n+2}}^{mp}(t,k)=V_{\alpha_{2n+2}}^{m}\left(pt,k\right),

it follows that φn(iC(𝕋)(f))(t)=f(pt)δ0\varphi_{n}\left(i_{C(\mathbb{T})}\left(f\right)\right)(t)=f\left(pt\right)\delta_{0} for any fC(𝕋)f\in C\left(\mathbb{T}\right). We compute that for any (t,k)×(t,k)\in\mathbb{R}\times\mathbb{Z},

φn(fδj)(t,k)\displaystyle\varphi_{n}\left(f\delta_{j}\right)\left(t,k\right) =φn(iC(𝕋)(f)Uα2nj)(t,k)\displaystyle=\varphi_{n}\left(i_{C(\mathbb{T})}(f)\ast U_{\alpha_{2n}}^{j}\right)(t,k)
=(φn(iC(𝕋)(f))φn(Uα2nj))(t,k)\displaystyle=\left(\varphi_{n}\left(i_{C(\mathbb{T})}\left(f\right)\right)\ast\varphi_{n}\left(U_{\alpha_{2n}}^{j}\right)\right)(t,k)
=(φn(iC(𝕋)(f))Uα2n+2pj)(t,k)\displaystyle=\left(\varphi_{n}\left(i_{C(\mathbb{T})}\left(f\right)\right)\ast U_{\alpha_{2n+2}}^{pj}\right)(t,k)
=(φn(iC(𝕋)(f))δpj)(t,k)\displaystyle=\left(\varphi_{n}\left(i_{C(\mathbb{T})}\left(f\right)\right)\ast\delta_{pj}\right)(t,k)
=fδpj(pt,k)\displaystyle=f\delta_{pj}\left(pt,k\right)

Extending by linearity, we have that for any finite sum mfmδm\sum_{m\in\mathbb{Z}}f_{m}\delta_{m},

φn(mfmδm)(t,k)=mfmδpm(pt,k).\varphi_{n}\left(\sum_{m\in\mathbb{Z}}f_{m}\delta_{m}\right)(t,k)=\sum_{m\in\mathbb{Z}}f_{m}\delta_{pm}(pt,k).

3. Forming Heisenberg Bimodules over Noncommutative Solenoids

3.1. The field of pp-adic numbers p\mathbb{Q}_{p}

We first take the standard algebraic approach to define the field of pp-adic numbers and refer to Chapter 1 of [15] for a more extensive exposition. For a fixed prime pp, recall that a pp-adic integer is a formal series a=j=0ajpja=\sum_{j=0}^{\infty}a_{j}p^{j} with integral coefficients aja_{j} satisfying 0ajp10\leq a_{j}\leq p-1. Under the usual addition and multiplication, the ring of pp-adic integers form an integral domain with additive identity 0=j=00pj0=\sum_{j=0}^{\infty}0\cdot p^{j}, and multiplicative identity 1=1+j=10pj1=1+\sum_{j=1}^{\infty}0\cdot p^{j}. If a0a\neq 0, then there exists a unique natural number vv such that av0a_{v}\neq 0 and aj=0a_{j}=0 for all j<vj<v. We call vv the order of aa and denote it by ord(a)\operatorname{ord}(a), with the usual convention that ord(0)=\operatorname{ord}(0)=\infty.

Lemma 3.1.

The group p×\mathbb{Z}_{p}^{\times} of invertible pp-adic integers consists exactly of the pp-adic integers of order 0. That is, ap×a\in\mathbb{Z}_{p}^{\times} if and only if a00a_{0}\neq 0.

Proof.

This is well known. ∎

The field of pp-adic numbers is then defined to be the field of fractions of p\mathbb{Z}_{p}. Each nonzero pp-adic number aa can be uniquely written as a=j=vajpja=\sum_{j=v}^{\infty}a_{j}p^{j} with v<v<\infty possibly negative, av0a_{v}\neq 0, and aj{0,1,,p1}a_{j}\in\left\{0,1,\dotsc,p-1\right\} for all jj. The fractional part of a pp-adic number is given by

{a}p=j=v1ajpj[1/p].\left\{a\right\}_{p}=\sum_{j=v}^{-1}a_{j}p^{j}\in\mathbb{Z}\left[1/p\right].

Naturally, {a}p=0\left\{a\right\}_{p}=0 if and only if aa is a pp-adic integer. Extending the definition of order for pp-adic integers, the order of aa is given by ord(a)=v\operatorname{ord}(a)=v. For each apa\in\mathbb{Q}_{p}, if ord(a)=v\operatorname{ord}\left(a\right)=v, then it is easy to deduce that ord(a1)=v\operatorname{ord}\left(a^{-1}\right)=-v.

Alternatively, taking an analytic approach, p\mathbb{Q}_{p} is the completion of \mathbb{Q} with respect to the pp-adic absolute value ||p|\cdot|_{p}: for a=pk(m/n)a=p^{k}\left(m/n\right)\in\mathbb{Q} with mm and nn both indivisible by pp (then kk is necessarily unique), |a|p=pk|a|_{p}=p^{-k}. We note that the inclusion of \mathbb{Q} into p\mathbb{Q}_{p} is an injective ring homomorphism with image exactly the set of pp-adic numbers with coefficients that are eventually periodic. Moreover, pp-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 pp-adic numbers. This also means that when convenient, we identify elements of [1/p]\mathbb{Z}\left[1/p\right] with the corresponding pp-adic numbers.

The following lemmas will be useful.

Lemma 3.3.

For x=j=vxjpjpx=\sum_{j=v}^{\infty}x_{j}p^{j}\in\mathbb{Q}_{p}, with v=ord(x)<v=\operatorname{ord}(x)<\infty,

{xs1s2}p(j=vk1+k21xjpj)s1s2mod,\left\{xs_{1}s_{2}\right\}_{p}\equiv\left(\sum_{j=v}^{k_{1}+k_{2}-1}x_{j}p^{j}\right)\cdot s_{1}s_{2}\mod\mathbb{Z},

where si=jipki[1p]s_{i}=\frac{j_{i}}{p^{k_{i}}}\in\mathbb{Z}\left[\frac{1}{p}\right] for i=1,2i=1,2.

Proof.

This is trivial if vk1+k2v\geq k_{1}+k_{2}, in which case xpx\in\mathbb{Z}_{p} and {xs1s2}p=(j=vk1+k21xjpj)s1s2=0\left\{xs_{1}s_{2}\right\}_{p}=\left(\sum_{j=v}^{k_{1}+k_{2}-1}x_{j}p^{j}\right)\cdot s_{1}s_{2}=0. For v<k1+k2v<k_{1}+k_{2}, observe that

{xs1s2}p={(j=vxjpj(k1+k2))j1j2}p\left\{xs_{1}s_{2}\right\}_{p}=\left\{\left(\sum_{j=v}^{\infty}x_{j}p^{j-\left(k_{1}+k_{2}\right)}\right)j_{1}j_{2}\right\}_{p}

and

(j=vk1+k21xjpj)s1s2=(j=vk1+k21xjpj(k1+k2))j1j2.\left(\sum_{j=v}^{k_{1}+k_{2}-1}x_{j}p^{j}\right)\cdot s_{1}s_{2}=\left(\sum_{j=v}^{k_{1}+k_{2}-1}x_{j}p^{j-\left(k_{1}+k_{2}\right)}\right)j_{1}j_{2}.

It is now visible that {xs1s2}p(j=0k1+k21xjpj)s1s2mod.\left\{xs_{1}s_{2}\right\}_{p}\equiv\left(\sum_{j=0}^{k_{1}+k_{2}-1}x_{j}p^{j}\right)\cdot s_{1}s_{2}\mod\mathbb{Z}.

Lemma 3.4.

Let x=j=vxjpjx=\sum_{j=v}^{\infty}x_{j}p^{j} be a pp-adic integer with 0ord(x)=v<0\leq\operatorname{ord}(x)=v<\infty and inverse x1=j=vyjpjx^{-1}=\sum_{j=-v}^{\infty}y_{j}p^{j}. Then for all nonnegative integer kk,

(j=vv+kyjpj+v)(j=vv+kxjpjv)1modpk+1.\left(\sum_{j=-v}^{-v+k}y_{j}p^{j+v}\right)\left(\sum_{j=v}^{v+k}x_{j}p^{j-v}\right)\equiv 1\mod p^{k+1}.
Proof.

This is equivalent to

(j=0Ny~jpj)(j=0Nx~jpj)1modpN+1,\left(\sum_{j=0}^{N}\widetilde{y}_{j}p^{j}\right)\left(\sum_{j=0}^{N}\widetilde{x}_{j}p^{j}\right)\equiv 1\mod p^{N+1},

for all nonnegative integer NN, which holds for any invertible pp-adic integer x=j=0x~jpjx=\sum_{j=0}^{\infty}\widetilde{x}_{j}p^{j} with inverse x1=j=0y~jpjx^{-1}=\sum_{j=0}^{\infty}\widetilde{y}_{j}p^{j}. ∎

3.2. Heisenberg bimodules of Rieffel

In this section, we only assume that α=(αn)n\alpha=\left(\alpha_{n}\right)_{n\in\mathbb{N}} is a sequence in Ξp\Xi_{p} with α00\alpha_{0}\neq 0. Using a construction of Rieffel [14] known as the Heisenberg bimodules, explicit 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}}-\mathcal{B}-equivalence bimodules are constructed in [6], where \mathcal{B} is isomorphic to a noncommutative solenoid. We will summarize the construction when applied to noncommutative solenoids and give a formula for βΩp\beta\in\Omega_{p} such that 𝒜β𝒮\mathcal{B}\cong\mathscr{A}_{\beta}^{\mathscr{S}} in Theorem 3.6 at the end of the section.

As before, we denote by Γ\Gamma the discrete group [1/p]×[1/p]\mathbb{Z}\left[1/p\right]\times\mathbb{Z}\left[1/p\right]. Consider the group M=p×M=\mathbb{Q}_{p}\times\mathbb{R}. Since both p\mathbb{Q}_{p} and \mathbb{R} are self-dual, MM is self-dual. Specifically, It is shown in [6] that every character of MM is given by

χ(x,r):{p×𝕋(q,t)χx(q)χr(t),\chi_{(x,r)}:\begin{cases}\mathbb{Q}_{p}\times\mathbb{R}&\rightarrow\mathbb{T}\\ \left(q,t\right)&\mapsto\chi_{x}\left(q\right)\chi_{r}\left(t\right),\end{cases}

for some (x,r)p×(x,r)\in\mathbb{Q}_{p}\times\mathbb{R}, where

χx:{p𝕋qe2πi{xq}pandχr:{𝕋te2πirt\chi_{x}:\begin{cases}\mathbb{Q}_{p}&\rightarrow\mathbb{T}\\ q&\mapsto e^{2\pi i\left\{xq\right\}_{p}}\end{cases}\quad\text{and}\quad\chi_{r}:\begin{cases}\mathbb{R}&\rightarrow\mathbb{T}\\ t&\mapsto e^{2\pi irt}\end{cases}

are characters of p\mathbb{Q}_{p} and \mathbb{R}, respectively. Again, {xq}p\{xq\}_{p} is the fractional part of the pp-adic number xqxq.

Following Rieffel [14], the Heisenberg multiplier η:(M×M^)×(M×M^)𝕋\eta:(M\times\hat{M})\times(M\times\hat{M})\rightarrow\mathbb{T} is defined to be

η:{[p×]2×[p×]2𝕋,([(q1,r1),(q2,r2)],[(q3,r3),(q4,r4)])e2πir1r4e2πi{q1q4}p.\eta:\begin{cases}\qquad\quad[\mathbb{Q}_{p}\times\mathbb{R}]^{2}\times[\mathbb{Q}_{p}\times\mathbb{R}]^{2}&\rightarrow\mathbb{T},\\ \left([(q_{1},r_{1}),(q_{2},r_{2})],[(q_{3},r_{3}),(q_{4},r_{4})]\right)&\mapsto e^{2\pi ir_{1}r_{4}}e^{2\pi i\left\{q_{1}q_{4}\right\}_{p}}.\end{cases}

The symmetrized version of η\eta, denoted by ρ\rho, is the following multiplier on [p×]2[\mathbb{Q}_{p}\times\mathbb{R}]^{2}:

ρ([(q1,r1),(q2,r2)],[(q3,r3),(q4,r4)])=η([(q1,r1),(q2,r2)],[(q3,r3),(q4,r4)])η([(q3,r3),(q4,r4)],[(q1,r1),(q2,r2)])¯,\rho\left([(q_{1},r_{1}),(q_{2},r_{2})],[(q_{3},r_{3}),(q_{4},r_{4})]\right)\\ =\eta\left([(q_{1},r_{1}),(q_{2},r_{2})],[(q_{3},r_{3}),(q_{4},r_{4})]\right)\overline{\eta\left([(q_{3},r_{3}),(q_{4},r_{4})],[(q_{1},r_{1}),(q_{2},r_{2})]\right)},

with (qi,ri)p×(q_{i},r_{i})\in\mathbb{Q}_{p}\times\mathbb{R} for i=1,2,3,4i=1,2,3,4.

Now, for each pair of xp{0}x\in\mathbb{Q}_{p}\setminus\{0\} and θ{0}\theta\in\mathbb{R}\setminus\{0\}, let ιx,θ:ΓM×M^M×M\iota_{x,\theta}:\Gamma\rightarrow M\times\hat{M}\cong M\times M be the embedding of Γ\Gamma as a lattice (discrete cocompact subgroup) into [p×]2\left[\mathbb{Q}_{p}\times\mathbb{R}\right]^{2} given by

ιx,θ(r1,r2)=[(xr1,θr1),(r2,r2)].\iota_{x,\theta}\left(r_{1},r_{2}\right)=\left[(x\cdot r_{1},\theta\cdot r_{1}),(r_{2},r_{2})\right].

We denote the image of ιx,θ\iota_{x,\theta} in M×M^M\times\hat{M} by Dx,θD_{x,\theta}. For each such embedding, the so-called annihilator of Dx,θD_{x,\theta} is defined as

Dx,θ:={[(q1,s1),(q2,s2)]M×M^:r1,r2[1p],ρ(ιx,θ(r1,r2),[(q1,s1),(q2,s2)])=1}.D_{x,\theta}^{\perp}:=\left\{[(q_{1},s_{1}),(q_{2},s_{2})]\in M\times\hat{M}\>:\forall r_{1},r_{2}\in\mathbb{Z}\left[\dfrac{1}{p}\right],\rho\left(\iota_{x,\theta}(r_{1},r_{2}),[(q_{1},s_{1}),(q_{2},s_{2})]\right)=1\right\}.

Since Dx,θD_{x,\theta} is a lattice in GG, so is Dx,θD_{x,\theta}^{\perp} [14, Lemma 3.1].

Applying the result in [14, Theorem 2.15] to M=p×M=\mathbb{Q}_{p}\times\mathbb{R}, Dx,θD_{x,\theta} and Dx,θD_{x,\theta}^{\perp}, we have that Cc(M)C_{c}(M), suitably completed, has the structure of a C(Dx,θ,η)C^{\ast}(D_{x,\theta},\eta)-C(Dx,θ,η¯)C^{\ast}(D_{x,\theta}^{\perp},\overline{\eta})-equivalence bimodule and implements the Morita equivalence between C(Dx,θ,η)C^{\ast}(D_{x,\theta},\eta) and C(Dx,θ,η¯)C^{\ast}(D_{x,\theta}^{\perp},\overline{\eta}). We now state the main theorem of this section.

Theorem 3.5.

Let pp be prime and let α=(αj)jΩp\alpha=(\alpha_{j})_{j\in\mathbb{N}}\in\Omega_{p} such that pαj+1=αj+xjp\alpha_{j+1}=\alpha_{j}+x_{j} with α00\alpha_{0}\neq 0 and x00x_{0}\neq 0. Then C(Dx,θ,η)C^{\ast}(D_{x,\theta},\eta) is isomorphic to the noncommutative solenoid 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}}, where x=j=0xjpjpx=\sum_{j=0}^{\infty}x_{j}p^{j}\in\mathbb{Z}_{p} and θ=α0\theta=\alpha_{0}. Since x00x_{0}\neq 0, we write (uniquely) x1=j=0yjpjpx^{-1}=\sum_{j=0}^{\infty}y_{j}p^{j}\in\mathbb{Z}_{p}. Via the Heisenberg equivalence bimodule of Rieffel, 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} is Morita equivalent to a noncommutative solenoid 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}}, where β=(βn)n\beta=\left(\beta_{n}\right)_{n\in\mathbb{N}} and

β=(1θ,1θp+y0p,1θp2+y0+y1pp2,,βn,),βn=1θpn+j=0n1yjpjpn.\beta=\left(\dfrac{1}{\theta},\dfrac{1}{\theta p}+\frac{y_{0}}{p},\dfrac{1}{\theta p^{2}}+\dfrac{y_{0}+y_{1}p}{p^{2}},\dotsc,\beta_{n},\dotsc\right),\quad\beta_{n}=\dfrac{1}{\theta p^{n}}+\dfrac{\sum_{j=0}^{n-1}y_{j}p^{j}}{p^{n}}.
Proof.

The statement of this theorem gives an explicit formula for the β\beta in [6, Theorem 5.6], with the additional assumption that x00x_{0}\neq 0. We note that in the proof of the aforementioned theorem, one only needs to assume α00\alpha_{0}\neq 0. This means that 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} is isomorphic C(Dx,θ,η)C^{\ast}\left(D_{x,\theta},\eta\right) even when α\alpha is taken from Ωp\Omega_{p}. We simply need to show that C(Dx,θ,η¯)C^{\ast}\left(D_{x,\theta}^{\perp},\overline{\eta}\right) is isomorphic to 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} to conclude that the two noncommutative solenoids 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} and 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} are Morita equivalent due to Rieffel’s method.

It is shown in [6, Lemma 5.5] that

Dx,θ={[(s1,s1),(x1s2,s2θ)]:s1,s2[1p]}.D_{x,\theta}^{\perp}=\left\{\left[(s_{1},-s_{1}),\left(x^{-1}s_{2},-\frac{s_{2}}{\theta}\right)\right]:s_{1},s_{2}\in\mathbb{Z}\left[\frac{1}{p}\right]\right\}.

For convenience, we choose to substitute s2s_{2} with s2-s_{2} and write

Dx,θ={[(s1,s1),(x1s2,s2θ)]:s1,s2[1p]}.D_{x,\theta}^{\perp}=\left\{\left[(s_{1},-s_{1}),\left(-x^{-1}s_{2},\frac{s_{2}}{\theta}\right)\right]:s_{1},s_{2}\in\mathbb{Z}\left[\frac{1}{p}\right]\right\}.

Let λx,θ\lambda_{x,\theta} denote the embedding of [1p]×[1p]\mathbb{Z}\left[\frac{1}{p}\right]\times\mathbb{Z}\left[\frac{1}{p}\right] into [p×]2\left[\mathbb{Q}_{p}\times\mathbb{R}\right]^{2}, so that the image of λx,θ\lambda_{x,\theta} is Dx,θD_{x,\theta}^{\perp}. We then have

η¯(λx,θ(s1,s2),λx,θ(s3,s4))\displaystyle\overline{\eta}\left(\lambda_{x,\theta}(s_{1},s_{2}),\lambda_{x,\theta}(s_{3},s_{4})\right) =η¯([(s1,s1),(x1s2,s2θ)],[(s3,s3),(x1s4,s4θ)])\displaystyle=\overline{\eta}\left(\left[(s_{1},-s_{1}),\left(-x^{-1}s_{2},\frac{s_{2}}{\theta}\right)\right],\left[(s_{3},-s_{3}),\left(-x^{-1}s_{4},\frac{s_{4}}{\theta}\right)\right]\right)
=e2πi1θs1s4e2πi{x1s1s4}p\displaystyle=e^{2\pi i\frac{1}{\theta}s_{1}s_{4}}e^{-2\pi i\{-x^{-1}s_{1}s_{4}\}_{p}}

Note that for a pp-adic number x=j=vxjpjx=\sum_{j=v}^{\infty}x_{j}p^{j} of order vv,

x=(pxv)pv+(p1xv+1)pv+1++(p1xj)pj+-x=(p-x_{v})p^{v}+(p-1-x_{v+1})p^{v+1}+\cdots+(p-1-x_{j})p^{j}+\cdots

It is clear that if xpx\in\mathbb{Z}_{p}, then {x}p={x}p=0\{x\}_{p}=\{-x\}_{p}=0. If xppx\in\mathbb{Q}_{p}\setminus\mathbb{Z}_{p}, then {x}p+{x}p=1\{x\}_{p}+\{-x\}_{p}=1. It follows that for x1s1s4p-x^{-1}s_{1}s_{4}\in\mathbb{Z}_{p}, either {x1s1s4}p={x1s1s4}p\{-x^{-1}s_{1}s_{4}\}_{p}=\{x^{-1}s_{1}s_{4}\}_{p} or {x1s1s4}p=1{x1s1s4}p\{-x^{-1}s_{1}s_{4}\}_{p}=1-\{x^{-1}s_{1}s_{4}\}_{p}. In either case, we have

η¯([(s1,s1),(x1s2,s2θ)],[(s3,s3),(x1s4,s4θ)])\displaystyle\overline{\eta}\left(\left[(s_{1},-s_{1}),\left(-x^{-1}s_{2},\frac{s_{2}}{\theta}\right)\right],\left[(s_{3},-s_{3}),\left(-x^{-1}s_{4},\frac{s_{4}}{\theta}\right)\right]\right) =e2πi1θs1s4e2πi{x1s1s4}p\displaystyle=e^{2\pi i\frac{1}{\theta}s_{1}s_{4}}e^{-2\pi i\{-x^{-1}s_{1}s_{4}\}_{p}}
=e2πi1θs1s4e2πi{x1s1s4}p\displaystyle=e^{2\pi i\frac{1}{\theta}s_{1}s_{4}}e^{2\pi i\{x^{-1}s_{1}s_{4}\}_{p}}

By Theorem 2.2, the multiplier Ψβ\Psi_{\beta} on [1p]×[1p]\mathbb{Z}\left[\frac{1}{p}\right]\times\mathbb{Z}\left[\frac{1}{p}\right] is given by:

Ψβ((j1pk1,j2pk2),(j3pk3,j4pk4))\displaystyle\Psi_{\beta}\left(\left(\frac{j_{1}}{p^{k_{1}}},\frac{j_{2}}{p^{k_{2}}}\right),\left(\frac{j_{3}}{p^{k_{3}}},\frac{j_{4}}{p^{k_{4}}}\right)\right) =exp(2πiβ(k1+k4)j1j4)\displaystyle=\exp\left(2\pi i\beta_{(k_{1}+k_{4})}j_{1}j_{4}\right)
=exp(2πi(1θpk1+k4+j=0k1+k41yjpjpk1+k4)j1j4)\displaystyle=\exp\left(2\pi i\left(\dfrac{1}{\theta p^{k_{1}+k_{4}}}+\dfrac{\sum_{j=0}^{k_{1}+k_{4}-1}y_{j}p^{j}}{p^{k_{1}+k_{4}}}\right)j_{1}j_{4}\right)
=exp(2πij1j4θpk1+k4)exp(2πij=0k1+k41yjpjpk1+k4j1j4)\displaystyle=\exp\left(2\pi i\dfrac{j_{1}j_{4}}{\theta p^{k_{1}+k_{4}}}\right)\cdot\exp\left(2\pi i\dfrac{\sum_{j=0}^{k_{1}+k_{4}-1}y_{j}p^{j}}{p^{k_{1}+k_{4}}}j_{1}j_{4}\right)
=exp(2πi1θs1s4)exp(2πi(j=0k1+k41yjpj)s1s4)\displaystyle=\exp\left(2\pi i\frac{1}{\theta}s_{1}s_{4}\right)\cdot\exp\left(2\pi i\left(\sum_{j=0}^{k_{1}+k_{4}-1}y_{j}p^{j}\right)s_{1}s_{4}\right)

By the Lemma 3.3, {x1s1s4}p(j=0k1+k41yjpj)s1s4 modulo \left\{x^{-1}s_{1}s_{4}\right\}_{p}\equiv\left(\sum_{j=0}^{k_{1}+k_{4}-1}y_{j}p^{j}\right)\cdot s_{1}s_{4}\text{ modulo }\mathbb{Z}, so we have

exp(2πi(j=0k1+k41yjpj)s1s4)=exp(2πi{x1s1s4}p).\displaystyle\exp\left(2\pi i\left(\sum_{j=0}^{k_{1}+k_{4}-1}y_{j}p^{j}\right)s_{1}s_{4}\right)=\exp\left(2\pi i\left\{x^{-1}s_{1}s_{4}\right\}_{p}\right).

This shows that η¯(λx,θ(s1,s2),λx,θ(s3,s4))=Ψβ((s1,s2),(s3,s4))\overline{\eta}\left(\lambda_{x,\theta}(s_{1},s_{2}),\lambda_{x,\theta}(s_{3},s_{4})\right)=\Psi_{\beta}\left(\left(s_{1},s_{2}\right),\left(s_{3},s_{4}\right)\right), for all s1s_{1}, s2s_{2}, s3s_{3}, and s4s_{4} in [1/p]\mathbb{Z}\left[1/p\right], as desired. Therefore, C(Dx,θ,η¯)C^{\ast}\left(D_{x,\theta}^{\perp},\overline{\eta}\right) is isomorphic to 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}}. It follows that 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} is Morita equivalent to 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}}. ∎

Thus, one sees that given any α\alpha in Ωp\Omega_{p} satisfying α00\alpha_{0}\neq 0 and xαx_{\alpha} invertible in the pp-adic integers, we can find a βΩp\beta\in\Omega_{p} (satisfying the same conditions), such that 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} and 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} are Morita equivalent. It is then not unexpected that if we started with β\beta, the formula in this theorem recovers α\alpha exactly.

We conclude this section with the following slightly more general result, in which we only assume the pp-adic integer xαx_{\alpha} associated to α\alpha is nonzero.

Theorem 3.6.

Let pp be prime and let α=(αj)jΩp\alpha=(\alpha_{j})_{j\in\mathbb{N}}\in\Omega_{p} such that pαj+1=αj+xjp\alpha_{j+1}=\alpha_{j}+x_{j} with α00\alpha_{0}\neq 0 and xα=j=0xjpj0x_{\alpha}=\sum_{j=0}^{\infty}x_{j}p^{j}\neq 0. Then C(Dx,θ,η)C^{\ast}(D_{x,\theta},\eta) is isomorphic to the noncommutative solenoid 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}}, where x=xαx=x_{\alpha} and θ=α0\theta=\alpha_{0}. Since xα0x_{\alpha}\neq 0, we write (uniquely) xα1=j=vyjpjpx_{\alpha}^{-1}=\sum_{j=-v}^{\infty}y_{j}p^{j}\in\mathbb{Q}_{p}. Via the Heisenberg equivalence bimodule of Rieffel, 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} is Morita equivalent to a noncommutative solenoid 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}}, where β=(βn)n\beta=\left(\beta_{n}\right)_{n\in\mathbb{N}} is given by

βn=1θpn+j=vn1yjpjpn.\beta_{n}=\dfrac{1}{\theta p^{n}}+\dfrac{\sum_{j=-v}^{n-1}y_{j}p^{j}}{p^{n}}.
Proof.

Using Lemma 3.3 and Lemma 3.4 to their full generality, the proof follows the same procedure presented in the proof of Theorem 3.5. ∎

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

A0{A_{0}}A1{A_{1}}A2{A_{2}}{\cdots}φ0\scriptstyle{\varphi_{0}}φ1\scriptstyle{\varphi_{1}}φ2\scriptstyle{\varphi_{2}}

and

B0{B_{0}}B1{B_{1}}B2{B_{2}}{\cdots}ψ0\scriptstyle{\psi_{0}}ψ1\scriptstyle{\psi_{1}}ψ2\scriptstyle{\psi_{2}}

be two directed systems of unital C*-algebras, whose *-morphisms are all unital maps. A sequence (Xn,in)n(X_{n},i_{n})_{n\in\mathbb{N}} is a directed system of equivalence bimodule adapted to the sequence (An)n(A_{n})_{n\in\mathbb{N}} and (Bn)n(B_{n})_{n\in\mathbb{N}} when XnX_{n} is an AnA_{n}-BnB_{n}-equivalence bimodule, whose AnA_{n}- and BnB_{n}-valued inner products are denoted respectively by ,An\langle\cdot,\cdot\rangle_{A_{n}} and ,Bn\langle\cdot,\cdot\rangle_{B_{n}}, for all nn\in\mathbb{N}, and such that the sequence

X0{X_{0}}X1{X_{1}}X2{X_{2}}{\cdots}i0\scriptstyle{i_{0}}i1\scriptstyle{i_{1}}i2\scriptstyle{i_{2}}

is a directed sequence of modules satisfying

in(f),in(g)Bn+1=ψn(f,gBn),for all f,gXn,\langle i_{n}(f),i_{n}(g)\rangle_{B_{n+1}}=\psi_{n}\left(\langle f,g\rangle_{B_{n}}\right),\quad\text{for all }f,g\in X_{n}, (4.1)

and

in(fb)=in(f)ψn(b),for all fXn,bBn,i_{n}(f\cdot b)=i_{n}(f)\cdot\psi_{n}(b),\quad\text{for all }f\in X_{n},b\in B_{n}, (4.2)

with analogous but symmetric equalities holding for the XnX_{n} viewed as left Hilbert AnA_{n}-modules:

in(f),in(g)An+1=φn(f,gAn),for all f,gXn,\langle i_{n}(f),i_{n}(g)\rangle_{A_{n+1}}=\varphi_{n}\left(\langle f,g\rangle_{A_{n}}\right),\quad\text{for all }f,g\in X_{n}, (4.3)

and

in(af)=φn(a)in(f)for all fXn,aAn,i_{n}(a\cdot f)=\varphi_{n}(a)\cdot i_{n}(f)\quad\text{for all }f\in X_{n},a\in A_{n}, (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 𝒜α𝒮=limAα2n\mathscr{A}_{\alpha}^{\mathscr{S}}=\varinjlim A_{\alpha_{2n}} and a projection PP from an m×mm\times m matrix algebra over Aα0A_{\alpha_{0}}, both satisfying certain conditions. We can then form a noncommutative solenoid 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} from the directed sequence of irrational rotation algebras Aβ2nPMm(Aα2n)PA_{\beta_{2n}}\cong PM_{m}\left(A_{\alpha_{2n}}\right)P with connecting maps given by Theorem 2.4. We will show that limPAα2nm\varinjlim PA_{\alpha_{2n}}^{m}, with the appropriate connecting maps, is a directed system of equivalence bimodules adapted to the sequences (Aα2n)n\left(A_{\alpha_{2n}}\right)_{n\in\mathbb{N}} and (PMm(Aα2n)P)n\left(PM_{m}\left(A_{\alpha_{2n}}\right)P\right)_{n\in\mathbb{N}}. 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 α=(αn)nΞp\alpha=\left(\alpha_{n}\right)_{n\in\mathbb{N}}\in\Xi_{p} be an irrational sequence and pαn+1=αn+xnp\alpha_{n+1}=\alpha_{n}+x_{n} for all nn\in\mathbb{N}. Let 𝒜α𝒮=limAα2n\mathscr{A}_{\alpha}^{\mathscr{S}}=\varinjlim A_{\alpha_{2n}} be an irrational noncommutative solenoid realized as the direct limit of the sequence of irrational rotation algebras (Aα2n)n\left(A_{\alpha_{2n}}\right)_{n\in\mathbb{N}}. We denote the unique tracial state on Aα2nA_{\alpha_{2n}} by τα2n\tau_{\alpha_{2n}}. We use the same notation τα2n\tau_{\alpha_{2n}} for the standard unnormalized trace on Mm(Aα2n)M_{m}\left(A_{\alpha_{2n}}\right), m1m\geq 1. Note that (1/m)τα2n\left(1/m\right)\tau_{\alpha_{2n}} is then the unique tracial state on Mm(Aα2n)M_{m}\left(A_{\alpha_{2n}}\right).

Let PP be a nontrivial projection in Mm(Aα0)M_{m}\left(A_{\alpha_{0}}\right) with τα0(P)=c0α0+d0(0,m)\tau_{\alpha_{0}}\left(P\right)=c_{0}\alpha_{0}+d_{0}\in(0,m) that satisfies the following condition:

Condition 4.3.

gcd(c0p,d0c0x0)=1\gcd\left(c_{0}p,d_{0}-c_{0}x_{0}\right)=1.

We follow the convention that gcd(a,0)=a\gcd\left(a,0\right)=a for any positive integer aa. For details of the existence and construction of such a projection, see [11]. When PP is considered as a projection in Mm(Aα2n)M_{m}\left(A_{\alpha_{2n}}\right) for any nn\in\mathbb{N}, it is not hard to see that τα2n(P)=τα0(P)=c0α0+d0\tau_{\alpha_{2n}}\left(P\right)=\tau_{\alpha_{0}}\left(P\right)=c_{0}\alpha_{0}+d_{0}, but α0=p2nα2nj=02n1xjpj\alpha_{0}=p^{2n}\alpha_{2n}-\sum_{j=0}^{2n-1}x_{j}p^{j}, so the trace of PP there is

τα2n(P)=(c0p2n)α2n+(d0c0j=02n1xjpj).\tau_{\alpha_{2n}}\left(P\right)=\left(c_{0}p^{2n}\right)\alpha_{2n}+\left(d_{0}-c_{0}\sum_{j=0}^{2n-1}x_{j}p^{j}\right).

We let

c2n=c0p2nandd2n=d0c0j=02n1xjpj,c_{2n}=c_{0}p^{2n}\quad\text{and}\quad d_{2n}=d_{0}-c_{0}\sum_{j=0}^{2n-1}x_{j}p^{j},

so that τα2n(P)=c2nα2n+d2n\tau_{\alpha_{2n}}\left(P\right)=c_{2n}\alpha_{2n}+d_{2n} in Mm(Aα2n)M_{m}\left(A_{\alpha_{2n}}\right).

The follow lemma states Condition 4.3 in a form that will be useful to us later.

Lemma 4.4.

If gcd(c0p,d0c0x0)=1\gcd\left(c_{0}p,d_{0}-c_{0}x_{0}\right)=1, then gcd(c2n,d2n)=1\gcd\left(c_{2n},d_{2n}\right)=1 for all nn\in\mathbb{N}.

Proof.

First, it is clear that gcd(c0p,d0+c0x0)=1\gcd\left(c_{0}p,d_{0}+c_{0}x_{0}\right)=1 implies gcd(c0,d0)=1\gcd(c_{0},d_{0})=1, as any common divisor of c0c_{0} and d0d_{0} divides both c0pc_{0}p and d0c0x0d_{0}-c_{0}x_{0}.

Fix n1n\geq 1. We show that if gcd(c0p2n,d0c0j=02n1xjpj)>1\gcd\left(c_{0}p^{2n},d_{0}-c_{0}\sum_{j=0}^{2n-1}x_{j}p^{j}\right)>1, then gcd(c0p,d0c0x0)>1\gcd\left(c_{0}p,d_{0}-c_{0}x_{0}\right)>1. It then follows that gcd(c0p,d0c0x0)=1\gcd(c_{0}p,d_{0}-c_{0}x_{0})=1 implies gcd(c0p2n,d0c0j=02n1xjpj)=1\gcd(c_{0}p^{2n},d_{0}-c_{0}\sum_{j=0}^{2n-1}x_{j}p^{j})=1 for all n1n\geq 1.

Assume gcd(c2n,d2n)=gcd(c0p2n,d0c0j=02n1xjpj)>1\gcd\left(c_{2n},d_{2n}\right)=\gcd\left(c_{0}p^{2n},d_{0}-c_{0}\sum_{j=0}^{2n-1}x_{j}p^{j}\right)>1, then either pp divides d0c0j=02n1xjpjd_{0}-c_{0}\sum_{j=0}^{2n-1}x_{j}p^{j}, or qq divides d0c0j=02n1xjpjd_{0}-c_{0}\sum_{j=0}^{2n-1}x_{j}p^{j} for some divisor qq of c0c_{0} that is strictly greater than 11. In the former case, for some integer kk,

kp=d0c0j=02n1xjpj=(d0c0x0)c0j=12n1xjpj,\displaystyle kp=d_{0}-c_{0}\sum_{j=0}^{2n-1}x_{j}p^{j}=\left(d_{0}-c_{0}x_{0}\right)-c_{0}\sum_{j=1}^{2n-1}x_{j}p^{j},

so d0c0b0=kp+c0j=12n1xjpj=p(k+c0j=12n1xjpj1)d_{0}-c_{0}b_{0}=kp+c_{0}\sum_{j=1}^{2n-1}x_{j}p^{j}=p\left(k+c_{0}\sum_{j=1}^{2n-1}x_{j}p^{j-1}\right). This makes pp a common divisor of c0pc_{0}p and d0c0x0d_{0}-c_{0}x_{0}. In the latter case, d0c0x0=(d0c0j=02n1xjpj)+c0j=12n1xjpjd_{0}-c_{0}x_{0}=\left(d_{0}-c_{0}\sum_{j=0}^{2n-1}x_{j}p^{j}\right)+c_{0}\sum_{j=1}^{2n-1}x_{j}p^{j}, making qq a common divisor of d0c0x0d_{0}-c_{0}x_{0} and c0pc_{0}p. This shows that gcd(c0p,d0c0x0)=1\gcd(c_{0}p,d_{0}-c_{0}x_{0})=1 implies gcd(c0p2n,d0c0j=02n1xjpj)=1\gcd(c_{0}p^{2n},d_{0}-c_{0}\sum_{j=0}^{2n-1}x_{j}p^{j})=1 for all n1n\geq 1. ∎

Remark 4.5.

We single out a special occurrence of how Condition 4.3 can be satisfied: if d0=0d_{0}=0, then Condition 4.3 is satisfied if and only if both c0=1c_{0}=1 and x00x_{0}\neq 0. This will be exactly the situation described in Proposition 4.24.

4.2. Review of Rieffel’s standard bimodule between irrational rotation algebras

Notation 4.6.

Throughout this section, we fix an irrational number α\alpha and let cc and dd be a pair of integers that generate \mathbb{Z} with cα+d0c\alpha+d\neq 0 and c0c\neq 0. Set γ=1/(cα+d)\gamma=1/\left(c\alpha+d\right). Moreover, we let aa and bb be any two integers such that adbc=1ad-bc=1, and let β=(aα+b)γ=(aα+b)/(cα+d)\beta=\left(a\alpha+b\right)\gamma=\left(a\alpha+b\right)/\left(c\alpha+d\right).

For any k1k\geq 1, let V(d,c;k)V\left(d,c;k\right) be the standard Mk(Aβ)M_{k}\left(A_{\beta}\right)-AαA_{\alpha}-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 PP be a projection in Mm(Aα)M_{m}\left(A_{\alpha}\right), m1m\geq 1, with unnormalized trace k(cα+d)k\left(c\alpha+d\right), then V(d,c;k)V\left(d,c;k\right) is isomorphic to PAαmPA_{\alpha}^{m} as a right AαA_{\alpha}-module.

Since PP is a nontrivial projection in the simple C*-algebra Mm(Aα)M_{m}\left(A_{\alpha}\right), PP is full and PAαmPA_{\alpha}^{m} is a PMm(Aα)PPM_{m}\left(A_{\alpha}\right)P-AαA_{\alpha}-equivalence bimodule. It follows from this observation and the lemma above that PMm(Aα)PMk(Aβ)PM_{m}\left(A_{\alpha}\right)P\cong M_{k}\left(A_{\beta}\right) as C*-algebras.

We now recall Rieffel’s construction of the standard Mk(Aβ)M_{k}\left(A_{\beta}\right)-AαA_{\alpha}-equivalence bimodule for the case k=1k=1, 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 G=×cG=\mathbb{R}\times\mathbb{Z}_{c} (If cc is a negative integer, we still define c\mathbb{Z}_{c} to be /c\mathbb{Z}/c\mathbb{Z}) and consider the following subgroups of GG:

H={(n,[dn]c):n},K={(nγ,[n]c):n}.H=\{\left(n,[dn]_{c}\right):n\in\mathbb{Z}\},\qquad K=\left\{\left(n\gamma,[n]_{c}\right):n\in\mathbb{Z}\right\}.

Let HH act on K\GK\backslash G (the right cosets of KK) by right translation, and let KK act on G/HG/H (the left cosets of HH) by left translation. Then the transformation group C*-algebras C(H,K\G)C^{\ast}\left(H,K\backslash G\right) and C(K,G/H)C^{\ast}\left(K,G/H\right) are isomorphic to AαA_{\alpha} and AβA_{\beta}, respectively. Furthermore, Cc(G)C_{c}\left(G\right), suitably completed and structured, provides an AβA_{\beta}-AαA_{\alpha}-equivalence bimodule.

Both isomorphisms rely on identifying the quotient group with the circle group 𝕋\mathbb{T} and realizing the corresponding group action of \mathbb{Z} (both HH and KK are isomorphic to \mathbb{Z}) gives rise to a rotation algebra. In particular, we employ the following identification of K\GK\backslash G and G/HG/H with 𝕋\mathbb{T}. The map given by

{G𝕋(t,[m]c)t/γmc,\begin{cases}G&\rightarrow\mathbb{T}\\ (t,[m]_{c})&\mapsto\dfrac{t/\gamma-m}{c},\end{cases}

is a group homomorphism with kernel exactly KK. We can then identify KGK\setminus G with 𝕋\mathbb{T}. Similarly, the map given by

{G𝕋(t,[m]c)tamc,\begin{cases}G&\rightarrow\mathbb{T}\\ (t,[m]_{c})&\mapsto\dfrac{t-am}{c},\end{cases}

is a group homomorphism with kernel exactly HH, so we can identify G/HG/H with 𝕋\mathbb{T}.

Remark 4.9.

We use an inverse identification of G/HG/H with 𝕋\mathbb{T} than what is presented in [13]. Over there, the identification is derived from the map G𝕋G\rightarrow\mathbb{T} given by (t,[m]c)(amt)/c(t,[m]_{c})\mapsto(am-t)/c. One of reasons for choosing this inverse identification is to later have UβVβ=e2πiβVβUβU_{\beta}V_{\beta}=e^{2\pi i\beta}V_{\beta}U_{\beta}, instead of e2πiβUβVβ=VβUβe^{2\pi i\beta}U_{\beta}V_{\beta}=V_{\beta}U_{\beta}, where UβU_{\beta} and VβV_{\beta} are the usual generating unitaries of AβA_{\beta}.

In the following two lemmas, we record the formulas for the generators of AβA_{\beta}, the left AβA_{\beta}-action on Cc(×c)C_{c}\left(\mathbb{R}\times\mathbb{Z}_{c}\right), and the AβA_{\beta}-valued inner product, all under the present identification of G/H×KG/H\times K with 𝕋×\mathbb{T}\times\mathbb{Z}. To ease notation, we just write nn to represent (nγ,[n]c)K\left(n\gamma,[n]_{c}\right)\in K. 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 a=1,b=0,c=p2ja=1,b=0,c=p^{2j}, and d=1d=1, for any jj\in\mathbb{N}.

For (t,[m]c)×c(t,[m]_{c})\in\mathbb{R}\times\mathbb{Z}_{c}, we denote by (t,[m]c)~\widetilde{(t,[m]_{c})} the class of (t,[m]c)(t,[m]_{c}) in G/HG/H. In the second lemma, the left AβA_{\beta}-action is define on the dense subspace Cc(G/H×K)C_{c}\left(G/H\times K\right) of C(K,G/H)AβC^{\ast}\left(K,G/H\right)\cong A_{\beta}, and (r,k)×(r,k)\in\mathbb{R}\times\mathbb{Z} is identified with G/H×KG/H\times K via the map (e2πir,k)((cr,[0]c)~,(kγ,[k]c))G/H×K(e^{2\pi ir},k)\mapsto\left(\widetilde{(cr,[0]_{c})},(k\gamma,[k]_{c})\right)\in G/H\times K.

Lemma 4.10.

The generators UβU_{\beta} and VβV_{\beta} of AβA_{\beta} are given by

Uβ((t,[m]c)~,n)={0, if n11, if n=1U_{\beta}(\widetilde{\left(t,[m]_{c}\right)},n)=\begin{cases}0,&\text{ if }n\neq 1\\ 1,&\text{ if }n=1\end{cases}

and

Vβ((t,[m]c)~,n)={0, if n0e2πi[(tam)/c], if n=0.V_{\beta}(\widetilde{\left(t,[m]_{c}\right)},n)=\begin{cases}0,&\text{ if }n\neq 0\\ e^{2\pi i\left[(t-am)/c\right]},&\text{ if }n=0.\end{cases}

We compute that

UβVβ((t,[m]c)~,n)={0, if n1,e2πi(tγam+a)/c, if n=1.U_{\beta}V_{\beta}\left(\widetilde{\left(t,[m]_{c}\right)},n\right)=\begin{cases}0,&\text{ if }n\neq 1,\\ e^{2\pi i(t-\gamma-am+a)/c},&\text{ if }n=1.\end{cases}

and

VβUβ((t,[m]c)~,n)={0, if n1,e2πi(tam)/c, if n=1.V_{\beta}U_{\beta}\left(\widetilde{\left(t,[m]_{c}\right)},n\right)=\begin{cases}0,&\text{ if }n\neq 1,\\ e^{2\pi i(t-am)/c},&\text{ if }n=1.\end{cases}

It follows that UβVβ=e2πi(aγ)/cVβUβU_{\beta}V_{\beta}=e^{2\pi i(a-\gamma)/c}V_{\beta}U_{\beta}. After writing 11 as adbcad-bc, a few steps of algebra will show that (aγ)/c=β(a-\gamma)/c=\beta, so we have UβVβ=e2πiβVβUβU_{\beta}V_{\beta}=e^{2\pi i\beta}V_{\beta}U_{\beta}.

Lemma 4.11.

For any F1,F2Cc(×c)F_{1},F_{2}\in C_{c}\left(\mathbb{R}\times\mathbb{Z}_{c}\right), fCc(G/H×K)f\in C_{c}\left(G/H\times K\right), (r,k)(,)(r,k)\in\left(\mathbb{R},\mathbb{Z}\right) and (t,[m]c)×c(t,[m]_{c})\in\mathbb{R}\times\mathbb{Z}_{c},

F1,F2Aβ(r,k)=mF1(cr+m,[dm]c)F2(cr+mkγ,[dmk]c)¯,\langle F_{1},F_{2}\rangle_{A_{\beta}}(r,k)=\sum_{m\in\mathbb{Z}}F_{1}\left(cr+m,[dm]_{c}\right)\overline{F_{2}\left(cr+m-k\gamma,[dm-k]_{c}\right)}, (4.5)

and

(fF)(t,[m]c)=nf((t,[m]c)~,(nγ,[n]c))F(tnγ,[mn]c).\left(f\cdot F\right)\left(t,[m]_{c}\right)=\sum_{n\in\mathbb{Z}}f\left(\widetilde{(t,[m]_{c})},(n\gamma,[n]_{c})\right)F\left(t-n\gamma,[m-n]_{c}\right). (4.6)

In particular, it follows that the generators UβU_{\beta} and VβV_{\beta} act on Cc(G)C_{c}(G) in the following way:

(UβF)(t,[m]c)=F(tγ,[m1]c),\left(U_{\beta}\cdot F\right)(t,[m]_{c})=F\left(t-\gamma,[m-1]_{c}\right), (4.7)

and

(VβF)(t,[m]c)=e2πi(tam)/cF(t,[m]c).\left(V_{\beta}\cdot F\right)(t,[m]_{c})=e^{2\pi i\left(t-am\right)/c}F(t,[m]_{c}). (4.8)

The right AαA_{\alpha}-module structure is given by the following lemmas.

Lemma 4.12.

The generators UαU_{\alpha} and VαV_{\alpha} of AαA_{\alpha} are given by

Uα((t,[m]c)~,n)={0, if n11, if n=1U_{\alpha}(\widetilde{\left(t,[m]_{c}\right)},n)=\begin{cases}0,&\text{ if }n\neq 1\\ 1,&\text{ if }n=1\end{cases}

and

Vα((t,[m]c)~,n)={0, if n0e2πi[(t/γm)/c], if n=0.V_{\alpha}(\widetilde{\left(t,[m]_{c}\right)},n)=\begin{cases}0,&\text{ if }n\neq 0\\ e^{2\pi i\left[(t/\gamma-m)/c\right]},&\text{ if }n=0.\end{cases}
Lemma 4.13.

For any F1,F2Cc(×c)F_{1},F_{2}\in C_{c}\left(\mathbb{R}\times\mathbb{Z}_{c}\right), gCc(H×K\G)g\in C_{c}\left(H\times K\backslash G\right), (r,k)(,)(r,k)\in\left(\mathbb{R},\mathbb{Z}\right) and (t,[m]c)×c(t,[m]_{c})\in\mathbb{R}\times\mathbb{Z}_{c},

F1,F2Aα(r,k)=mf((crm)γ,[m]c)¯g((crm)γ+k,[dkm]c),\langle F_{1},F_{2}\rangle_{A_{\alpha}}(r,k)=\sum_{m\in\mathbb{Z}}\overline{f\left(\left(cr-m\right)\gamma,[-m]_{c}\right)}g\left(\left(cr-m\right)\gamma+k,[dk-m]_{c}\right), (4.9)

and

(Fg)(t,[m]c)=nF(tn,[mdn]c)g((tn,[mdn]c)~,(n,[dn]c)).\left(F\cdot g\right)\left(t,[m]_{c}\right)=\sum_{n\in\mathbb{Z}}F\left(t-n,[m-dn]_{c}\right)g\left(\widetilde{(t-n,[m-dn]_{c})},\left(n,[dn]_{c}\right)\right). (4.10)

In particular, it follows that the generators UαU_{\alpha} and VαV_{\alpha} act on Cc(G)C_{c}(G) in the following way:

(FUα)(t,[m]c)=F(t1,[md]c),\left(F\cdot U_{\alpha}\right)(t,[m]_{c})=F\left(t-1,[m-d]_{c}\right), (4.11)

and

(FVα)(t,[m]c)=e2πi(t/γm)/cF(t,[m]c).\left(F\cdot V_{\alpha}\right)(t,[m]_{c})=e^{2\pi i\left(t/\gamma-m\right)/c}F(t,[m]_{c}). (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 PP be a projection in Mm(Aα0)M_{m}\left(A_{\alpha_{0}}\right) with τα2n(P)=c0α0+d0=c2nα2n+d2n\tau_{\alpha_{2n}}\left(P\right)=c_{0}\alpha_{0}+d_{0}=c_{2n}\alpha_{2n}+d_{2n}. Assume PP satisfies Condition 4.3. For each nn\in\mathbb{N}, let a2na_{2n} and b2nb_{2n} be a pair of integers satisfying a2nd2nb2nc2n=1a_{2n}d_{2n}-b_{2n}c_{2n}=1 and set

β2n=a2nα2n+b2nc2nα2n+d2n.\beta_{2n}=\dfrac{a_{2n}\alpha_{2n}+b_{2n}}{c_{2n}\alpha_{2n}+d_{2n}}.

In particular, we can uniquely choose a2na_{2n} and b2nb_{2n} so that β2n\beta_{2n} is between 0 and 11. Then for each nn\in\mathbb{N}, Cc(×c2n)C_{c}\left(\mathbb{R}\times\mathbb{Z}_{c_{2n}}\right), suitably completed, has the structure of a Aβ2nA_{\beta_{2n}}-Aα2nA_{\alpha_{2n}}-equivalence bimodule.

Moreover, we can form a noncommutative solenoid 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} by taking the direct limit of irrational rotation algebras Aβ2nA_{\beta_{2n}} with the appropriate connecting maps.

Proof.

Since each connecting map φn:Aα2nAα2n+2\varphi_{n}:A_{\alpha_{2n}}\rightarrow A_{\alpha_{2n+2}} is a unital inclusion, we can view PP as a projection in Mm(Aα2n)M_{m}\left(A_{\alpha_{2n}}\right) for all nn. By Lemma 4.4, c2nc_{2n} and d2nd_{2n} together generate \mathbb{Z}. The Aβ2nA_{\beta_{2n}}-Aα2nA_{\alpha_{2n}}-equivalence bimodule structure is then the one from Theorem 4.8.

Since c2nc_{2n} and d2nd_{2n} are coprime, we can find integers a2na_{2n} and b2nb_{2n} such that a2nd2nb2nc2n=1a_{2n}d_{2n}-b_{2n}c_{2n}=1. Moreover, if a2na_{2n} and b2nb_{2n} are a pair of such integers, as are a2n+c2na_{2n}+\ell c_{2n} and b2n+d2nb_{2n}+\ell d_{2n} for any integer \ell. If β2n=(a2nα2n+b2n)/(c2nα2n+d2n)\beta_{2n}=\left(a_{2n}\alpha_{2n}+b_{2n}\right)/\left(c_{2n}\alpha_{2n}+d_{2n}\right), then

(α2n+c2n)+(b2n+d2n)c2nα2n+d2n=β2n+.\dfrac{\left(\alpha_{2n}+\ell c_{2n}\right)+\left(b_{2n}+\ell d_{2n}\right)}{c_{2n}\alpha_{2n}+d_{2n}}=\beta_{2n}+\ell.

It follows that there is a unique pair of a2na_{2n} and b2nb_{2n} putting β2n\beta_{2n} is between 0 and 11.

We show that the family of irrational rotation algebras (Aβ2n)n\left(A_{\beta_{2n}}\right)_{n\in\mathbb{N}} form a noncommutative solenoid by taking the direct limit with the prescribed connecting maps. To this end, we simply need to check that p2β2n+2β2nmodp^{2}\beta_{2n+2}\equiv\beta_{2n}\mod\mathbb{Z} for all nn\in\mathbb{N}.

To ease notation, we denote a2na_{2n}, b2nb_{2n}, c2nc_{2n}, d2nd_{2n}, a2n+2a_{2n+2}, b2n+2b_{2n+2}, c2n+2c_{2n+2}, and d2n+2d_{2n+2} by aa, bb, cc, dd, AA, BB, CC, and DD, respectively. By definition of α\alpha, C=cp2C=cp^{2} and D=dckD=d-ck for some of k{0,,p21}k\in\{0,\dotsc,p^{2}-1\}. We have

β2n=aα2n+bcα2n+d,andβ2n+2=Aα2n+2+BCα2n+2+D=Aα2n+2+B(cp2)α2n+2+(dkc).\beta_{2n}=\dfrac{a\alpha_{2n}+b}{c\alpha_{2n}+d},\quad\text{and}\quad\beta_{2n+2}=\dfrac{A\alpha_{2n+2}+B}{C\alpha_{2n+2}+D}=\dfrac{A\alpha_{2n+2}+B}{(cp^{2})\alpha_{2n+2}+(d-kc)}.

We show that p2β2n+2β2np^{2}\beta_{2n+2}-\beta_{2n} is an integer. With a few steps of algebra, one checks that

p2β2n+2β2n=(Aa)α2n+(Ak+Bp2b)cα2n+d.p^{2}\beta_{2n+2}-\beta_{2n}=\dfrac{(A-a)\alpha_{2n}+(Ak+Bp^{2}-b)}{c\alpha_{2n}+d}.

Note that AdAckBcp2=1Ad-Ack-Bcp^{2}=1, so Ak=(AdBcp21)/cAk=\left(Ad-Bcp^{2}-1\right)/c, and arrive at the following simplification:

(Aa)α2n+(Ak+Bp2b)cα2n+d=Aac.\dfrac{(A-a)\alpha_{2n}+(Ak+Bp^{2}-b)}{c\alpha_{2n}+d}=\dfrac{A-a}{c}.

Since 1=adbc=Ad(Ak+Bp2)c1=ad-bc=Ad-\left(Ak+Bp^{2}\right)c, it must be the case that Aa=cA-a=\ell c for some integer \ell. Thus, p2β2n+2β2nmodp^{2}\beta_{2n+2}\equiv\beta_{2n}\mod\mathbb{Z}. Note that if both β2n\beta_{2n} and β2n+2\beta_{2n+2} are chosen to be between 0 and 11, then p2β2n+2β2n{0,,p21}p^{2}\beta_{2n+2}-\beta_{2n}\in\{0,\dotsc,p^{2}-1\}. ∎

Remark 4.15.

In the above proposition, the sequence (β2n)n\left(\beta_{2n}\right)_{n\in\mathbb{N}} determines a unique element β=(βk)k\beta=\left(\beta_{k}\right)_{k\in\mathbb{N}} of Ξp\Xi_{p} by Remark 2.3. By the same remark, all other choices for the a2na_{2n}’s and b2nb_{2n}’s give rise to sequences in Ωp\Omega_{p} that lead to the same noncommutative solenoid 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}}.

Let X2n=Cc(×c2n)¯X_{2n}=\overline{C_{c}\left(\mathbb{R}\times\mathbb{Z}_{c_{2n}}\right)} denote the Aβ2nA_{\beta_{2n}}-Aα2nA_{\alpha_{2n}}-equivalence bimodule given in the above proposition. When working with X2nX_{2n}, results are almost always formulated in terms for the following dense subset. For fCc()f\in C_{c}\left(\mathbb{R}\right) and j{0,,c2n1}j\in\{0,\dotsc,c_{2n}-1\}, let fδjf\delta_{j} be the function in Cc(×c2n)C_{c}\left(\mathbb{R}\times\mathbb{Z}_{c_{2n}}\right) given by

fδj(t,[m])={f(t) if [m]=[j] in c2n,0 otherwise.f\delta_{j}\left(t,[m]\right)=\begin{cases}f\left(t\right)&\text{ if }[m]=[j]\text{ in }\mathbb{Z}_{c_{2n}},\\ 0&\text{ otherwise}.\end{cases}

One can regard δj\delta_{j} as the indicator function on [j]c2n[j]\in\mathbb{Z}_{c_{2n}}. Then the set

{fδj:jfCc() and j{0,,c2n1}}\left\{f\delta_{j}\>:j\in f\in C_{c}\left(\mathbb{R}\right)\text{ and }j\in\{0,\dotsc,c_{2n}-1\}\right\}

spans a dense subalgebra of Cc(×c2n)C_{c}\left(\mathbb{R}\times\mathbb{Z}_{c_{2n}}\right).

We now define the key embedding of X2nX_{2n} in X2n+2X_{2n+2} for each nn that gives us the directed systems of equivalence bimodules.

Definition 4.16.

We define ιn:X2nX2n+2\iota_{n}:X_{2n}\rightarrow X_{2n+2} on functions of the form fδjf\delta_{j} and extend by linearity (Since c2n\mathbb{Z}_{c_{2n}} is finite, extending by linearity is especially simple):

ιn(fδj)(t,[m])=1pi=0p1f(tp)δjp+ic0p2n+1([m]),\iota_{n}(f\delta_{j})(t,[m])=\dfrac{1}{\sqrt{p}}\sum_{i=0}^{p-1}f\left(\dfrac{t}{p}\right)\delta_{jp+ic_{0}p^{2n+1}}([m]), (4.13)

for tt\in\mathbb{R} and [m]c2n+2[m]\in\mathbb{Z}_{c_{2n+2}}. That is, ιn(fδj)(t,[m])\iota_{n}(f\delta_{j})(t,[m]) evaluates to (1/p)f(t/p)\left(1/\sqrt{p}\right)f(t/p) when m=jp+ic0p2n+1+p2n+2m=jp+ic_{0}p^{2n+1}+\ell p^{2n+2} for i{0,,p1}i\in\{0,\dotsc,p-1\} and \ell\in\mathbb{Z}, and zero otherwise.

It remains to show that (X2n,ιn)n\left(X_{2n},\iota_{n}\right)_{n\in\mathbb{N}} satisfies Definition 4.1. Specifically, we need to check that Equations 4.1 - 4.4 are satisfied.

Notation 4.17.

In the remainder of this section, for 𝒜α𝒮=lim(Aα2n,ψn)n\mathscr{A}_{\alpha}^{\mathscr{S}}=\varinjlim\left(A_{\alpha_{2n}},\psi_{n}\right)_{n\in\mathbb{N}}, we let 𝒜β𝒮=lim(Aβ2n,φn)n\mathscr{A}_{\beta}^{\mathscr{S}}=\varinjlim\left(A_{\beta_{2n}},\varphi_{n}\right)_{n\in\mathbb{N}} be the irrational noncommutative solenoid constructed in Proposition 4.14. Both embeddings φn:Aβ2nAβ2n+2\varphi_{n}:A_{\beta_{2n}}\rightarrow A_{\beta_{2n+2}} and ψn:Aα2nAα2n+2\psi_{n}:A_{\alpha_{2n}}\rightarrow A_{\alpha_{2n+2}} are as given in Lemma 2.6. Lastly, for any fCc()f\in C_{c}\left(\mathbb{R}\right), we let fξ(t)f_{\xi}(t) be f(t+ξ)f(t+\xi) for any fixed ξ\xi\in\mathbb{R}.

Lemma 4.18.

For each nn\in\mathbb{N}, ιn:X2nX2n+2\iota_{n}:X_{2n}\rightarrow X_{2n+2} preserves left module action. That is,

ιn(fF)=φn(f)ιn(F),\iota_{n}\left(f\cdot F\right)=\varphi_{n}\left(f\right)\cdot\iota_{n}\left(F\right), (4.14)

for all fAβ2nf\in A_{\beta_{2n}} and FX2nF\in X_{2n}.

Proof.

We check that ιn\iota_{n} respects left module action by showing that for any fCc()f\in C_{c}\left(\mathbb{R}\right) and j{0,,c2n1}j\in\{0,\dotsc,c_{2n}-1\},

ιn(Uβ2n(fδj))=Uβ2n+2pιn(fδj), and ιn(Vβ2n(fδj))=Vβ2n+2pιn(fδj).\iota_{n}\left(U_{\beta_{2n}}\cdot\left(f\delta_{j}\right)\right)=U_{\beta_{2n+2}}^{p}\cdot\iota_{n}\left(f\delta_{j}\right),\text{ and }\iota_{n}\left(V_{\beta_{2n}}\cdot\left(f\delta_{j}\right)\right)=V_{\beta_{2n+2}}^{p}\cdot\iota_{n}\left(f\delta_{j}\right).

Since τα2n(P)=c0α0+d0=c2nα2n+d2n\tau_{\alpha_{2n}}\left(P\right)=c_{0}\alpha_{0}+d_{0}=c_{2n}\alpha_{2n}+d_{2n}, we have γ2n=1/(c2nα2n+d2n)=1/(c0α0+d0)\gamma_{2n}=1/\left(c_{2n}\alpha_{2n}+d_{2n}\right)=1/\left(c_{0}\alpha_{0}+d_{0}\right) (see Notation 4.6), and we simply denote this number by γ\gamma.

For any (t,[m])×c2n+2\left(t,[m]\right)\in\mathbb{R}\times\mathbb{Z}_{c_{2n+2}}, it follows from Equation 4.7 that

ιn(Uβ2n(fδj))(t,[m])\displaystyle\iota_{n}\left(U_{\beta_{2n}}\cdot\left(f\delta_{j}\right)\right)\left(t,[m]\right) =ιn(fγδj+1)(t,[m])\displaystyle=\iota_{n}\left(f_{-\gamma}\delta_{j+1}\right)\left(t,[m]\right)
=1pi=0p1fγδ(j+1)p+ic0p2n+1(tp,[m]).\displaystyle=\dfrac{1}{\sqrt{p}}\sum_{i=0}^{p-1}f_{-\gamma}\delta_{(j+1)p+ic_{0}p^{2n+1}}\left(\dfrac{t}{p},[m]\right).

On the other hand,

Uβ2n+2pιn(fδj)(t,[m])\displaystyle U_{\beta_{2n+2}}^{p}\cdot\iota_{n}\left(f\delta_{j}\right)\left(t,[m]\right) =ιn(fδj)(tpγ,[mp])\displaystyle=\iota_{n}\left(f\delta_{j}\right)\left(t-p\gamma,[m-p]\right)
=1pi=0p1fδjp+ic0p2n+1(tpγp,[mp])\displaystyle=\dfrac{1}{\sqrt{p}}\sum_{i=0}^{p-1}f\delta_{jp+ic_{0}p^{2n+1}}\left(\frac{t-p\gamma}{p},[m-p]\right)
=1pi=0p1fγδjp+ic0p2n+1+p(tp,[m]).\displaystyle=\dfrac{1}{\sqrt{p}}\sum_{i=0}^{p-1}f_{-\gamma}\delta_{jp+ic_{0}p^{2n+1}+p}\left(\frac{t}{p},[m]\right).

For any (t,[m])×c2n(t,[m])\in\mathbb{R}\times\mathbb{Z}_{c_{2n}}, by Equation 4.8,

(Vβ2n(fδj))(t,[m])={e2πi(ta2nj)/c2nf(t)if [m]=[j] in c2n,0otherwise.\left(V_{\beta_{2n}}\cdot\left(f\delta_{j}\right)\right)(t,[m])=\begin{cases}e^{2\pi i\left(t-a_{2n}j\right)/c_{2n}}f(t)&\text{if }[m]=[j]\text{ in }\mathbb{Z}_{c_{2n}},\\ 0&\text{otherwise}.\end{cases}

Then for (t,[m])×c2n+2\left(t,[m]\right)\in\mathbb{R}\times\mathbb{Z}_{c_{2n+2}},

ιn(Vβ2n(fδj))(t,[m])=(1/p)e2πi(t/pa2nj)/c2nf(t/p)\iota_{n}\left(V_{\beta_{2n}}\cdot\left(f\delta_{j}\right)\right)\left(t,[m]\right)=\left(1/\sqrt{p}\right)e^{2\pi i(t/p-a_{2n}j)/c_{2n}}f(t/p)

when m=jp+ic0p2n+1+p2n+2m=jp+ic_{0}p^{2n+1}+\ell p^{2n+2} for any i{0,,p1}i\in\{0,\dotsc,p-1\} and \ell\in\mathbb{Z}. Otherwise it evaluates to zero.

On the other hand, using the fact that pc2n=c0p2n+1=c2n+2/ppc_{2n}=c_{0}p^{2n+1}=c_{2n+2}/p, we have

Vβ2n+2pιn(fδj)(t,[m])\displaystyle V_{\beta_{2n+2}}^{p}\cdot\iota_{n}\left(f\delta_{j}\right)\left(t,[m]\right) =e2πi[(ta2n+2m)/c2n+2]pιn(fδj)(t,[m])\displaystyle=e^{2\pi i\left[(t-a_{2n+2}m)/c_{2n+2}\right]\cdot p}\iota_{n}\left(f\delta_{j}\right)\left(t,[m]\right)
=1pe2πi(ta2n+2m)/pc2ni=0p1fδjp+ic0p2n+1(tp,[m]),\displaystyle=\dfrac{1}{\sqrt{p}}e^{2\pi i(t-a_{2n+2}m)/pc_{2n}}\sum_{i=0}^{p-1}f\delta_{jp+ic_{0}p^{2n+1}}\left(\frac{t}{p},[m]\right),

which is nonzero for m=jp+ic0p2n+1+p2n+2m=jp+ic_{0}p^{2n+1}+\ell p^{2n+2} for any i{0,,p1}i\in\{0,\dotsc,p-1\} and \ell\in\mathbb{Z}. Notice that the evaluation is independent of the choice of representative for [m]c2n+2[m]_{c_{2n+2}}, so we let =0\ell=0 and

(1/p)e2πi[ta2n+2(jp+ic0p2n+1)/pc2n]\displaystyle\left(1/\sqrt{p}\right)e^{2\pi i\left[t-a_{2n+2}\left(jp+ic_{0}p^{2n+1}\right)/pc_{2n}\right]} =(1/p)e2πi[ta2n+2(jp+ic0p2n+1)/(c0p2n+1)]\displaystyle=\left(1/\sqrt{p}\right)e^{2\pi i\left[t-a_{2n+2}\left(jp+ic_{0}p^{2n+1}\right)/\left(c_{0}p^{2n+1}\right)\right]}

Moreover, since a2n+2=a2n+zc0p2na_{2n+2}=a_{2n}+zc_{0}p^{2n} for some zz\in\mathbb{Z},

ta2n+2(jp+ic0p2n+1)c0p2n+1\displaystyle\dfrac{t-a_{2n+2}\left(jp+ic_{0}p^{2n+1}\right)}{c_{0}p^{2n+1}} =t(a2n+zc0p2n)(jp+ic0p2n+1)c0p2n+1\displaystyle=\dfrac{t-\left(a_{2n}+zc_{0}p^{2n}\right)\left(jp+ic_{0}p^{2n+1}\right)}{c_{0}p^{2n+1}}
=ta2njpc0p2n+1+K2n\displaystyle=\dfrac{t-a_{2n}jp}{c_{0}p^{2n+1}}+K_{2n}
=t/pa2njc0p2n+K2n\displaystyle=\dfrac{t/p-a_{2n}j}{c_{0}p^{2n}}+K_{2n}
=t/pa2njc2n+K2n\displaystyle=\dfrac{t/p-a_{2n}j}{c_{2n}}+K_{2n}

with integer K2n=a2nzjzic0p2nK_{2n}=-a_{2n}-zj-zic_{0}p^{2n}. This shows that

e2πi(ta2n+2(jp+ic0p2n+1)/pc2n)=e2πi(t/pa2nj)/c2ne^{2\pi i\left(t-a_{2n+2}\left(jp+ic_{0}p^{2n+1}\right)/pc_{2n}\right)}=e^{2\pi i\left(t/p-a_{2n}j\right)/c_{2n}}

and allows us to the conclude that ιn(Vβ2n(fδj))=Vβ2n+2pιn(fδj)\iota_{n}\left(V_{\beta_{2n}}\cdot\left(f\delta_{j}\right)\right)=V_{\beta_{2n+2}}^{p}\cdot\iota_{n}\left(f\delta_{j}\right). ∎

Lemma 4.19.

For each nn\in\mathbb{N}, ιn:X2nX2n+2\iota_{n}:X_{2n}\rightarrow X_{2n+2} preserves right module action. That is,

ιn(Fg)=ιn(F)ψn(g),\iota_{n}\left(F\cdot g\right)=\iota_{n}\left(F\right)\cdot\psi_{n}\left(g\right), (4.15)

for all gAα2ng\in A_{\alpha_{2n}} and FX2nF\in X_{2n}.

Proof.

We check that ιn\iota_{n} respects left module action by showing that for any fCc()f\in C_{c}\left(\mathbb{R}\right) and j{0,,c2n1}j\in\{0,\dotsc,c_{2n}-1\},

ιn((fδj)Uα2n)=ιn(fδj)Uα2n+2p, and ιn((fδj)Vα2n)=ιn(fδj)Vα2n+2p.\iota_{n}\left(\left(f\delta_{j}\right)\cdot U_{\alpha_{2n}}\right)=\iota_{n}\left(f\delta_{j}\right)\cdot U_{\alpha_{2n+2}}^{p},\text{ and }\iota_{n}\left(\left(f\delta_{j}\right)\cdot V_{\alpha_{2n}}\right)=\iota_{n}\left(f\delta_{j}\right)\cdot V_{\alpha_{2n+2}}^{p}.

Since τα2n(P)=c0α0+d0=c2nα2n+d2n\tau_{\alpha_{2n}}\left(P\right)=c_{0}\alpha_{0}+d_{0}=c_{2n}\alpha_{2n}+d_{2n}, we have γ2n=1/(c2nα2n+d2n)=1/(c0α0+d0)\gamma_{2n}=1/\left(c_{2n}\alpha_{2n}+d_{2n}\right)=1/\left(c_{0}\alpha_{0}+d_{0}\right) (see Notation 4.6), and we simply denote this number by γ\gamma.

For any (t,[m])×c2n+2\left(t,[m]\right)\in\mathbb{R}\times\mathbb{Z}_{c_{2n+2}}, it follows from Equation 4.11 that

ιn((fδj)Uα2n)(t,[m])\displaystyle\iota_{n}\left(\left(f\delta_{j}\right)\cdot U_{\alpha_{2n}}\right)\left(t,[m]\right) =ιn(f1δj+d2n)(t,[m])\displaystyle=\iota_{n}\left(f_{-1}\delta_{j+d_{2n}}\right)\left(t,[m]\right)
=1pi=0p1f1δ(j+d2n)p+ic0p2n+1(tp,[m])\displaystyle=\dfrac{1}{\sqrt{p}}\sum_{i=0}^{p-1}f_{-1}\delta_{\left(j+d_{2n}\right)p+ic_{0}p^{2n+1}}\left(\dfrac{t}{p},[m]\right)

On the other hand,

(ιn(fδj)Uα2n+2p)(t,[m])\displaystyle\left(\iota_{n}\left(f\delta_{j}\right)\cdot U_{\alpha_{2n+2}}^{p}\right)\left(t,[m]\right) =ιn(fδj)(tp,[mpd2n+2])\displaystyle=\iota_{n}\left(f\delta_{j}\right)\left(t-p,[m-pd_{2n+2}]\right)
=1pi=0p1fδjp+ic0p2n+1(tpp,[mpd2n+2])\displaystyle=\dfrac{1}{\sqrt{p}}\sum_{i=0}^{p-1}f\delta_{jp+ic_{0}p^{2n+1}}\left(\dfrac{t-p}{p},[m-pd_{2n+2}]\right)
=1pi=0p1f1δjp+ic0p2n+1+pd2n+2(tp,[m])\displaystyle=\dfrac{1}{\sqrt{p}}\sum_{i=0}^{p-1}f_{-1}\delta_{jp+ic_{0}p^{2n+1}+pd_{2n+2}}\left(\dfrac{t}{p},[m]\right)

Since d2n+2=d2nc0x2np2nx2n+1p2n+1d_{2n+2}=d_{2n}-c_{0}x_{2n}p^{2n}-x_{2n+1}p^{2n+1},

jp+ic0p2n+1+pd2n+2jp+ic0p2n+1+pd2n+c0x2np2n+1modp2n+2jp+ic_{0}p^{2n+1}+pd_{2n+2}\equiv jp+ic_{0}p^{2n+1}+pd_{2n}+c_{0}x_{2n}p^{2n+1}\mod p^{2n+2}

For i=0,,p1i=0,\dotsc,p-1, ic0p2n+1ic_{0}p^{2n+1} goes through the same equivalence classes of c0p2n+2=c2n+2\mathbb{Z}_{c_{0}p^{2n+2}}=\mathbb{Z}_{c_{2n+2}} as (i+x2n)c0p2n+1\left(i+x_{2n}\right)c_{0}p^{2n+1}, so

δ(j+d2n)p+ic0p2n+1=δjp+ic0p2n+1+pd2n+2,\delta_{\left(j+d_{2n}\right)p+ic_{0}p^{2n+1}}=\delta_{jp+ic_{0}p^{2n+1}+pd_{2n+2}},

and ιn((fδj)Uα2n)=ιn(fδj)Uα2n+2p\iota_{n}\left(\left(f\delta_{j}\right)\cdot U_{\alpha_{2n}}\right)=\iota_{n}\left(f\delta_{j}\right)\cdot U_{\alpha_{2n+2}}^{p}.

For any (t,[m])×c2n\left(t,[m]\right)\in\mathbb{R}\times\mathbb{Z}_{c_{2n}}, by Equation 4.12, we have

((fδj)Vα2n)(t,[m])={e2πi(t/γj)/c2nf(t)if [m]=[j] in c2n,0otherwise.\left(\left(f\delta_{j}\right)\cdot V_{\alpha_{2n}}\right)\left(t,[m]\right)=\begin{cases}e^{2\pi i\left(t/\gamma-j\right)/c_{2n}}f(t)&\text{if }[m]=[j]\text{ in }\mathbb{Z}_{c_{2n}},\\ 0&\text{otherwise}.\end{cases}

Then for (t,[m])×c2n+2\left(t,[m]\right)\in\mathbb{R}\times\mathbb{Z}_{c_{2n+2}},

ιn((fδj)Vα2n)(t,[m])=(1/p)e2πi(t/(pγ)j)/c2nf(t/p)\iota_{n}\left(\left(f\delta_{j}\right)\cdot V_{\alpha_{2n}}\right)\left(t,[m]\right)=\left(1/\sqrt{p}\right)e^{2\pi i\left(t/\left(p\gamma\right)-j\right)/c_{2n}}f(t/p)

when m=jp+ic0p2n+1+p2n+2m=jp+ic_{0}p^{2n+1}+\ell p^{2n+2} for any i{0,,p1}i\in\{0,\dotsc,p-1\} and \ell\in\mathbb{Z}. Otherwise it evaluates to zero.

On the other hand, since pc2n=c0p2n+1=c2n+2/ppc_{2n}=c_{0}p^{2n+1}=c_{2n+2}/p, we have

(ιn(fδj)Vα2n+2p)(t,[m])\displaystyle\left(\iota_{n}\left(f\delta_{j}\right)\cdot V_{\alpha_{2n+2}}^{p}\right)\left(t,[m]\right) =e2πi[t/γm/c2n+2]pιn(fδj)(t,[m])\displaystyle=e^{2\pi i\left[t/\gamma-m/c_{2n+2}\right]\cdot p}\iota_{n}\left(f\delta_{j}\right)\left(t,[m]\right)
=e2πi(t/γm/pc2n)ιn(fδj)(t,[m])\displaystyle=e^{2\pi i\left(t/\gamma-m/pc_{2n}\right)}\iota_{n}\left(f\delta_{j}\right)\left(t,[m]\right)
=1pe2πi(t/γm/pc2n)i=0p1fδjp+ic0p2n+1(tp,[m]),\displaystyle=\dfrac{1}{\sqrt{p}}e^{2\pi i\left(t/\gamma-m/pc_{2n}\right)}\sum_{i=0}^{p-1}f\delta_{jp+ic_{0}p^{2n+1}}\left(\frac{t}{p},[m]\right),

which is nonzero for m=jp+ic0p2n+1+p2n+2m=jp+ic_{0}p^{2n+1}+\ell p^{2n+2} for any i{0,,p1}i\in\{0,\dotsc,p-1\} and \ell\in\mathbb{Z}. Choose =0\ell=0 (the computations are independent of this choice), we have that for m=jp+ic0p2n+1m=jp+ic_{0}p^{2n+1},

e2πi(t/γm/pc2n)\displaystyle e^{2\pi i\left(t/\gamma-m/pc_{2n}\right)} =e2πi[(t/γjp+ic0p2n+1)/pc2n]\displaystyle=e^{2\pi i\left[\left(t/\gamma-jp+ic_{0}p^{2n+1}\right)/pc_{2n}\right]}
=e2πi[(t/(pγ)j+ic0p2n)/c2n]\displaystyle=e^{2\pi i\left[\left(t/\left(p\gamma\right)-j+ic_{0}p^{2n}\right)/c_{2n}\right]}
=e2πi[(t/(pγ)j)/c2n]\displaystyle=e^{2\pi i\left[\left(t/\left(p\gamma\right)-j\right)/c_{2n}\right]}

The last equality follow from c0p2n=c2nc_{0}p^{2n}=c_{2n}. In conclusion, ιn((fδj)Vα2n)=ιn(fδj)Vα2n+2p\iota_{n}\left(\left(f\delta_{j}\right)\cdot V_{\alpha_{2n}}\right)=\iota_{n}\left(f\delta_{j}\right)\cdot V_{\alpha_{2n+2}}^{p}. ∎

We now prove the following major lemma of this paper.

Lemma 4.20.

For each nn\in\mathbb{N}, ιn:X2nX2n+2\iota_{n}:X_{2n}\rightarrow X_{2n+2} preserves left inner product. That is,

φn(F1,F2Aβ2n)=ιn(F1),ιn(F2)Aβ2n+2,\varphi_{n}\left(\langle F_{1},F_{2}\rangle_{A_{\beta_{2n}}}\right)=\langle\iota_{n}\left(F_{1}\right),\iota_{n}\left(F_{2}\right)\rangle_{A_{\beta_{2n+2}}}, (4.16)

for all F1,F2X2nF_{1},F_{2}\in X_{2n}.

We first give a formula for ιn(fδj,gδjAβ2n)\iota_{n}\left(\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}\right) in the following lemma.

Lemma 4.21.

Fix f,gCc()f,g\in C_{c}\left(\mathbb{R}\right) and j,j{0,,c2n1}j,j^{\prime}\in\{0,\dotsc,c_{2n}-1\}. Let rr\in\mathbb{R} and k=(jj)p+c0p2n+1k=(j-j^{\prime})p+\ell c_{0}p^{2n+1} for any \ell\in\mathbb{Z},

φn(fδj,gδjAβ2n)(r,k)\displaystyle\varphi_{n}\left(\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}\right)(r,k)
=mf(c0p2n+1r+ja2n+mc0p2n)g(c0p2n+1r+ja2n+mc0p2n[(jj)+c0p2n]γ)¯\displaystyle=\sum_{m\in\mathbb{Z}}f\left(c_{0}p^{2n+1}r+ja_{2n}+mc_{0}p^{2n}\right)\overline{g\left(c_{0}p^{2n+1}r+ja_{2n}+mc_{0}p^{2n}-[(j-j^{\prime})+\ell c_{0}p^{2n}]\gamma\right)}

and ιn(fδj,gδjAβ2n)(r,k)=0\iota_{n}\left(\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}\right)(r,k)=0 otherwise.

Proof.

We begin by computing fδj,gδjAβ2n\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}} for f,gCc()f,g\in C_{c}(\mathbb{R}) and j,j{0,,c2n1}j,j^{\prime}\in\{0,\dotsc,c_{2n}-1\}. For any (r,k)×(r,k)\in\mathbb{R}\times\mathbb{Z}, it follow from Equation 4.5 that

fδj,gδjAβ2n(r,k)=mfδj(c0p2nr+m,[d2nm]c0p2n)gδj(c0p2nr+mkγ,[d2nmk]c0p2n)¯\displaystyle\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}(r,k)=\sum_{m\in\mathbb{Z}}f\delta_{j}\left(c_{0}p^{2n}r+m,[d_{2n}m]_{c_{0}p^{2n}}\right)\overline{g\delta_{j^{\prime}}\left(c_{0}p^{2n}r+m-k\gamma,[d_{2n}m-k]_{c_{0}p^{2n}}\right)}

First, the summand associated to mm is only nonzero when d2nmjmodc0p2nd_{2n}m\equiv j\mod c_{0}p^{2n}. Since a2nd2n1modc0p2na_{2n}d_{2n}\equiv 1\mod c_{0}p^{2n}, d2nmjmodc0p2nd_{2n}m\equiv j\mod c_{0}p^{2n} if and only if mja2nmodc0p2nm\equiv ja_{2n}\mod c_{0}p^{2n}, so

fδj,gδjAβ2n(r,k)=mf(c0p2nr+ja2n+mc0p2n)×gδj(c0p2nr+ja2n+mc0p2nkγ,[jk]c0p2n)¯.\displaystyle\begin{split}\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}(r,k)&=\sum_{m\in\mathbb{Z}}f\left(c_{0}p^{2n}r+ja_{2n}+mc_{0}p^{2n}\right)\\ &\times\overline{g\delta_{j^{\prime}}\left(c_{0}p^{2n}r+ja_{2n}+mc_{0}p^{2n}-k\gamma,[j-k]_{c_{0}p^{2n}}\right)}.\end{split}

Since gδjg\delta_{j^{\prime}} is only nonzero when jkjmodc0p2nj-k\equiv j^{\prime}\mod c_{0}p^{2n}, fδj,gδjAβ2n(r,k)\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}(r,k) is only nonzero when k=(jj)+c0p2nk=(j-j^{\prime})+\ell c_{0}p^{2n} for \ell\in\mathbb{Z}, in which case

fδj,gδjAβ2n(r,k)=mf(c0p2nr+ja2n+mc0p2n)×g(c0p2nr+ja2n+mc0p2n[(jj)+c0p2n]γ)¯,\displaystyle\begin{split}\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}(r,k)&=\sum_{m\in\mathbb{Z}}f\left(c_{0}p^{2n}r+ja_{2n}+mc_{0}p^{2n}\right)\\ &\times\overline{g\left(c_{0}p^{2n}r+ja_{2n}+mc_{0}p^{2n}-\left[(j-j^{\prime})+\ell c_{0}p^{2n}\right]\gamma\right)},\end{split}

and fδj,gδjAβ2n(r,k)=0\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}(r,k)=0 otherwise. The formula for φn(fδj,gδjAβ2n)(r,k)\varphi_{n}\left(\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}\right)(r,k) then follows from the embedding of Aβ2nA_{\beta_{2n}} into Aβ2n+2A_{\beta_{2n+2}} characterized in Lemma 2.6. ∎

Proof of Lemma 4.20.

We now compute ιn(fδj),ιn(gδj)Aβ2n+2\langle\iota_{n}\left(f\delta_{j}\right),\iota_{n}\left(g\delta_{j^{\prime}}\right)\rangle_{A_{\beta_{2n+2}}}. For a fixed pair of jj and jj^{\prime} in {0,,c2n1}\{0,\dotsc,c_{2n}-1\} and [m]c2n+2[m]\in\mathbb{Z}_{c_{2n+2}}, we have

ιn(fδj)(t,[m])=1pi=0p1f(tp)δjp+ic0p2n+1([m]),\iota_{n}(f\delta_{j})(t,[m])=\dfrac{1}{\sqrt{p}}\sum_{i=0}^{p-1}f\left(\dfrac{t}{p}\right)\delta_{jp+ic_{0}p^{2n+1}}([m]),

and

ιn(gδj)(t,[m])=1pi=0p1g(tp)δjp+ic0p2n+1([m]).\iota_{n}(g\delta_{j^{\prime}})(t,[m])=\dfrac{1}{\sqrt{p}}\sum_{i^{\prime}=0}^{p-1}g\left(\dfrac{t}{p}\right)\delta_{j^{\prime}p+i^{\prime}c_{0}p^{2n+1}}([m]).

For rr\in\mathbb{R} and kk\in\mathbb{Z}, we compute

ιn(fδj),\displaystyle\langle\iota_{n}\left(f\delta_{j}\right), ιn(gδj)Aβ2n+2(r,k)\displaystyle\iota_{n}\left(g\delta_{j^{\prime}}\right)\rangle_{A_{\beta_{2n+2}}}(r,k)
=1pi=0p1f(rp)δjp+ic0p2n+1,1pi=0p1g(rp)δjp+ic0p2n+1Aβ2n+2(r,k)\displaystyle=\bigg{\langle}\dfrac{1}{\sqrt{p}}\sum_{i=0}^{p-1}f\left(\frac{r}{p}\right)\delta_{jp+ic_{0}p^{2n+1}},\dfrac{1}{\sqrt{p}}\sum_{i^{\prime}=0}^{p-1}g\left(\frac{r}{p}\right)\delta_{j^{\prime}p+i^{\prime}c_{0}p^{2n+1}}\bigg{\rangle}_{A_{\beta_{2n+2}}}(r,k)
=1pi=0p1i=0p1f(rp)δjp+ic0p2n+1,g(rp)δjp+ic0p2n+1Aβ2n+2(r,k).\displaystyle=\dfrac{1}{p}\sum_{i=0}^{p-1}\sum_{i^{\prime}=0}^{p-1}\big{\langle}f\left(\frac{r}{p}\right)\delta_{jp+ic_{0}p^{2n+1}},g\left(\frac{r}{p}\right)\delta_{j^{\prime}p+i^{\prime}c_{0}p^{2n+1}}\big{\rangle}_{A_{\beta_{2n+2}}}(r,k).

We focus on the inner product involving ii and ii^{\prime} and compute

f(rp)δjp+ic0p2n+1,\displaystyle\big{\langle}f\left(\frac{r}{p}\right)\delta_{jp+ic_{0}p^{2n+1}}, g(rp)δjp+ic0p2n+1Aβ2n+2(r,k)\displaystyle g\left(\frac{r}{p}\right)\delta_{j^{\prime}p+i^{\prime}c_{0}p^{2n+1}}\big{\rangle}_{A_{\beta_{2n+2}}}(r,k)
=\displaystyle= mf(c0p2n+2r+(jp+ic0p2n+1)a2n+2+mc0p2n+2p)\displaystyle\sum_{m^{\prime}\in\mathbb{Z}}f\left(\dfrac{c_{0}p^{2n+2}r+(jp+ic_{0}p^{2n+1})a_{2n+2}+m^{\prime}c_{0}p^{2n+2}}{p}\right)
×g(c0p2n+2r+(jp+ic0p2n+1)a2n+2+mc0p2n+2kγp)¯\displaystyle\times\overline{g\left(\dfrac{c_{0}p^{2n+2}r+(jp+ic_{0}p^{2n+1})a_{2n+2}+m^{\prime}c_{0}p^{2n+2}-k\gamma}{p}\right)}
=\displaystyle= mf(c0p2n+1r+(j+ic0p2n)a2n+2+mc0p2n+1)\displaystyle\sum_{m^{\prime}\in\mathbb{Z}}f\left(c_{0}p^{2n+1}r+(j+ic_{0}p^{2n})a_{2n+2}+m^{\prime}c_{0}p^{2n+1}\right)
×g(c0p2n+1r+(j+ic0p2n)a2n+2+mc0p2n+1[(jj)+(ii)c0p2n+c0p2n+1]γ)¯\displaystyle\times\overline{g\left(c_{0}p^{2n+1}r+(j+ic_{0}p^{2n})a_{2n+2}+m^{\prime}c_{0}p^{2n+1}-[(j-j^{\prime})+(i-i^{\prime})c_{0}p^{2n}+\ell c_{0}p^{2n+1}]\gamma\right)}

when k=(jj)p+(ii)c0p2n+1+c0p2n+2=(jj)p+[(ii)+p]c0p2n+1k=(j-j^{\prime})p+(i-i^{\prime})c_{0}p^{2n+1}+\ell c_{0}p^{2n+2}=(j-j^{\prime})p+[(i-i^{\prime})+\ell p]c_{0}p^{2n+1} for \ell\in\mathbb{Z}, and the inner product evaluates to zero otherwise.

Since ii and ii^{\prime} are both integers between 0 and p1p-1, (ii)+p(i-i^{\prime})+\ell p must be K+pK+\ell p for some K{0,,p1}K\in\{0,\dotsc,p-1\}. Specifically, there are pKp-K pairs of (i,i)(i,i^{\prime}), with iii\geq i^{\prime}, such that (ii)+p=K+p(i-i^{\prime})+\ell p=K+\ell p, and KK pairs of (i,i)(i,i^{\prime}), with i<ii<i^{\prime}, such that (ii)+(+1)p=K+p(i-i^{\prime})+(\ell+1)p=K+\ell p. Together, there are exactly pp nonzero inner products contributing to the value of ιn(fδj),ιn(gδj)Aβ2n+2(r,k)\langle\iota_{n}(f\delta_{j}),\iota_{n}(g\delta_{j^{\prime}})\rangle_{A_{\beta_{2n+2}}}(r,k) for k=(jj)p+[K+p]c0p2n+1k=(j-j^{\prime})p+[K+\ell p]c_{0}p^{2n+1}. Therefore, for each fixed ii, there are pp nonzero inner products, each with value

mf(c0p2n+1r+(j+ic0p2n)a2n+2+mc0p2n+1)\displaystyle\sum_{m^{\prime}\in\mathbb{Z}}f\left(c_{0}p^{2n+1}r+(j+ic_{0}p^{2n})a_{2n+2}+m^{\prime}c_{0}p^{2n+1}\right)
×g(c0p2n+1r+(j+ic0p2n)a2n+2+mc0p2n+1[(jj)+[K+p]c0p2n]γ)¯.\displaystyle\hskip 56.9055pt\times\overline{g\left(c_{0}p^{2n+1}r+(j+ic_{0}p^{2n})a_{2n+2}+m^{\prime}c_{0}p^{2n+1}-[(j-j^{\prime})+\left[K+\ell p\right]c_{0}p^{2n}]\gamma\right)}.

Therefore,

ιn(fδj),ιn(gδj)Aβ2n+2(r,(jj)p+[K+p]c0p2n+1)\displaystyle\langle\iota_{n}\left(f\delta_{j}\right),\iota_{n}\left(g\delta_{j^{\prime}}\right)\rangle_{A_{\beta_{2n+2}}}(r,(j-j^{\prime})p+\left[K+\ell p\right]c_{0}p^{2n+1})
=\displaystyle= 1ppi=0p1(mf(c0p2n+1r+(j+ic0p2n)a2n+2+mc0p2n+1)\displaystyle\dfrac{1}{p}\cdot p\sum_{i=0}^{p-1}\Big{(}\sum_{m^{\prime}\in\mathbb{Z}}f\left(c_{0}p^{2n+1}r+(j+ic_{0}p^{2n})a_{2n+2}+m^{\prime}c_{0}p^{2n+1}\right)
×g(c0p2n+1r+(j+ic0p2n)a2n+2+mc0p2n+1[(jj)+[K+p]c0p2n]γ)¯)\displaystyle\times\overline{g\left(c_{0}p^{2n+1}r+(j+ic_{0}p^{2n})a_{2n+2}+m^{\prime}c_{0}p^{2n+1}-[(j-j^{\prime})+\left[K+\ell p\right]c_{0}p^{2n}]\gamma\right)}\Big{)}
=\displaystyle= i=0p1(mf(c0p2n+1r+ja2n+2+(ia2n+2+mp)c0p2n)\displaystyle\sum_{i=0}^{p-1}\Big{(}\sum_{m^{\prime}\in\mathbb{Z}}f\left(c_{0}p^{2n+1}r+ja_{2n+2}+\left(ia_{2n+2}+m^{\prime}p\right)c_{0}p^{2n}\right)
×g(c0p2n+1r+ja2n+2+(ia2n+2+mp)c0p2n[(jj)+[K+p]c0p2n]γ)¯)\displaystyle\times\overline{g\left(c_{0}p^{2n+1}r+ja_{2n+2}+\left(ia_{2n+2}+m^{\prime}p\right)c_{0}p^{2n}-[(j-j^{\prime})+\left[K+\ell p\right]c_{0}p^{2n}]\gamma\right)}\Big{)}

Since a2n+2a_{2n+2} is relatively prime to pp, as it is relatively prime to c2n=c0p2nc_{2n}=c_{0}p^{2n} and ia2n+2ia_{2n+2} goes through the modular classes of pp exactly once as ii goes from 0 to p1p-1. Therefore,

i=0p1mia2n+2+mp=,\bigcup_{i=0}^{p-1}\bigcup_{m^{\prime}\in\mathbb{Z}}ia_{2n+2}+m^{\prime}p=\mathbb{Z},

so

ιn(fδj),ιn(gδj)Aβ2n+2(r,(jj)p+[K+p]c0p2n+1)\displaystyle\langle\iota_{n}\left(f\delta_{j}\right),\iota_{n}\left(g\delta_{j^{\prime}}\right)\rangle_{A_{\beta_{2n+2}}}(r,(j-j^{\prime})p+\left[K+\ell p\right]c_{0}p^{2n+1})
=\displaystyle= mf(c0p2n+1r+ja2n+2+mc0p2n)\displaystyle\sum_{m\in\mathbb{Z}}f\left(c_{0}p^{2n+1}r+ja_{2n+2}+mc_{0}p^{2n}\right)
×g(c0p2n+1r+ja2n+2+mc0p2n[(jj)+[K+p]c0p2n]γ)¯.\displaystyle\times\overline{g\left(c_{0}p^{2n+1}r+ja_{2n+2}+mc_{0}p^{2n}-[(j-j^{\prime})+\left[K+\ell p\right]c_{0}p^{2n}]\gamma\right)}.

We now need to relate a2n+2a_{2n+2} and a2na_{2n}. In the proof of Proposition 4.14, we showed that

a2n+2a2nc2n=z,\dfrac{a_{2n+2}-a_{2n}}{c_{2n}}=z\in\mathbb{Z},

so a2n+2=a2n+zc0p2na_{2n+2}=a_{2n}+zc_{0}p^{2n} and we have

ιn(fδj),ιn(gδj)Aβ2n+2(r,(jj)p+[K+p]c0p2n+1)\displaystyle\langle\iota_{n}\left(f\delta_{j}\right),\iota_{n}\left(g\delta_{j^{\prime}}\right)\rangle_{A_{\beta_{2n+2}}}(r,(j-j^{\prime})p+\left[K+\ell p\right]c_{0}p^{2n+1})
=\displaystyle= mf(c0p2n+1r+j(a2n+zc0p2n)+mc0p2n)\displaystyle\sum_{m\in\mathbb{Z}}f\left(c_{0}p^{2n+1}r+j\left(a_{2n}+zc_{0}p^{2n}\right)+mc_{0}p^{2n}\right)
×g(c0p2n+1r+j(a2n+zc0p2n)+mc0p2n[(jj)+[K+p]c0p2n]γ)¯\displaystyle\times\overline{g\left(c_{0}p^{2n+1}r+j\left(a_{2n}+zc_{0}p^{2n}\right)+mc_{0}p^{2n}-[(j-j^{\prime})+\left[K+\ell p\right]c_{0}p^{2n}]\gamma\right)}
=\displaystyle= mf(c0p2n+1r+ja2n+(jz+m)c0p2n)\displaystyle\sum_{m\in\mathbb{Z}}f\left(c_{0}p^{2n+1}r+ja_{2n}+\left(jz+m\right)c_{0}p^{2n}\right)
×g(c0p2n+1r+ja2n+(jz+m)c0p2n[(jj)+[K+p]c0p2n]γ)¯\displaystyle\times\overline{g\left(c_{0}p^{2n+1}r+ja_{2n}+\left(jz+m\right)c_{0}p^{2n}-[(j-j^{\prime})+\left[K+\ell p\right]c_{0}p^{2n}]\gamma\right)}
=\displaystyle= mf(c0p2n+1r+ja2n+mc0p2n)\displaystyle\sum_{m\in\mathbb{Z}}f\left(c_{0}p^{2n+1}r+ja_{2n}+mc_{0}p^{2n}\right)
×g(c0p2n+1r+ja2n+mc0p2n[(jj)+[K+p]c0p2n]γ)¯\displaystyle\times\overline{g\left(c_{0}p^{2n+1}r+ja_{2n}+mc_{0}p^{2n}-[(j-j^{\prime})+\left[K+\ell p\right]c_{0}p^{2n}]\gamma\right)}

Since jj and zz are fixed, jz+mjz+m goes through the integers exactly once as mm runs through \mathbb{Z}, so the last equality follows.

Lastly, as KK goes from 0 to p1p-1, and \ell goes over all integers, the sum K+pK+\ell p will go over the integers exactly once, so we have that, for k=(jj)p+c0p2n+1k=(j-j^{\prime})p+\ell c_{0}p^{2n+1},

ιn(fδj),ιn(gδj)Aβ2n+2(r,k)\displaystyle\langle\iota_{n}\left(f\delta_{j}\right),\iota_{n}\left(g\delta_{j^{\prime}}\right)\rangle_{A_{\beta_{2n+2}}}(r,k)
=mf(c0p2n+1r+ja2n+mc0p2n)g(c0p2n+1r+ja2n+mc0p2n[(jj)+c0p2n]γ)¯\displaystyle=\sum_{m\in\mathbb{Z}}f\left(c_{0}p^{2n+1}r+ja_{2n}+mc_{0}p^{2n}\right)\overline{g\left(c_{0}p^{2n+1}r+ja_{2n}+mc_{0}p^{2n}-[(j-j^{\prime})+\ell c_{0}p^{2n}]\gamma\right)}

and ιn(fδj),ιn(gδj)Aβ2n+2(r,k)=0\langle\iota_{n}\left(f\delta_{j}\right),\iota_{n}\left(g\delta_{j^{\prime}}\right)\rangle_{A_{\beta_{2n+2}}}(r,k)=0 otherwise, which is exactly φn(fδj,gδjAβ2n)(r,k)\varphi_{n}\left(\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}\right)(r,k) in Lemma 4.21. We conclude that

φn(fδj,gδjAβ2n)=ιn(fδj),ιn(gδj)Aβ2n+2.\varphi_{n}\left(\langle f\delta_{j},g\delta_{j^{\prime}}\rangle_{A_{\beta_{2n}}}\right)=\langle\iota_{n}\left(f\delta_{j}\right),\iota_{n}\left(g\delta_{j^{\prime}}\right)\rangle_{A_{\beta_{2n+2}}}.

The main result of this section is the following.

Theorem 4.22.

Fix prime pp and let α=(αn)nΞp\alpha=\left(\alpha_{n}\right)_{n\in\mathbb{N}}\in\Xi_{p} be an irrational sequence with pαn+1=αn+xnp\alpha_{n+1}=\alpha_{n}+x_{n} for all nn\in\mathbb{N}. Let PP be a projection in Mm(Aα0)M_{m}\left(A_{\alpha_{0}}\right), m1m\geq 1, that satisfies Condition 4.3. Then (X2n,ιn)n\left(X_{2n},\iota_{n}\right)_{n\in\mathbb{N}} is a directed system of equivalence bimodules adapted to the sequence (PMm(Aα2n)P,φn)n=(Aβ2n,φn)n\left(PM_{m}\left(A_{\alpha_{2n}}\right)P,\varphi_{n}\right)_{n\in\mathbb{N}}=\left(A_{\beta_{2n}},\varphi_{n}\right)_{n\in\mathbb{N}} and (Aα2n,ψn)n\left(A_{\alpha_{2n}},\psi_{n}\right)_{n\in\mathbb{N}}.

Proof.

We showed in Proposition 4.14 that X2nX_{2n} is an Aβ2nA_{\beta_{2n}}-Aα2nA_{\alpha_{2n}}-equivalence bimodule for each nn. By Lemma 4.18 and 4.20, each embedding preserves left module action and left inner product, that is,

ιn(fF)=φn(f)ιn(F), and φn(F1,F2Aβ2n)=ιn(F1),ιn(F2)Aβ2n+2,\iota_{n}\left(f\cdot F\right)=\varphi_{n}\left(f\right)\cdot\iota_{n}\left(F\right),\text{ and }\varphi_{n}\left(\langle F_{1},F_{2}\rangle_{A_{\beta_{2n}}}\right)=\langle\iota_{n}\left(F_{1}\right),\iota_{n}\left(F_{2}\right)\rangle_{A_{\beta_{2n+2}}},

for fAβ2nf\in A_{\beta_{2n}} and FF, F1F_{1}, and F2F_{2} in X2nX_{2n}. Moreover, Lemma 4.19 shows that each embedding respects the right modules action.

It only remains to show that ψn(F1,F2Aα2n)=ιn(F1),ιn(F2)Aα2n+2\psi_{n}(\langle F_{1},F_{2}\rangle_{A_{\alpha_{2n}}})=\langle\iota_{n}(F_{1}),\iota_{n}(F_{2})\rangle_{A_{\alpha_{2n+2}}} for all F1,F2X2nF_{1},F_{2}\in X_{2n}. First, observe that for any F,G,HX2nF,G,H\in X_{2n}, we have

F,GAβ2nH=FG,HAα2n,\langle F,G\rangle_{A_{\beta_{2n}}}\cdot H=F\cdot\langle G,H\rangle_{A_{\alpha_{2n}}},

so

ιn(F,GAβ2nH)=ιn(FG,HAα2n),\iota_{n}\left(\langle F,G\rangle_{A_{\beta_{2n}}}\cdot H\right)=\iota_{n}\left(F\cdot\langle G,H\rangle_{A_{\alpha_{2n}}}\right),

or

φn(F,GAβ2n)ιn(H)=ιn(F)ψn(G,HAα2n).\varphi_{n}\left(\langle F,G\rangle_{A_{\beta_{2n}}}\right)\cdot\iota_{n}\left(H\right)=\iota_{n}\left(F\right)\cdot\psi_{n}\left(\langle G,H\rangle_{A_{\alpha_{2n}}}\right).

By Lemma 4.20, the left side of the equality is equal to ιn(F),ιn(G)Aβ2n+2ιn(H)\langle\iota_{n}\left(F\right),\iota_{n}\left(G\right)\rangle_{A_{\beta_{2n+2}}}\cdot\iota_{n}\left(H\right), which is equal to ιn(F)ιn(G),ιn(H)Aα2n+2\iota_{n}\left(F\right)\cdot\langle\iota_{n}\left(G\right),\iota_{n}\left(H\right)\rangle_{A_{\alpha_{2n+2}}}. Therefore, for any fixed G,HX2nG,H\in X_{2n},

ιn(F)ιn(G),ιn(H)Aα2n+2=ιn(F)ψn(G,HAα2n)\iota_{n}\left(F\right)\cdot\langle\iota_{n}\left(G\right),\iota_{n}\left(H\right)\rangle_{A_{\alpha_{2n+2}}}=\iota_{n}\left(F\right)\cdot\psi_{n}\left(\langle G,H\rangle_{A_{\alpha_{2n}}}\right)

for all FX2nF\in X_{2n}. This is sufficient for us to conclude that

ιn(G),ιn(H)Aα2n+2=ψn(G,HAα2n)\langle\iota_{n}\left(G\right),\iota_{n}\left(H\right)\rangle_{A_{\alpha_{2n+2}}}=\psi_{n}\left(\langle G,H\rangle_{A_{\alpha_{2n}}}\right)

for all G,HX2nG,H\in X_{2n} as follows. We have shown in Lemma 4.20 that ιn(X2n),ιn(X2n)Aβ2n+2=ψn(Aβ2n)\langle\iota_{n}\left(X_{2n}\right),\iota_{n}\left(X_{2n}\right)\rangle_{A_{\beta_{2n+2}}}=\psi_{n}\left(A_{\beta_{2n}}\right). One checks that ιn(X2n),ιn(X2n)Aα2n+2ψn(Aα2n)\langle\iota_{n}\left(X_{2n}\right),\iota_{n}\left(X_{2n}\right)\rangle_{A_{\alpha_{2n+2}}}\subseteq\psi_{n}\left(A_{\alpha_{2n}}\right). Therefore, we can regard ιn(X2n)\iota_{n}\left(X_{2n}\right) as a Aβ2nA_{\beta_{2n}}-Aα2nA_{\alpha_{2n}}-equivalence bimodule. By [9, Proposition 3.8], the map ϕ:Aα2n𝒦(ιn(X2n))\phi:A_{\alpha_{2n}}\rightarrow\mathcal{K}\left(\iota_{n}\left(X_{2n}\right)\right) from Aα2nA_{\alpha_{2n}} to the compact operators on ιn(X2n)\iota_{n}\left(X_{2n}\right) given by aϕa(x)=xaa\mapsto\phi_{a}(x)=x\cdot a is an isomorphism of C*-algebras. Let a=ιn(G),ιn(H)Aα2n+2a=\langle\iota_{n}\left(G\right),\iota_{n}\left(H\right)\rangle_{A_{\alpha_{2n+2}}} and a=ψn(G,HAα2n)a^{\prime}=\psi_{n}\left(\langle G,H\rangle_{A_{\alpha_{2n}}}\right), we have ϕa=ϕa\phi_{a}=\phi_{a^{\prime}} on all of ιn(X2n)\iota_{n}\left(X_{2n}\right), which implies that a=aa=a^{\prime}, as we desired to show. ∎

Corollary 4.23.

Let 𝒜α𝒮=limAα2n\mathscr{A}_{\alpha}^{\mathscr{S}}=\varinjlim A_{\alpha_{2n}} and 𝒜β𝒮=limAβ2n\mathscr{A}_{\beta}^{\mathscr{S}}=\varinjlim A_{\beta_{2n}} be given as in Theorem 4.22. Then 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} and 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} are Morita equivalent.

Proof.

This follows directly from [7, Theorem 3.4] by taking the direct limit of (X2n,ιn)n\left(X_{2n},\iota_{n}\right)_{n\in\mathbb{N}}. ∎

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 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} with α=(αn)nΞp\alpha=\left(\alpha_{n}\right)_{n\in\mathbb{N}}\in\Xi_{p} and pαn+1=αn+xnp\alpha_{n+1}=\alpha_{n}+x_{n}, xn{0,,p1}x_{n}\in\{0,\dots,p-1\} for all nn. Assume also that α00\alpha_{0}\neq 0 and x00x_{0}\neq 0, so the associated pp-adic integer xα=j=0xjpjx_{\alpha}=\sum_{j=0}^{\infty}x_{j}p^{j} has inverse xα1=j=0yjpjpx_{\alpha}^{-1}=\sum_{j=0}^{\infty}y_{j}p^{j}\in\mathbb{Z}_{p}.

For such a fixed 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}}, the Morita equivalent solenoid 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} that arises from the Heisenberg bimodules construction (see Theorem 3.5) is given by

β=(1α0pn+j=0n1yjpjpn)n.\beta=\left(\dfrac{1}{\alpha_{0}p^{n}}+\dfrac{\sum_{j=0}^{n-1}y_{j}p^{j}}{p^{n}}\right)_{n\in\mathbb{N}}.

Following Remark 4.5 and Notation 4.2, the projection PAα0P\in A_{\alpha_{0}} with τα0(P)=α0\tau_{\alpha_{0}}\left(P\right)=\alpha_{0} satisfies Condition 4.3, so by Proposition 4.14, the noncommutative solenoid 𝒜β𝒮=limPAα2nP\mathscr{A}_{\beta}^{\mathscr{S}}=\varinjlim PA_{\alpha_{2n}}P is determined by

β2n=β2n=a2nα2n+b2nc2nα2n+d2n,\beta_{2n}=\beta_{2n}=\dfrac{a_{2n}\alpha_{2n}+b_{2n}}{c_{2n}\alpha_{2n}+d_{2n}},

with

a2n=j=02n1yjpj,b2n=((j=02n1yjpj)(j=02n1xjpj)+1)p2n,a_{2n}=\sum_{j=0}^{2n-1}y_{j}p^{j},\ b_{2n}=\left(\left(\sum_{j=0}^{2n-1}y_{j}p^{j}\right)\left(-\sum_{j=0}^{2n-1}x_{j}p^{j}\right)+1\right)p^{-2n},
c2n=p2n,and d2n=j=02n1xjpj.c_{2n}=p^{2n},\ \text{and }d_{2n}=-\sum_{j=0}^{2n-1}x_{j}p^{j}.

By Lemma 3.4, b2nb_{2}n is an integer and one checks that a2nd2nb2nc2n=1a_{2n}d_{2n}-b_{2n}c_{2n}=1 for all nn. Observe that, since pnαn=α0+j=0n1xjpjp^{n}\alpha_{n}=\alpha_{0}+\sum_{j=0}^{n-1}x_{j}p^{j},

β2n\displaystyle\beta_{2n} =a2nα2n+b2np2nα2n+(j=02n1xjpj)\displaystyle=\dfrac{a_{2n}\alpha_{2n}+b_{2n}}{p^{2n}\alpha_{2n}+\left(-\sum_{j=0}^{2n-1}x_{j}p^{j}\right)}
=(j=02n1yjpj)α2n+((j=02n1yjpj)(j=02n1xjpj)+1)p2np2nα2n+(j=02n1xjpj)\displaystyle=\dfrac{\left(\sum_{j=0}^{2n-1}y_{j}p^{j}\right)\alpha_{2n}+\left(\left(\sum_{j=0}^{2n-1}y_{j}p^{j}\right)\left(-\sum_{j=0}^{2n-1}x_{j}p^{j}\right)+1\right)p^{-2n}}{p^{2n}\alpha_{2n}+\left(-\sum_{j=0}^{2n-1}x_{j}p^{j}\right)}
=(j=02n1yjpj)α2np2n+((j=02n1yjpj)(j=02n1xjpj)+1)α0p2n\displaystyle=\dfrac{\left(\sum_{j=0}^{2n-1}y_{j}p^{j}\right)\alpha_{2n}p^{2n}+\left(\left(\sum_{j=0}^{2n-1}y_{j}p^{j}\right)\left(-\sum_{j=0}^{2n-1}x_{j}p^{j}\right)+1\right)}{\alpha_{0}p^{2n}}
=(j=02n1yjpj)(α0+j=02n1xjpj)+((j=02n1yjpj)(j=02n1xjpj)+1)α0p2n\displaystyle=\dfrac{\left(\sum_{j=0}^{2n-1}y_{j}p^{j}\right)\left(\alpha_{0}+\sum_{j=0}^{2n-1}x_{j}p^{j}\right)+\left(\left(\sum_{j=0}^{2n-1}y_{j}p^{j}\right)\left(-\sum_{j=0}^{2n-1}x_{j}p^{j}\right)+1\right)}{\alpha_{0}p^{2n}}
=α0(j=02n1yjpj)+1α0p2n\displaystyle=\dfrac{\alpha_{0}\left(\sum_{j=0}^{2n-1}y_{j}p^{j}\right)+1}{\alpha_{0}p^{2n}}
=1α0p2n+j=02n1yjpjp2n,\displaystyle=\dfrac{1}{\alpha_{0}p^{2n}}+\dfrac{\sum_{j=0}^{2n-1}y_{j}p^{j}}{p^{2n}},

which are exactly the even entries of β\beta calculated from the Heisenberg bimodules construction. By Remark 2.3, since the two β\beta’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 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} with α=(αn)nΞp\alpha=\left(\alpha_{n}\right)_{n\in\mathbb{N}}\in\Xi_{p} and pαn+1=αn+xnp\alpha_{n+1}=\alpha_{n}+x_{n}, xn{0,,p1}x_{n}\in\{0,\dots,p-1\} for all nn. Assume also that α00\alpha_{0}\neq 0 and x00x_{0}\neq 0. The Morita equivalent solenoid 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} formed from the directed system of equivalence bimodule construction, using projection PAα0P\in A_{\alpha_{0}} with τα0(P)=α0\tau_{\alpha_{0}}\left(P\right)=\alpha_{0}, is the same as the one from the Heisenberg bimodule construction.

Remark 4.25.

It remains an open question as to whether or not we can extend this result to the 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} given in Theorem 3.6, which generalizes Theorem 3.5 used in the above discussion.

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 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} be an irrational noncommutative solenoid, then it does not contain a unital C*-subalgebra isomorphic to Mm(Aθ)M_{m}\left(A_{\theta}\right) for any mm strictly greater than 11.

Proof.

In [6], Latrémolière and Packer showed that an irrational noncommutative solenoid is simple and has a unique tracial state, which we denote by τ\tau. Moreover, the range of this trace on K0(𝒜α𝒮)K_{0}\left(\mathscr{A}_{\alpha}^{\mathscr{S}}\right) is given by {z+yαk:z,y,k}\{z+y\alpha_{k}\>:z,y\in\mathbb{Z},k\in\mathbb{N}\}. The restriction of τ\tau to any unital C*-subalgebra B of 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} is again a tracial state on BB. For a contradiction, suppose that BMm(Aθ)B\cong M_{m}\left(A_{\theta}\right) for m2m\geq 2, then the restriction of τ\tau must be the unique tracial state on Mm(Aθ)M_{m}\left(A_{\theta}\right). Since diag(1,0,,0)Mm(Aθ)\operatorname{diag}\left(1,0,\dotsc,0\right)\in M_{m}\left(A_{\theta}\right) is a projection with τ(diag(1,0,,0))=1/m\tau\left(\operatorname{diag}\left(1,0,\dotsc,0\right)\right)=1/m, the range τ\tau on K0(𝒜α𝒮)K_{0}\left(\mathscr{A}_{\alpha}^{\mathscr{S}}\right) must contain the fraction 1/m1/m and we arrive at a contradiction. ∎

It follows from the lemma below that every projection in Mm(𝒜α𝒮)M_{m}\left(\mathscr{A}_{\alpha}^{\mathscr{S}}\right) is unitarily equivalent to a projection PMm(Aα2k)P\in M_{m}\left(A_{\alpha_{2k}}\right) for some kk\in\mathbb{N}.

Lemma 5.2.

Let AA be the direct limit of the direct sequence of C*-algebras (An,φn)n=1\left(A_{n},\varphi_{n}\right)_{n=1}^{\infty} and suppose that φn:AnA\varphi^{n}:A_{n}\rightarrow A is the natural map for each nn. If pp is a projection in AA, then there is a k1k\geq 1 and a projection qAkq\in A_{k} such that pp and φk(q)\varphi^{k}\left(q\right) are unitarily equivalent.

Proof.

By [16, Proposition 5.2.6], it is sufficient to find a projection qAkq\in A_{k} such that pφk(q)\left\lVert p-\varphi^{k}\left(q\right)\right\rVert is strictly less than 11. Such a projection exists by [16, Corollary 5.1.7]. ∎

Theorem 5.3.

Let pp, qq be prime numbers. Let αΞp\alpha\in\Xi_{p} and βΞq\beta\in\Xi_{q}. Then the following statements are equivalent.

  1. (1)

    𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} and 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} are Morita equivalent.

  2. (2)

    p=qp=q and there exists a projection PMm(Aα2k)P\in M_{m}\left(A_{\alpha_{2k}}\right) for some k,mk,m\in\mathbb{N} that satisfies Condition 4.3, and 𝒜β𝒮limnkPMm(Aα2n)P\mathscr{A}_{\beta}^{\mathscr{S}}\cong\varinjlim_{n\geq k}PM_{m}\left(A_{\alpha_{2n}}\right)P.

Remark 5.4.

It is clear that truncating the first kk terms of α=(αn)n\alpha=\left(\alpha_{n}\right)_{n\in\mathbb{N}} does not change the resulting noncommutative solenoid. For this reason, we say that a projection PMm(Aα2k)P\in M_{m}\left(A_{\alpha_{2k}}\right) satisfies Condition 4.3 to mean: τα2n(P)=a2nα2n+b2n\tau_{\alpha_{2n}}\left(P\right)=a_{2n}\alpha_{2n}+b_{2n} and gcd(pc2n,d2nc2nx2n)=1\gcd\left(pc_{2n},d_{2n}-c_{2n}x_{2n}\right)=1. Essentially, our results in the previous section apply to any projection in Mm(Aαk)M_{m}\left(A_{\alpha_{k}}\right) for any kk\in\mathbb{N} by truncating the first kk terms of α\alpha starting from zero. Without loss of generality, we can always assume kk is even.

Proof of Theorem 5.3.

(2)(2) implying (1)(1) follows from the directed system of equivalence bimodules construction, as in Theorem 4.22 and its Corollary, with projection PP and truncated sequence (Aα2n,φn)nk\left(A_{\alpha_{2n}},\varphi_{n}\right)_{n\geq k}.

We show that (1)(1) implies (2)(2) as follows. In [8, Theorem 3.7], the KK-theory of noncommutative solenoids is computed: in particular, for αΞp\alpha\in\Xi_{p},

K1(𝒜α𝒮)=[1p]×[1p].K_{1}\left(\mathscr{A}_{\alpha}^{\mathscr{S}}\right)=\mathbb{Z}\left[\dfrac{1}{p}\right]\times\mathbb{Z}\left[\dfrac{1}{p}\right].

It is clear that for pp and qq primes, [1/p]×[1/p][1/q]×[1/q]\mathbb{Z}\left[1/p\right]\times\mathbb{Z}\left[1/p\right]\cong\mathbb{Z}\left[1/q\right]\times\mathbb{Z}\left[1/q\right] if and only if p=qp=q. Since Morita equivalent C*-algebras have isomorphic K1K_{1} groups, we must have p=qp=q.

Since both 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} and 𝒜β𝒮\mathscr{A}_{\beta}^{\mathscr{S}} are unital, 𝒜β𝒮PMm(𝒜α𝒮)P\mathscr{A}_{\beta}^{\mathscr{S}}\cong PM_{m}\left(\mathscr{A}_{\alpha}^{\mathscr{S}}\right)P for a suitable m1m\geq 1 and a projection PMm(𝒜α𝒮)P\in M_{m}\left(\mathscr{A}_{\alpha}^{\mathscr{S}}\right) [11, Proposition 2.1]. Recall that 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} is the direct limit of the sequence of increasing irrational rotation algebras associated to the α2n\alpha_{2n}’s. By Lemma 5.2, we can then assume that PMm(Aα2k)P\in M_{m}\left(A_{\alpha_{2k}}\right) for a large enough kk. It is evident that

𝒜β𝒮PMm(𝒜α𝒮)PlimnkPMm(Aα2n)P.\mathscr{A}_{\beta}^{\mathscr{S}}\cong PM_{m}\left(\mathscr{A}_{\alpha}^{\mathscr{S}}\right)P\cong\varinjlim_{n\geq k}PM_{m}\left(A_{\alpha_{2n}}\right)P.

It remains to show that PP must satisfy Condition 4.3. Assume, for the sake of contradiction, that PP does not satisfy the Condition. Then by Lemma 4.4, there exists NkN\geq k such that τα2N(P)=(cα2N+d)\tau_{\alpha_{2N}}\left(P\right)=\ell\left(c\alpha_{2N}+d\right), with gcd(c,d)=1\gcd\left(c,d\right)=1. As discussed in the paragraph following Lemma 4.7, PMm(Aα2n)PPM_{m}\left(A_{\alpha_{2n}}\right)P is then isomorphic to an ×\ell\times\ell matrix algebra over some irrational rotation algebra, say AβA_{\beta}. Moreover, it is easy to check that \ell is a common divisor of c2nc_{2n} and d2nd_{2n} for all nNn\geq N. So the direct limit C*-algebra contains M(Aβ)M_{\ell}\left(A_{\beta}\right) 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 PAα2nPPA_{\alpha_{2n}}P can be a direct limit C*-algebra of matrix algebras that are strictly increasing in size.

Example 5.5.

For α00\alpha_{0}\neq 0, consider the irrational noncommutative solenoid 𝒜α𝒮\mathscr{A}_{\alpha}^{\mathscr{S}} given by

α=(α0,α0p,α0p2,)Ξp.\alpha=\left(\alpha_{0},\dfrac{\alpha_{0}}{p},\dfrac{\alpha_{0}}{p^{2}},\dotsc\right)\in\Xi_{p}.

Let PP be the projection in Aα0A_{\alpha_{0}} with trace α0\alpha_{0}. As a projection in Aα2n,τα2n(p)=p2nα2nA_{\alpha_{2n}},\tau_{\alpha_{2n}}\left(p\right)=p^{2n}\alpha_{2n}. By Lemma 4.7, PAα2nPMp2n(A1α2n)PA_{\alpha_{2n}}P\cong M_{p^{2n}}\left(A_{\frac{1}{\alpha_{2n}}}\right). Therefore, by Lemma 5.1, PMn(𝒜α)PPM_{n}(\mathscr{A}_{\alpha})P cannot be a noncommutative solenoid. However, it would be interesting to further investigate the structure of the C*-algebras limPAα2nPlimMp2n(A1α2n)\varinjlim PA_{\alpha_{2n}}P\cong\varinjlim M_{p^{2n}}\left(A_{\frac{1}{\alpha_{2n}}}\right), 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 C\mathrm{C}^{*}-algebras, 2019.
  • [4] Kazunori Kodaka. Endomorphisms of certain irrational rotation CC^{*}-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 CC^{*}-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 CC^{*}-algebras. Math. Ann., 222(1):7–22, 1976.
  • [11] Marc A. Rieffel. CC^{\ast}-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 CC^{\ast}-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 CC^{\ast}-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 pp-adic analysis, volume 198 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2000.
  • [16] N. E. Wegge-Olsen. KK-theory and CC^{*}-algebras. Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, 1993. A friendly approach.
  • [17] Dana P. Williams. Crossed products of CC{{}^{\ast}}-algebras, volume 134 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2007.