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

Pseudo-Cartan Inclusions

David R. Pitts Department of Mathematics
University of Nebraska-Lincoln
Lincoln
NE 68588-0130 U.S.A.
dpitts2@unl.edu
Abstract.

A pseudo-Cartan inclusion is a regular inclusion having a Cartan envelope. Unital pseudo-Cartan inclusions were classified in [PittsStReInII]; we extend this classification to include the non-unital case. The class of pseudo-Cartan inclusions coincides with the class of regular inclusions having the faithful unique pseudo-expectation property and can also be described using the ideal intersection property. We describe the twisted groupoid associated with the Cartan envelope of a pseudo-Cartan inclusion. These results significantly extend previous results obtained for the unital setting.

We explore properties of pseudo-Cartan inclusions and the relationship between a pseudo-Cartan inclusion and its Cartan envelope. For example, if 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} is a pseudo-Cartan inclusion with Cartan envelope 𝒜{\mathcal{B}}\subseteq{\mathcal{A}}, then 𝒞{\mathcal{C}} is simple if and only if 𝒜{\mathcal{A}} is simple. Also every regular *-automorphism of 𝒞{\mathcal{C}} uniquely extends to a *-automorphism of 𝒜{\mathcal{A}}. We show that the inductive limit of pseudo-Cartan inclusions with suitable connecting maps is a pseudo-Cartan inclusion, and the minimal tensor product of pseudo-Cartan inclusions is a pseudo-Cartan inclusion. Further, we describe the Cartan envelope of pseudo-Cartan inclusions arising from these constructions. We conclude with some applications and a few open questions.

2020 Mathematics Subject Classification:
46L05

1. Introduction

A famous theorem of Gelfand states that an abelian CC^{*}-algebra 𝒞{\mathcal{C}} may be recognized as an algebra of continuous functions: 𝒞C0(X){\mathcal{C}}\simeq C_{0}(X), where X=𝒞^X=\hat{\mathcal{C}} is the space of non-zero multiplicative linear functionals on 𝒞{\mathcal{C}} equipped with the weak-* topology. In some cases, non-commutative CC^{*}-algebras can also be represented as functions on spaces. For example, Strătilă and Voiculescu [StratilaVoiculescuReAFAl], showed that for a unital AF-algebra 𝒜=lim𝒜n{\mathcal{A}}=\varinjlim{\mathcal{A}}_{n} (each 𝒜n{\mathcal{A}}_{n} is a finite dimensional CC^{*}-algebra), there exists a maximal abelian CC^{*}-subalgebra 𝒜{\mathcal{B}}\subseteq{\mathcal{A}} such that for every nn, 𝒜n{\mathcal{B}}\cap{\mathcal{A}}_{n} is maximal abelian in 𝒜n{\mathcal{A}}_{n}. Strătilă and Voiculescu used {\mathcal{B}} to construct a “coordinate system” XX for 𝒜{\mathcal{A}} which behaves in much the same way as a system of matrix units for Mn()M_{n}({\mathbb{C}}); furthermore, the AF-algebra 𝒜{\mathcal{A}} can be viewed as a collection of functions on XX. Shortly after Strătilă and Voiculescu’s work, Feldman and Moore [FeldmanMooreErEqReII] showed that if 𝒟{\mathcal{D}} is a WW^{*}-Cartan subalgebra of the (separably acting) von Neumann algebra {\mathcal{M}} and 𝒟L(X,μ){\mathcal{D}}\simeq L^{\infty}(X,\mu), then there is a “measured” equivalence relation RR on X×XX\times X along with a 2-cocycle σ\sigma on RR such that {\mathcal{M}} is, loosely speaking, isomorphic to an algebra of functions, M(R,σ)M(R,\sigma), on RR with product given by a “generalized matrix multiplication”; further, the isomorphism of {\mathcal{M}} onto M(R,σ)M(R,\sigma) carries 𝒟{\mathcal{D}} onto an algebra of functions supported on {(x,x):xX}\{(x,x):x\in X\}.

Efforts to adapt the ideas of Feldman and Moore to CC^{*}-algebras began with work of Renault in [RenaultGrApC*Al], continued with Kumjian’s work on CC^{*}-diagonals in [KumjianOnC*Di], and culminated with Renault’s definition of a Cartan MASA in a CC^{*}-algebra found in [RenaultCaSuC*Al]. Given a Cartan MASA 𝒟{\mathcal{D}} in the CC^{*}-algebra 𝒞{\mathcal{C}}, the Kumjian-Renault theory produces a Hausdorff, étale, and effective groupoid GG, along with a central groupoid extension G(0)×𝕋ΣGG^{(0)}\times{\mathbb{T}}\hookrightarrow\Sigma\twoheadrightarrow G; this is the CC^{*}-algebraic analog of Feldman-Moore’s equivalence relation RR and 2-cocycle σ\sigma. Moreover, there is a unique isomorphism of 𝒞{\mathcal{C}} onto a CC^{*}-algebra, Cr(Σ,G)C^{*}_{r}(\Sigma,G), which carries 𝒟{\mathcal{D}} onto C0(G(0))C_{0}(G^{(0)}). The algebra Cr(Σ,G)C^{*}_{r}(\Sigma,G) can be viewed as an algebra of functions on Σ\Sigma with a product generalizing matrix multiplication, so here again, 𝒞{\mathcal{C}} may be viewed as an algebra of functions on a space. Conversely, every such central extension can be used to construct a Cartan MASA in a CC^{*}-algebra, so the theory gives a complete description of Cartan inclusions 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}}. These results are presented in [RenaultCaSuC*Al]; for a less terse treatment, see [SimsGroupoidsBook]. While Renault and Kumjian worked in the context of separable CC^{*}-algebras, Raad showed in [RaadGeReThCaSu] that this assumption can be removed.

Despite the success of the Kumjian-Renault theory, there are desirable settings in the CC^{*}-algebra context where the axioms for a Cartan MASA in the algebra 𝒞{\mathcal{C}} are not satisfied; examples include the virtual Cartan inclusions introduced in [PittsStReInI, Definition 1.1] and the weak Cartan inclusions defined in [ExelPittsChGrC*AlNoHaEtGr, Definition 2.11.5]. For these classes, the inclusion 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} is regular and 𝒟{\mathcal{D}} is abelian. In such settings, it is possible to follow the Kumjian-Renault procedure to construct a groupoid GG and a central groupoid extension as above. However, when this is done, the resulting groupoid GG may fail to be Hausdorff, see [PittsStReInII, Theorem 4.4]. One approach to dealing with such situations is to accept the fact that non-Hausdorff groupoids may arise, and develop a theory which includes them. This is the approach taken in [ExelPittsChGrC*AlNoHaEtGr].

An alternate approach to studying regular inclusions of the form 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} with 𝒟{\mathcal{D}} abelian, is to seek embeddings of 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} into a Cartan inclusion 𝒜{\mathcal{B}}\subseteq{\mathcal{A}} so that the dynamics of 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} are embedded into the dynamics of 𝒜{\mathcal{B}}\subseteq{\mathcal{A}}. (Such embeddings are called regular embeddings.) The idea is that the coordinates from the larger inclusion 𝒜{\mathcal{B}}\subseteq{\mathcal{A}} might then be used to study 𝒞{\mathcal{C}}. In [PittsStReInI], we gave a characterization of those unital regular inclusions which can be embedded into a CC^{*}-diagonal or a Cartan inclusion. However, the CC^{*}-diagonal 𝒜{\mathcal{B}}\subseteq{\mathcal{A}} obtained from the construction in [PittsStReInI] may bear little relation to the inclusion 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}}, and the embedding of 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} into 𝒜{\mathcal{B}}\subseteq{\mathcal{A}} is not sufficiently rigid to transfer information from 𝒜{\mathcal{B}}\subseteq{\mathcal{A}} to 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}}.

Part of the motivation for [PittsStReInII] was to remedy this defect. In [PittsStReInII], the notion of the Cartan envelope is defined and two characterizations of those unital regular inclusions having a Cartan envelope are given. The Cartan envelope described in [PittsStReInII] is unique when it exists and is a minimal Cartan pair generated by the original inclusion 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}}. Also, in [PittsStReInII, Section 7], a description of the twisted groupoid associated to the Cartan envelope for 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} is given.

Many regular inclusions are not unital. Examples include inclusions from graph algebras (or higher rank graph algebras), from crossed product constructions, or arise from tensor product constructions (such as those in Section 6.3 below). It is therefore of interest to extend results on Cartan envelopes to the non-unital setting.

Three primary goals of the present paper are: first, to establish results on Cartan envelopes which apply in the non-unital context; second, to give results showing that the Cartan envelope of a regular inclusion carries information about the original inclusion; and third, to give constructions of pseudo-Cartan inclusions along with their Cartan envelopes. Extending the characterization of regular inclusions having a Cartan envelope to include non-unital regular inclusions is unexpectedly subtle. On first glance, one might expect that passing from the unital setting to the non-unital case would simply be a matter of adjoining units and then applying the results on Cartan envelopes for unital inclusions from [PittsStReInII]. However, the unitization process need not preserve normalizing elements (see Definition 2.1(d)), nor does it preserve regular mappings (Definition 2.1(i)). For a concrete example, suppose 𝔍C0(){\mathfrak{J}}\unlhd C_{0}({\mathbb{R}}) is the ideal of functions in C0()C_{0}({\mathbb{R}}) which vanish on [1,1][-1,1]. Any fC0()f\in C_{0}({\mathbb{R}}) is a normalizer for the inclusion 𝔍C0(){\mathfrak{J}}\subseteq C_{0}({\mathbb{R}}), yet ff is not a normalizer for the unitized inclusion 𝔍~(C0())\tilde{\mathfrak{J}}\subseteq(C_{0}({\mathbb{R}}))^{\sim} unless |f||f| is constant on [1,1][-1,1]. In particular, this shows that while the identity mapping on C0()C_{0}({\mathbb{R}}) is a regular map, its extension to (C0())(C_{0}({\mathbb{R}}))^{\sim} (that is, Id(C0())\operatorname{Id}_{(C_{0}({\mathbb{R}}))^{\sim}}) is not regular. Furthermore, [PittsNoApUnInC*Al, Example 3.1] gives an example of an inclusion 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} which is not regular, but whose unitization 𝒟~𝒞~\tilde{\mathcal{D}}\subseteq\tilde{\mathcal{C}} is a Cartan inclusion. Finally, a vexing issue we have been unable to resolve is whether regularity for an inclusion is preserved when units are adjoined. The ill behavior of regularity properties under the unitization process leads to serious technical difficulties when considering Cartan envelopes for non-unital inclusions. Using tools found in [PittsNoApUnInC*Al], we develop techniques to overcome those obstacles and are able to characterize those regular inclusions having a Cartan envelope; we shall describe the characterization in a moment.

Our work on Cartan envelopes yields an alternate set of axioms for a Cartan MASA 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}}. Recall that an inclusion 𝒜{\mathcal{B}}\subseteq{\mathcal{A}} of CC^{*}-algebras has the ideal intersection property (iip) if every non-zero ideal of 𝒜{\mathcal{A}} has non-trivial intersection with {\mathcal{B}}. We also use c=𝒜{\mathcal{B}}^{c}={\mathcal{A}}\cap{\mathcal{B}}^{\prime} for the relative commutant of {\mathcal{B}} in 𝒜{\mathcal{A}}. The columns of the following table give Renault’s axioms for a Cartan MASA and (equivalent) alternate axioms for a Cartan MASA.

Cartan Inclusions: Cartan Inclusions:
Renault’s Axioms Alternate Axioms
𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} is regular 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} is regular
𝒟{\mathcal{D}} is a MASA in 𝒞{\mathcal{C}} 𝒟{\mathcal{D}} and 𝒟c{\mathcal{D}}^{c} are abelian, and both
the inclusions 𝒟𝒟c{\mathcal{D}}\subseteq{\mathcal{D}}^{c} and 𝒟c𝒞{\mathcal{D}}^{c}\subseteq{\mathcal{C}}
have the ideal intersection property
\exists a faithful conditional expectation \exists a conditional expectation
E:𝒞𝒟E:{\mathcal{C}}\rightarrow{\mathcal{D}} E:𝒞𝒟E:{\mathcal{C}}\rightarrow{\mathcal{D}}

(Renault’s original definition of Cartan MASA also assumes 𝒟{\mathcal{D}} contains an approximate identity for 𝒞{\mathcal{C}}, but [PittsNoApUnInC*Al, Theorem 2.4] shows that assumption can be removed.) By keeping the first two entries of the right-hand column but not requiring the third, we obtain a class of inclusions which we call pseudo-Cartan inclusions.

The characterization theorem, Theorem 3, extends the characterizations of the existence of a Cartan envelope found for unital inclusions in [PittsStReInII] to include the non-unital case. Theorem 3 shows the equivalence of the following: the inclusion 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} (again with 𝒟{\mathcal{D}} abelian) has a Cartan envelope; 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} is a pseudo-Cartan inclusion; and 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} has the faithful unique pseudo-expectation property (see Definition 2.3). That an inclusion is a pseudo-Cartan inclusion if and only if it has the faithful unique pseudo-expectation property is the reason for the term pseudo-Cartan inclusion. A description of the Kumjian-Renault twist G(0)×𝕋ΣGG^{(0)}\times{\mathbb{T}}\hookrightarrow\Sigma\twoheadrightarrow G for the Cartan envelope of 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} is given in Theorem 5.2. This description exhibits Σ\Sigma as a family of linear functionals on 𝒞{\mathcal{C}} and GG is the quotient of Σ\Sigma by a certain action of 𝕋{\mathbb{T}}.

The class of pseudo-Cartan inclusions contains the following classes: Cartan inclusions; the virtual Cartan inclusions of [PittsStReInI]; the weak Cartan inclusions introduced in [ExelPittsChGrC*AlNoHaEtGr]; and inclusions of the form JC0(X)J\subseteq C_{0}(X), where XX is a locally compact Hausdorff space and JJ is an essential ideal in C0(X)C_{0}(X).

Section 2 sets terminology, notation, and also gives a number of foundational results needed throughout the paper. These results concern weakly non-degenerate inclusions, pseudo-expectations, and the ideal intersection property. Some of the results in Section 2, such as Theorem 2.3, illustrate the technical challenges which can arise when verifying that desirable properties of an inclusion are preserved under the unitization process.

After presenting our results on existence and uniqueness of Cartan envelopes in Section 3, we explore properties of pseudo-Cartan inclusions and relationships between a pseudo-Cartan inclusion and its Cartan envelope. Section 4 contains the definition of pseudo-Cartan inclusions along with a few properties shared by 𝒞{\mathcal{C}} and 𝒜{\mathcal{A}} when 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} is a pseudo-Cartan inclusion having Cartan envelope 𝒜{\mathcal{B}}\subseteq{\mathcal{A}}. For example, Proposition 4.3 shows 𝒞{\mathcal{C}} is simple if and only if 𝒜{\mathcal{A}} is simple. While a pseudo-Cartan inclusion  can be (minimally) embedded into a Cartan inclusion, Proposition 4.4 gives a reverse process: it shows how to construct a family of pseudo-Cartan inclusions from a given Cartan MASA 𝒜{\mathcal{B}}\subseteq{\mathcal{A}}, and also examines the Cartan envelope for pseudo-Cartan inclusions arising from this construction.

Section 5 constructs the Kumjian-Renault groupoid model for the Cartan envelope of a given pseudo-Cartan inclusion 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} starting with a family of linear functionals on 𝒞{\mathcal{C}}. In particular, this allows us to explicitly see the embedding of 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} into its CC^{*}-envelope 𝒜{\mathcal{B}}\subseteq{\mathcal{A}} by realizing 𝒞{\mathcal{C}} as an algebra of functions (with a convolution product as multiplication).

Section 6 contains a rigidity property of the Cartan envelope and explores some permanence properties of pseudo-Cartan inclusions and their Cartan envelopes. The rigidity property is Proposition 6.1: it shows that given a pseudo-Cartan inclusion 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} with Cartan envelope 𝒜{\mathcal{B}}\subseteq{\mathcal{A}}, any regular *-automorphism of 𝒞{\mathcal{C}} extends uniquely to a *-automorphism of 𝒜{\mathcal{A}}. We show in Theorem 6.2 that an inductive limit of pseudo-Cartan inclusions with suitable connecting maps is again a pseudo-Cartan inclusion inclusion and the Cartan envelope of such an inductive limit is the inductive limit of the Cartan envelopes of the approximating pseudo-Cartan inclusions. The proof of Theorem 6.2 uses an important mapping property of Cartan envelopes, Theorem 6.1. Theorem 6.3 shows that the minimal tensor product of two pseudo-Cartan inclusions is a pseudo-Cartan inclusion, and the Cartan envelope of their minimal tensor product is the minimal tensor product of their Cartan envelopes. Because quotients of pseudo-Cartan inclusions by regular ideals is the subject of a forthcoming paper (by Brown, Fuller, Reznikoff and the author), we do not consider quotients here.

We provide some applications of our work in Section 7. We show that when 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} is a unital pseudo-Cartan inclusion, the CC^{*}-envelope of a Banach algebra 𝒟𝒜𝒞{\mathcal{D}}\subseteq{\mathcal{A}}\subseteq{\mathcal{C}} is the CC^{*}-subalgebra of 𝒞{\mathcal{C}} generated by 𝒜{\mathcal{A}} and that 𝒟{\mathcal{D}} norms 𝒞{\mathcal{C}} in the sense of [PopSinclairSmithNoC*Al]. We combine these results to show that if, for i=1,2i=1,2: 𝒟i𝒞i{\mathcal{D}}_{i}\subseteq{\mathcal{C}}_{i} are pseudo-Cartan inclusions such that their unitizations 𝒟~i𝒞~i\tilde{\mathcal{D}}_{i}\subseteq\tilde{\mathcal{C}}_{i} are also pseudo-Cartan inclusions; 𝒜i{\mathcal{A}}_{i} are Banach algebras with 𝒟i𝒜i𝒞i{\mathcal{D}}_{i}\subseteq{\mathcal{A}}_{i}\subseteq{\mathcal{C}}_{i}; and u:𝒜1𝒜2u:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2} is an isometric isomorphism, then uu uniquely extends to a *-isomorphism of C(𝒜1)C^{*}({\mathcal{A}}_{1}) onto C(A2)C^{*}(A_{2}). While these applications have antecedents in the literature, our results are significant generalizations. We include a few open questions in Section 8.

We require [PittsStReInII, Theorem 6.9] at two points in our work: in Section 5.1(d), and in the proof of Lemma 7. While [PittsStReInII, Theorem 6.9] is correctly stated in [PittsStReInII], due to an error in the statement of [PittsStReInII, Lemma 2.3], the proof of [PittsStReInII, Theorem 6.9] is insufficient. The correct statement of [PittsStReInII, Lemma 2.3] and a complete proof of [PittsStReInII, Theorem 6.9] may be found in Appendix A.

We thank Adam Fuller for several useful comments and Alex Kumjian for a helpful comment regarding the history of the adaptation of the Feldman-Moore theory to the CC^{*}-algebraic context.

2. Foundations

2.1. Terminology and Notation

For a normed linear space XX, we shall use X#X^{\#} to denote the collection of all bounded linear functionals on XX.

Let 𝒜{\mathcal{A}} be a Banach algebra. When 𝒜{\mathcal{A}} is unital, we denote its unit by II or I𝒜I_{\mathcal{A}}. Unless explicitly stated otherwise, when referring to an ideal JJ in 𝒜{\mathcal{A}}, we will always assume JJ is closed and two-sided. When JJ is an ideal in 𝒜{\mathcal{A}}, we write J𝒜J\unlhd{\mathcal{A}}. The ideal JJ is called an essential ideal if for any non-zero ideal KAK\unlhd A, JK{0}J\cap K\neq\{0\}.

We will follow the notation and conventions regarding unitization of a CC^{*}-algebra found in [BlackadarOpAl, II.1.2.1]. For a CC^{*}-algebra 𝒜{\mathcal{A}}, let 𝒜:=𝒜×{\mathcal{A}}^{\dagger}:={\mathcal{A}}\times{\mathbb{C}} with the usual operations and norm which make it into a unital CC^{*}-algebra. Define 𝒜~\tilde{\mathcal{A}} and a *-monomorphism u𝒜:𝒜𝒜~u_{\mathcal{A}}:{\mathcal{A}}\rightarrow\tilde{\mathcal{A}} by

(2.1.1) 𝒜~\displaystyle\tilde{\mathcal{A}} :={𝒜if 𝒜 is unital𝒜if 𝒜 is not unital;\displaystyle:=\begin{cases}{\mathcal{A}}&\text{if ${\mathcal{A}}$ is unital}\\ {\mathcal{A}}^{\dagger}&\text{if ${\mathcal{A}}$ is not unital;}\end{cases}
and for x𝒜x\in{\mathcal{A}},
(2.1.2) u𝒜(x)\displaystyle u_{\mathcal{A}}(x) :={xif 𝒜 is unital(x,0)if 𝒜 is not unital.\displaystyle:=\begin{cases}x&\text{if ${\mathcal{A}}$ is unital}\\ (x,0)&\text{if ${\mathcal{A}}$ is not unital.}\end{cases}

We shall consider the trivial algebra {0}\{0\} to be unital. Also, when there is no danger of confusion, we will identify 𝒜{\mathcal{A}} with u𝒜(𝒜)u_{\mathcal{A}}({\mathcal{A}}), and regard 𝒜𝒜~{\mathcal{A}}\subseteq\tilde{\mathcal{A}}.

Suppose 𝒜1{\mathcal{A}}_{1} and 𝒜2{\mathcal{A}}_{2} are CC^{*}-algebras. Given a bounded linear map ψ:𝒜1𝒜2\psi:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2}, the map ψ~:𝒜~1𝒜~2\tilde{\psi}:\tilde{\mathcal{A}}_{1}\rightarrow\tilde{\mathcal{A}}_{2} given by

(2.1.3) ψ~(x)={u𝒜2(ψ(x))if 𝒜1 is unital;u𝒜2(ψ(a))+λI𝒜~2if 𝒜1 is not unital and x=(a,λ)𝒜~1\tilde{\psi}(x)=\begin{cases}u_{{\mathcal{A}}_{2}}(\psi(x))&\text{if ${\mathcal{A}}_{1}$ is unital;}\\ u_{{\mathcal{A}}_{2}}(\psi(a))+\lambda I_{\tilde{\mathcal{A}}_{2}}&\text{if ${\mathcal{A}}_{1}$ is not unital and $x=(a,\lambda)\in\tilde{\mathcal{A}}_{1}$}\end{cases}

will be called the standard extension of ψ\psi to 𝒜~1\tilde{\mathcal{A}}_{1}. We will frequently use the following property of the standard extension without comment:

(2.1.4) u𝒜2ψ=ψ~u𝒜1.u_{{\mathcal{A}}_{2}}\circ\psi=\tilde{\psi}\circ u_{{\mathcal{A}}_{1}}.

It is not generally the case that ψ~(I𝒜~1)=IA~2\tilde{\psi}(I_{\tilde{\mathcal{A}}_{1}})=I_{\tilde{A}_{2}}. When ψ\psi is a *-homomorphism, so is ψ~\tilde{\psi}.

Observation \the\numberby.

If ψ\psi is contractive and completely positive, then ψ~\tilde{\psi} is also contractive and completely postive.

Proof.

When 𝒜1{\mathcal{A}}_{1} is not unital, this follows from [BrownOzawaC*AlFiDiAp, Proposition 2.2.1] applied to u𝒜2ψu_{{\mathcal{A}}_{2}}\circ\psi, and from the definition of ψ~\tilde{\psi} when 𝒜1{\mathcal{A}}_{1} is unital. ∎

Fundamental to our study are inclusions, which we usually consider as a CC^{*}-algebra {\mathcal{B}} contained in another CC^{*}-algebra 𝒜{\mathcal{A}}. At times, particularly in this section, it will be helpful to explicitly mention the embedding of {\mathcal{B}} into 𝒜{\mathcal{A}}. While this makes notation more cumbersome, the benefit of clarity outweighs the notational burdens.

Definition \the\numberby.
  1. (a)

    An inclusion is a triple (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f), where 𝒜{\mathcal{A}} and {\mathcal{B}} are CC^{*}-algebras and f:𝒜f:{\mathcal{B}}\rightarrow{\mathcal{A}} is a *-monomorphism. We include the possibility that ={0}{\mathcal{B}}=\{0\}. (Starting in Section 3, we will always assume {\mathcal{B}} is abelian, but until then we do not make that restriction.)

  2. (b)

    Let (𝒜1,1,f1)({\mathcal{A}}_{1},{\mathcal{B}}_{1},f_{1}) be an inclusion. An inclusion (𝒜2,2,f2)({\mathcal{A}}_{2},{\mathcal{B}}_{2},f_{2}) together with a *-monomorphism α:𝒜1𝒜2\alpha:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2} such that α(f1(1))f2(2)\alpha(f_{1}({\mathcal{B}}_{1}))\subseteq f_{2}({\mathcal{B}}_{2}) will be called an expansion of (𝒜1,1,f1)({\mathcal{A}}_{1},{\mathcal{B}}_{1},f_{1}) and will be denoted by (𝒜2,2,f2 :: α)({\mathcal{A}}_{2},{\mathcal{B}}_{2},f_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha). Because f2f_{2} is one-to-one, (𝒜2,2,f2 :: α)({\mathcal{A}}_{2},{\mathcal{B}}_{2},f_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an expansion of (𝒜1,1,f1)({\mathcal{A}}_{1},{\mathcal{B}}_{1},f_{1}) if and only if α:𝒜1𝒜2\alpha:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2} is a *-monomorphism and there exists a *-monomorphism α¯:12\underline{\alpha}:{\mathcal{B}}_{1}\rightarrow{\mathcal{B}}_{2} such that

    (2.1.5) αf1=f2α¯.\alpha\circ f_{1}=f_{2}\circ\underline{\alpha}.
  3. (c)

    We will call the two inclusions (𝒜1,1,f1)({\mathcal{A}}_{1},{\mathcal{B}}_{1},f_{1}) and (𝒜2,2,f2)({\mathcal{A}}_{2},{\mathcal{B}}_{2},f_{2}) isomorphic inclusions if there is a *-isomorphism α:𝒜1𝒜2\alpha:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2} such that α(f1(1))=f2(2)\alpha(f_{1}({\mathcal{B}}_{1}))=f_{2}({\mathcal{B}}_{2}).

  4. (d)

    The inclusion (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is called a unital inclusion if 𝒜{\mathcal{A}} and {\mathcal{B}} are unital and f(I)=I𝒜f(I_{\mathcal{B}})=I_{\mathcal{A}}.

  5. (e)

    Given the inclusion (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f), the standard extension of ff to f~:~𝒜~\tilde{f}:\tilde{\mathcal{B}}\rightarrow\tilde{\mathcal{A}} described in (2.1.3) is a *-monomorphism because {\mathcal{B}} is an essential ideal in ~\tilde{\mathcal{B}}. Thus (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) is an inclusion. We will call f~\tilde{f} the the standard embedding of ~\tilde{\mathcal{B}} into 𝒜~\tilde{\mathcal{A}}. (Caution: (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) need not be a unital inclusion.)

    Because

    (2.1.6) f~u=u𝒜f,\tilde{f}\circ u_{\mathcal{B}}=u_{\mathcal{A}}\circ f,

    (𝒜~,~,f~ :: u𝒜)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}\hbox{\,:\hskip-1.0pt:\,}u_{\mathcal{A}}) is an expansion of (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f). We will call (𝒜~,~,f~ :: u𝒜)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}\hbox{\,:\hskip-1.0pt:\,}u_{\mathcal{A}}) the standard expansion of (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f).

  6. (f)

    For 1i<j31\leq i<j\leq 3, let (j,i,fji)({\mathcal{B}}_{j},{\mathcal{B}}_{i},f_{ji}) be inclusions. If f31=f32f21f_{31}=f_{32}\circ f_{21}, we will say the two inclusions (2,1,f21)({\mathcal{B}}_{2},{\mathcal{B}}_{1},f_{21}) and (3,2,f32)({\mathcal{B}}_{3},{\mathcal{B}}_{2},f_{32}) are intermediate inclusions for (3,1,f31)({\mathcal{B}}_{3},{\mathcal{B}}_{1},f_{31}).

    (2.1.7) 3\textstyle{{\mathcal{B}}_{3}}1\textstyle{{\mathcal{B}}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f31\scriptstyle{f_{31}}f21\scriptstyle{f_{21}}2.\textstyle{{\mathcal{B}}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\,\,.}f32\scriptstyle{f_{32}}

Notation \the\numberby. Let (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) be an inclusion.

  • When there is little danger of confusion, we will often identify {\mathcal{B}} with its image f()f({\mathcal{B}}), so that 𝒜{\mathcal{B}}\subseteq{\mathcal{A}}. When this identification is made, we will suppress ff and call (𝒜,)({\mathcal{A}},{\mathcal{B}}) an inclusion.

  • When we write (𝒜~,~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}}), we always mean (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}). Since f~\tilde{f} is not necessarily a unital mapping, (𝒜~,~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}}) need not be a unital inclusion in the sense of Definition 2.1(d).

  • We will sometimes write (𝒜,,)({\mathcal{A}},{\mathcal{B}},\subseteq) for inclusions when {\mathcal{B}} is a CC^{*}-subalgebra of 𝒜{\mathcal{A}}.

  • If (𝒜2,2,f2 :: α)({\mathcal{A}}_{2},{\mathcal{B}}_{2},f_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an expansion of (𝒜1,1,f1)({\mathcal{A}}_{1},{\mathcal{B}}_{1},f_{1}) and i{\mathcal{B}}_{i} has been identified with fi(i)f_{i}({\mathcal{B}}_{i}) as above, we will again suppress writing fif_{i} and say (𝒜2,2 :: α)({\mathcal{A}}_{2},{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an expansion of (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}).

  • If (2,1,f21)({\mathcal{B}}_{2},{\mathcal{B}}_{1},f_{21}) and (3,2,f32)({\mathcal{B}}_{3},{\mathcal{B}}_{2},f_{32}) are inclusions which are intermediate for (3,1,f31)({\mathcal{B}}_{3},{\mathcal{B}}_{1},f_{31}), we will sometimes say that 2{\mathcal{B}}_{2} is intermediate to (3,1)({\mathcal{B}}_{3},{\mathcal{B}}_{1}) and may also indicate this using the notation 123{\mathcal{B}}_{1}\subseteq{\mathcal{B}}_{2}\subseteq{\mathcal{B}}_{3}.

There are: various types of inclusions; objects associated to an inclusion; and properties an inclusion may have. We describe some of them here.

Definitions, Notations, and Comments \the\numberby. Let (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) be an inclusion.

  1. (a)

    We will use rCom(𝒜,,f)\operatorname{\textsc{rCom}}({\mathcal{A}},{\mathcal{B}},f) for the relative commutant of f()f({\mathcal{B}}) in 𝒜{\mathcal{A}}, that is,

    rCom(𝒜,,f):={a𝒜:af(b)=f(b)a for all b}.\operatorname{\textsc{rCom}}({\mathcal{A}},{\mathcal{B}},f):=\{a\in{\mathcal{A}}:af(b)=f(b)a\text{ for all }b\in{\mathcal{B}}\}.

    When {\mathcal{B}} is identified with f()f({\mathcal{B}}), we will frequently use the notation c{\mathcal{B}}^{c} or rCom(𝒜,)\operatorname{\textsc{rCom}}({\mathcal{A}},{\mathcal{B}}) instead of rCom(𝒜,,f)\operatorname{\textsc{rCom}}({\mathcal{A}},{\mathcal{B}},f).

  2. (b)

    We say that (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the has the ideal intersection property if every non-zero ideal in 𝒜{\mathcal{A}} has non-trivial intersection with f()f({\mathcal{B}}). We will also use the term essential inclusion as a synonym for an inclusion with the ideal intersection property. Likewise, when (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is an essential inclusion, we will sometimes call the map ff an essential map.

    (The literature contains several synonyms for the ideal intersection property, for example in [KwasniewskiMeyerApAlExPrUnPsEx], an inclusion with the ideal intersection property is said to detect ideals, and in [PittsZarikianUnPsExC*In], such an inclusion is called CC^{*}-essential.)

  3. (c)

    (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the approximate unit property (abbreviated AUP) if there is an approximate unit (uλ)(u_{\lambda}) for {\mathcal{B}} such that f(uλ)f(u_{\lambda}) is an approximate unit for 𝒜{\mathcal{A}}. In remarks following [ExelOnKuC*DiOpId, Definition 2.1], Exel notes that the Krein-Milman Theorem and [AkemannShultzPeC*Al, Lemma 2.32] yield the following characterization of the approximate unit property.

    Fact \the\numberby (see [ExelOnKuC*DiOpId]).

    An inclusion (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the approximate unit property if and only if no pure state of 𝒜{\mathcal{A}} annihilates f()f({\mathcal{B}}).

  4. (d)

    A normalizer for (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is an element of the set,

    𝒩(𝒜,,f):={v𝒜:vf()vvf()vf()}.{\mathcal{N}}({\mathcal{A}},{\mathcal{B}},f):=\{v\in{\mathcal{A}}:vf({\mathcal{B}})v^{*}\cup v^{*}f({\mathcal{B}})v\subseteq f({\mathcal{B}})\}.

    If {\mathcal{B}} is identified with f()f({\mathcal{B}}), we will write 𝒩(𝒜,){\mathcal{N}}({\mathcal{A}},{\mathcal{B}}) rather than 𝒩(𝒜,,f){\mathcal{N}}({\mathcal{A}},{\mathcal{B}},f).

  5. (e)

    Closely related to normalizers are intertwiners. An element w𝒜w\in{\mathcal{A}} is called an intertwiner if the sets f()wf({\mathcal{B}})w and wf()wf({\mathcal{B}}) coincide. We will write (𝒜,,f){\mathcal{I}}({\mathcal{A}},{\mathcal{B}},f) (or (𝒜,){\mathcal{I}}({\mathcal{A}},{\mathcal{B}}) when ff is suppressed) for the set of all intertwiners.

  6. (f)

    (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is said to be a regular inclusion if span¯𝒩(𝒜,,f)=𝒜\overline{\operatorname{span}}\,{\mathcal{N}}({\mathcal{A}},{\mathcal{B}},f)={\mathcal{A}}.

  7. (g)

    If f()f({\mathcal{B}}) is maximal abelian in 𝒜{\mathcal{A}}, we will call (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) a MASA inclusion.

  8. (h)

    (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is a Cartan inclusion, also called a Cartan pair, if it is a regular MASA inclusion and there exists a faithful conditional expectation Δ:𝒜f()\Delta:{\mathcal{A}}\rightarrow f({\mathcal{B}}). Cartan inclusions necessarily have the AUP, [PittsNoApUnInC*Al, Theorem 2.6].

    The conditional expectation Δ\Delta is unique; see [RenaultCaSuC*Al, Corollary 5.10] or [PittsStReInI, Theorem 3.5]. Also, Δ\Delta is invariant under 𝒩(𝒜,,f){\mathcal{N}}({\mathcal{A}},{\mathcal{B}},f) in the sense that for every v𝒩(𝒜,,f)v\in{\mathcal{N}}({\mathcal{A}},{\mathcal{B}},f) and x𝒜x\in{\mathcal{A}},

    (2.1.10) Δ(vxv)=vΔ(x)v.\Delta(v^{*}xv)=v^{*}\Delta(x)v.

    While we do not have an explicit reference for (2.1.10), it is certainly known; for example, it follows from Renault’s (twisted) groupoid model for a Cartan pair.

    Comment \the\numberby. We will repeatedly use the fact that the regular inclusion (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is a Cartan inclusion if and only if (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) is a Cartan inclusion, see [PittsNoApUnInC*Al, Proposition 3.2]. Example 3.1 of [PittsNoApUnInC*Al] shows the necessity of the regularity hypothesis on (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f).

  9. (i)

    If (𝒜1,1,f1)({\mathcal{A}}_{1},{\mathcal{B}}_{1},f_{1}) is another inclusion, a *-homomorphism α:𝒜𝒜1\alpha:{\mathcal{A}}\rightarrow{\mathcal{A}}_{1} is a regular homomorphism if

    α(𝒩(𝒜,,f))𝒩(𝒜1,1,f1).\alpha({\mathcal{N}}({\mathcal{A}},{\mathcal{B}},f))\subseteq{\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1},f_{1}).

    We will sometimes indicate this by saying α:(𝒜,,f)(𝒜1,1,f1)\alpha:({\mathcal{A}},{\mathcal{B}},f)\rightarrow({\mathcal{A}}_{1},{\mathcal{B}}_{1},f_{1}) is a regular homomorphism.

When 2{\mathcal{B}}_{2} is intermediate to (3,1)({\mathcal{B}}_{3},{\mathcal{B}}_{1}), it may happen that ~2\tilde{\mathcal{B}}_{2} is not intermediate to (~3,~1)(\tilde{\mathcal{B}}_{3},\tilde{\mathcal{B}}_{1}). Here is an example of this behavior (the example also provides an example of an inclusion whose “unitization” is not unital).

Example \the\numberby. Let S0S_{0} and S1S_{1} be isometries on a Hilbert space such that S0S0+S1S1=IS_{0}S_{0}^{*}+S_{1}S_{1}^{*}=I, let 𝒮{\mathcal{S}} be the inverse semigroup of partial isometries consisting of all finite products of elements of the set {S0,S1,S0,S1}\{S_{0},S_{1},S_{0}^{*},S_{1}^{*}\}, and let 𝒟{\mathcal{D}} be the CC^{*}-algebra generated by the projections in 𝒮{\mathcal{S}}. There is a unique multiplicative linear functional τ:𝒟\tau:{\mathcal{D}}\rightarrow{\mathbb{C}} on 𝒟{\mathcal{D}} such that for every nn\in{\mathbb{N}}, τ(S1nS1n)=1\tau(S_{1}^{n}S_{1}^{n}{}^{*})=1. (Indeed, 𝒟{\mathcal{D}} is isomorphic to the continuous functions on the Cantor middle thirds set, and under this isomorphism, τ\tau corresponds to evaluation at 1.) Let

3=1=kerτ,2=𝒟,{\mathcal{B}}_{3}={\mathcal{B}}_{1}=\ker\tau,\quad{\mathcal{B}}_{2}={\mathcal{D}},

and let σ:𝒟kerτ\sigma:{\mathcal{D}}\rightarrow\ker\tau be the map

σ(d)=S0dS0.\sigma(d)=S_{0}dS_{0}^{*}.

Put

f31:=σ2|1,f21:=σ|1andf32:=σ.f_{31}:=\sigma^{2}|_{{\mathcal{B}}_{1}},\quad f_{21}:=\sigma|_{{\mathcal{B}}_{1}}\quad\text{and}\quad f_{32}:=\sigma.

Since f31=f32f21f_{31}=f_{32}\circ f_{21}, (2,1,f21)({\mathcal{B}}_{2},{\mathcal{B}}_{1},f_{21}) and (3,2,f32)({\mathcal{B}}_{3},{\mathcal{B}}_{2},f_{32}) are intermediate inclusions for (3,1,f31)({\mathcal{B}}_{3},{\mathcal{B}}_{1},f_{31}). Let qq denote the quotient map of ~3\tilde{\mathcal{B}}_{3} onto ~3/3\tilde{\mathcal{B}}_{3}/{\mathcal{B}}_{3}\simeq{\mathbb{C}}. Since 1{\mathcal{B}}_{1} and 3{\mathcal{B}}_{3} are not unital, qf~310q\circ\tilde{f}_{31}\neq 0. On the other hand, since 2{\mathcal{B}}_{2} is unital, qf~32f~21=0q\circ\tilde{f}_{32}\circ\tilde{f}_{21}=0. Therefore, f~31f~32f~21\tilde{f}_{31}\neq\tilde{f}_{32}\circ\tilde{f}_{21}, so (~2,~1,f~21)(\tilde{\mathcal{B}}_{2},\tilde{\mathcal{B}}_{1},\tilde{f}_{21}) and (~3,~2,f~32)(\tilde{\mathcal{B}}_{3},\tilde{\mathcal{B}}_{2},\tilde{f}_{32}) are not intermediate inclusions for (~3,~1,f~31)(\tilde{\mathcal{B}}_{3},\tilde{\mathcal{B}}_{1},\tilde{f}_{31}). Finally, note that (2,1,f21)({\mathcal{B}}_{2},{\mathcal{B}}_{1},f_{21}) is an example of an inclusion where (~2,~1,f~21)(\tilde{\mathcal{B}}_{2},\tilde{\mathcal{B}}_{1},\tilde{f}_{21}) is not a unital inclusion.

2.2. Weakly Non-Degenerate Inclusions

The approximate unit property for an inclusion (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has been used in the literature as a definition of non-degeneracy, see for example [CrytserNagySiCrEtGrC*Al, Definition 1.4] or [ExelPittsChGrC*AlNoHaEtGr, Definition 2.3.1(iii)]. However, when J𝒜J\unlhd{\mathcal{A}} is a proper and essential ideal, the inclusion (𝒜,J,)({\mathcal{A}},J,\subseteq) cannot have the AUP, so using the AUP as a non-degeneracy condition excludes such examples. For our purposes, the following weaker version of non-degeneracy is more appropriate.

Definition \the\numberby.

Let (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) be an inclusion.

  1. (a)

    Define the annihilator of {\mathcal{B}} in 𝒜{\mathcal{A}} to be the set,

    Ann(𝒜,,f):={a𝒜:af(b)=0=f(b)a for all b}.\operatorname{Ann}({\mathcal{A}},{\mathcal{B}},f):=\{a\in{\mathcal{A}}:af(b)=0=f(b)a\text{ for all }b\in{\mathcal{B}}\}.

    Notice that Ann(𝒜,,f)\operatorname{Ann}({\mathcal{A}},{\mathcal{B}},f) is an ideal of rCom(𝒜,,f)\operatorname{\textsc{rCom}}({\mathcal{A}},{\mathcal{B}},f).

  2. (b)

    An inclusion (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) such that Ann(𝒜,,f)={0}\operatorname{Ann}({\mathcal{A}},{\mathcal{B}},f)=\{0\} will be called a weakly non-degenerate inclusion.

As usual, when {\mathcal{B}} is identified with f()𝒜f({\mathcal{B}})\subseteq{\mathcal{A}}, we will write Ann(𝒜,)\operatorname{Ann}({\mathcal{A}},{\mathcal{B}}) instead of Ann(𝒜,,f)\operatorname{Ann}({\mathcal{A}},{\mathcal{B}},f) and will say (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate when (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is weakly non-degenerate.

Remark \the\numberby. We chose the term “weakly non-degenerate” instead of “non-degenerate” to avoid confusion with existing literature. Because of the similarity between the terms “weakly non-degenerate” and “non-degenerate” we will always use the term “approximate unit property” instead of “non-degenerate.” Finally, to avoid possible confusion, for a representation π\pi of a CC^{*}-algebra 𝒜{\mathcal{A}} on a Hilbert space {\mathcal{H}}, we will sometimes explicitly write π(𝒜)¯=\overline{\pi({\mathcal{A}}){\mathcal{H}}}={\mathcal{H}} instead of using the term “non-degenerate representation.”

Here are some conditions ensuring an inclusion is weakly non-degenerate.

Lemma \the\numberby.

Let (𝒜,)({\mathcal{A}},{\mathcal{B}}) be an inclusion.

  1. (a)

    If (𝒜,)({\mathcal{A}},{\mathcal{B}}) has the AUP, then (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate.

  2. (b)

    If 𝒜{\mathcal{B}}\unlhd{\mathcal{A}} is an essential ideal, then (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate. In particular, when 𝒜{\mathcal{A}} is not unital, (𝒜~,𝒜,u𝒜)(\tilde{\mathcal{A}},{\mathcal{A}},u_{\mathcal{A}}) is weakly non-degenerate.

  3. (c)

    If {\mathcal{B}} is contained in the center of 𝒜{\mathcal{A}} and (𝒜,)({\mathcal{A}},{\mathcal{B}}) has the ideal intersection property, then (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate.

  4. (d)

    Suppose E:𝒜E:{\mathcal{A}}\rightarrow{\mathcal{B}} is a faithful conditional expectation. Then (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate.

  5. (e)

    Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is an inclusion such that 𝒟{\mathcal{D}} is abelian and (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) has the ideal intersection property. Then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is weakly non-degenerate. In particular, any MASA inclusion is weakly non-degenerate.

Proof.

(a) Let (uλ)(u_{\lambda})\subseteq{\mathcal{B}} be an approximate unit for 𝒜{\mathcal{A}}. If xAnn(𝒜,)x\in\operatorname{Ann}({\mathcal{A}},{\mathcal{B}}), then x=limxuλ=0x=\lim xu_{\lambda}=0.

(b) Since {\mathcal{B}} is an ideal of 𝒜{\mathcal{A}}, Ann(𝒜,)\operatorname{Ann}({\mathcal{A}},{\mathcal{B}}) is also an ideal in 𝒜{\mathcal{A}}. Clearly Ann(𝒜,)={0}\operatorname{Ann}({\mathcal{A}},{\mathcal{B}})\cap{\mathcal{B}}=\{0\}, and as {\mathcal{B}} is an essential ideal, Ann(𝒜,)={0}\operatorname{Ann}({\mathcal{A}},{\mathcal{B}})=\{0\}.

(c) By hypothesis, c=𝒜{\mathcal{B}}^{c}={\mathcal{A}}, so Ann(𝒜,)\operatorname{Ann}({\mathcal{A}},{\mathcal{B}}) is an ideal of 𝒜{\mathcal{A}}. As Ann(𝒜,)\operatorname{Ann}({\mathcal{A}},{\mathcal{B}}) has trivial intersection with {\mathcal{B}}, the ideal intersection property gives Ann(𝒜,)={0}\operatorname{Ann}({\mathcal{A}},{\mathcal{B}})=\{0\}. Therefore, (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate.

(d) Let xAnn(𝒜,)x\in\operatorname{Ann}({\mathcal{A}},{\mathcal{B}}). Since xxAnn(𝒜,)x^{*}x\in\operatorname{Ann}({\mathcal{A}},{\mathcal{B}}) and E(xx)E(x^{*}x)\in{\mathcal{B}},

E(xx)2=E(xxE(xx))=0,E(x^{*}x)^{2}=E(x^{*}xE(x^{*}x))=0,

with the first equality following from Tomiyama’s theorem, see [BrownOzawaC*AlFiDiAp, Theorem 1.5.10]. This gives x=0x=0 by faithfulness of EE. So (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate.

(e) Applying part (c) to (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) gives Ann(𝒟c,𝒟)={0}\operatorname{Ann}({\mathcal{D}}^{c},{\mathcal{D}})=\{0\}. But Ann(𝒞,𝒟)\operatorname{Ann}({\mathcal{C}},{\mathcal{D}}) is an ideal in 𝒟c{\mathcal{D}}^{c} having trivial intersection with 𝒟{\mathcal{D}}. So Ann(𝒞,𝒟)={0}\operatorname{Ann}({\mathcal{C}},{\mathcal{D}})=\{0\}, showing (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is weakly non-degenerate. ∎

In general, for an inclusion (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f), there are multiple embeddings of ~\tilde{\mathcal{B}} into 𝒜~\tilde{\mathcal{A}} which extend u𝒜fu_{\mathcal{A}}\circ f, and it may happen that the image of I~I_{\tilde{\mathcal{B}}} under the standard embedding is not I𝒜~I_{\tilde{\mathcal{A}}}. Fortunately, these behaviors cannot occur when (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is weakly non-degenerate, as we shall see in parts (e) and (f) of Lemma 2.2 below. Before proving Lemma 2.2, we require some preparation. The first is the following well-known fact (see [HamanaReEmCStAlMoCoCStAl, Lemma 4.8] or [EffrosAsNoOr, (2.2)]).

Lemma \the\numberby.

Suppose =12{\mathcal{H}}={\mathcal{H}}_{1}\oplus{\mathcal{H}}_{2} is the direct sum of the Hilbert spaces 1{\mathcal{H}}_{1} and 2{\mathcal{H}}_{2}. Let A(1)A\in{\mathcal{B}}({\mathcal{H}}_{1}) be positive and invertible, B(2,1)B\in{\mathcal{B}}({\mathcal{H}}_{2},{\mathcal{H}}_{1}), and 0C(2)0\leq C\in{\mathcal{B}}({\mathcal{H}}_{2}). Then

T:=[ABBC](12)T:=\begin{bmatrix}A&B\\ B^{*}&C\end{bmatrix}\in{\mathcal{B}}({\mathcal{H}}_{1}\oplus{\mathcal{H}}_{2})

is a positive operator if and only if BA1BCB^{*}A^{-1}B\leq C. Moreover, if T0T\geq 0 and C=0C=0, then B=0B=0.

Proof.

Both statements follow from the factorization,

[ABBC]=[I1A1B0I2][A00C(A1/2B)(A1/2B)][I1A1B0I2].\begin{bmatrix}A&B\\ B^{*}&C\end{bmatrix}=\begin{bmatrix}I_{{\mathcal{H}}_{1}}&A^{-1}B\\ 0&I_{{\mathcal{H}}_{2}}\end{bmatrix}^{*}\begin{bmatrix}A&0\\ 0&C-(A^{-1/2}B)^{*}(A^{-1/2}B)\end{bmatrix}\begin{bmatrix}I_{{\mathcal{H}}_{1}}&A^{-1}B\\ 0&I_{{\mathcal{H}}_{2}}\end{bmatrix}.

Among inclusions, weakly non-degenerate inclusions are rather well-behaved. Here are a number of desirable properties which weakly non-degenerate inclusions possess.

Lemma \the\numberby.

Suppose (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is a weakly non-degenerate inclusion. The following statements hold.

  1. (a)

    If {\mathcal{B}} is unital, then 𝒜{\mathcal{A}} is unital and f(I)f(I_{\mathcal{B}}) is the unit for 𝒜{\mathcal{A}}.

  2. (b)

    Suppose a𝒜a\in{\mathcal{A}} satisfies f(b)a=af(b)=f(b)f(b)a=af(b)=f(b) for every bb\in{\mathcal{B}}. Then 𝒜{\mathcal{A}} is unital and a=I𝒜a=I_{\mathcal{A}}.

  3. (c)

    Let (uλ)(u_{\lambda}) be an approximate unit for {\mathcal{B}}. Suppose a𝒜a\in{\mathcal{A}} satisfies: 0a0\leq a, a1\left\lVert a\right\rVert\leq 1, and f(uλ)af(u_{\lambda})\leq a for every λ\lambda. Then 𝒜{\mathcal{A}} is unital and a=I𝒜a=I_{\mathcal{A}}.

  4. (d)

    (𝒜~,,u𝒜f)(\tilde{\mathcal{A}},{\mathcal{B}},u_{\mathcal{A}}\circ f) is weakly non-degenerate.

  5. (e)

    (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) is a unital inclusion; in particular it is weakly non-degenerate.

  6. (f)

    If σ:~𝒜~\sigma:\tilde{\mathcal{B}}\rightarrow\tilde{\mathcal{A}} is a *-monomorphism such that σu=u𝒜f\sigma\circ u_{\mathcal{B}}=u_{\mathcal{A}}\circ f, then σ=f~\sigma=\tilde{f}.

Proof.

(a) We may suppose 𝒜(){\mathcal{A}}\subseteq{\mathcal{B}}({\mathcal{H}}). Let I=II=I_{\mathcal{H}}, p:=f(I)p:=f(I_{\mathcal{B}}), and choose a𝒜a\in{\mathcal{A}} with 0a0\leq a. Relative to the decomposition, =p+(Ip){\mathcal{H}}=p{\mathcal{H}}+(I-p){\mathcal{H}}, aa and a+pa+p have the form,

a=[pappa(Ip)(Ip)ap(Ip)a(Ip)]anda+p=[pap+ppa(Ip)(Ip)ap(Ip)a(Ip)].a=\begin{bmatrix}pap&pa(I-p)\\ (I-p)ap&(I-p)a(I-p)\end{bmatrix}\quad\text{and}\quad a+p=\begin{bmatrix}pap+p&pa(I-p)\\ (I-p)ap&(I-p)a(I-p)\end{bmatrix}.

Let x:=(Ip)a(Ip)=aappa+papx:=(I-p)a(I-p)=a-ap-pa+pap. Then xAnn(𝒜,,f)x\in\operatorname{Ann}({\mathcal{A}},{\mathcal{B}},f). As (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is weakly non-degenerate, (Ip)a(Ip)=0(I-p)a(I-p)=0. Since 0aa+p0\leq a\leq a+p, applying Lemma 2.2 to a+pa+p gives (Ip)ap=pa(Ip)=0(I-p)ap=pa(I-p)=0. Therefore,

ap=pa=pap=a.ap=pa=pap=a.

As 𝒜{\mathcal{A}} is the linear span of its positive elements, we conclude that p=I𝒜p=I_{\mathcal{A}}.

(b) Let p=aap=a^{*}a. Then for every bb\in{\mathcal{B}}, pf(b)=f(b)p=f(b)pf(b)=f(b)p=f(b) because {\mathcal{B}} is a self-adjoint subspace of 𝒜{\mathcal{A}}. Therefore, p2pAnn(𝒜,,f)={0}p^{2}-p\in\operatorname{Ann}({\mathcal{A}},{\mathcal{B}},f)=\{0\}, so pp is a projection in 𝒜{\mathcal{A}}. Consider 𝒞:=f()+p{\mathcal{C}}:=f({\mathcal{B}})+{\mathbb{C}}p. Then 𝒞{\mathcal{C}} is a unital subalgebra of 𝒜{\mathcal{A}}, and since Ann(𝒜,𝒞,)Ann(𝒜,,f)\operatorname{Ann}({\mathcal{A}},{\mathcal{C}},\subseteq)\subseteq\operatorname{Ann}({\mathcal{A}},{\mathcal{B}},f), (𝒜,𝒞,)({\mathcal{A}},{\mathcal{C}},\subseteq) is weakly non-degenerate. By part (a), p=I𝒜p=I_{\mathcal{A}}. Noting that paAnn(𝒜,,f)p-a\in\operatorname{Ann}({\mathcal{A}},{\mathcal{B}},f), we obtain I𝒜=p=aI_{\mathcal{A}}=p=a, as desired.

(c) Once again, we assume 𝒜(){\mathcal{A}}\subseteq{\mathcal{B}}({\mathcal{H}}). Since (f(uλ))(f(u_{\lambda})) is a bounded increasing net of positive semi-definite operators, it converges in the strong operator topology. Let p:=sotlimf(uλ)p:=\textsc{sot}\lim f(u_{\lambda}). Then pp is a projection and for every bb\in{\mathcal{B}},

(2.2.2) f(b)p=pf(b)=f(b).f(b)p=pf(b)=f(b).

Since f(uλ)af(u_{\lambda})\leq a for each λ\lambda, we obtain pap\leq a. Decomposing aa with respect to =(Ip)p{\mathcal{H}}=(I-p){\mathcal{H}}\oplus p{\mathcal{H}} we find a=[pappappappap]a=\begin{bmatrix}p^{\perp}ap^{\perp}&p^{\perp}ap\\ pap^{\perp}&pap\end{bmatrix}. As apa\geq p, we have papp0pap\geq p\geq 0. Since pap1\left\lVert pap\right\rVert\leq 1, we obtain p=papp=pap. Then

0ap+p=[pap+ppappappapp]=[pap+ppappap0].0\leq a-p+p^{\perp}=\begin{bmatrix}p^{\perp}ap^{\perp}+p^{\perp}&p^{\perp}ap\\ pap^{\perp}&pap-p\end{bmatrix}=\begin{bmatrix}p^{\perp}ap^{\perp}+p^{\perp}&p^{\perp}ap\\ pap^{\perp}&0\end{bmatrix}.

Lemma 2.2 gives pap=0p^{\perp}ap=0. Therefore a=[pap00p]a=\begin{bmatrix}p^{\perp}ap^{\perp}&0\\ 0&p\end{bmatrix}. It follows that for bb\in{\mathcal{B}},

af(b)=f(b)a=f(b).af(b)=f(b)a=f(b).

An application of part (b) completes the proof.

(d) There is nothing to do when 𝒜{\mathcal{A}} is unital, so assume 𝒜{\mathcal{A}} has no unit. Let (a,λ)Ann(𝒜~,,u𝒜f)(a,\lambda)\in\operatorname{Ann}(\tilde{\mathcal{A}},{\mathcal{B}},u_{\mathcal{A}}\circ f). It suffices to show λ=0\lambda=0, for once this is established, it follows that aAnn(𝒜,,f)a\in\operatorname{Ann}({\mathcal{A}},{\mathcal{B}},f), whence a=0a=0. To show λ=0\lambda=0, we argue by contradiction.

Suppose λ0\lambda\neq 0. Then by scaling, we may assume λ=1\lambda=-1. Note that for bb\in{\mathcal{B}}, u𝒜(f(b))=(f(b),0)u_{\mathcal{A}}(f(b))=(f(b),0). Thus

(0,0)=(a,1)(f(b),0)=(af(b)f(b),0)=(f(b)af(b),0)=(f(b),0)(a,1).(0,0)=(a,-1)(f(b),0)=(af(b)-f(b),0)=(f(b)a-f(b),0)=(f(b),0)(a,-1).

So for all bb\in{\mathcal{B}}, af(b)=f(b)=f(b)aaf(b)=f(b)=f(b)a. Part (b) now gives a=I𝒜a=I_{\mathcal{A}}, contrary to assumption on 𝒜{\mathcal{A}}. Thus, λ=0\lambda=0, completing the proof.

(e) Let a=f~(I~)a=\tilde{f}(I_{\tilde{\mathcal{B}}}). Then a1\left\lVert a\right\rVert\leq 1 and a0a\geq 0. For every bb\in{\mathcal{B}}, au𝒜(f(b))=u𝒜(f(b))a=u𝒜(f(b))au_{\mathcal{A}}(f(b))=u_{\mathcal{A}}(f(b))a=u_{\mathcal{A}}(f(b)). Applying part (d), then part (b), yields a=I𝒜~a=I_{\tilde{\mathcal{A}}}. Therefore (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) is a unital inclusion and hence is weakly non-degenerate.

(f) If {\mathcal{B}} is unital, so is 𝒜{\mathcal{A}} by part (a). Therefore, σ=σu=u𝒜f=f=f~\sigma=\sigma\circ u_{\mathcal{B}}=u_{\mathcal{A}}\circ f=f=\tilde{f}, and all is well.

Now assume {\mathcal{B}} is not unital. We claim that f~(I~)=σ(I~)\tilde{f}(I_{\tilde{\mathcal{B}}})=\sigma(I_{\tilde{\mathcal{B}}}). For bb\in{\mathcal{B}},

σ(I~)u𝒜(f(b))=σ(I~)σ(u(b))=σ(u(b))=u𝒜(f(b));\sigma(I_{\tilde{\mathcal{B}}})u_{\mathcal{A}}(f(b))=\sigma(I_{\tilde{\mathcal{B}}})\sigma(u_{\mathcal{B}}(b))=\sigma(u_{\mathcal{B}}(b))=u_{\mathcal{A}}(f(b));

likewise

u𝒜(f(b))σ(I~)=u𝒜(f(b)).u_{\mathcal{A}}(f(b))\sigma(I_{\tilde{\mathcal{B}}})=u_{\mathcal{A}}(f(b)).

Similarly, f~(I~)u𝒜(f(b))=u𝒜(f(b))f~(I~)=u𝒜(f(b))\tilde{f}(I_{\tilde{\mathcal{B}}})u_{\mathcal{A}}(f(b))=u_{\mathcal{A}}(f(b))\tilde{f}(I_{\tilde{\mathcal{B}}})=u_{\mathcal{A}}(f(b)). Therefore,

f~(I~)σ(I~)Ann(𝒜~,,u𝒜f).\tilde{f}(I_{\tilde{\mathcal{B}}})-\sigma(I_{\tilde{\mathcal{B}}})\in\operatorname{Ann}(\tilde{\mathcal{A}},{\mathcal{B}},u_{\mathcal{A}}\circ f).

Part (d) gives f~(I~)=σ(I~)\tilde{f}(I_{\tilde{\mathcal{B}}})=\sigma(I_{\tilde{\mathcal{B}}}), so the claim holds.

For (b,λ)~(b,\lambda)\in\tilde{\mathcal{B}},

σ(b,λ)=σ(u(b))+λf~(I~)=u𝒜(f(b))+λf~(I~)=f~(b,λ),\sigma(b,\lambda)=\sigma(u_{\mathcal{B}}(b))+\lambda\tilde{f}(I_{\tilde{\mathcal{B}}})=u_{\mathcal{A}}(f(b))+\lambda\tilde{f}(I_{\tilde{\mathcal{B}}})=\tilde{f}(b,\lambda),

so σ=f~\sigma=\tilde{f}. ∎

We now show that the poor behavior for intermediate inclusions exhibited in Example 2.1 cannot occur when (3,1)({\mathcal{B}}_{3},{\mathcal{B}}_{1}) is weakly non-degenerate: if 2{\mathcal{B}}_{2} is intermediate to the weakly non-degenerate inclusion (3,1)({\mathcal{B}}_{3},{\mathcal{B}}_{1}), then ~2\tilde{\mathcal{B}}_{2} is intermediate to (~3,~1)(\tilde{\mathcal{B}}_{3},\tilde{\mathcal{B}}_{1}).

Corollary \the\numberby.

For 1i<j31\leq i<j\leq 3, let (j,i,fji)({\mathcal{B}}_{j},{\mathcal{B}}_{i},f_{ji}) be inclusions such that f31=f32f21f_{31}=f_{32}\circ f_{21}. If (3,1,f31)({\mathcal{B}}_{3},{\mathcal{B}}_{1},f_{31}) is weakly non-degenerate, then f~31=f~32f~21\tilde{f}_{31}=\tilde{f}_{32}\circ\tilde{f}_{21}.

Proof.

Since (3,1,f31)({\mathcal{B}}_{3},{\mathcal{B}}_{1},f_{31}) is weakly non-degenerate, both (3,2,f32)({\mathcal{B}}_{3},{\mathcal{B}}_{2},f_{32}) and (2,1,f21)({\mathcal{B}}_{2},{\mathcal{B}}_{1},f_{21}) are weakly non-degenerate. Thus by Lemma 2.2(e),

f~32(f~21(I~1))=f~32(I~2)=I~3=f~31(I~1).\tilde{f}_{32}(\tilde{f}_{21}(I_{\tilde{\mathcal{B}}_{1}}))=\tilde{f}_{32}(I_{\tilde{\mathcal{B}}_{2}})=I_{\tilde{\mathcal{B}}_{3}}=\tilde{f}_{31}(I_{\tilde{\mathcal{B}}_{1}}).

Given x~1x\in\tilde{\mathcal{B}}_{1}, there exist b11b_{1}\in{\mathcal{B}}_{1} and λ\lambda\in{\mathbb{C}} so that x=u1(b1)+λI~1x=u_{{\mathcal{B}}_{1}}(b_{1})+\lambda I_{\tilde{\mathcal{B}}_{1}}. Using using (2.1.6),

f~32(f~21(x))\displaystyle\tilde{f}_{32}(\tilde{f}_{21}(x)) =f~32(f~21(u1(b1)))+λI~3=f~32(u2(f21(b1)))+λI~3\displaystyle=\tilde{f}_{32}(\tilde{f}_{21}(u_{{\mathcal{B}}_{1}}(b_{1})))+\lambda I_{\tilde{\mathcal{B}}_{3}}=\tilde{f}_{32}(u_{{\mathcal{B}}_{2}}(f_{21}(b_{1})))+\lambda I_{\tilde{\mathcal{B}}_{3}}
=u3(f32((f21(b1))))+λI~3=u3(f31(b1))+λI~3\displaystyle=u_{{\mathcal{B}}_{3}}(f_{32}((f_{21}(b_{1}))))+\lambda I_{\tilde{\mathcal{B}}_{3}}=u_{{\mathcal{B}}_{3}}(f_{31}(b_{1}))+\lambda I_{\tilde{\mathcal{B}}_{3}}
=f~31(x),\displaystyle=\tilde{f}_{31}(x),

as desired. ∎

The following extends [PaulsenCoBoMaOpAl, Corollary 3.19] from unital inclusions to weakly non-degenerate inclusions.

Proposition \the\numberby.

Let (𝒜1,,f1)({\mathcal{A}}_{1},{\mathcal{B}},f_{1}) and (𝒜2,,f2)({\mathcal{A}}_{2},{\mathcal{B}},f_{2}) be weakly non-degenerate inclusions and suppose

Φ:𝒜1𝒜2\Phi:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2}

is a contractive completely positive map such that Φf1=f2\Phi\circ f_{1}=f_{2}. The following statements hold.

  1. (a)

    The standard extension, Φ~:𝒜~1𝒜~2\tilde{\Phi}:\tilde{\mathcal{A}}_{1}\rightarrow\tilde{\mathcal{A}}_{2}, of Φ\Phi is a unital completely positive map.

  2. (b)

    For every x𝒜1x\in{\mathcal{A}}_{1} and h,kh,k\in{\mathcal{B}},

    Φ(f1(h)xf1(k))=f2(h)Φ(x)f2(k).\Phi(f_{1}(h)xf_{1}(k))=f_{2}(h)\Phi(x)f_{2}(k).
Proof.

For i=1,2i=1,2, parts (d) and (e) of Lemma 2.2 show that (𝒜~i,,u𝒜ifi)(\tilde{\mathcal{A}}_{i},{\mathcal{B}},u_{{\mathcal{A}}_{i}}\circ f_{i}) are weakly non-degenerate inclusions and (𝒜~i,~,f~i)(\tilde{\mathcal{A}}_{i},\tilde{\mathcal{B}},\tilde{f}_{i}) are unital inclusions.

(a) Observation 2.1 shows Φ~\tilde{\Phi} is completely positive. To see Φ~\tilde{\Phi} is a unital map, let (hλ)(h_{\lambda}) be an approximate unit for {\mathcal{B}}. We have u𝒜1(f1(hλ))=f~1(u(hλ))I𝒜~1u_{{\mathcal{A}}_{1}}(f_{1}(h_{\lambda}))=\tilde{f}_{1}(u_{\mathcal{B}}(h_{\lambda}))\leq I_{\tilde{\mathcal{A}}_{1}}, so

u𝒜2(f2(hλ))=u𝒜2(Φ(f1(h(uλ))))=Φ~(u𝒜1(f1(hλ)))Φ~(I𝒜~1).u_{{\mathcal{A}}_{2}}(f_{2}(h_{\lambda}))=u_{{\mathcal{A}}_{2}}(\Phi(f_{1}(h(u_{\lambda}))))=\tilde{\Phi}(u_{{\mathcal{A}}_{1}}(f_{1}(h_{\lambda})))\leq\tilde{\Phi}(I_{\tilde{\mathcal{A}}_{1}}).

Applying Proposition 2.2(c) to (𝒜~2,,u𝒜2f2)(\tilde{\mathcal{A}}_{2},{\mathcal{B}},u_{{\mathcal{A}}_{2}}\circ f_{2}), we obtain Φ~(I𝒜~1)=I𝒜~2\tilde{\Phi}(I_{\tilde{\mathcal{A}}_{1}})=I_{\tilde{\mathcal{A}}_{2}}, so Φ~\tilde{\Phi} is a unital map.

(b) By [PaulsenCoBoMaOpAl, Corollary 3.19], Φ~\tilde{\Phi} is a ~\tilde{\mathcal{B}}-bimodule map, that is, for h~,k~~\tilde{h},\tilde{k}\in\tilde{\mathcal{B}} and x~𝒜~1\tilde{x}\in\tilde{\mathcal{A}}_{1},

Φ~(f~1(h~)x~f~1(k~))=f~2(h~)Φ~(x~)f~2(k~).\tilde{\Phi}(\tilde{f}_{1}(\tilde{h})\tilde{x}\tilde{f}_{1}(\tilde{k}))=\tilde{f}_{2}(\tilde{h})\tilde{\Phi}(\tilde{x})\tilde{f}_{2}(\tilde{k}).

Thus for h,kh,k\in{\mathcal{B}} and x𝒜1x\in{\mathcal{A}}_{1}, (and using (2.1.4))

(u𝒜2Φ)(f1(h)xf1(k)))\displaystyle(u_{{\mathcal{A}}_{2}}\circ\Phi)(f_{1}(h)xf_{1}(k))) =(Φ~u𝒜1)(f1(h)xf1(k))\displaystyle=(\tilde{\Phi}\circ u_{{\mathcal{A}}_{1}})(f_{1}(h)xf_{1}(k))
=Φ~(f~1(u(h))u𝒜1(x)f~1(u(k))))\displaystyle=\tilde{\Phi}(\tilde{f}_{1}(u_{\mathcal{B}}(h))u_{{\mathcal{A}}_{1}}(x)\tilde{f}_{1}(u_{\mathcal{B}}(k))))
=f~2(u(h))Φ~(u𝒜1(x))f~2(u(k))\displaystyle=\tilde{f}_{2}(u_{\mathcal{B}}(h))\tilde{\Phi}(u_{{\mathcal{A}}_{1}}(x))\tilde{f}_{2}(u_{{\mathcal{B}}}(k))
=(u𝒜2f2)(h)Φ~(u𝒜1(x))(u𝒜2f2)(k)\displaystyle=(u_{{\mathcal{A}}_{2}}\circ f_{2})(h)\tilde{\Phi}(u_{{\mathcal{A}}_{1}}(x))(u_{{\mathcal{A}}_{2}}\circ f_{2})(k)
=(u𝒜2f2)(h)(u𝒜2Φ)(x)(u𝒜2f2)(k).\displaystyle=(u_{{\mathcal{A}}_{2}}\circ f_{2})(h)(u_{{\mathcal{A}}_{2}}\circ\Phi)(x)(u_{{\mathcal{A}}_{2}}\circ f_{2})(k).

As u𝒜2u_{{\mathcal{A}}_{2}} is a *-monomorphism, the proposition follows. ∎

2.3. Pseudo-Expectations

We shall require pseudo-expectations, and we give a brief discussion of them here. Hamana [HamanaInEnC*Al] showed that given a unital CC^{*}-algebra 𝒜{\mathcal{A}} there is a CC^{*}-algebra I(𝒜)I({\mathcal{A}}) and a one-to-one unital *-homomorphism ι:𝒜I(A)\iota:{\mathcal{A}}\rightarrow I(A) such that:

  • I(𝒜)I({\mathcal{A}}) is an injective object in the category of operator systems and unital completely positive maps; and

  • the only unital completely positive map ϕ:I(𝒜)I(𝒜)\phi:I({\mathcal{A}})\rightarrow I({\mathcal{A}}) satisfying ϕι=ι\phi\circ\iota=\iota is the identity mapping on I(𝒜)I({\mathcal{A}}).

Hamana calls the pair (I(𝒜),ι)(I({\mathcal{A}}),\iota) an injective envelope of 𝒜{\mathcal{A}}. The injective envelope of 𝒜{\mathcal{A}} is monotone closed and has the following uniqueness property: if (1,ι1)({\mathcal{I}}_{1},\iota_{1}) and (2,ι2)({\mathcal{I}}_{2},\iota_{2}) are injective envelopes for 𝒜{\mathcal{A}}, there exists a unique *-isomorphism α:12\alpha:{\mathcal{I}}_{1}\rightarrow{\mathcal{I}}_{2} such that αι1=ι2\alpha\circ\iota_{1}=\iota_{2}.

Remark \the\numberby. Let {\mathfrak{C}} be the category of unital, abelian CC^{*}-algebras and unital *-homomorphisms. Hadwin and Paulsen note that an object 𝒜{\mathcal{A}} in {\mathfrak{C}} is injective if and only if 𝒜{\mathcal{A}} is injective in the category of operator systems and completely positive unital maps, see [HadwinPaulsenInPrAnTo, Theorem 2.4].

When the CC^{*}-algebra 𝒜{\mathcal{A}} is non-unital, an injective envelope for 𝒜{\mathcal{A}} is defined to be an injective envelope (I(𝒜~),ι)(I(\tilde{\mathcal{A}}),\iota) for 𝒜~\tilde{\mathcal{A}}.

Remark \the\numberby.

To simplify notation, we will often say (I(𝒜),ι)(I({\mathcal{A}}),\iota) is an injective envelope of 𝒜{\mathcal{A}} regardless of whether 𝒜{\mathcal{A}} has a unit: when 𝒜{\mathcal{A}} is not unital, it is to be understood that (I(𝒜),ι)(I({\mathcal{A}}),\iota) means (I(A~),ι)(I(\tilde{A}),\iota), where (I(𝒜~),ι)(I(\tilde{\mathcal{A}}),\iota) is an injective envelope for A~\tilde{A}. When there is a need for clarity, we will sometimes write (I(𝒜~),ι)(I(\tilde{\mathcal{A}}),\iota) for an injective envelope of 𝒜{\mathcal{A}}.

We need the following fact about injective envelopes; it is due to Hamana. Let (I(𝒜),ι)(I({\mathcal{A}}),\iota) be an injective envelope for the CC^{*}-algebra 𝒜{\mathcal{A}}, let J𝒜J\unlhd{\mathcal{A}} be a closed ideal of 𝒜{\mathcal{A}}, let (eλ)(e_{\lambda}) be an approximate unit for 𝒜{\mathcal{A}}, and let P=supI(𝒜)s.a.ι(eλ)P=\sup_{I({\mathcal{A}})_{s.a.}}\iota(e_{\lambda}), where I(𝒜)s.a.I({\mathcal{A}})_{s.a.} is the partially ordered set of self-adjoint elements of I(𝒜)I({\mathcal{A}}). By [HamanaCeReMoCoCStAl, Lemma 1.1, parts (i) and (iii)], PP is a central projection of I(𝒜)I({\mathcal{A}}) and (PI(𝒜),ι|J)(PI({\mathcal{A}}),\iota|_{J}) is an injective envelope for JJ. We will call PP the support projection for JJ.

The proof of the next lemma follows from the uniqueness property of injective envelopes.

Lemma \the\numberby (c.f. [PittsStReInI, Proposition 1.11]).

Suppose 𝒜{\mathcal{A}} is an abelian CC^{*}-algebra and let (I(𝒜),ι)(I({\mathcal{A}}),\iota) be an injective envelope for 𝒜{\mathcal{A}}. For i=1,2i=1,2, let JiJ_{i} be closed ideals of 𝒜{\mathcal{A}} with support projections QiQ_{i}. If ϕ:J1J2\phi:J_{1}\rightarrow J_{2} is a *-isomorphism, then there is a unique *-isomorphism ϕ¯:Q1I(𝒜)Q2I(𝒜)\overline{\phi}:Q_{1}I({\mathcal{A}})\rightarrow Q_{2}I({\mathcal{A}}) such that ϕ¯ι|J1=ιϕ\overline{\phi}\circ\iota|_{J_{1}}=\iota\circ\phi.

The following is an interesting example of a weakly non-degenerate inclusion.

Lemma \the\numberby.

Suppose (I(~),ι)(I(\tilde{\mathcal{B}}),\iota) is an injective envelope for the CC^{*}-algebra {\mathcal{B}}. Then (I(~),,ιu)(I(\tilde{\mathcal{B}}),{\mathcal{B}},\iota\circ u_{\mathcal{B}}) is a weakly non-degenerate inclusion.

Proof.

We suppress ι\iota and uu_{\mathcal{B}}, so that ~I(~){\mathcal{B}}\subseteq\tilde{\mathcal{B}}\subseteq I(\tilde{\mathcal{B}}). Let (uλ)(u_{\lambda}) be an approximate unit for {\mathcal{B}}, and let p:=supI(~)uλp:=\sup_{I(\tilde{\mathcal{B}})}u_{\lambda}, that is, pp is the least upper bound of {uλ}\{u_{\lambda}\} taken in the partially ordered set I(~)saI(\tilde{\mathcal{B}})_{sa}. As {\mathcal{B}} is a hereditary subalgebra of ~\tilde{\mathcal{B}}[HamanaCeReMoCoCStAl, Lemma 1.1(i)] shows pp is a projection and 𝔓:=pI(~)p{\mathfrak{P}}:=pI(\tilde{\mathcal{B}})p is an injective envelope for {\mathcal{B}}.

We claim that p=II(~)p=I_{I(\tilde{\mathcal{B}})}. As 𝔓I(~){\mathcal{B}}\subseteq{\mathfrak{P}}\subseteq I(\tilde{\mathcal{B}}), p=supI(~)uλsup𝔓uλpp=\sup_{I(\tilde{\mathcal{B}})}u_{\lambda}\leq\sup_{\mathfrak{P}}u_{\lambda}\leq p so that

supI(~)uλ=sup𝔓uλ.\sup_{I(\tilde{\mathcal{B}})}u_{\lambda}=\sup_{\mathfrak{P}}u_{\lambda}.

By uniqueness of injective envelopes, there is a *-isomorphism θ:𝔓I()\theta:{\mathfrak{P}}\rightarrow I({\mathcal{B}}) with θ|~=id|~\theta|_{\tilde{\mathcal{B}}}={\operatorname{id}}|_{\tilde{\mathcal{B}}}. Then

II(~)=θ(p)=θ(sup𝔓uλ)=supI(~)θ(uλ)=supI(~)uλ=p.I_{I(\tilde{\mathcal{B}})}=\theta(p)=\theta(\sup_{{\mathfrak{P}}}u_{\lambda})=\sup_{I(\tilde{\mathcal{B}})}\theta(u_{\lambda})=\sup_{I(\tilde{\mathcal{B}})}u_{\lambda}=p.

Thus the claim holds.

For aAnn(I(~),)a\in\operatorname{Ann}(I(\tilde{\mathcal{B}}),{\mathcal{B}})[HamanaReEmCStAlMoCoCStAl, Corollary 4.10] shows

aa=apa=a(supI(~)uλ)a=supI(~)auλa=0.a^{*}a=a^{*}pa=a^{*}(\sup_{I(\tilde{\mathcal{B}})}u_{\lambda})a=\sup_{I(\tilde{\mathcal{B}})}a^{*}u_{\lambda}a=0.

Definition \the\numberby (c.f. [PittsStReInI]).

Let (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) be an inclusion and let (I(),ι)(I({\mathcal{B}}),\iota) be an injective envelope for {\mathcal{B}}. A pseudo-expectation for (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) relative to (I(),ι)(I({\mathcal{B}}),\iota) is a contractive and completely positive linear map E:𝒜I()E:{\mathcal{A}}\rightarrow I({\mathcal{B}}) such that Ef=ιuE\circ f=\iota\circ u_{\mathcal{B}}.

When {\mathcal{B}} is identified with f()𝒜f({\mathcal{B}})\subseteq{\mathcal{A}} and u()~u_{\mathcal{B}}({\mathcal{B}})\subseteq\tilde{\mathcal{B}}, we will write E|=ι|E|_{\mathcal{B}}=\iota|_{\mathcal{B}} instead of Ef=ιuE\circ f=\iota\circ u_{\mathcal{B}}. When these identifications are made, we will simply say EE is a pseudo-expectation for (𝒜,)({\mathcal{A}},{\mathcal{B}}).

Notation \the\numberby. For an inclusion (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f), PsExp(𝒜,,f)\operatorname{\operatorname{PsExp}}({\mathcal{A}},{\mathcal{B}},f) will denote the collection of all pseudo-expectations for (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) (relative to a fixed injective envelope (I(),ι)(I({\mathcal{B}}),\iota) for {\mathcal{B}}). As usual, when f()f({\mathcal{B}}) is identified with {\mathcal{B}}, we write PsExp(𝒜,)\operatorname{\operatorname{PsExp}}({\mathcal{A}},{\mathcal{B}}) instead of PsExp(𝒜,,f)\operatorname{\operatorname{PsExp}}({\mathcal{A}},{\mathcal{B}},f).

In general, there are many pseudo-expectations. In some cases however, there is a unique pseudo expectation, and this property plays an essential role in our study.

Definition \the\numberby.

We will say that the inclusion (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the unique pseudo-expectation property if PsExp(𝒜,,f)\operatorname{\operatorname{PsExp}}({\mathcal{A}},{\mathcal{B}},f) is a singleton set. When (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the unique pseudo-expectation property, and the (unique) pseudo-expectation is faithful, we say (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the faithful unique pseudo-expectation property.

Note that by Remark 2.3, if 𝒜{\mathcal{A}} is abelian, and (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the unique pseudo-expectation property, the pseudo-expectation is multiplicative.

That PsExp(𝒜,)\operatorname{\operatorname{PsExp}}({\mathcal{A}},{\mathcal{B}}) is not empty follows from the following general fact.

Lemma \the\numberby.

Suppose {\mathfrak{I}} is an injective CC^{*}-algebra and (𝒜,)({\mathcal{A}},{\mathcal{B}}) is an inclusion. If ϕ:\phi:{\mathcal{B}}\rightarrow{\mathfrak{I}} is a contractive and completely positive map, then ϕ\phi extends to a contractive and completely positive map Φ:𝒜\Phi:{\mathcal{A}}\rightarrow{\mathfrak{I}}.

Proof.

We may assume 𝒜(){\mathcal{A}}\subseteq{\mathcal{B}}({\mathcal{H}}) satisfies 𝒜¯=\overline{{\mathcal{A}}{\mathcal{H}}}={\mathcal{H}} for some Hilbert space {\mathcal{H}}. Thus 𝒜~=𝒜+I\tilde{\mathcal{A}}={\mathcal{A}}+{\mathbb{C}}I_{\mathcal{H}}. Let (uλ)(u_{\lambda}) be an approximate unit for {\mathcal{B}} and put Q:=sotlimuλQ:=\textsc{sot}\lim u_{\lambda}. Then (Q𝒜~Q,+Q)(Q\tilde{\mathcal{A}}Q,{\mathcal{B}}+{\mathbb{C}}Q) is a unital inclusion. As ~+Q\tilde{\mathcal{B}}\simeq{\mathcal{B}}+{\mathbb{C}}Q, we may apply [BrownOzawaC*AlFiDiAp, Proposition 2.2.1] to obtain a unital and completely positive map ϕ~:+Q\tilde{\phi}:{\mathcal{B}}+{\mathbb{C}}Q\rightarrow{\mathfrak{I}} which extends ϕ\phi. Injectivity of {\mathfrak{I}} shows that we may extend ϕ~\tilde{\phi} to a unital, completely positive map Δ:Q𝒜~QI()\Delta:Q\tilde{\mathcal{A}}Q\rightarrow I({\mathcal{B}}). Now observe that the map Φ:𝒜\Phi:{\mathcal{A}}\rightarrow{\mathfrak{I}} given by Φ(a)=Δ(QaQ)\Phi(a)=\Delta(QaQ) is a contractive and completely positive map extending ϕ\phi. ∎

The next several results explore properties of pseudo-expectations. We begin with the relationship between pseudo-expectations for (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) and pseudo-expectations for (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}).

Lemma \the\numberby.

Let (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) be an inclusion, and let (I(),ι)(I({\mathcal{B}}),\iota) be an injective envelope for {\mathcal{B}}. The following statements hold.

  1. (a)

    Let (hλ)(h_{\lambda}) be an approximate unit for {\mathcal{B}}. Suppose x𝒜~x\in\tilde{\mathcal{A}} satisfies 0x0\leq x, x1\left\lVert x\right\rVert\leq 1, and u𝒜(f(hλ))xu_{\mathcal{A}}(f(h_{\lambda}))\leq x for every λ\lambda. If EE is a pseudo-expectation for (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f), then E~(x)=II()\tilde{E}(x)=I_{I({\mathcal{B}})}.

  2. (b)

    If EE is a pseudo-expectation for (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f), then

    E~(I𝒜~)=II()=E~(f~(I~))\tilde{E}(I_{\tilde{\mathcal{A}}})=I_{I({\mathcal{B}})}=\tilde{E}(\tilde{f}(I_{\tilde{\mathcal{B}}}))

    and E~\tilde{E} is a pseudo-expectation for (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}).

  3. (c)

    Suppose Φ:𝒜~I()\Phi:\tilde{\mathcal{A}}\rightarrow I({\mathcal{B}}) is a pseudo-expectation for (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}), and let E:=Φu𝒜E:=\Phi\circ u_{\mathcal{A}}. Then:

    1. (i)

      EE is a pseudo-expectation for (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f),

    2. (ii)

      Φ(I𝒜~)=Φ(f~(I~))=II()\Phi(I_{\tilde{\mathcal{A}}})=\Phi(\tilde{f}(I_{\tilde{\mathcal{B}}}))=I_{I({\mathcal{B}})}; and

    3. (iii)

      E~=Φ\tilde{E}=\Phi.

Proof.

(a) We have already noted in Observation 2.1 that E~\tilde{E} is contractive and completely positive. Let z=E~(x)z=\tilde{E}(x). Then

ι(u(hλ))=E(f(hλ))=E~(u𝒜(f(hλ)))E~(x)=z.\iota(u_{\mathcal{B}}(h_{\lambda}))=E(f(h_{\lambda}))=\tilde{E}(u_{\mathcal{A}}(f(h_{\lambda})))\leq\tilde{E}(x)=z.

Since (I(),,ιu)(I({\mathcal{B}}),{\mathcal{B}},\iota\circ u_{\mathcal{B}}) is weakly non-degenerate (Lemma 2.3), Proposition 2.2(c) shows z=II()z=I_{I({\mathcal{B}})}.

(b) Taking x=I𝒜~x=I_{\tilde{\mathcal{A}}} in part (a) shows E~\tilde{E} is a unital completely positive map and hence is contractive. It remains to show E~f~=ι\tilde{E}\circ\tilde{f}=\iota. To do this, note that part (a) applied with x=I~x=I_{\tilde{\mathcal{B}}} gives II()=E~(f~(I~))I_{I({\mathcal{B}})}=\tilde{E}(\tilde{f}(I_{\tilde{\mathcal{B}}})). Since ~=I~+f~(u())\tilde{\mathcal{B}}={\mathbb{C}}I_{\tilde{\mathcal{B}}}+\tilde{f}(u_{\mathcal{B}}({\mathcal{B}})) and E~f~u=Ef=ιu\tilde{E}\circ\tilde{f}\circ u_{\mathcal{B}}=E\circ f=\iota\circ u_{\mathcal{B}}, it follows that E~f~=ι\tilde{E}\circ\tilde{f}=\iota, so part (b) holds.

(c) Since Φ\Phi is contractive and completely positive, so is EE. As Ef=Φu𝒜f=Φf~u=ιuE\circ f=\Phi\circ u_{\mathcal{A}}\circ f=\Phi\circ\tilde{f}\circ u_{\mathcal{B}}=\iota\circ u_{\mathcal{B}}, EE is a pseudo-expectation for (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f).

Part (a) shows Φ\Phi is a unital map. Since 𝒜~=I𝒜~+u𝒜(𝒜)\tilde{\mathcal{A}}={\mathbb{C}}I_{\tilde{\mathcal{A}}}+u_{\mathcal{A}}({\mathcal{A}}), it follows that Φ=E~\Phi=\tilde{E}. Part (b) now implies item (ii), which completes the proof of part (c). ∎

It is worth noting a bijection between PsExp(𝒜~,~)\operatorname{\operatorname{PsExp}}(\tilde{\mathcal{A}},\tilde{\mathcal{B}}) and PsExp(𝒜,)\operatorname{\operatorname{PsExp}}({\mathcal{A}},{\mathcal{B}}).

Corollary \the\numberby.

The map,

PsExp(𝒜~,~)Φ\displaystyle\operatorname{\operatorname{PsExp}}(\tilde{\mathcal{A}},\tilde{\mathcal{B}})\ni\Phi Φ|𝒜PsExp(𝒜,)\displaystyle\mapsto\Phi|_{\mathcal{A}}\in\operatorname{\operatorname{PsExp}}({\mathcal{A}},{\mathcal{B}})
is a bijection with inverse
PsExp(𝒜,)E\displaystyle\operatorname{\operatorname{PsExp}}({\mathcal{A}},{\mathcal{B}})\ni E E~PsExp(𝒜~,~).\displaystyle\mapsto\tilde{E}\in\operatorname{\operatorname{PsExp}}(\tilde{\mathcal{A}},\tilde{\mathcal{B}}).
Proof.

Apply Lemma 2.3. ∎

Combining Proposition 2.2 and Lemma 2.3 we obtain the following.

Proposition \the\numberby.

Suppose (𝒜,)({\mathcal{A}},{\mathcal{B}}) is a weakly non-degenerate inclusion and let (I(),ι)(I({\mathcal{B}}),\iota) be an injective envelope for {\mathcal{B}}. If E:𝒜I()E:{\mathcal{A}}\rightarrow I({\mathcal{B}}) is a pseudo-expectation, then for every x𝒜x\in{\mathcal{A}} and h,kh,k\in{\mathcal{B}},

E(hxk)=ι(h)E(x)ι(k).E(hxk)=\iota(h)E(x)\iota(k).

We do not know whether the faithfulness of a pseudo-expectation EE for an inclusion (𝒜,)({\mathcal{A}},{\mathcal{B}}) implies E~\tilde{E} is a faithful pseudo-expectation for (𝒜~,~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}}), but we suspect it is not true in general. Our next few results concern the relationship between faithfulness for a pseudo-expectation EE and faithfulness of E~\tilde{E}.

Lemma \the\numberby.

Let EE be a pseudo-expectation for (𝒜,)({\mathcal{A}},{\mathcal{B}}). If E~\tilde{E} is faithful, then (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate.

Proof.

By Lemma 2.3(c) E~(I𝒜~I~)=0\tilde{E}(I_{\tilde{\mathcal{A}}}-I_{\tilde{\mathcal{B}}})=0, so faithfulness of E~\tilde{E} gives I𝒜~=I~I_{\tilde{\mathcal{A}}}=I_{\tilde{\mathcal{B}}}. Thus (𝒜~,~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}}) is a unital inclusion and hence is weakly non-degenerate.

When xAnn(𝒜,)x\in\operatorname{Ann}({\mathcal{A}},{\mathcal{B}}) and bb\in{\mathcal{B}}, Proposition 2.3 implies

E(xx)ι(b)=E(xxb)=0=E(bxx)=ι(b)E(xx).E(x^{*}x)\iota(b)=E(x^{*}xb)=0=E(bx^{*}x)=\iota(b)E(x^{*}x).

Therefore E(xx)Ann(I(),,ιu)E(x^{*}x)\in\operatorname{Ann}(I({\mathcal{B}}),{\mathcal{B}},\iota\circ u_{\mathcal{B}}). By Lemma 2.3, E(xx)=0E(x^{*}x)=0, so faithfulness of EE gives x=0x=0. Thus (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate. ∎

Proposition \the\numberby.

Let (𝒜,)({\mathcal{A}},{\mathcal{B}}) be an inclusion and suppose EE is a pseudo-expectation for (𝒜,)({\mathcal{A}},{\mathcal{B}}). Then E~\tilde{E} is a faithful pseudo-expectation for (𝒜~,~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}}) if and only if (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate and EE is faithful.

Proof.

(\Rightarrow) Suppose E~\tilde{E} is faithful. Clearly EE is faithful, and Lemma 2.3 shows (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate.

(\Leftarrow) Suppose EE is faithful and (𝒜,)({\mathcal{A}},{\mathcal{B}}) is weakly non-degenerate. Since E~=E\tilde{E}=E when 𝒜{\mathcal{A}} is unital, we may as well assume 𝒜{\mathcal{A}} is not unital.

Suppose (x,λ)𝒜~(x,\lambda)\in\tilde{\mathcal{A}} and E~((x,λ)(x,λ))=0\tilde{E}((x,\lambda)^{*}(x,\lambda))=0. Then

0\displaystyle 0 =E(xx)+λ¯E(x)+λE(x)+|λ|2II()\displaystyle=E(x^{*}x)+\overline{\lambda}E(x)+\lambda E(x)^{*}+|\lambda|^{2}I_{I({\mathcal{B}})}
E(x)E(x)+λ¯E(x)+λE(x)+|λ|2II()\displaystyle\geq E(x)^{*}E(x)+\overline{\lambda}E(x)+\lambda E(x)^{*}+|\lambda|^{2}I_{I({\mathcal{B}})}
=(E(x)+λII())(E(x)+λII()),\displaystyle=(E(x)+\lambda I_{I({\mathcal{B}})})^{*}(E(x)+\lambda I_{I({\mathcal{B}})}),

whence

E(x)=λII().E(x)=-\lambda I_{I({\mathcal{B}})}.

We claim that λ=0\lambda=0.

We argue by contradiction. Suppose λ0\lambda\neq 0. By scaling, we may assume λ=1\lambda=-1, so E(x)=II()E(x)=I_{I({\mathcal{B}})}. Then

E(xx)=II()E(x^{*}x)=I_{I({\mathcal{B}})}

because 0=E~((x,1)(x,1))0=\tilde{E}((x,-1)^{*}(x,-1)).

Given bb\in{\mathcal{B}}, Proposition 2.3 shows

E((xbb)(xbb))=ι(b)E(xx)ι(b)ι(b)E(x)ι(b)ι(b)E(x)ι(b)+ι(bb)=0.E((xb-b)^{*}(xb-b))=\iota(b)^{*}E(x^{*}x)\iota(b)-\iota(b)^{*}E(x)\iota(b)-\iota(b)^{*}E(x^{*})\iota(b)+\iota(b^{*}b)=0.

Faithfulness of EE gives xb=bxb=b. Similar considerations yield bx=bbx=b. Lemma 2.2(b) shows that xx is the identity for 𝒜{\mathcal{A}}, contradicting the assumption that 𝒜{\mathcal{A}} is not unital. Hence λ=0\lambda=0.

Thus (x,λ)=(x,0)(x,\lambda)=(x,0) and using faithfulness of EE once again, we find x=0x=0. Therefore E~\tilde{E} is faithful. ∎

For a unital inclusion (𝒜,)({\mathcal{A}},{\mathcal{B}})[PittsZarikianUnPsExC*In, Corollary 3.14] shows that the faithful unique pseudo-expectation property implies c{\mathcal{B}}^{c} is abelian. We now note that this useful structural fact holds when (𝒜,)({\mathcal{A}},{\mathcal{B}}) is not assumed unital.

Proposition \the\numberby.

If the inclusion (𝒜,)({\mathcal{A}},{\mathcal{B}}) has the faithful unique pseudo-expectation property, then BcB^{c} is abelian.

Proof.

The proof is an adaptation of the arguments establishing [PittsZarikianUnPsExC*In, Theorem 3.12 and Corollary 3.14].

We may suppose that for some Hilbert space {\mathcal{H}}, 𝒜(){\mathcal{A}}\subseteq{\mathcal{B}}({\mathcal{H}}) and 𝒜¯=\overline{{\mathcal{A}}{\mathcal{H}}}={\mathcal{H}}. Then 𝒜~=𝒜+I\tilde{\mathcal{A}}={\mathcal{A}}+{\mathbb{C}}I. Let E:𝒜I()E:{\mathcal{A}}\rightarrow I({\mathcal{B}}) be the faithful unique pseudo-expectation, and put

Φ:=E~.\Phi:=\tilde{E}.

While Φ\Phi is the unique pseudo-expectation for (𝒜~,~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}}), we do not know whether it is faithful, so we cannot apply [PittsZarikianUnPsExC*In, Corollary 3.14]. Instead we argue by contradiction.

Assume 𝒜{\mathcal{B}}^{\prime}\cap{\mathcal{A}} is not abelian. Then there exists x𝒜x\in{\mathcal{B}}^{\prime}\cap{\mathcal{A}} so that x2=0x^{2}=0 and x=1\left\lVert x\right\rVert=1. Proceed as in the proof of [PittsZarikianUnPsExC*In, Theorem 3.12] to obtain a one-parameter family of unital completely positive maps θλ:()()\theta_{\lambda}:{\mathcal{B}}({\mathcal{H}})\rightarrow{\mathcal{B}}({\mathcal{H}}) (λ[0,1])\lambda\in[0,1])) such that

θλ|=id|,θλ(xx)=λ(xx+xx),andθλ(xx)=(1λ)(xx+xx).\theta_{\lambda}|_{\mathcal{B}}={\operatorname{id}}|_{\mathcal{B}},\quad\theta_{\lambda}(x^{*}x)=\lambda(x^{*}x+xx^{*}),\quad\text{and}\quad\theta_{\lambda}(xx^{*})=(1-\lambda)(x^{*}x+xx^{*}).

Let

𝒮:=+xx+xx+I{\mathcal{S}}:={\mathcal{B}}+{\mathbb{C}}x^{*}x+{\mathbb{C}}xx^{*}+{\mathbb{C}}I

and note that 𝒮𝒜~{\mathcal{S}}\subseteq\tilde{\mathcal{A}} is an operator system such that θλ(𝒮)𝒮\theta_{\lambda}({\mathcal{S}})\subseteq{\mathcal{S}}. Let Φλ(0)=Φθλ\Phi_{\lambda}^{(0)}=\Phi\circ\theta_{\lambda}. Then Φλ(0)\Phi_{\lambda}^{(0)} is a unital completely positive map from 𝒮{\mathcal{S}} into I()I({\mathcal{B}}). Using the injectivity of I()I({\mathcal{B}}), we may extend Φλ(0)\Phi_{\lambda}^{(0)} to a unital completely positive map Φλ:𝒜~I()\Phi_{\lambda}:\tilde{\mathcal{A}}\rightarrow I({\mathcal{B}}). Putting Eλ:=Φλ|𝒜E_{\lambda}:=\Phi_{\lambda}|_{\mathcal{A}}, we find EλE_{\lambda} is a pseudo-expectation for (𝒜,)({\mathcal{A}},{\mathcal{B}}).

As xx𝒮x^{*}x\in{\mathcal{S}}, for λ0\lambda\neq 0, the faithfulness of EE gives

Eλ(xx)=Φ(θλ(xx))=λE(xx+xx)0.E_{\lambda}(x^{*}x)=\Phi(\theta_{\lambda}(x^{*}x))=\lambda E(x^{*}x+xx^{*})\neq 0.

Therefore, when λ,μ(0,1]\lambda,\mu\in(0,1] and λμ\lambda\neq\mu, EλE_{\lambda} and EμE_{\mu} are distinct pseudo-expectations for (𝒜,)({\mathcal{A}},{\mathcal{B}}). This contradicts the hypothesis that (𝒜,)({\mathcal{A}},{\mathcal{B}}) has a unique pseudo-expectation, and completes the proof. ∎

Our next goal is to show that when (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is an inclusion and 𝒟{\mathcal{D}} is abelian, then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property if and only if (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) does. We begin with the commutative case.

Lemma \the\numberby.

Suppose (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) is an inclusion with 𝒜{\mathcal{A}} abelian. If (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the faithful unique pseudo-expectation property, then (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) has the faithful unique pseudo-expectation property.

Proof.

Let EE be the pseudo-expectation for (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f). Then E~\tilde{E} is the unique pseudo-expectation for (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) and E~(I𝒜~f~(I~))=0\tilde{E}(I_{\tilde{\mathcal{A}}}-\tilde{f}(I_{\tilde{\mathcal{B}}}))=0 (Lemma 2.3).

We wish to show (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) is a unital inclusion. Let p:=I𝒜~f~(I~)p:=I_{\tilde{\mathcal{A}}}-\tilde{f}(I_{\tilde{\mathcal{B}}}). Then pp is a projection in 𝒜~\tilde{\mathcal{A}}. Suppose 0x𝒜0\leq x\in{\mathcal{A}} is such that u𝒜(x)p𝒜~u_{\mathcal{A}}(x)\in p\tilde{\mathcal{A}}. Then

0E(x)=E~(u𝒜(x))=E~(pu𝒜(x)p)xE~(p)=0,0\leq E(x)=\tilde{E}(u_{\mathcal{A}}(x))=\tilde{E}(pu_{\mathcal{A}}(x)p)\leq\left\lVert x\right\rVert\tilde{E}(p)=0,

so E(x)=0E(x)=0. Faithfulness of EE yields x=0x=0. Since any CC^{*}-algebra is the span of its positive elements, u𝒜(𝒜)p𝒜~={0}u_{\mathcal{A}}({\mathcal{A}})\cap p\tilde{\mathcal{A}}=\{0\}. Thus p𝒜~p\tilde{\mathcal{A}} is an ideal of 𝒜~\tilde{\mathcal{A}} having trivial intersection with u𝒜(𝒜)u_{\mathcal{A}}({\mathcal{A}}). Since u𝒜(𝒜)u_{\mathcal{A}}({\mathcal{A}}) is an essential ideal of 𝒜~\tilde{\mathcal{A}}, we conclude that p𝒜~={0}p\tilde{\mathcal{A}}=\{0\}, whence p=0p=0. Thus, (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) is a unital inclusion having the unique pseudo-expectation property.

By [PittsZarikianUnPsExC*In, Corollary 3.21], E~\tilde{E} is a *-homomorphism. Since EE is faithful, kerE~\ker\tilde{E} is an ideal of 𝒜~\tilde{\mathcal{A}} having trivial intersection with u𝒜(𝒜)u_{\mathcal{A}}({\mathcal{A}}). Using the fact that u𝒜(𝒜)u_{\mathcal{A}}({\mathcal{A}}) is an essential ideal in 𝒜~\tilde{\mathcal{A}} once again, we conclude that kerE~={0}\ker\tilde{E}=\{0\}, that is, E~\tilde{E} is faithful. ∎

Theorem \the\numberby.

Suppose (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) is an inclusion with 𝒟{\mathcal{D}} abelian. Then (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) has the faithful unique pseudo-expectation property if and only if (𝒞~,𝒟~,f~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}},\tilde{f}) has the faithful unique pseudo-expectation property.

Proof.

(\Leftarrow) Let ΦPsExp(𝒞~,𝒟~,f~)\Phi\in\operatorname{\operatorname{PsExp}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}},\tilde{f}) be the unique (and faithful) pseudo-expectation. Lemma 2.3 shows Φu𝒞\Phi\circ u_{\mathcal{C}} is the unique, and necessarily faithful, pseudo-expectation for (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f).

(\Rightarrow) Now suppose (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) has the faithful unique pseudo-expectation property and let EE be the pseudo-expectation. For notational purposes, let

g:=u𝒞f,:=rCom(𝒞,𝒟,f),and+:=rCom(𝒞~,𝒟,g),g:=u_{\mathcal{C}}\circ f,\quad{\mathcal{B}}:=\operatorname{\textsc{rCom}}({\mathcal{C}},{\mathcal{D}},f),\quad\text{and}\quad{\mathcal{B}}^{+}:=\operatorname{\textsc{rCom}}(\tilde{\mathcal{C}},{\mathcal{D}},g),

that is,

\displaystyle{\mathcal{B}} :={x𝒞:f(d)x=xf(d) for all d𝒟}\displaystyle:=\{x\in{\mathcal{C}}:f(d)x=xf(d)\text{ for all }d\in{\mathcal{D}}\}
and
+\displaystyle{\mathcal{B}}^{+} :={x𝒞~:g(d)x=xg(d) for all d𝒟}.\displaystyle:=\{x\in\tilde{\mathcal{C}}:g(d)x=xg(d)\text{ for all }d\in{\mathcal{D}}\}.

By Proposition 2.3, {\mathcal{B}} is abelian; hence +{\mathcal{B}}^{+} is also abelian. As 𝒟{\mathcal{D}} is abelian, we obtain the two inclusions,

(,𝒟,f)and(+,𝒟,g).({\mathcal{B}},{\mathcal{D}},f)\quad\text{and}\quad({\mathcal{B}}^{+},{\mathcal{D}},g).

We establish faithfulness of E~:𝒞~I(𝒟)\tilde{E}:\tilde{\mathcal{C}}\rightarrow I({\mathcal{D}}) by considering two cases: {\mathcal{B}} is unital; and {\mathcal{B}} is not unital.

Suppose first {\mathcal{B}} is unital. By the definition of pseudo-expectation, Ef=ιu𝒟E\circ f=\iota\circ u_{\mathcal{D}}. Therefore, Δ:=E|\Delta:=E|_{\mathcal{B}} is a faithful pseudo-expectation for (,𝒟,f)({\mathcal{B}},{\mathcal{D}},f). Lemma 2.3(b) shows Δ~\tilde{\Delta} is a pseudo-expectation for (~,𝒟~,f~)(\tilde{\mathcal{B}},\tilde{\mathcal{D}},\tilde{f}). But ~=\tilde{\mathcal{B}}={\mathcal{B}} by hypothesis, so Δ~\tilde{\Delta} is a faithful pseudo-expectation for (~,𝒟~,f~)(\tilde{\mathcal{B}},\tilde{\mathcal{D}},\tilde{f}). By Lemma 2.3, (,𝒟,f)({\mathcal{B}},{\mathcal{D}},f) is weakly non-degenerate. Recalling that Ann(𝒞,𝒟,f)\operatorname{Ann}({\mathcal{C}},{\mathcal{D}},f)\subseteq{\mathcal{B}}, we find Ann(𝒞,𝒟,f)=Ann(,𝒟,f)\operatorname{Ann}({\mathcal{C}},{\mathcal{D}},f)=\operatorname{Ann}({\mathcal{B}},{\mathcal{D}},f); therefore (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) is weakly non-degenerate. Proposition 2.3 now shows that E~\tilde{E} is a faithful pseudo-expectation for (𝒞~,𝒟~,f~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}},\tilde{f}). Thus (𝒞~,𝒟~,f~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}},\tilde{f}) has the faithful unique pseudo-expectation property (Corollary 2.3).

Now suppose {\mathcal{B}} is not unital. By definition of {\mathcal{B}}, 𝒞{\mathcal{C}} cannot be unital. Thus 𝒞~=𝒞\tilde{\mathcal{C}}={\mathcal{C}}^{\dagger}, ~=\tilde{\mathcal{B}}={\mathcal{B}}^{\dagger}, and for d𝒟d\in{\mathcal{D}} and (x,λ)𝒞~(x,\lambda)\in\tilde{\mathcal{C}},

g(d)=(f(d),0)andE~(x,λ)=E(x)+λII().g(d)=(f(d),0)\quad\text{and}\quad\tilde{E}(x,\lambda)=E(x)+\lambda I_{I({\mathcal{B}})}.

Then +={(b,λ)𝒞:b}={\mathcal{B}}^{+}=\{(b,\lambda)\in{\mathcal{C}}^{\dagger}:b\in{\mathcal{B}}\}={\mathcal{B}}^{\dagger}. Therefore,

(~,𝒟,uf)=(+,𝒟,g)and hence(~,𝒟~,f~)=(+,𝒟~,g~).(\tilde{\mathcal{B}},{\mathcal{D}},u_{\mathcal{B}}\circ f)=({\mathcal{B}}^{+},{\mathcal{D}},g)\quad\text{and hence}\quad(\tilde{\mathcal{B}},\tilde{\mathcal{D}},\tilde{f})=({\mathcal{B}}^{+},\tilde{\mathcal{D}},\tilde{g}).

We claim that (,𝒟,f)({\mathcal{B}},{\mathcal{D}},f) has the faithful unique pseudo-expectation property. Lemma 2.3 implies that E~\tilde{E} is the unique pseudo-expectation for (𝒞~,𝒟~,f~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}},\tilde{f}). If Ψ:+I(𝒟)\Psi:{\mathcal{B}}^{+}\rightarrow I({\mathcal{D}}) is a pseudo-expectation for (+,𝒟~,f~)({\mathcal{B}}^{+},\tilde{\mathcal{D}},\tilde{f}), then Ψ\Psi is a unital map (Lemma 2.3(c)). Injectivity of I()I({\mathcal{B}}) then implies that Ψ\Psi extends to a pseudo-expectation PP for (𝒞~,𝒟~,f~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}},\tilde{f}). Thus P=E~P=\tilde{E}, whence E~|+\tilde{E}|_{{\mathcal{B}}^{+}} is the unique pseudo-expectation for (+,𝒟~,f~)=(~,𝒟~,f~)({\mathcal{B}}^{+},\tilde{\mathcal{D}},\tilde{f})=(\tilde{\mathcal{B}},\tilde{\mathcal{D}},\tilde{f}). Corollary 2.3 shows (,𝒟,f)({\mathcal{B}},{\mathcal{D}},f) has the unique pseudo-expectation property and E|E|_{\mathcal{B}} is the pseudo-expectation. Since EE is faithful on 𝒞{\mathcal{C}}, E|E|_{\mathcal{B}} is faithful. Thus (,𝒟,f)({\mathcal{B}},{\mathcal{D}},f) has the faithful unique pseudo-expectation property and Δ:=E|\Delta:=E|_{{\mathcal{B}}} is its pseudo-expectation.

By Lemma 2.3, (~,𝒟~,f~)(\tilde{\mathcal{B}},\tilde{\mathcal{D}},\tilde{f}) has the faithful unique pseudo expectation property. Thus, Δ~\tilde{\Delta} is faithful. Proposition 2.3 shows (,𝒟,f)({\mathcal{B}},{\mathcal{D}},f) is weakly non-degenerate. Because Ann(𝒞,𝒟,f)\operatorname{Ann}({\mathcal{C}},{\mathcal{D}},f)\subseteq{\mathcal{B}}, Ann(𝒞,𝒟,f)=Ann(,𝒟,f)={0}\operatorname{Ann}({\mathcal{C}},{\mathcal{D}},f)=\operatorname{Ann}({\mathcal{B}},{\mathcal{D}},f)=\{0\}. Therefore, (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) is weakly non-degenerate. Another application of Proposition 2.3 now shows E~\tilde{E} is faithful. Corollary 2.3 shows (𝒞~,𝒟~,f~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}},\tilde{f}) has the faithful unique pseudo-expectation property, completing the proof. ∎

Corollary \the\numberby.

Let (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) be an inclusion such that 𝒟{\mathcal{D}} is abelian. If (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) has the faithful unique pseudo-expectation property, then (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) is weakly non-degenerate.

Proof.

Combine Theorem 2.3 with Proposition 2.3. ∎

Corollary \the\numberby.

Suppose (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) is an inclusion such that 𝒟{\mathcal{D}} is abelian and let {\mathcal{B}} be a CC^{*}-algebra such that f(𝒟)𝒞f({\mathcal{D}})\subseteq{\mathcal{B}}\subseteq{\mathcal{C}}. If (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) has the faithful unique pseudo-expectation property, then (,𝒟,f)({\mathcal{B}},{\mathcal{D}},f) has the faithful unique pseudo-expectation property.

Proof.

Corollary 2.3 shows (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f) is weakly non-degenerate, so (𝒞~,𝒟~,f~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}},\tilde{f}) is a unital inclusion having the faithful unique pseudo-expectation property. Corollary 2.2 shows (~,𝒟~,f~)(\tilde{\mathcal{B}},\tilde{\mathcal{D}},\tilde{f}) is a unital inclusion with f~(𝒟~)~𝒞~\tilde{f}(\tilde{\mathcal{D}})\subseteq\tilde{\mathcal{B}}\subseteq\tilde{\mathcal{C}}. Thus, if E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) is the pseudo-expectation for (𝒞,𝒟,f)({\mathcal{C}},{\mathcal{D}},f), then E~\tilde{E} is the faithful and unique pseudo-expectation for (𝒞~,𝒟~,f~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}},\tilde{f}). Then by [PittsZarikianUnPsExC*In, Proposition 2.6], (~,𝒟~,f~)(\tilde{\mathcal{B}},\tilde{\mathcal{D}},\tilde{f}) has the faithful unique pseudo-expectation property and (E~)|~(\tilde{E})|_{\tilde{\mathcal{B}}} is the unique pseudo-expectation for (~,𝒟~,f~)(\tilde{\mathcal{B}},\tilde{\mathcal{D}},\tilde{f}). Another application of Theorem 2.3 shows (,𝒟,f)({\mathcal{B}},{\mathcal{D}},f) has the faithful unique pseudo-expectation property. ∎

We now extend [PittsStReInII, Proposition 5.5(b)] from the unital setting to include the non-unital case. The utility of the following comes from the fact that it is sometimes easier to establish the faithful unique pseudo-expectation property than to show a subalgebra is a MASA.

Proposition \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular inclusion with 𝒟{\mathcal{D}} abelian. The following statements are equivalent.

  1. (a)

    (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a Cartan inclusion.

  2. (b)

    (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property and there is a conditional expectation E:𝒞𝒟E:{\mathcal{C}}\rightarrow{\mathcal{D}}.

Proof.

First notice that if (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) satisfies either condition (a) or (b), then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is weakly non-degenerate. Indeed, if (a) holds, then Lemma 2.2(d) implies (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is weakly non-degenerate; when (b) holds, apply Corollary 2.3. Thus in both cases, (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a unital inclusion (Lemma 2.2(e)).

Our arguments establishing (a)\Leftrightarrow(b) differ depending on whether 𝒞{\mathcal{C}} is unital or non-unital, so we consider those cases separately.

Case 1: Assume 𝒞{\mathcal{C}} is unital.

(a)\Rightarrow(b) By hypothesis, (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a MASA inclusion, hence (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a unital inclusion. Then [PittsStReInII, Proposition 5.5(b)] gives (b).

(b)\Rightarrow(a) For any d𝒟d\in{\mathcal{D}}, we have

E(I𝒞)d=E(d)=d=dE(I𝒞),E(I_{\mathcal{C}})d=E(d)=d=dE(I_{\mathcal{C}}),

so E(I𝒞)E(I_{\mathcal{C}}) is the unit for 𝒟{\mathcal{D}}. As (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is weakly non-degenerate, it is a unital inclusion. Now [PittsStReInII, Proposition 5.5(b)] shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a Cartan inclusion.

Case 2: Assume 𝒞{\mathcal{C}} is not unital. Then 𝒟{\mathcal{D}} is not unital by Lemma 2.2.

(a)\Rightarrow(b) By [PittsNoApUnInC*Al, Proposition 3.2], (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a Cartan inclusion. Case 1 shows: i) (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property (see Corollary 2.3); and ii) there exists a conditional expectation E~:𝒞~𝒟~\tilde{E}:\tilde{\mathcal{C}}\rightarrow\tilde{\mathcal{D}}. We wish to show E~|𝒞\tilde{E}|_{\mathcal{C}} is a conditional expectation of 𝒞{\mathcal{C}} onto 𝒟{\mathcal{D}}. Let (uλ)(u_{\lambda}) be an approximate unit for 𝒟{\mathcal{D}}. Then (uλ)(u_{\lambda}) is an approximate unit for 𝒞{\mathcal{C}} by [PittsNoApUnInC*Al, Theorem 2.5]. So for x𝒞x\in{\mathcal{C}}, the fact that 𝒟{\mathcal{D}} is an ideal in 𝒟~\tilde{\mathcal{D}} gives

E~(x)=limλE~(xuλ)=limλE~(x)uλ𝒟.\tilde{E}(x)=\lim_{\lambda}\tilde{E}(xu_{\lambda})=\lim_{\lambda}\tilde{E}(x)u_{\lambda}\in{\mathcal{D}}.

Thus E~|𝒞\tilde{E}|_{\mathcal{C}} is a conditional expectation of 𝒞{\mathcal{C}} onto 𝒟{\mathcal{D}}. This establishes (b).

(b)\Rightarrow(a) By Proposition 2.3 and Corollary 2.3, (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) has the faithful unique pseudo-expectation property. Noting that E~\tilde{E} is a conditional expectation of 𝒞~\tilde{\mathcal{C}} onto 𝒟~\tilde{\mathcal{D}}, we conclude from Case 1 that (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a Cartan inclusion. Applying [PittsNoApUnInC*Al, Proposition 3.2] again, we find (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a Cartan inclusion. This completes the proof. ∎

2.4. The Ideal Intersection Property

The purpose of this section is to show that the inclusions (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) and (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) both have the ideal intersection property or both do not, and also to explore some relationships between the ideal intersection property and the faithful unique pseudo-expectation property.

Lemma \the\numberby.

Let (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) be an inclusion. Then (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the ideal intersection property if and only if (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) has the ideal intersection property.

Proof.

Suppose (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the ideal intersection property and let J𝒜~J\unlhd\tilde{\mathcal{A}} satisfy Jf~(~)={0}J\cap\tilde{f}(\tilde{\mathcal{B}})=\{0\}. As

(Ju𝒜(𝒜))u𝒜(f())=(Ju𝒜(𝒜))f~(u())Jf~(~)={0},(J\cap u_{\mathcal{A}}({\mathcal{A}}))\cap u_{\mathcal{A}}(f({\mathcal{B}}))=(J\cap u_{\mathcal{A}}({\mathcal{A}}))\cap\tilde{f}(u_{\mathcal{B}}({\mathcal{B}}))\subseteq J\cap\tilde{f}(\tilde{\mathcal{B}})=\{0\},

the ideal intersection property for (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) gives Ju𝒜(𝒜)={0}J\cap u_{\mathcal{A}}({\mathcal{A}})=\{0\}. Since u𝒜(𝒜)u_{\mathcal{A}}({\mathcal{A}}) is an essential ideal in 𝒜~\tilde{\mathcal{A}}, we obtain J=0J=0. Thus (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) has the ideal intersection property.

Now suppose (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) has the ideal intersection property and let J𝒜J\unlhd{\mathcal{A}} satisfy Jf()={0}J\cap f({\mathcal{B}})=\{0\}. Then u𝒜(J)𝒜~u_{\mathcal{A}}(J)\unlhd\tilde{\mathcal{A}}, and we claim

(2.4.1) u𝒜(J)f~(~)={0}.u_{\mathcal{A}}(J)\cap\tilde{f}(\tilde{\mathcal{B}})=\{0\}.

When {\mathcal{B}} is unital, ~=\tilde{\mathcal{B}}={\mathcal{B}} and f~=u𝒜f\tilde{f}=u_{\mathcal{A}}\circ f, so (2.4.1) holds because u𝒜u_{\mathcal{A}} is one-to-one.

Suppose then that {\mathcal{B}} is not unital. Recall that {0}\{0\} is a unital algebra, so in particular, {0}{\mathcal{B}}\neq\{0\}. If xu𝒜(J)f~(~)x\in u_{\mathcal{A}}(J)\cap\tilde{f}(\tilde{\mathcal{B}}), then xx has the form

x=f~(b,λ)=u𝒜(f(b))+λI𝒜~x=\tilde{f}(b,\lambda)=u_{\mathcal{A}}(f(b))+\lambda I_{\tilde{\mathcal{A}}}

for some bb\in{\mathcal{B}} and λ\lambda\in{\mathbb{C}}. Since Jf()={0}J\cap f({\mathcal{B}})=\{0\}, (2.4.1) will follow once we show λ=0\lambda=0.

Arguing by contradiction, suppose λ0\lambda\neq 0. By scaling, we may assume λ=1\lambda=-1, so that x=u𝒜(f(b))I𝒜~x=u_{\mathcal{A}}(f(b))-I_{\tilde{\mathcal{A}}}, for some bb\in{\mathcal{B}}. For any hh\in{\mathcal{B}}, we have xu𝒜(f(h))u𝒜(J)u𝒜(f())={0}xu_{\mathcal{A}}(f(h))\in u_{\mathcal{A}}(J)\cap u_{\mathcal{A}}(f({\mathcal{B}}))=\{0\}, so xu𝒜(f(h))=0xu_{\mathcal{A}}(f(h))=0. Taking h=bh=b^{*}, we conclude that

bb=b.bb^{*}=b^{*}.

Since {\mathcal{B}} is not unital, this forces b=0b=0. Therefore x=I𝒜~u𝒜(J)x=-I_{\tilde{\mathcal{A}}}\in u_{\mathcal{A}}(J), whence u𝒜(J)=𝒜~u_{\mathcal{A}}(J)=\tilde{\mathcal{A}}. This is impossible if 𝒜{\mathcal{A}} is not unital because u𝒜(J)u𝒜(𝒜)𝒜~u_{\mathcal{A}}(J)\subseteq u_{\mathcal{A}}({\mathcal{A}})\neq\tilde{\mathcal{A}}. On the other hand, if 𝒜{\mathcal{A}} is unital, we again reach a contradiction because {0}=u𝒜(Jf())=𝒜~u𝒜(f())=u𝒜(f())\{0\}=u_{\mathcal{A}}(J\cap f({\mathcal{B}}))=\tilde{\mathcal{A}}\cap u_{\mathcal{A}}(f({\mathcal{B}}))=u_{\mathcal{A}}(f({\mathcal{B}})), contrary to the fact that {0}{\mathcal{B}}\neq\{0\}. Therefore, λ=0\lambda=0, completing the proof of (2.4.1).

Since (𝒜~,~,f~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}},\tilde{f}) has the ideal intersection property,  (2.4.1) gives J={0}J=\{0\}. Thus (𝒜,,f)({\mathcal{A}},{\mathcal{B}},f) has the ideal intersection property. ∎

Our next result concerns the ideal intersection property for intermediate inclusions in the abelian case. It is the same as [PittsStReInII, Lemma 5.4] except the hypothesis that the CC^{*}-algebras involved have a common unit is dropped and we explicitly consider the inclusion mappings.

Lemma \the\numberby.

For 1i<j31\leq i<j\leq 3, Let (𝒟j,𝒟i,fji)({\mathcal{D}}_{j},{\mathcal{D}}_{i},f_{ji}) be inclusions such that f31=f32f21f_{31}=f_{32}\circ f_{21} and 𝒟3{\mathcal{D}}_{3} is abelian. Then (𝒟3,𝒟1,f31)({\mathcal{D}}_{3},{\mathcal{D}}_{1},f_{31}) has the ideal intersection property if and only if both (𝒟3,𝒟2,f32)({\mathcal{D}}_{3},{\mathcal{D}}_{2},f_{32}) and (𝒟2,𝒟1,f21)({\mathcal{D}}_{2},{\mathcal{D}}_{1},f_{21}) have the ideal intersection property.

Proof.

The implication ()(\Leftarrow) is left to the reader.

()(\Rightarrow) Suppose (𝒟3,𝒟1,f31)({\mathcal{D}}_{3},{\mathcal{D}}_{1},f_{31}) has the ideal intersection property. Lemma 2.2(c) shows (𝒟3,𝒟1,f31)({\mathcal{D}}_{3},{\mathcal{D}}_{1},f_{31}) is weakly non-degenerate. Therefore, both (𝒟2,𝒟1,f21)({\mathcal{D}}_{2},{\mathcal{D}}_{1},f_{21}) and (𝒟3,𝒟2,f32)({\mathcal{D}}_{3},{\mathcal{D}}_{2},f_{32}) are weakly non-degenerate. Then for 1i<j31\leq i<j\leq 3, each of (𝒟~j,𝒟~i,f~ji)(\tilde{\mathcal{D}}_{j},\tilde{\mathcal{D}}_{i},\tilde{f}_{ji}) is a unital inclusion, hence f~31=f~32f~21\tilde{f}_{31}=\tilde{f}_{32}\circ\tilde{f}_{21} (Corollary 2.2). By Lemma 2.2(a), these inclusions have a common unit in the sense that

I𝒟~3=f~32(I𝒟~2)andI𝒟~2=f~21(I𝒟~1).I_{\tilde{\mathcal{D}}_{3}}=\tilde{f}_{32}(I_{\tilde{\mathcal{D}}_{2}})\quad\text{and}\quad I_{\tilde{\mathcal{D}}_{2}}=\tilde{f}_{21}(I_{\tilde{\mathcal{D}}_{1}}).

By [PittsStReInII, Lemma 5.4], (𝒟~2,𝒟~1,f~21)(\tilde{\mathcal{D}}_{2},\tilde{\mathcal{D}}_{1},\tilde{f}_{21}) and (𝒟~3,𝒟~2,f~32)(\tilde{\mathcal{D}}_{3},\tilde{\mathcal{D}}_{2},\tilde{f}_{32}) have the ideal intersection property. An application of Lemma 2.4 completes the proof. ∎

Our final result of this section extends a portion of [PittsZarikianUnPsExC*In, Corollary 3.22] to include non-unital settings.

Proposition \the\numberby.

Suppose (𝒟2,𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1}) is an inclusion with 𝒟2{\mathcal{D}}_{2} abelian and let (I(𝒟1),ι1)(I({\mathcal{D}}_{1}),\iota_{1}) be an injective envelope for 𝒟1{\mathcal{D}}_{1}. Then (𝒟2,𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1}) has the ideal intersection property if and only if (𝒟2,𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1}) has the faithful unique pseudo-expectation property.

When this occurs, the unique pseudo-expectation ι2:𝒟2I(𝒟1)\iota_{2}:{\mathcal{D}}_{2}\rightarrow I({\mathcal{D}}_{1}) is a *-monomorphism and (I(𝒟1),ι2)(I({\mathcal{D}}_{1}),\iota_{2}) is an injective envelope for 𝒟2{\mathcal{D}}_{2}.

Proof.

(\Rightarrow) Suppose (𝒟2,𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1}) has the ideal intersection property. An application of Lemma 2.2(e) shows (𝒟2,𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1}) is weakly non-degenerate. By Proposition 2.2(e) and Lemma 2.4, (𝒟~2,𝒟~1)(\tilde{\mathcal{D}}_{2},\tilde{\mathcal{D}}_{1}) is a unital inclusion having the ideal intersection property, so [PittsZarikianUnPsExC*In, Corollary 3.22] shows (𝒟~2,𝒟~1)(\tilde{\mathcal{D}}_{2},\tilde{\mathcal{D}}_{1}) has the faithful unique pseudo-expectation property. Corollary 2.3 shows (𝒟2,𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1}) has the faithful unique pseudo-expectation property.

(\Leftarrow) Suppose (𝒟2,𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1}) has the faithful unique pseudo-expectation property. By Corollary 2.3, (𝒟2,𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1}) is weakly non-degenerate, whence (𝒟~2,𝒟~1)(\tilde{\mathcal{D}}_{2},\tilde{\mathcal{D}}_{1}) is a unital inclusion. Apply [PittsZarikianUnPsExC*In, Corollary 3.22] to see (𝒟~2,𝒟~1)(\tilde{\mathcal{D}}_{2},\tilde{\mathcal{D}}_{1}) has the ideal intersection property. Lemma 2.4 shows (𝒟2,𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1}) has the ideal intersection property.

Turning to the last statement, suppose (𝒟2,𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1}) has the faithful unique pseudo-expectation property with pseudo-expectation ι2\iota_{2}. Applying Corollary 2.3 and Lemma 2.2(e), we see that (𝒟~2,𝒟~1)(\tilde{\mathcal{D}}_{2},\tilde{\mathcal{D}}_{1}) is a unital inclusion. Corollary 2.3 shows ι~2:𝒟~2I(𝒟1)\tilde{\iota}_{2}:\tilde{\mathcal{D}}_{2}\rightarrow I({\mathcal{D}}_{1}) is the unique pseudo-expectation for (𝒟~2,𝒟~1)(\tilde{\mathcal{D}}_{2},\tilde{\mathcal{D}}_{1}) and by Theorem 2.3, ι~2\tilde{\iota}_{2} is faithful. By [PittsZarikianUnPsExC*In, Corollary 3.22], ι~2\tilde{\iota}_{2} is a *-monomorphism, hence so is ι2\iota_{2}.

Finally, since ι1(𝒟1)ι2(𝒟2)I(𝒟1)\iota_{1}({\mathcal{D}}_{1})\subseteq\iota_{2}({\mathcal{D}}_{2})\subseteq I({\mathcal{D}}_{1}), the minimality of injective envelopes shows (I(𝒟1),ι2)(I({\mathcal{D}}_{1}),\iota_{2}) is an injective envelope for 𝒟2{\mathcal{D}}_{2}. ∎

3. Cartan Envelopes

Up to this point, we have mostly considered general inclusions. We now restrict attention to inclusions where the subalgebra is abelian.

Standing Assumption \the\numberby. Unless explicitly stated otherwise, for the remainder of this work, whenever (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is an inclusion, we shall always assume 𝒟{\mathcal{D}} is abelian.

In [PittsStReInII], we defined the notion of a Cartan envelope for unital and regular inclusions and characterized when such inclusions have a Cartan envelope, see [PittsStReInII, Theorem 5.2]. The purpose of this section  is to extend the characterization of regular inclusions having a Cartan envelope from unital regular inclusions to all regular inclusions. This is accomplished in Theorem 3; it is among our main results. While the statement of Theorem 3 parallels that of [PittsStReInII, Theorem 5.2], discarding the hypothesis that (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is unital presents challenges which cannot be overcome by simply adjoining a unit.

We begin by recording a few useful facts about inclusions satisfying Assumption 3.

Lemma \the\numberby ([PittsNoApUnInC*Al, Corollary 2.2]).

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be an inclusion and fix a normalizer v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). Then vv𝒟¯\overline{vv^{*}{\mathcal{D}}} and vv𝒟¯\overline{v^{*}v{\mathcal{D}}} are ideals in 𝒟{\mathcal{D}} and the map vvdvdvvv^{*}d\mapsto v^{*}dv uniquely extends to a *-isomorphism θv:vv𝒟¯vv𝒟¯\theta_{v}:\overline{vv^{*}{\mathcal{D}}}\rightarrow\overline{v^{*}v{\mathcal{D}}}; moreover, for every hvv𝒟¯h\in\overline{vv^{*}{\mathcal{D}}}, vθv(h)=hvv\theta_{v}(h)=hv.

Our proofs of the statements in the next lemma depend on results from [PittsNoApUnInC*Al].

Lemma \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is an inclusion and let 𝒟c{\mathcal{D}}^{c} be the relative commutant of 𝒟{\mathcal{D}} in 𝒞{\mathcal{C}}. The following statements hold.

  1. (a)

    The identity mapping on 𝒞{\mathcal{C}} is a regular map from (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) into (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}), that is,

    𝒩(𝒞,𝒟)𝒩(𝒞,𝒟c).{\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}^{c}).
  2. (b)

    Suppose (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) is an inclusion and α:(𝒞,𝒟)(𝒞1,𝒟1)\alpha:({\mathcal{C}},{\mathcal{D}})\rightarrow({\mathcal{C}}_{1},{\mathcal{D}}_{1}) is a regular *-monomorphism. Then

    α(𝒟)𝒟1c.\alpha({\mathcal{D}})\subseteq{\mathcal{D}}_{1}^{c}.

    In addition, if (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) has the AUP, then α(𝒟)𝒟1\alpha({\mathcal{D}})\subseteq{\mathcal{D}}_{1}.

  3. (c)

    If (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP, then 𝒩(𝒞,𝒟)𝒩(𝒞~,𝒟~){\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\subseteq{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}).

  4. (d)

    Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular and weakly non-degenerate inclusion with 𝒞{\mathcal{C}} not unital. If 𝒩(𝒞,𝒟)𝒩(𝒞~,𝒟~){\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\subseteq{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}), then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP.

  5. (e)

    Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is regular and has the AUP. Then (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is regular.

Proof.

a) Let v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). Let d𝒟d\in{\mathcal{D}} and x𝒟cx\in{\mathcal{D}}^{c}. Using Lemma 3, we find that for every hvv𝒟¯h\in\overline{vv^{*}{\mathcal{D}}},

(vxv)dh=vxθv(dh)v=vθv(dh)xv=dh(vxv).(vxv^{*})dh=vx\theta_{v}(dh)v^{*}=v\theta_{v}(dh)xv^{*}=dh(vxv^{*}).

Taking h=uλh=u_{\lambda} where (uλ)(u_{\lambda}) is an approximate unit for vv𝒟¯\overline{vv^{*}{\mathcal{D}}}, and using the fact that vuλv0\left\lVert v^{*}u_{\lambda}-v^{*}\right\rVert\rightarrow 0, we obtain vxv𝒟cvxv^{*}\in{\mathcal{D}}^{c}. Likewise vxv𝒟cv^{*}xv\in{\mathcal{D}}^{c}, so (a) holds.

b) Let 0d𝒟0\leq d\in{\mathcal{D}}. Since d1/2𝒟𝒩(𝒞,𝒟)d^{1/2}\in{\mathcal{D}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), we have α(d1/2)𝒩(𝒞1,𝒟1)\alpha(d^{1/2})\in{\mathcal{N}}({\mathcal{C}}_{1},{\mathcal{D}}_{1}). For every h𝒟1h\in{\mathcal{D}}_{1}[PittsNoApUnInC*Al, Proposition 2.1] gives,

(3.2) α(d)h=α(d1/2)α(d1/2)h=hα(d)𝒟1.\alpha(d)h=\alpha(d^{1/2})^{*}\alpha(d^{1/2})h=h\alpha(d)\in{\mathcal{D}}_{1}.

Thus α(d)𝒟1c\alpha(d)\in{\mathcal{D}}_{1}^{c}. Since 𝒟{\mathcal{D}} is the linear span of its positive elements, α(𝒟)𝒟1c\alpha({\mathcal{D}})\subseteq{\mathcal{D}}_{1}^{c}.

Let (uλ)𝒟1(u_{\lambda})\subseteq{\mathcal{D}}_{1} be an approximate unit for 𝒞1{\mathcal{C}}_{1}. Replacing hh in (3.2) with uλu_{\lambda} and taking the limit along λ\lambda gives α(d)𝒟1\alpha(d)\in{\mathcal{D}}_{1}.

c) Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP. By Lemma 2.2(e), (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a unital inclusion. The conclusion is obvious when 𝒞{\mathcal{C}} is unital. So assume 𝒞{\mathcal{C}} is not unital and let (uλ)(u_{\lambda}) be a net in 𝒟{\mathcal{D}} which is an approximate unit for 𝒞{\mathcal{C}}. For v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}),  [PittsNoApUnInC*Al, Proposition 2.1] shows vvuλv^{*}vu_{\lambda} and vvuλvv^{*}u_{\lambda} belong to 𝒟{\mathcal{D}}. Taking the limit shows vvv^{*}v and vvvv^{*} are elements of 𝒟{\mathcal{D}}. Thus, if (d,λ)𝒟~(d,\lambda)\in\tilde{\mathcal{D}},

(3.3) (v,0)(d,λ)(v,0)=(vdv+λvv,0)and(v,0)(d,λ)(v,0)=(vdv+λvv,0).(v,0)^{*}(d,\lambda)(v,0)=(v^{*}dv+\lambda v^{*}v,0)\quad\text{and}\quad(v,0)(d,\lambda)(v,0)^{*}=(vdv^{*}+\lambda vv^{*},0).

This gives 𝒩(𝒞,𝒟)𝒩(𝒞~,𝒟~){\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\subseteq{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}).

d) Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is regular, weakly non-degenerate, and 𝒩(𝒞,𝒟)𝒩(𝒞~,𝒟~){\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\subseteq{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}). Since 𝒞{\mathcal{C}} is not unital, neither is 𝒟{\mathcal{D}}, so that (𝒞~,𝒟~)=(𝒞,𝒟)(\tilde{\mathcal{C}},\tilde{\mathcal{D}})=({\mathcal{C}}^{\dagger},{\mathcal{D}}^{\dagger}). The calculation in (3.3) with d=0d=0 and λ=1\lambda=1 shows that for every v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), vv𝒟v^{*}v\in{\mathcal{D}}. Then [PittsNoApUnInC*Al, Observation 1.3(2)] shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP.

e) Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is regular and has the AUP. There is nothing to do if 𝒞{\mathcal{C}} is unital, for then (𝒞~,𝒟~)=(𝒞,𝒟)(\tilde{\mathcal{C}},\tilde{\mathcal{D}})=({\mathcal{C}},{\mathcal{D}}). Assume then, that 𝒞{\mathcal{C}} is not unital. Part (c) shows

𝒞~={(0,λ):λ}+span¯{(v,0):v𝒩(𝒞,𝒟)},\tilde{\mathcal{C}}=\{(0,\lambda):\lambda\in{\mathbb{C}}\}+\overline{\operatorname{span}}\{(v,0):v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\},

so (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is regular. ∎

While the following corollary is immediate from parts (c) and (d) of Lemma 3, it is worth making explicit. In particular, u𝒞u_{\mathcal{C}} is not regular when 𝒟{\mathcal{D}} is an essential and proper ideal in the abelian CC^{*}-algebra 𝒞{\mathcal{C}}; this also shows that (c)⇏\not\Rightarrow(b) in Corollary 3.

Corollary \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular and weakly non-degenerate inclusion with 𝒞{\mathcal{C}} not unital. Consider the following statements.

  1. (a)

    (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP;

  2. (b)

    the inclusion mapping u𝒞:(𝒞,𝒟)(𝒞~,𝒟~)u_{\mathcal{C}}:({\mathcal{C}},{\mathcal{D}})\rightarrow(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a regular *-monomorphism;

  3. (c)

    (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a regular inclusion.

The following implications hold: (a)\Leftrightarrow(b)\Rightarrow(c).

Proof.

(a)\Leftrightarrow(b) Use 3(c) and 3(d).

(b)\Rightarrow(c) By hypothesis, {(v,0):v𝒩(𝒞,𝒟)}𝒩(𝒞~,𝒟~)\{(v,0):v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\}\subseteq{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}). As I𝒞~𝒩(𝒞~,𝒟~){\mathbb{C}}I_{\tilde{\mathcal{C}}}\subseteq{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}), it follows that span𝒩(𝒞~,𝒟~)\operatorname{span}{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is dense in 𝒞~\tilde{\mathcal{C}}.

Example \the\numberby.

Lemma 3(b) shows α(𝒟)\alpha({\mathcal{D}})\subseteq{\mathcal{B}} whenever (𝒜,)({\mathcal{A}},{\mathcal{B}}) has the AUP and α\alpha is a regular map. However, if (𝒜,)({\mathcal{A}},{\mathcal{B}}) does not have the AUP, this can fail. For an elementary example of this behavior, let

𝒞=𝒜=C0(),𝒟=={hC0():h(x)=0 for all x>1},{\mathcal{C}}={\mathcal{A}}=C_{0}({\mathbb{R}}),\quad{\mathcal{D}}={\mathcal{B}}=\{h\in C_{0}({\mathbb{R}}):h(x)=0\text{ for all }x>1\},

and define α:𝒞𝒜\alpha:{\mathcal{C}}\rightarrow{\mathcal{A}} by

α(h)(x)=h(x/2).\alpha(h)(x)=h(x/2).

Since 𝒩(𝒞,𝒟)=𝒩(𝒜,)=C0(){\mathcal{N}}({\mathcal{C}},{\mathcal{D}})={\mathcal{N}}({\mathcal{A}},{\mathcal{B}})=C_{0}({\mathbb{R}}), α\alpha is a regular map, but α(𝒟)\alpha({\mathcal{D}}) is not contained in {\mathcal{B}}.

The following extends parts of [PittsStReInI, Theorem 3.5] to include inclusions which may not be unital.

Theorem \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular MASA inclusion. Then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the unique pseudo-expectation property. Furthermore, (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property if and only if (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the ideal intersection property.

Proof.

Since (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular MASA inclusion, it has the AUP ([PittsNoApUnInC*Al, Theorem 2.5]), whence (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a unital inclusion (Lemma 2.2(e)), and regular by Lemma 3(e).

We claim that (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a MASA inclusion. If 𝒟{\mathcal{D}} is unital, then so is 𝒞{\mathcal{C}} by Lemma 2.2(a); thus in this case (𝒞~,𝒟~)=(𝒞,𝒟)(\tilde{\mathcal{C}},\tilde{\mathcal{D}})=({\mathcal{C}},{\mathcal{D}}) and all is well. Suppose then that 𝒟{\mathcal{D}} is not unital. Since 𝒟{\mathcal{D}} is a MASA, 𝒞{\mathcal{C}} cannot be unital. Thus (𝒞~,𝒟~)=(𝒞,𝒟)(\tilde{\mathcal{C}},\tilde{\mathcal{D}})=({\mathcal{C}}^{\dagger},{\mathcal{D}}^{\dagger}) and a routine argument shows 𝒟~\tilde{\mathcal{D}} is a MASA in 𝒞~\tilde{\mathcal{C}}.

We have established that (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a regular, unital, MASA inclusion. By  [PittsStReInI, Theorem 3.5] (𝒞~,𝒞~)(\tilde{\mathcal{C}},\tilde{\mathcal{C}}) has the unique pseudo-expectation property, so Corollary 2.3 shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) also has the unique pseudo-expectation property.

Turning to the second statement, let E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be the pseudo-expectation. Then E~\tilde{E} is the pseudo-expectation for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}).

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the ideal intersection property. Lemma 2.4 implies (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) has the ideal intersection property. By [PittsStReInI, Theorem 3.15], the left kernel :={x𝒞~:E~(xx)=0}{\mathcal{L}}:=\{x\in\tilde{\mathcal{C}}:\tilde{E}(x^{*}x)=0\} is an ideal of 𝒞~\tilde{\mathcal{C}} having trivial intersection with 𝒟~\tilde{\mathcal{D}}. Therefore, E~\tilde{E} is faithful, whence EE is faithful.

Now suppose EE is faithful. By Proposition 2.3, E~\tilde{E} is faithful. If J𝒞~J\unlhd\tilde{\mathcal{C}} has trivial intersection with 𝒟~\tilde{\mathcal{D}}, [PittsStReInI, Theorem 3.15] gives J={0}J\subseteq{\mathcal{L}}=\{0\}. So (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) has the ideal intersection property. By Lemma 2.4, (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the ideal intersection property, completing the proof. ∎

Definition \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is an inclusion and (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

  1. (a)

    We call (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) an essential expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) if (,𝒟,α|𝒟)({\mathcal{B}},{\mathcal{D}},\alpha|_{{\mathcal{D}}}) is an inclusion having the ideal intersection property.

  2. (b)

    If the *-monomorphism α\alpha is a regular map, we will say (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a regular expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

  3. (c)

    If (𝒜,)({\mathcal{A}},{\mathcal{B}}) is a Cartan pair, we say (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

Remarks \the\numberby.

  1. (a)

    It may seem odd that in Definition 3(a), we only require that (,𝒟,α|𝒟)({\mathcal{B}},{\mathcal{D}},\alpha|_{{\mathcal{D}}}) has the ideal intersection property instead of also placing that requirement on (𝒜,,α)({\mathcal{A}},{\mathcal{B}},\alpha). In the context of most interest to us, that is, when (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) and (𝒜,)({\mathcal{A}},{\mathcal{B}}) both have the faithful unique pseudo-expectation property, we will see in Observation 4.3 that this is automatic.

  2. (b)

    In some cases, regular maps produce expansions. For example, Lemma 3(b) shows that when α:(𝒞1,𝒟1)(𝒞2,𝒟2)\alpha:({\mathcal{C}}_{1},{\mathcal{D}}_{1})\rightarrow({\mathcal{C}}_{2},{\mathcal{D}}_{2}) is a regular *-monomorphism, and (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) is a Cartan inclusion, then (𝒞2,𝒟2 :: α)({\mathcal{C}}_{2},{\mathcal{D}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) is automatically a regular expansion of (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}).

The next two lemmas give some useful properties of essential expansions.

Lemma \the\numberby.

Suppose (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an essential expansion for the inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) and there is a faithful conditional expectation Δ:𝒜\Delta:{\mathcal{A}}\rightarrow{\mathcal{B}}. Then (𝒜,𝒟,α|𝒟)({\mathcal{A}},{\mathcal{D}},\alpha|_{\mathcal{D}}) and (𝒜,𝒞,α)({\mathcal{A}},{\mathcal{C}},\alpha) are weakly non-degenerate inclusions. In particular, if 𝒞{\mathcal{C}} is unital, then 𝒜{\mathcal{A}} is unital and α(I𝒞)=I𝒜\alpha(I_{\mathcal{C}})=I_{\mathcal{A}}.

Proof.

By hypothesis, (,𝒟,α|𝒟)({\mathcal{B}},{\mathcal{D}},\alpha|_{\mathcal{D}}) has the ideal intersection property, so it is weakly non-degenerate by Lemma 2.2(c). Suppose 0a𝒜0\leq a\in{\mathcal{A}} belongs to Ann(𝒜,𝒟,α|𝒟)\operatorname{Ann}({\mathcal{A}},{\mathcal{D}},\alpha|_{\mathcal{D}}). Then for every d𝒟d\in{\mathcal{D}},

0=Δ(α(d)a)=α(d)Δ(a)=Δ(a)α(d)=Δ(aα(d)).0=\Delta(\alpha(d)a)=\alpha(d)\Delta(a)=\Delta(a)\alpha(d)=\Delta(a\alpha(d)).

Thus Δ(a)Ann(,𝒟,α|𝒟)\Delta(a)\in\operatorname{Ann}({\mathcal{B}},{\mathcal{D}},\alpha|_{\mathcal{D}}). Faithfulness of Δ\Delta gives a=0a=0. As Ann(𝒜,𝒟,α)\operatorname{Ann}({\mathcal{A}},{\mathcal{D}},\alpha) is the span of its postive elements, we obtain Ann(𝒜,𝒟,α)={0}\operatorname{Ann}({\mathcal{A}},{\mathcal{D}},\alpha)=\{0\}, so (𝒜,𝒟,α|𝒟)({\mathcal{A}},{\mathcal{D}},\alpha|_{\mathcal{D}}) is weakly non-degenerate. That (𝒜,𝒞,α)({\mathcal{A}},{\mathcal{C}},\alpha) is weakly non-degenerate follows from the fact that Ann(𝒜,𝒞,α)Ann(𝒜,𝒟,α|𝒟)\operatorname{Ann}({\mathcal{A}},{\mathcal{C}},\alpha)\subseteq\operatorname{Ann}({\mathcal{A}},{\mathcal{D}},\alpha|_{\mathcal{D}}). If 𝒞{\mathcal{C}} is unital, Lemma 2.2(a) shows 𝒜{\mathcal{A}} is unital and α(I𝒞)=I𝒜\alpha(I_{\mathcal{C}})=I_{\mathcal{A}}. ∎

Lemma \the\numberby.

Let (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) be an inclusion. Suppose (𝒞2,𝒟2 :: α)({\mathcal{C}}_{2},{\mathcal{D}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an essential expansion of (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) has the faithful unique pseudo-expectation property. Then

  1. (a)

    (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) has the faithful unique pseudo-expectation property; and

  2. (b)

    α(𝒟1c)𝒟2c\alpha({\mathcal{D}}_{1}^{c})\subseteq{\mathcal{D}}_{2}^{c}.

Proof.

Before giving the proofs of (a) and (b), we make a few remarks. Let (I(𝒟1),ι1)(I({\mathcal{D}}_{1}),\iota_{1}) be an injective envelope for 𝒟1{\mathcal{D}}_{1}. By Proposition 2.4, (𝒟2,𝒟1,α|𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1},\alpha|_{{\mathcal{D}}_{1}}) has the faithful unique pseudo-expectation property; let ι2:𝒟2I(𝒟1)\iota_{2}:{\mathcal{D}}_{2}\rightarrow I({\mathcal{D}}_{1}) be the pseudo-expectation for (𝒟2,𝒟1,α|𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1},\alpha|_{{\mathcal{D}}_{1}}) relative to (I(𝒟1),ι1)(I({\mathcal{D}}_{1}),\iota_{1}). Then (I(𝒟1),ι2)(I({\mathcal{D}}_{1}),\iota_{2}) is an injective envelope for 𝒟2{\mathcal{D}}_{2} (Proposition 2.4). Let

E2:𝒞2I(𝒟1)E_{2}:{\mathcal{C}}_{2}\rightarrow I({\mathcal{D}}_{1})

be the pseudo-expectation for (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) relative to (I(𝒟1),ι2)(I({\mathcal{D}}_{1}),\iota_{2}).

(a) Let us first show (𝒞2,𝒟1,α|𝒟1)({\mathcal{C}}_{2},{\mathcal{D}}_{1},\alpha|_{{\mathcal{D}}_{1}}) has the faithful unique pseudo-expectation property. Suppose F:𝒞2I(𝒟1)F:{\mathcal{C}}_{2}\rightarrow I({\mathcal{D}}_{1}) is a pseudo-expectation for (𝒞2,𝒟1,α|𝒟1)({\mathcal{C}}_{2},{\mathcal{D}}_{1},\alpha|_{{\mathcal{D}}_{1}}) relative to (I(𝒟1),ι1)(I({\mathcal{D}}_{1}),\iota_{1}). Then for d𝒟1d\in{\mathcal{D}}_{1}, F(α(d))=ι1(d)=ι2(α(d))F(\alpha(d))=\iota_{1}(d)=\iota_{2}(\alpha(d)). Thus F|𝒟2F|_{{\mathcal{D}}_{2}} is a pseudo-expectation for (𝒟2,𝒟1,α|𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1},\alpha|_{{\mathcal{D}}_{1}}) relative to (I(𝒟1),ι1)(I({\mathcal{D}}_{1}),\iota_{1}). Therefore F|𝒟2=ι2F|_{{\mathcal{D}}_{2}}=\iota_{2}, that is, FF is a pseudo-expectation for (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) relative to (I(𝒟1),ι2)(I({\mathcal{D}}_{1}),\iota_{2}). This forces F=E2F=E_{2}, showing (𝒞2,𝒟1,α|𝒟1)({\mathcal{C}}_{2},{\mathcal{D}}_{1},\alpha|_{{\mathcal{D}}_{1}}) has the faithful unique pseudo-expectation property.

Corollary 2.3 applied to (𝒞2,𝒟1,α|𝒟1)({\mathcal{C}}_{2},{\mathcal{D}}_{1},\alpha|_{{\mathcal{D}}_{1}}) with =α(𝒞1){\mathcal{B}}=\alpha({\mathcal{C}}_{1}) implies (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) has the faithful unique pseudo-expectation property (the pseudo-expectation is E2αE_{2}\circ\alpha).

(b) By Proposition 2.3, rCom(𝒞2,α(𝒟1))\operatorname{\textsc{rCom}}({\mathcal{C}}_{2},\alpha({\mathcal{D}}_{1})) is abelian. As α(𝒟1)𝒟2\alpha({\mathcal{D}}_{1})\subseteq{\mathcal{D}}_{2}, we find

rCom(𝒞2,𝒟2)rCom(𝒞2,α(𝒟1)).\operatorname{\textsc{rCom}}({\mathcal{C}}_{2},{\mathcal{D}}_{2})\subseteq\operatorname{\textsc{rCom}}({\mathcal{C}}_{2},\alpha({\mathcal{D}}_{1})).

Since (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) has the faithful unique pseudo-expectation property, rCom(𝒞2,𝒟2)\operatorname{\textsc{rCom}}({\mathcal{C}}_{2},{\mathcal{D}}_{2}) is a MASA in 𝒞2{\mathcal{C}}_{2}, so rCom(𝒞2,𝒟2)=rCom(𝒞2,α(𝒟1))\operatorname{\textsc{rCom}}({\mathcal{C}}_{2},{\mathcal{D}}_{2})=\operatorname{\textsc{rCom}}({\mathcal{C}}_{2},\alpha({\mathcal{D}}_{1})). As

α(𝒟1c)rCom(𝒞2,α(𝒟1))=𝒟2c,\alpha({\mathcal{D}}_{1}^{c})\subseteq\operatorname{\textsc{rCom}}({\mathcal{C}}_{2},\alpha({\mathcal{D}}_{1}))={\mathcal{D}}_{2}^{c},

part (b) holds. ∎

The definition of Cartan envelope given in [PittsStReInII] extends to any regular inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Here is the definition, along with the definitions of other useful expansions.

Definition \the\numberby.

(cf. [PittsStReInII, Definition 5.1]) Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular inclusion.

  1. (a)

    A package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular expansion (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) such that

    1. (i)

      there is a faithful conditional expectation Δ:𝒜\Delta:{\mathcal{A}}\rightarrow{\mathcal{B}};

    2. (ii)

      (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) generates (𝒜,)({\mathcal{A}},{\mathcal{B}}) in the sense that =C(Δ(α(𝒞))){\mathcal{B}}=C^{*}(\Delta(\alpha({\mathcal{C}}))) and 𝒜=C(α(𝒞)){\mathcal{A}}=C^{*}(\alpha({\mathcal{C}})\cup{\mathcal{B}}).

    If in addition, (𝒜,)({\mathcal{A}},{\mathcal{B}}) is a Cartan inclusion, we say (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

  2. (b)

    If (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) which is also an essential expansion for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) (see Definition 3(a)), then we call (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) an envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

  3. (c)

    We say (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) if it is an envelope and a Cartan package.

We now show that when (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an essential and regular Cartan expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), the inclusion generated by (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a Cartan envelope.

Proposition \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular inclusion and (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an essential, regular, and Cartan expansion for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Let Δ:𝒜\Delta:{\mathcal{A}}\rightarrow{\mathcal{B}} be the conditional expectation and put

1:=C(Δ(α(𝒞)))and𝒜1:=C(α(𝒞)1).{\mathcal{B}}_{1}:=C^{*}(\Delta(\alpha({\mathcal{C}})))\quad\text{and}\quad{\mathcal{A}}_{1}:=C^{*}(\alpha({\mathcal{C}})\cup{\mathcal{B}}_{1}).

Then (𝒜1,1 :: α)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) and Δ|𝒜1:𝒜11\Delta|_{{\mathcal{A}}_{1}}:{\mathcal{A}}_{1}\rightarrow{\mathcal{B}}_{1} is the conditional expectation.

Proof.

By construction, α(𝒞)𝒜1\alpha({\mathcal{C}})\subseteq{\mathcal{A}}_{1} and α(𝒟)1\alpha({\mathcal{D}})\subseteq{\mathcal{B}}_{1}, so (𝒜1,1 :: α)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). That Δ|𝒜1:𝒜11\Delta|_{{\mathcal{A}}_{1}}:{\mathcal{A}}_{1}\rightarrow{\mathcal{B}}_{1} is a faithful conditional expectation follows from faithfulness of Δ\Delta, the definitions of 𝒜1{\mathcal{A}}_{1} and 1{\mathcal{B}}_{1}, and the fact that 1{\mathcal{B}}_{1}\subseteq{\mathcal{B}}.

Claim \the\numberby.

α:(𝒞,𝒟)(𝒜1,1)\alpha:({\mathcal{C}},{\mathcal{D}})\rightarrow({\mathcal{A}}_{1},{\mathcal{B}}_{1}) is a regular map.

Proof of Claim 3. For v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), we must show that α(v)\alpha(v) normalizes 1{\mathcal{B}}_{1}. As (𝒜,)({\mathcal{A}},{\mathcal{B}}) is a Cartan inclusion, it has the faithful unique pseudo-expectation property by Proposition 2.3. Lemma 3 shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property and

(3.5) α(𝒟c)c=.\alpha({\mathcal{D}}^{c})\subseteq{\mathcal{B}}^{c}={\mathcal{B}}.

An argument similar to that used for establishing [PittsStReInII, (5.14)] shows that for nn\in{\mathbb{N}}, and every collection of nn elements {xj}j=1n𝒞\{x_{j}\}_{j=1}^{n}\subseteq{\mathcal{C}},

(3.6) α(v)(j=1nΔ(α(xj)))α(v)1.\alpha(v)^{*}\left(\prod_{j=1}^{n}\Delta(\alpha(x_{j}))\right)\alpha(v)\in{\mathcal{B}}_{1}.

However, the proofs of (3.6) and [PittsStReInII, (5.14)] have some technical differences, so we include the proof of (3.6) here. The argument is by induction. When n=1n=1, the invariance of Δ\Delta under 𝒩(𝒜,){\mathcal{N}}({\mathcal{A}},{\mathcal{B}}) (see (2.1.10)) gives,

α(v)Δ(α(x1))α(v)=Δ(α(v)α(x1)α(v))Δ(α(𝒞))1.\alpha(v)^{*}\Delta(\alpha(x_{1}))\alpha(v)=\Delta(\alpha(v^{*})\alpha(x_{1})\alpha(v))\in\Delta(\alpha({\mathcal{C}}))\subseteq{\mathcal{B}}_{1}.

Suppose now that (3.6) holds for some nn\in{\mathbb{N}} and every collection of nn elements of 𝒞{\mathcal{C}}. Let {xj}j=1n+1𝒞\{x_{j}\}_{j=1}^{n+1}\subseteq{\mathcal{C}} and set y=j=1nΔ(α(xj))y=\prod_{j=1}^{n}\Delta(\alpha(x_{j})). For hvv𝒟c¯h\in\overline{vv^{*}{\mathcal{D}}^{c}}, Lemma 3(b) and Lemma 3 show α(h)α(vv𝒟c¯)α(vv)¯\alpha(h)\in\alpha(\overline{vv^{*}{\mathcal{D}}^{c}})\subseteq\overline{\alpha(vv^{*}){\mathcal{B}}}; the induction hypothesis gives α(v)yα(v)1\alpha(v^{*})y\alpha(v)\in{\mathcal{B}}_{1}. Given h0vv𝒟ch_{0}\in vv^{*}{\mathcal{D}}^{c}, write h0=vvkh_{0}=vv^{*}k where k𝒟ck\in{\mathcal{D}}^{c}. Using Lemma 3,

α(v)yΔ(α(xn+1))α(h0)α(v)\displaystyle\alpha(v^{*})\>y\Delta(\alpha(x_{n+1}))\alpha(h_{0})\>\alpha(v) =α(v)yα(v)θα(v)(Δ(α(xn+1))α(h0))\displaystyle=\alpha(v^{*})y\alpha(v)\>\theta_{\alpha(v)}(\Delta(\alpha(x_{n+1}))\alpha(h_{0}))
=α(v)yα(v)θα(v)(α(vv)Δ(α(xn+1k)))\displaystyle=\alpha(v^{*})y\alpha(v)\>\theta_{\alpha(v)}(\alpha(vv^{*})\Delta(\alpha(x_{n+1}k)))
=(α(v)yα(v))(α(v)Δ(α(xn+1k))α(v))\displaystyle=(\alpha(v^{*})y\alpha(v))\>\left(\alpha(v)^{*}\Delta(\alpha(x_{n+1}k))\alpha(v)\right)
=(α(v)yα(v))Δ(α(v)α(xn+1k)α(v))1.\displaystyle=(\alpha(v^{*})y\alpha(v))\>\Delta(\alpha(v)^{*}\alpha(x_{n+1}k)\alpha(v))\in{\mathcal{B}}_{1}.

Therefore, for every hvv𝒟c¯h\in\overline{vv^{*}{\mathcal{D}}^{c}},

(3.7) α(v)yΔ(α(xn+1))α(h)α(v)1.\alpha(v^{*})\>y\Delta(\alpha(x_{n+1}))\alpha(h)\>\alpha(v)\in{\mathcal{B}}_{1}.

Note that v=limn(vv)1/nvv=\lim_{n\rightarrow\infty}(vv^{*})^{1/n}v. Thus for any d𝒟cd\in{\mathcal{D}}^{c}, choosing h=d(vv)1/nvv𝒟c¯h=d(vv^{*})^{1/n}\in\overline{vv^{*}{\mathcal{D}}^{c}} in (3.7), we find

(3.8) α(v)yΔ(α(xn+1))α(d)α(v)=limnα(v)yΔ(α(xn+1))α(d(vv)1/n)α(v)1.\alpha(v)^{*}y\Delta(\alpha(x_{n+1}))\alpha(d)\alpha(v)=\lim_{n}\alpha(v)^{*}y\Delta(\alpha(x_{n+1}))\alpha(d(vv^{*})^{1/n})\alpha(v)\in{\mathcal{B}}_{1}.

As 𝒟c{\mathcal{D}}^{c} is abelian (by (3.5)), (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) is a MASA inclusion. Lemma 3(a) shows that (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) is a regular MASA inclusion, and therefore has the AUP by [PittsNoApUnInC*Al, Theorem 2.5]. Let (uλ)𝒟c(u_{\lambda})\subseteq{\mathcal{D}}^{c} be an approximate unit for 𝒞{\mathcal{C}}. Taking d=uλd=u_{\lambda} in (3.8), we conclude

α(v)yΔ(α(xn+1))α(v)\displaystyle\alpha(v)^{*}y\Delta(\alpha(x_{n+1}))\alpha(v) =limλα(v)yΔ(α(xn+1uλ))α(v)\displaystyle=\lim_{\lambda}\alpha(v)^{*}y\Delta(\alpha(x_{n+1}u_{\lambda}))\alpha(v)
=limλα(v)yΔ(α(xn+1))α(uλ)α(v)1.\displaystyle=\lim_{\lambda}\alpha(v)^{*}y\Delta(\alpha(x_{n+1}))\alpha(u_{\lambda})\alpha(v)\in{\mathcal{B}}_{1}.

Thus (3.6) holds.

Since 1{\mathcal{B}}_{1} is generated by Δ(α(𝒞))\Delta(\alpha({\mathcal{C}})), (3.6) implies α(v)1α(v)1\alpha(v)^{*}{\mathcal{B}}_{1}\alpha(v)\subseteq{\mathcal{B}}_{1}. As this holds for every v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), we may replace vv with vv^{*} to obtain α(v)1α(v)1\alpha(v){\mathcal{B}}_{1}\alpha(v)^{*}\subseteq{\mathcal{B}}_{1}. Therefore, α(v)𝒩(𝒜1,1)\alpha(v)\in{\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}). \diamondsuit

By Claim 3, α(𝒩(𝒞,𝒟))1𝒩(𝒜1,1)\alpha({\mathcal{N}}({\mathcal{C}},{\mathcal{D}}))\cup{\mathcal{B}}_{1}\subseteq{\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}), whence (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}) is a regular inclusion. Also,

α(𝒟)1.\alpha({\mathcal{D}})\subseteq{\mathcal{B}}_{1}\subseteq{\mathcal{B}}.

Applying Lemma 2.4, (,1)({\mathcal{B}},{\mathcal{B}}_{1}) and (1,𝒟,α|𝒟)({\mathcal{B}}_{1},{\mathcal{D}},\alpha|_{\mathcal{D}}) have the ideal intersection property. Thus, (𝒜, :: )({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\,\subseteq) is an essential expansion of (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}). By Lemma 3, (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}) has the faithful unique pseudo-expectation property.

Proposition 2.3 shows (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}) is a Cartan inclusion. Therefore, (𝒜1,1 :: α)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). As (1,𝒟,α|𝒟)({\mathcal{B}}_{1},{\mathcal{D}},\alpha|_{\mathcal{D}}) has the ideal intersection property, (𝒜1,1 :: α)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha) is also an envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Thus (𝒜1,1 :: α)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). ∎

The following will be used in the proof of Theorem 3; we will establish its converse in Proposition 4.2.

Proposition \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular inclusion such that (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) has the ideal intersection property and 𝒟c{\mathcal{D}}^{c} is abelian. If (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}), then (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is also a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

Proof.

Let Δ:𝒜\Delta:{\mathcal{A}}\rightarrow{\mathcal{B}} be the conditional expectation.

Since 𝒟𝒟c{\mathcal{D}}\subseteq{\mathcal{D}}^{c}, (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Lemma 3(a) and the regularity of α:(𝒞,𝒟c)(𝒜,)\alpha:({\mathcal{C}},{\mathcal{D}}^{c})\rightarrow({\mathcal{A}},{\mathcal{B}}) show α:(𝒞,𝒟)(𝒜,)\alpha:({\mathcal{C}},{\mathcal{D}})\rightarrow({\mathcal{A}},{\mathcal{B}}) is a regular mapping. Since (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}), =C(Δ(α(𝒞))){\mathcal{B}}=C^{*}(\Delta(\alpha({\mathcal{C}}))) and 𝒜=C(α(𝒞)){\mathcal{A}}=C^{*}({\mathcal{B}}\cup\alpha({\mathcal{C}})). Thus, (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

We have α(𝒟)α(𝒟c)\alpha({\mathcal{D}})\subseteq\alpha({\mathcal{D}}^{c})\subseteq{\mathcal{B}}. By definition of Cartan envelope, (,𝒟c,α)({\mathcal{B}},{\mathcal{D}}^{c},\alpha) has the ideal intersection property. Since (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) has the ideal intersection property, so does (,𝒟,α)({\mathcal{B}},{\mathcal{D}},\alpha) (Lemma 2.4). Hence (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). ∎

We turn to some technical preparations for the proof of the uniqueness statement for Cartan envelopes found in Theorem 3 below. We will use Lemma 3 repeatedly, often without comment. Proposition 3 shows how to produce useful *-semigroups of intertwiners from *-semigroups of normalizers; these semigroups are key to the uniqueness statement in Theorem 3. Lemma 3 exhibits some intertwiners associated to a normalizer and is the foundation for the proof of Proposition 3. Recall from Definition 2.1(e) that (𝒞,𝒟){\mathcal{I}}({\mathcal{C}},{\mathcal{D}}) denotes the collection of all intertwiners for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

When (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is an inclusion and v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), it is convenient to use the notation,

(3.9) R(v):=vv𝒟¯andS(v):=vv𝒟¯.R(v):=\overline{vv^{*}{\mathcal{D}}}\quad\text{and}\quad S(v):=\overline{v^{*}v{\mathcal{D}}}.

For an abelian CC^{*}-algebra 𝒟{\mathcal{D}} and an ideal J𝒟J\unlhd{\mathcal{D}}, let

(3.10) Jc:={dJ: there exists hJ such that dh=d}.J_{c}:=\{d\in J:\text{ there exists }h\in J\text{ such that }dh=d\}.
Observation \the\numberby.

JcJ_{c} is an algebraic ideal of 𝒟{\mathcal{D}} whose closure is JJ. (In general, JcJ_{c} is not closed.)

Sketch of Proof.

For k1,k2𝒟k_{1},k_{2}\in{\mathcal{D}} with ki0k_{i}\geq 0, we let k1k2k_{1}\vee k_{2} be the element of 𝒟{\mathcal{D}} whose Gelfand transform is 𝒟^σmax{σ(k1),σ(k2)}\hat{\mathcal{D}}\ni\sigma\mapsto\max\{\sigma(k_{1}),\sigma(k_{2})\}. Also, for h𝒟h\in{\mathcal{D}}, let hh^{\prime} be the element of 𝒟{\mathcal{D}} whose Gelfand transform is the function

𝒟^σmin{1,|σ(h)|}.\hat{\mathcal{D}}\ni\sigma\mapsto\min\{1,|\sigma(h)|\}.

Note that if dd and hh belong to JJ, then dh=ddh=d implies dh=ddh^{\prime}=d; observe that hJh^{\prime}\in J because h=f(h)h^{\prime}=f(h), where f(z)=min{1,|z|}f(z)=\min\{1,|z|\}. Thus if d1,d2Jcd_{1},d_{2}\in J_{c} and h1,h2Jh_{1},h_{2}\in J satisfy dihi=did_{i}h_{i}=d_{i}, then (d1+d2)(h1h2)=d1+d2(d_{1}+d_{2})(h_{1}^{\prime}\vee h_{2}^{\prime})=d_{1}+d_{2}, so JcJ_{c} is closed under addition. That JcJ_{c} is an ideal in 𝒟{\mathcal{D}} is now obvious. The fact that JcJ_{c} is dense in JJ will follow from Lemma 3(a) below. Indeed, for dJd\in J, R(d)JR(d)\subseteq J and R(d)cJcR(d)_{c}\subseteq J_{c}. Since dR(d)d\in R(d), Lemma 3(a) (whose proof is independent of Observation 3) gives dR(d)c¯J¯cd\in\overline{R(d)_{c}}\subseteq\overline{J}_{c}, so JJc¯J\subseteq\overline{J_{c}}. ∎

Lemma \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is an inclusion. Let v,w𝒩(𝒞,𝒟)v,w\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}).

  1. (a)

    Suppose vv𝒟vv^{*}\in{\mathcal{D}} (resp. vv𝒟v^{*}v\in{\mathcal{D}}). Then given ε>0\varepsilon>0, there exists kR(v)ck\in R(v)_{c}, (resp. kS(v)ck\in S(v)_{c}) such that vkv<ε\left\lVert v-kv\right\rVert<\varepsilon (resp. vvk<ε\left\lVert v-vk\right\rVert<\varepsilon).

  2. (b)

    If kR(v)ck\in R(v)_{c} (resp. kS(v)ck\in S(v)_{c}), then there exists sR(v)cs\in R(v)_{c} (resp. sS(v)cs\in S(v)_{c}) such that k=vvsk=vv^{*}s (resp. k=vvsk=v^{*}vs).

  3. (c)

    If kR(v)ck\in R(v)_{c} (resp. kS(v)ck\in S(v)_{c}), then kv(𝒞,𝒟)kv\in{\mathcal{I}}({\mathcal{C}},{\mathcal{D}}) (resp. vk(𝒞,𝒟)vk\in{\mathcal{I}}({\mathcal{C}},{\mathcal{D}})).

  4. (d)

    For hR(v)ch\in R(v)_{c} (resp. hS(v)ch\in S(v)_{c}), θv(h)R(v)c\theta_{v}(h^{*})\in R(v^{*})_{c} (resp. θv1(h)S(v)c\theta_{v}^{-1}(h^{*})\in S(v^{*})_{c}).

  5. (e)

    Suppose hR(v)ch\in R(v)_{c} and kR(w)ck\in R(w)_{c} (resp. hS(v)ch\in S(v)_{c} and kS(w)c)k\in S(w)_{c}). Then θv1(θv(h)k)R(vw)c\theta_{v}^{-1}(\theta_{v}(h)k)\in R(vw)_{c} (resp. θw(hθw1(k))S(vw)c\theta_{w}(h\theta_{w}^{-1}(k))\in S(vw)_{c}).

Proof.

We give the proofs for vvvv^{*}; the proofs for vvv^{*}v are similar.

(a) Suppose vv𝒟vv^{*}\in{\mathcal{D}} and let ε>0\varepsilon>0. Since vv^C0(𝒟^)\widehat{vv^{*}}\in C_{0}(\hat{\mathcal{D}}), there exists a compact set C𝒟^C\subseteq\hat{\mathcal{D}} such that σ(vv)<ε2/4\sigma(vv^{*})<\varepsilon^{2}/4 whenever σ𝒟^C\sigma\in\hat{\mathcal{D}}\setminus C. Let

K:=C{σ𝒟^:σ(vv)ε2/4}.K:=C\cap\{\sigma\in\hat{\mathcal{D}}:\sigma(vv^{*})\geq\varepsilon^{2}/4\}.

Then KK is a compact subset of 𝒟^\hat{\mathcal{D}}. By local compactness of 𝒟^\hat{\mathcal{D}}, we may find open sets U,V𝒟^U,V\subseteq\hat{\mathcal{D}} such that:

  • U¯\overline{U} and V¯\overline{V} are compact;

  • KUU¯{σ𝒟^:σ(vv)ε2/9}K\subseteq U\subseteq\overline{U}\subseteq\{\sigma\in\hat{\mathcal{D}}:\sigma(vv^{*})\geq\varepsilon^{2}/9\}; and

  • U¯VV¯{σ𝒟^:σ(vv)ε2/16}\overline{U}\subseteq V\subseteq\overline{V}\subseteq\{\sigma\in\hat{\mathcal{D}}:\sigma(vv^{*})\geq\varepsilon^{2}/16\}.

Urysohn’s lemma ensures there exists h,k{d𝒟:d0 and d1}h,k\in\{d\in{\mathcal{D}}:d\geq 0\text{ and }\left\lVert d\right\rVert\leq 1\} such that k^1\hat{k}\equiv 1 on KK, k^\hat{k} vanishes off UU, h^1\hat{h}\equiv 1 on U¯\overline{U}, and h^\hat{h} vanishes off VV. Then h,kR(v)h,k\in R(v) and hk=khk=k, so that kR(v)ck\in R(v)_{c}.

Note that for σ𝒟^K\sigma\in\hat{\mathcal{D}}\setminus K, σ(vv)<ε2/4\sigma(vv^{*})<\varepsilon^{2}/4. Also, (kvv)(kvv)=k2vv2kvv+vv𝒟(kv-v)(kv-v)^{*}=k^{2}vv^{*}-2kvv^{*}+vv^{*}\in{\mathcal{D}}, so

kvv2=supσ𝒟^σ((kvv)(kvv))=supσ𝒟^σ(vv)(σ(k)1)2ε2/4<ε2.\left\lVert kv-v\right\rVert^{2}=\sup_{\sigma\in\hat{\mathcal{D}}}\sigma(({kv-v})(kv-v)^{*})=\sup_{\sigma\in\hat{\mathcal{D}}}\sigma(vv^{*})(\sigma(k)-1)^{2}\leq\varepsilon^{2}/4<\varepsilon^{2}.

(b) Let

range(v)={σ𝒟^:σ(vv)0}.\operatorname{range}(v)=\{\sigma\in\hat{\mathcal{D}}:\sigma(vv^{*})\neq 0\}.

Then

R(v)={d𝒟:d^ vanishes off range(v)}C0(range(v)).R(v)=\{d\in{\mathcal{D}}:\hat{d}\text{ vanishes off }\operatorname{range}(v)\}\cong C_{0}(\operatorname{range}(v)).

As kR(v)ck\in R(v)_{c}, we may find hR(v)h\in R(v) so that kh=kkh=k. We may therefore find a compact set Frange(v)F\subseteq\operatorname{range}(v) such that |σ(h)|<1/2|\sigma(h)|<1/2 for σ𝒟^F\sigma\in\hat{\mathcal{D}}\setminus F. As kh=kkh=k, k^\hat{k} vanishes off FF. Define a function δ\delta on 𝒟^\hat{\mathcal{D}} by

δ(σ)={σ(k)σ(vv)for σF;0for σ𝒟^F.\delta(\sigma)=\begin{cases}\displaystyle\frac{\sigma(k)}{\sigma(vv^{*})}&\text{for }\sigma\in F;\\ 0&\text{for }\sigma\in\hat{\mathcal{D}}\setminus F.\end{cases}

Since vv^\widehat{vv^{*}} is bounded away from 0 on FF, δC0(𝒟^)\delta\in C_{0}(\hat{\mathcal{D}}), and we define s𝒟s\in{\mathcal{D}} by s^=δ\hat{s}=\delta. By construction, k=vvsk=vv^{*}s, and since s^h^=s^\hat{s}\hat{h}=\hat{s}, we find sR(v)cs\in R(v)_{c}.

(c) Let kR(v)ck\in R(v)_{c} and choose hR(v)=vv𝒟¯h\in R(v)=\overline{vv^{*}{\mathcal{D}}} such that kh=kkh=k. Set u:=kvu:=kv. Then for d𝒟d\in{\mathcal{D}} we have

du=k(dhv)=kvθv(dh)=uθv(dh).du=k(dhv)=kv\theta_{v}(dh)=u\theta_{v}(dh).

Thus, 𝒟uu𝒟{\mathcal{D}}u\subseteq u{\mathcal{D}}.

For the reverse inclusion, note that θv(kh)=θv(k)θv(h)=θv(k)\theta_{v}(kh)=\theta_{v}(k)\theta_{v}(h)=\theta_{v}(k), so that θv(k)S(v)c\theta_{v}(k)\in S(v)_{c}. Thus, for d𝒟d\in{\mathcal{D}} we have

ud=kv(θv(h)d)=θv1(θv(h)d)kv𝒟u.ud=kv\,(\theta_{v}(h)d)=\theta_{v}^{-1}(\theta_{v}(h)d)kv\in{\mathcal{D}}u.

Therefore kv(𝒞,𝒟)kv\in{\mathcal{I}}({\mathcal{C}},{\mathcal{D}}), as claimed.

(d) By part (b), we may write h=vvsh=vv^{*}s for sR(v)cs\in R(v)_{c}. Then θv(h)=vsvS(v)c=R(v)c\theta_{v}(h^{*})=v^{*}s^{*}v\in S(v)_{c}=R(v^{*})_{c}.

(e) Using part (b) we may choose sR(v)cs\in R(v)_{c}, tR(w)ct\in R(w)_{c} so that h=(vv)2sh=(vv^{*})^{2}s and k=wwtk=ww^{*}t. Then

θv1(θv(h)k)\displaystyle\theta_{v}^{-1}(\theta_{v}(h)k) =θv1(θv((vv)2s)k)=θv1(v(vvs)vk)\displaystyle=\theta_{v}^{-1}(\theta_{v}((vv^{*})^{2}s)k)=\theta_{v}^{-1}(v^{*}(vv^{*}s)vk)
=vvsvkv=(vv)sv(tww)v\displaystyle=vv^{*}svkv^{*}=(vv^{*})sv(tww^{*})v^{*}
=s(vv)vtwwv\displaystyle=s(vv^{*})vtww^{*}v^{*}
=s(vtv)(vw)(vw)(vw)(vw)𝒟R(vw).\displaystyle=s\,(vtv^{*})(vw)(vw)^{*}\in(vw)(vw)^{*}{\mathcal{D}}\subseteq R(vw).

By definition, we may find xR(v)x\in R(v) and yR(w)y\in R(w) so that h=hxh=hx and k=kyk=ky. Since R(v)c¯=R(v)\overline{R(v)_{c}}=R(v) and R(w)c¯=R(w)\overline{R(w)_{c}}=R(w) (see Observation 3), we may find sequences (xk)(x_{k}) in R(v)cR(v)_{c} and (yk)(y_{k}) in R(w)cR(w)_{c} which converge to xx and yy respectively. The previous calculations applied to xkx_{k} and yky_{k} show θv1(θv(xk)yk)R(vw)\theta_{v}^{-1}(\theta_{v}(x_{k})y_{k})\in R(vw). Taking limits yields θv1(θv(x)y)R(vw)\theta_{v}^{-1}(\theta_{v}(x)y)\in R(vw). But then

θv1(θv(h)k)θv1(θv(x)y)=θv1(θv(h)k),\theta_{v}^{-1}(\theta_{v}(h)k)\theta_{v}^{-1}(\theta_{v}(x)y)=\theta_{v}^{-1}(\theta_{v}(h)k),

showing that θv1(θv(h)k)R(vw)c\theta_{v}^{-1}(\theta_{v}(h)k)\in R(vw)_{c}. ∎

Proposition \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be an inclusion. Let 𝒫𝒩(𝒞,𝒟){\mathcal{P}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) be a *-semigroup and put

𝒬:={vh:v𝒫 and hS(v)c}and𝒮:=𝒬𝒟.{\mathcal{Q}}:=\{vh:v\in{\mathcal{P}}\text{ and }h\in S(v)_{c}\}\quad\text{and}\quad{\mathcal{S}}:={\mathcal{Q}}\cup{\mathcal{D}}.

The following statements hold.

  1. (a)

    𝒬{\mathcal{Q}} and 𝒮{\mathcal{S}} are *-semigroups and both are contained in 𝒩(𝒞,𝒟)(𝒞,𝒟){\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\cap{\mathcal{I}}({\mathcal{C}},{\mathcal{D}}).

  2. (b)

    When (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the approximate unit property and the *-algebra generated by 𝒫𝒟{\mathcal{P}}\cup{\mathcal{D}} is dense in 𝒞{\mathcal{C}}, span𝒮\operatorname{span}{\mathcal{S}} is a dense *-subalgebra of 𝒞{\mathcal{C}}.

Proof.

(a) That 𝒬𝒩(𝒞,𝒟){\mathcal{Q}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) is clear and Lemma 3(c) shows 𝒬(𝒞,𝒟){\mathcal{Q}}\subseteq{\mathcal{I}}({\mathcal{C}},{\mathcal{D}}). For v,w𝒫v,w\in{\mathcal{P}}, hS(v)ch\in S(v)_{c} and kS(w)ck\in S(w)_{c},

vhwk=vhθw1(k)w=vwθw(hθw1(k))𝒬,vhwk=vh\theta_{w}^{-1}(k)w=vw\theta_{w}(h\theta_{w}^{-1}(k))\in{\mathcal{Q}},

by Lemma 3(e). Next, Lemma 3(d) implies 𝒬{\mathcal{Q}} is closed under adjoints: indeed, (vh)=hv=vθv1(h)𝒬(vh)^{*}=h^{*}v^{*}=v^{*}\theta_{v}^{-1}(h^{*})\in{\mathcal{Q}}. Thus 𝒬{\mathcal{Q}} is a *-semigroup.

Turning our attention to 𝒮{\mathcal{S}}, let v𝒫v\in{\mathcal{P}} and hS(v)ch\in S(v)_{c}. For d𝒟d\in{\mathcal{D}}, the fact that vh(𝒞,𝒟)vh\in{\mathcal{I}}({\mathcal{C}},{\mathcal{D}}) gives d(vh)𝒬d(vh)\in{\mathcal{Q}} and (vh)d𝒬(vh)d\in{\mathcal{Q}}. Thus, 𝒬𝒟𝒟𝒬𝒬{\mathcal{Q}}{\mathcal{D}}\cup{\mathcal{D}}{\mathcal{Q}}\subseteq{\mathcal{Q}}. It follows that 𝒮{\mathcal{S}} is a *-semigroup. Since 𝒩(𝒞,𝒟){\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) and (𝒞,𝒟){\mathcal{I}}({\mathcal{C}},{\mathcal{D}}) are *-semigroups containing 𝒟{\mathcal{D}}, 𝒮𝒩(𝒞,𝒟)(𝒞,𝒟){\mathcal{S}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\cap{\mathcal{I}}({\mathcal{C}},{\mathcal{D}}).

(b) Note that (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular inclusion because 𝒫𝒟{\mathcal{P}}\cup{\mathcal{D}} generates a *-semigroup of normalizers. As (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP, [PittsNoApUnInC*Al, Observation 1.3] shows vv𝒟v^{*}v\in{\mathcal{D}} for every v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). Lemma 3(a) shows v𝒮¯v\in\overline{{\mathcal{S}}} for every v𝒫v\in{\mathcal{P}}. Thus span𝒮¯\overline{\operatorname{span}{\mathcal{S}}} contains 𝒫𝒟{\mathcal{P}}\cup{\mathcal{D}}, whence span𝒮¯=𝒞\overline{\operatorname{span}{\mathcal{S}}}={\mathcal{C}}. Finally, as 𝒮{\mathcal{S}} is a *-semigroup, span𝒮\operatorname{span}{\mathcal{S}} is a *-algebra. ∎

By taking 𝒫=𝒩(𝒞,𝒟){\mathcal{P}}={\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), we see that for a broad class of inclusions, the supply of intertwiners is large.

Corollary \the\numberby.

If (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular inclusion with the AUP, then (𝒞,𝒟)𝒩(𝒞,𝒟){\mathcal{I}}({\mathcal{C}},{\mathcal{D}})\cap{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) has dense span in 𝒞{\mathcal{C}}.

We come to a main result, which is the promised characterization of those regular inclusions which have a Cartan envelope. This characterization significantly extends [PittsStReInII, Theorem 5.2], which dealt with unital inclusions.

Theorem \the\numberby (c.f. [PittsStReInII, Theorem 5.2]).

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular inclusion. The following are equivalent.

  1. (a)

    (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has a Cartan envelope.

  2. (b)

    (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property.

  3. (c)

    𝒟c{\mathcal{D}}^{c} is abelian and both (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) and (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) have the ideal intersection property.

When (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) satisfies any of conditions (a)–(c), (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is weakly non-degenerate and the following statements hold.

Uniqueness:

If for j=1,2j=1,2, (𝒜j,j :: αj)({\mathcal{A}}_{j},{\mathcal{B}}_{j}\hbox{\,:\hskip-1.0pt:\,}\alpha_{j}) are Cartan envelopes for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), there exists a unique regular *-isomorphism ψ:(𝒜1,1)(𝒜2,2)\psi:({\mathcal{A}}_{1},{\mathcal{B}}_{1})\rightarrow({\mathcal{A}}_{2},{\mathcal{B}}_{2}) such that ψα1=α2\psi\circ\alpha_{1}=\alpha_{2}.

Minimality:

If (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), there is an ideal 𝔍𝒜{\mathfrak{J}}\subseteq{\mathcal{A}} such that 𝔍α(𝒞)={0}{\mathfrak{J}}\cap\alpha({\mathcal{C}})=\{0\} and, letting q:𝒜𝒜/𝔍q:{\mathcal{A}}\rightarrow{\mathcal{A}}/{\mathfrak{J}} denote the quotient map, (𝒜/𝔍,/(𝔍) :: qα)({\mathcal{A}}/{\mathfrak{J}},{\mathcal{B}}/({\mathfrak{J}}\cap{\mathcal{B}})\hbox{\,:\hskip-1.0pt:\,}q\circ\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

The proof of Theorem 3 is organized as follows. We show (a)\Rightarrow(b)\Rightarrow(c)\Rightarrow(a) and weak non-degeneracy for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), then the uniqueness statement. The proof of the minimality statement may be found after the proof of Proposition 3. During all parts of the proof of Theorem 3, assume that an injective envelope (I(𝒟),ι)(I({\mathcal{D}}),\iota) for 𝒟{\mathcal{D}} has been fixed.

(a)\Rightarrow(b).

Since (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has a Cartan envelope, Lemma 3 shows it is weakly non-degenerate.

Suppose (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). By Proposition 2.3, (𝒜,)({\mathcal{A}},{\mathcal{B}}) has the faithful unique pseudo-expectation property. By definition, (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an essential expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), so Lemma 3 shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property. ∎

(b)\Rightarrow(c).

By Proposition 2.3, 𝒟c{\mathcal{D}}^{c} is abelian, so 𝒟𝒟c𝒞{\mathcal{D}}\subseteq{\mathcal{D}}^{c}\subseteq{\mathcal{C}}. Corollary 2.3 shows (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) has the faithful unique pseudo-expectation property, whence (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) has the ideal intersection property by Proposition 2.4.

Let E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be the faithful unique pseudo-expectation. Since ι2:=E|𝒟c\iota_{2}:=E|_{{\mathcal{D}}^{c}} is a pseudo-expectation for (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}), Proposition 2.4 shows ι2:=E|𝒟c\iota_{2}:=E|_{{\mathcal{D}}^{c}} is the only pseudo-expectation for (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}), and (I(𝒟),ι2)(I({\mathcal{D}}),\iota_{2}) is an injective envelope for 𝒟c{\mathcal{D}}^{c}.

It follows from Lemma 3(a) that (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) is regular MASA inclusion, whence (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) has the unique pseudo-expectation property by Theorem 3. Letting F:𝒞I(𝒟)F:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be the pseudo-expectation for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) relative to (I(𝒟),ι2)(I({\mathcal{D}}),\iota_{2}), we see that F|𝒟cF|_{{\mathcal{D}}^{c}} is a pseudo-expectation for (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}), so F|𝒟c=ι2F|_{{\mathcal{D}}^{c}}=\iota_{2}. But then F|𝒟=ι2|𝒟=ιF|_{\mathcal{D}}=\iota_{2}|_{{\mathcal{D}}}=\iota, hence FF is a pseudo-expectatation for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). This forces F=EF=E. As EE is faithful, we conclude that (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) has the faithful unique pseudo-expectation property. Theorem 3 now shows (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) has the ideal intersection property. ∎

(c)\Rightarrow(a).

For notational convenience, let :=𝒟c{\mathcal{M}}:={\mathcal{D}}^{c}. By Lemma 2.2(e), (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is weakly non-degenerate; thus Lemma 2.2(e) shows (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a unital inclusion, and Corollary 2.2 gives 𝒟~~𝒞~\tilde{\mathcal{D}}\subseteq\tilde{\mathcal{M}}\subseteq\tilde{\mathcal{C}}. By [PittsNoApUnInC*Al, Theorem 2.5], (𝒞,)({\mathcal{C}},{\mathcal{M}}) has the AUP, so

𝒩(𝒞,𝒟)3(a)𝒩(𝒞,)3(c)𝒩(𝒞~,~).{\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\stackrel{{\scriptstyle\ref{relcom}\eqref{relcom1}}}{{\subseteq}}{\mathcal{N}}({\mathcal{C}},{\mathcal{M}})\stackrel{{\scriptstyle\ref{relcom}\eqref{relcom3}}}{{\subseteq}}{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{M}}).

Therefore, (𝒞,)({\mathcal{C}},{\mathcal{M}}) is a regular MASA inclusion; it has the ideal intersection property by hypothesis. It follows from Lemma 2.4 that (𝒞~,~)(\tilde{\mathcal{C}},\tilde{\mathcal{M}}) is also a regular MASA inclusion with the ideal intersection property.

By [PittsStReInII, Theorem 5.2], (𝒞~,~)(\tilde{\mathcal{C}},\tilde{\mathcal{M}}) has a Cartan envelope (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha). Since (,α(~))({\mathcal{B}},\alpha(\tilde{\mathcal{M}})) and (~,)(\tilde{\mathcal{M}},{\mathcal{M}}) have the ideal intersection property, so does (,α())({\mathcal{B}},\alpha({\mathcal{M}})) (Lemma 2.4). Thus (𝒜, :: α|𝒞)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha|_{{\mathcal{C}}}) is an essential and regular Cartan expansion for (𝒞,)({\mathcal{C}},{\mathcal{M}}). Proposition 3 shows that (𝒞,)({\mathcal{C}},{\mathcal{M}}) has a Cartan envelope. Finally, Proposition 3 implies (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has a Cartan envelope. ∎

Proof of the Uniqueness Statement in Theorem 3.

The proof follows the same pattern as the proof of [PittsStReInII, Proposition 5.24]. Suppose for i=1,2i=1,2 that (𝒜i,i :: αi)({\mathcal{A}}_{i},{\mathcal{B}}_{i}\hbox{\,:\hskip-1.0pt:\,}\alpha_{i}) are Cartan envelopes for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). The equivalence of statements (a) and (b) in Theorem 3 shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property. Since (𝒜i,i :: α)({\mathcal{A}}_{i},{\mathcal{B}}_{i}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an essential expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), (i,𝒟,αi|𝒟)({\mathcal{B}}_{i},{\mathcal{D}},\alpha_{i}|_{{\mathcal{D}}}) has the ideal intersection property, so it has the faithful unique pseudo-expectation property (Proposition 2.4). Thus there exist unique *-monomorphisms ιi:iI(𝒟)\iota_{i}:{\mathcal{B}}_{i}\rightarrow I({\mathcal{D}}) such that ιi(αi|𝒟)=ι\iota_{i}\circ(\alpha_{i}|_{\mathcal{D}})=\iota. As ιiΔiαi\iota_{i}\circ\Delta_{i}\circ\alpha_{i} is a pseudo-expectation for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), we obtain

(3.11) ι1Δ1α1=E=ι2Δ2α2.\iota_{1}\circ\Delta_{1}\circ\alpha_{1}=E=\iota_{2}\circ\Delta_{2}\circ\alpha_{2}.

But i=C(Δi(αi(𝒞))){\mathcal{B}}_{i}=C^{*}(\Delta_{i}(\alpha_{i}({\mathcal{C}}))), so ιi(i)=C(E(𝒞))\iota_{i}({\mathcal{B}}_{i})=C^{*}(E({\mathcal{C}})). As ιi\iota_{i} is a *-monomorphism and is the unique pseudo-expectation for (i,𝒟,αi|𝒟)({\mathcal{B}}_{i},{\mathcal{D}},\alpha_{i}|_{{\mathcal{D}}}), we conclude ψ:=ι21ι1\psi_{\mathcal{B}}:=\iota_{2}^{-1}\circ\iota_{1} is the unique *-isomorphism of 1{\mathcal{B}}_{1} onto 2{\mathcal{B}}_{2} satisfying

ψα1|𝒟=α2|𝒟.\psi_{\mathcal{B}}\circ\alpha_{1}|_{{\mathcal{D}}}=\alpha_{2}|_{\mathcal{D}}.

The following diagram illustrates the maps just discussed.

(3.12) 𝒜1\textstyle{{\mathcal{A}}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ1\scriptstyle{\Delta_{1}}𝒜2\textstyle{{\mathcal{A}}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ2\scriptstyle{\Delta_{2}}𝒞\textstyle{{\mathcal{C}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α1\scriptstyle{\alpha_{1}}α2\scriptstyle{\alpha_{2}}E\scriptstyle{E}I(𝒟)\textstyle{I({\mathcal{D}})}𝒟\textstyle{{\mathcal{D}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}α1|𝒟\scriptstyle{\alpha_{1}|_{{\mathcal{D}}}}α2|𝒟\scriptstyle{\alpha_{2}|_{{\mathcal{D}}}}1\textstyle{{\mathcal{B}}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι1\scriptstyle{\iota_{1}}ψ:=ι21ι1\scriptstyle{\psi_{\mathcal{B}}:=\iota_{2}^{-1}\circ\iota_{1}}2\textstyle{{\mathcal{B}}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι2\scriptstyle{\iota_{2}}

We shall extend ψ\psi_{\mathcal{B}} to the desired isomorphism of 𝒜1{\mathcal{A}}_{1} onto 𝒜2{\mathcal{A}}_{2}. We do this in stages.

By the regularity of the map αi\alpha_{i},

𝒫i:=αi(𝒩(𝒞,𝒟)){\mathcal{P}}_{i}:=\alpha_{i}({\mathcal{N}}({\mathcal{C}},{\mathcal{D}}))

is a *-semigroup contained in 𝒩(𝒜i,i){\mathcal{N}}({\mathcal{A}}_{i},{\mathcal{B}}_{i}). Since Cartan inclusions have the AUP, Lemma 3(b) shows that for v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), αi(vv)i\alpha_{i}(v^{*}v)\in{\mathcal{B}}_{i}. Also, (3.11) gives

(3.13) ψΔ1α1=Δ2α2.\psi_{\mathcal{B}}\circ\Delta_{1}\circ\alpha_{1}=\Delta_{2}\circ\alpha_{2}.

Put

𝒬i:={αi(v)h:v𝒩(𝒞,𝒟),h(αi(vv)i¯)c}and𝒮i:=i𝒬i.{\mathcal{Q}}_{i}:=\{\alpha_{i}(v)h:v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}),h\in(\overline{\alpha_{i}(v^{*}v){\mathcal{B}}_{i}})_{c}\}\quad\text{and}\quad{\mathcal{S}}_{i}:={\mathcal{B}}_{i}\cup{\mathcal{Q}}_{i}.

Since (𝒜i,i)({\mathcal{A}}_{i},{\mathcal{B}}_{i}) has the AUP and 𝒜i{\mathcal{A}}_{i} is generated by i𝒫i{\mathcal{B}}_{i}\cup{\mathcal{P}}_{i}, Proposition 3 shows that 𝒬i{\mathcal{Q}}_{i} and 𝒮i{\mathcal{S}}_{i} are *-semigroups contained in 𝒩(𝒜i,i){\mathcal{N}}({\mathcal{A}}_{i},{\mathcal{B}}_{i}) and span𝒮i\operatorname{span}{\mathcal{S}}_{i} is a dense *-subalgebra of 𝒜i{\mathcal{A}}_{i}.111The proof of [PittsStReInII, Proposition 5.24] states without proof that the set ={αi(v)h:v𝒩(𝒞,𝒟),hαi(vv)𝒟i¯}{\mathcal{M}}=\{\alpha_{i}(v)h:v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}),h\in\overline{\alpha_{i}(v^{*}v){\mathcal{D}}_{i}}\} is a *-semigroup and uses {\mathcal{M}} instead of 𝒬{\mathcal{Q}}. (Note that span𝒬i\operatorname{span}{\mathcal{Q}}_{i} is also a *-subalgebra of span𝒮i\operatorname{span}{\mathcal{S}}_{i}.)

The first stage in extending ψ\psi_{\mathcal{B}} is to extend it to an isomorphism ψ𝒬:span𝒬1span𝒬2\psi_{\mathcal{Q}}:\operatorname{span}{\mathcal{Q}}_{1}\rightarrow\operatorname{span}{\mathcal{Q}}_{2}. Let nn\in{\mathbb{N}} and suppose for 1kn1\leq k\leq n, vk𝒩(𝒞,𝒟)v_{k}\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) and hk(α1(vkvk)1¯)ch_{k}\in(\overline{\alpha_{1}(v_{k}^{*}v_{k}){\mathcal{B}}_{1}})_{c}. Then

0=k=1nα1(vk)hk\displaystyle 0=\sum_{k=1}^{n}\alpha_{1}(v_{k})h_{k} Δ1(k,=1nhkα1(vkv)h)=0\displaystyle\iff\Delta_{1}\left(\sum_{k,\ell=1}^{n}h_{k}^{*}\alpha_{1}(v_{k}^{*}v_{\ell})h_{\ell}\right)=0
ψ(k,=1nhkΔ1(α1(vkv))h)=0\displaystyle\iff\psi_{\mathcal{B}}\left(\sum_{k,\ell=1}^{n}h_{k}^{*}\Delta_{1}(\alpha_{1}(v_{k}^{*}v_{\ell}))h_{\ell}\right)=0
(3.13)k,=1nψ(hk)Δ2(α2(vkv))ψ(h)=0\displaystyle\stackrel{{\scriptstyle\eqref{!pscharu1}}}{{\iff}}\sum_{k,\ell=1}^{n}\psi_{\mathcal{B}}(h_{k}^{*})\Delta_{2}(\alpha_{2}(v_{k}^{*}v_{\ell}))\psi_{\mathcal{B}}(h_{\ell})=0
Δ2(k,=1nψ(hk)α2(vkv)ψ(h))=0\displaystyle\iff\Delta_{2}\left(\sum_{k,\ell=1}^{n}\psi_{\mathcal{B}}(h_{k}^{*})\alpha_{2}(v_{k}^{*}v_{\ell})\psi_{\mathcal{B}}(h_{\ell})\right)=0
k=1nα2(vk)ψ(hk)=0.\displaystyle\iff\sum_{k=1}^{n}\alpha_{2}(v_{k})\psi_{\mathcal{B}}(h_{k})=0.

Therefore, ψ\psi_{\mathcal{B}} extends uniquely to a *-isomorphism ψ𝒬:span(𝒬1)span(𝒬2)\psi_{\mathcal{Q}}:\operatorname{span}({\mathcal{Q}}_{1})\rightarrow\operatorname{span}({\mathcal{Q}}_{2}) given by

k=1nα1(vk)hkk=1nα2(vk)ψ(hk).\sum_{k=1}^{n}\alpha_{1}(v_{k})h_{k}\mapsto\sum_{k=1}^{n}\alpha_{2}(v_{k})\psi(h_{k}).

For v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) and h(α1(vv)1¯)ch\in(\overline{\alpha_{1}(v^{*}v){\mathcal{B}}_{1}})_{c},

ψ(Δ1(α1(v)h))\displaystyle\psi_{\mathcal{B}}(\Delta_{1}(\alpha_{1}(v)h)) =ι21(ι1(Δ1(α1(v))h))=ι21(E(v)ι1(h))\displaystyle=\iota_{2}^{-1}(\iota_{1}(\Delta_{1}(\alpha_{1}(v))h))=\iota_{2}^{-1}(E(v)\iota_{1}(h))
=ι21(E(v))ψ(h)=ι21(ι2(Δ2(α2(v))))ψ(h)\displaystyle=\iota_{2}^{-1}(E(v))\psi_{\mathcal{B}}(h)=\iota_{2}^{-1}(\iota_{2}(\Delta_{2}(\alpha_{2}(v))))\psi_{\mathcal{B}}(h)
=Δ2(α2(v)ψ(h))=Δ2(ψ𝒬(α1(v)h)).\displaystyle=\Delta_{2}(\alpha_{2}(v)\psi_{\mathcal{B}}(h))=\Delta_{2}(\psi_{\mathcal{Q}}(\alpha_{1}(v)h)).

It follows that

(3.14) ψΔ1|span𝒬1=Δ2ψ𝒬.\psi_{\mathcal{B}}\circ\Delta_{1}|_{\operatorname{span}{\mathcal{Q}}_{1}}=\Delta_{2}\circ\psi_{\mathcal{Q}}.

Next we extend ψ𝒬\psi_{\mathcal{Q}} to a *-isomorphism ψ𝒮:span𝒮1span𝒮2\psi_{\mathcal{S}}:\operatorname{span}{\mathcal{S}}_{1}\rightarrow\operatorname{span}{\mathcal{S}}_{2}. For b11b_{1}\in{\mathcal{B}}_{1} and t1span𝒬1t_{1}\in\operatorname{span}{\mathcal{Q}}_{1} we have

0=b1+t1Δ1((b1+t1)(b1+t1))=0b1b1+b1Δ1(t1)+Δ1(t1)b1+Δ1(t1t1)=0(3.14)ψ(b1b1)+ψ(b1)Δ2(ψ𝒬(t1))+Δ2(ψ𝒬(t1))ψ(b1)+Δ2(ψ𝒬(t1t1))=0Δ2((ψ(b1)+ψ𝒬(t1))(ψ(b1)+ψ𝒬(t1)))=0ψ(b1)+ψQ(t1)=0.\displaystyle\begin{split}0=b_{1}+t_{1}\iff&\Delta_{1}((b_{1}+t_{1})^{*}(b_{1}+t_{1}))=0\\ \iff&b_{1}^{*}b_{1}+b_{1}^{*}\Delta_{1}(t_{1})+\Delta_{1}(t_{1})^{*}b_{1}+\Delta_{1}(t_{1}^{*}t_{1})=0\\ \stackrel{{\scriptstyle\eqref{psiIDelta}}}{{\iff}}&\psi_{\mathcal{B}}(b_{1}^{*}b_{1})+\psi_{\mathcal{B}}(b_{1})^{*}\Delta_{2}(\psi_{\mathcal{Q}}(t_{1}))+\Delta_{2}(\psi_{\mathcal{Q}}(t_{1}^{*}))\psi_{\mathcal{B}}(b_{1})\\ &\qquad+\Delta_{2}(\psi_{\mathcal{Q}}(t_{1}^{*}t_{1}))=0\\ \iff&\Delta_{2}((\psi_{\mathcal{B}}(b_{1})+\psi_{\mathcal{Q}}(t_{1}))^{*}(\psi_{\mathcal{B}}(b_{1})+\psi_{\mathcal{Q}}(t_{1})))=0\\ \iff&\psi_{\mathcal{B}}(b_{1})+\psi_{Q}(t_{1})=0.\end{split}

Therefore, the map span𝒮1=1+span𝒬1b1+t1ψ(b1)+ψQ(t1)span𝒮2\operatorname{span}{\mathcal{S}}_{1}={\mathcal{B}}_{1}+\operatorname{span}{\mathcal{Q}}_{1}\ni b_{1}+t_{1}\mapsto\psi_{\mathcal{B}}(b_{1})+\psi_{Q}(t_{1})\in\operatorname{span}{\mathcal{S}}_{2} is well-defined, so we obtain a *-isomorphism ψ𝒮:span𝒮1span𝒮2\psi_{\mathcal{S}}:\operatorname{span}{\mathcal{S}}_{1}\rightarrow\operatorname{span}{\mathcal{S}}_{2}.

Now let 𝒮~i:=u𝒜i(𝒮i)I𝒜~i𝒜~i\tilde{\mathcal{S}}_{i}:=u_{{\mathcal{A}}_{i}}({\mathcal{S}}_{i})\cup{\mathbb{C}}I_{\tilde{\mathcal{A}}_{i}}\subseteq\tilde{\mathcal{A}}_{i}. Then 𝒮~i\tilde{\mathcal{S}}_{i} is a MASA skeleton for (𝒜~i,~i)(\tilde{\mathcal{A}}_{i},\tilde{\mathcal{B}}_{i}) in the sense of [PittsStReInI, Definition 3.1] (see also [PittsStReInI, Definition 1.7]). Note that ψ𝒮\psi_{\mathcal{S}} extends uniquely to a *-isomorphism ψ~𝒮:span𝒮~1span𝒮~2\tilde{\psi}_{\mathcal{S}}:\operatorname{span}\tilde{\mathcal{S}}_{1}\rightarrow\operatorname{span}\tilde{\mathcal{S}}_{2}. Since (𝒜~i,~i)(\tilde{\mathcal{A}}_{i},\tilde{\mathcal{B}}_{i}) is a Cartan inclusion, we may argue as in the proof of [PittsStReInII, Proposition 5.24] to conclude that ψ~𝒮\tilde{\psi}_{\mathcal{S}} extends to a regular *-isomorphism ψ~\tilde{\psi} of (𝒜~1,~1)(\tilde{\mathcal{A}}_{1},\tilde{\mathcal{B}}_{1}) onto (𝒜~2,~2)(\tilde{\mathcal{A}}_{2},\tilde{\mathcal{B}}_{2}). Thus ψ:=ψ~|𝒜1\psi:=\tilde{\psi}|_{{\mathcal{A}}_{1}} is an isomorphism of 𝒜1{\mathcal{A}}_{1} onto 𝒜2{\mathcal{A}}_{2}. The constructions show ψα1=α2\psi\circ\alpha_{1}=\alpha_{2}.

Suppose ψ:𝒜1𝒜2\psi^{\prime}:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2} is a *-isomorphism such that ψα1=α2\psi^{\prime}\circ\alpha_{1}=\alpha_{2}. The uniqueness assertion for ψ\psi_{\mathcal{B}} shows ψ|1=ψ|1\psi^{\prime}|_{{\mathcal{B}}_{1}}=\psi|_{{\mathcal{B}}_{1}}. Examining each stage of the construction of ψ|𝒮\psi|_{\mathcal{S}}, we obtain ψ|𝒮1=ψ|𝒮1\psi^{\prime}|_{{\mathcal{S}}_{1}}=\psi|_{{\mathcal{S}}_{1}}. As span𝒮i\operatorname{span}{\mathcal{S}}_{i} is dense in 𝒜i{\mathcal{A}}_{i}, this forces ψ=ψ\psi^{\prime}=\psi and completes the proof of the uniqueness assertion. ∎

Before turning to the proof of minimality, we need a bit of terminology and a fact about abelian CC^{*}-algebras. Suppose 𝒜{\mathcal{A}} is an arbitrary CC^{*}-algebra and J𝒜J\unlhd{\mathcal{A}}. Set

J:={a𝒜:aj=ja=0 for all jJ}J^{\perp}:=\{a\in{\mathcal{A}}:aj=ja=0\text{ for all }j\in J\}

and recall that JJ is called a regular ideal if J=JJ^{\perp\perp}=J, where JJ^{\perp\perp} means (J)(J^{\perp})^{\perp}. (Because JJ is an ideal, so is JJ^{\perp}.)

Next, let XX be a locally compact Hausdorff space and suppose JC0(X)J\unlhd C_{0}(X). For an open set GXG\subseteq X, we write GG^{\perp} for the interior of XGX\setminus G and write supp(J):={xX:f(x)0 for some fJ}\operatorname{supp}(J):=\{x\in X:f(x)\neq 0\text{ for some }f\in J\}. The proof of the following is left to the reader.

Fact \the\numberby.

Let JC0(X)J\unlhd C_{0}(X) and put G:=supp(J)G:=\operatorname{supp}(J). Then supp(J)=G\operatorname{supp}(J^{\perp\perp})=G^{\perp\perp}, GG is a dense open subset of GG^{\perp\perp}, and (J,J)(J^{\perp\perp},J) has the ideal intersection property.

We require the following invariance type result; it is the extension of [PittsStReInI, Proposition 3.14] to our context.

Proposition \the\numberby (c.f. [PittsStReInI, Proposition 3.14]).

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular inclusion having the unique faithful pseudo-expectation property, and let E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be the pseudo-expectation. Fix v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), and let θ¯v\overline{\theta}_{v} be the result of applying Lemma 2.3 to θv:vv𝒟¯vv𝒟¯\theta_{v}:\overline{vv^{*}{\mathcal{D}}}\rightarrow\overline{v^{*}v{\mathcal{D}}}. Then for every x𝒞x\in{\mathcal{C}},

(3.15) E(vxv)=θ¯v(E(vvx)).E(v^{*}xv)=\overline{\theta}_{v}(E(vv^{*}x)).
Proof.

To simplify notation, throughout the proof we will identify 𝒜{\mathcal{A}} with its image under u𝒜u_{\mathcal{A}} in 𝒜~\tilde{\mathcal{A}}, and we will write

:=𝒟c.{\mathcal{M}}:={\mathcal{D}}^{c}.

We wish to apply [PittsStReInI, Proposition 3.14] to (𝒞~,~)(\tilde{\mathcal{C}},\tilde{\mathcal{M}}). To do so, we show that (𝒞~,~)(\tilde{\mathcal{C}},\tilde{\mathcal{M}}) is a regular and unital MASA inclusion with the faithful unique pseudo-expectation property, that (I(𝒟),E~|~)(I({\mathcal{D}}),\tilde{E}|_{\tilde{\mathcal{M}}}) is an injective envelope for ~\tilde{\mathcal{M}}, and that E~:𝒞~I(𝒟)\tilde{E}:\tilde{\mathcal{C}}\rightarrow I({\mathcal{D}}) is the pseudo-expectation for (𝒞~,~)(\tilde{\mathcal{C}},\tilde{\mathcal{M}}).

By Corollary 2.3, (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is weakly non-degenerate, so Lemma 2.2(e) shows (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a unital inclusion. Also, Corollary 2.2 gives 𝒟~~𝒞~\tilde{\mathcal{D}}\subseteq\tilde{\mathcal{M}}\subseteq\tilde{\mathcal{C}}. In particular, (𝒞~,~)(\tilde{\mathcal{C}},\tilde{\mathcal{M}}) is a unital inclusion. Proposition 2.3 shows {\mathcal{M}} is abelian, Corollary 2.3 shows (,𝒟)({\mathcal{M}},{\mathcal{D}}) has the faithful unique pseudo-expectation property, and by Proposition 2.4, (I(𝒟),E|)(I({\mathcal{D}}),E|_{{\mathcal{M}}}) is an injective envelope for {\mathcal{M}}. Thus, (I(𝒟),E~|~)(I({\mathcal{D}}),\tilde{E}|_{\tilde{\mathcal{M}}}) is an injective envelope for ~\tilde{\mathcal{M}}. Next, (𝒞,)({\mathcal{C}},{\mathcal{M}}) is a regular MASA inclusion by Lemma 3(a). Therefore (𝒞,)({\mathcal{C}},{\mathcal{M}}) has the AUP by  [PittsNoApUnInC*Al, Corollary 2.6], so (𝒞~,~)(\tilde{\mathcal{C}},\tilde{\mathcal{M}}) is a unital, regular MASA inclusion (Lemma 3(e)). Corollary 2.3 shows E~\tilde{E} is the pseudo-expectation for (𝒞~,~)(\tilde{\mathcal{C}},\tilde{\mathcal{M}}) and E~\tilde{E} is faithful by Theorem 2.3.

Using parts (a) and (e) of Lemma 3, we obtain

𝒩(𝒞,𝒟)𝒩(𝒞,)𝒩(𝒞~,~).{\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{M}})\subseteq{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{{\mathcal{M}}}).

Thus by [PittsStReInI, Proposition 3.14], for v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) and x𝒞x\in{\mathcal{C}},

E~(vxv)=θ¯v(E~(vvx)),\tilde{E}(v^{*}xv)=\overline{\theta}_{v}(\tilde{E}(vv^{*}x)),

that is, E(vxv)=θ¯v(E(vvx))E(v^{*}xv)=\overline{\theta}_{v}(E(vv^{*}x)). ∎

Proof of the Minimality Statement in Theorem 3.

Suppose that (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) and Δ:𝒜\Delta:{\mathcal{A}}\rightarrow{\mathcal{B}} is the conditional expectation. By injectivity of I(𝒟)I({\mathcal{D}}), there exists a *-homomorphism γ~:~I(𝒟)\tilde{\gamma}:\tilde{\mathcal{B}}\rightarrow I({\mathcal{D}}) such that

(3.16) ι=γ~α~|𝒟~.\iota=\tilde{\gamma}\circ\tilde{\alpha}|_{\tilde{\mathcal{D}}}.

To show γ~\tilde{\gamma} is the unique such *-homomorphism, suppose τ~:~I(𝒟)\tilde{\tau}:\tilde{\mathcal{B}}\rightarrow I({\mathcal{D}}) is another *-homomorphism such that ι=τ~α~|𝒟~\iota=\tilde{\tau}\circ\tilde{\alpha}|_{\tilde{\mathcal{D}}}. Note that γ~Δ~α~|𝒟~=ι\tilde{\gamma}\circ\tilde{\Delta}\circ\tilde{\alpha}|_{\tilde{\mathcal{D}}}=\iota, so γ~Δ~α~\tilde{\gamma}\circ\tilde{\Delta}\circ\tilde{\alpha} is a pseudo-expectation for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}). Likewise τ~Δ~α~\tilde{\tau}\circ\tilde{\Delta}\circ\tilde{\alpha} is a pseudo-expectation for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}). Since (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the unique pseudo-expectation property, so does (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) (Corollary 2.3). Therefore,

γ~Δ~α~=τ~Δ~α~.\tilde{\gamma}\circ\tilde{\Delta}\circ\tilde{\alpha}=\tilde{\tau}\circ\tilde{\Delta}\circ\tilde{\alpha}.

The definition of Cartan package shows =C(Δ(α(𝒞))){\mathcal{B}}=C^{*}(\Delta(\alpha({\mathcal{C}}))), hence ~=C(Δ~(α~(𝒞~))\tilde{\mathcal{B}}=C^{*}(\tilde{\Delta}(\tilde{\alpha}(\tilde{\mathcal{C}})). Thus τ~=γ~\tilde{\tau}=\tilde{\gamma}, which shows γ~\tilde{\gamma} is the unique *-homomorphism satisfying (3.16).

Let E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be the (unique and faithful) pseudo-expectation. Since γ~Δ~α~\tilde{\gamma}\circ\tilde{\Delta}\circ\tilde{\alpha} is the pseudo-expectation for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}), Lemma 2.3 shows E~=γ~Δ~α~\tilde{E}=\tilde{\gamma}\circ\tilde{\Delta}\circ\tilde{\alpha} and E=γ~Δ~α~u𝒞E=\tilde{\gamma}\circ\tilde{\Delta}\circ\tilde{\alpha}\circ u_{\mathcal{C}}. Define

γ:=γ~u.\gamma:=\tilde{\gamma}\circ u_{\mathcal{B}}.

Using (2.1.4), we find

(3.17) E=γ~uΔα=γΔα.E=\tilde{\gamma}\circ u_{\mathcal{B}}\circ\Delta\circ\alpha=\gamma\circ\Delta\circ\alpha.
Claim \the\numberby.

Put

𝔍:={a𝒜:Δ(aa)kerγ}.{\mathfrak{J}}:=\{a\in{\mathcal{A}}:\Delta(a^{*}a)\in\ker\gamma\}.

Then 𝔍{\mathfrak{J}} is a regular ideal in 𝒜{\mathcal{A}} such that 𝔍=kerγ{\mathfrak{J}}\cap{\mathcal{B}}=\ker\gamma.

Proof of Claim 3.

To verify Claim 3, we use [BrownFullerPittsReznikoffReIdIdInQu, Proposition 3.19(ii)]. To do this, we require the following facts:

  1. (a)

    kerγ\ker\gamma is a regular ideal of {\mathcal{B}}; and

  2. (b)

    kerγ\ker\gamma is α(𝒩(𝒞,𝒟))\alpha({\mathcal{N}}({\mathcal{C}},{\mathcal{D}})) invariant, that is, for every v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}),

    (3.18) α(v)(kerγ)α(v)kerγ.\alpha(v^{*})(\ker\gamma)\alpha(v)\subseteq\ker\gamma.

Our first step in verifying (a) is to show kerγ\ker\gamma is the unique ideal in {\mathcal{B}} maximal with respect to having trivial intersection with α(𝒟)\alpha({\mathcal{D}}). By [PittsZarikianUnPsExC*In, Corollary 3.21], kerγ~\ker\tilde{\gamma} is the unique maximal α~(𝒟~)\tilde{\alpha}(\tilde{\mathcal{D}})-disjoint ideal of ~\tilde{\mathcal{B}}. Suppose JJ\unlhd{\mathcal{B}} satisfies Jα(𝒟)={0}J\cap\alpha({\mathcal{D}})=\{0\}. Then u(J)~u_{\mathcal{B}}(J)\unlhd\tilde{\mathcal{B}}. Let us show

(3.19) u(J)α~(𝒟~)={0}.u_{\mathcal{B}}(J)\cap\tilde{\alpha}(\tilde{\mathcal{D}})=\{0\}.

We do this by considering two cases.

Case 1: 𝒟{\mathcal{D}} is unital. Since (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is weakly non-degenerate, 𝒞{\mathcal{C}} is unital. Let us show {\mathcal{B}} is unital and that e:=α(I𝒞)e:=\alpha(I_{\mathcal{C}}) is its unit. Since α\alpha is a regular *-monomorphism, e𝒩(𝒜,)e\in{\mathcal{N}}({\mathcal{A}},{\mathcal{B}}), hence e=eee=e^{*}e commutes with {\mathcal{B}} by [PittsNoApUnInC*Al, Proposition 2.1]. Thus ee\in{\mathcal{B}} because {\mathcal{B}} is a MASA in 𝒜{\mathcal{A}}. Let n1n\geq 1 and x1,,xn𝒞x_{1},\dots,x_{n}\in{\mathcal{C}}. Then

eΔ(α(x1))=Δ(eα(x1))=Δ(α(x1)); likewise Δ(α(xn))e=Δ(α(xn)).e\Delta(\alpha(x_{1}))=\Delta(e\alpha(x_{1}))=\Delta(\alpha(x_{1}));\text{ likewise }\Delta(\alpha(x_{n}))e=\Delta(\alpha(x_{n})).

Hence

e(k=1nΔ(α(xk)))=(k=1nΔ(α(xk)))e=k=1nΔ(α(xk)).e\left(\prod_{k=1}^{n}\Delta(\alpha(x_{k}))\right)=\left(\prod_{k=1}^{n}\Delta(\alpha(x_{k}))\right)e=\prod_{k=1}^{n}\Delta(\alpha(x_{k})).

By the definition of a package, the span of finite products from Δ(𝒞)\Delta({\mathcal{C}}) is dense in {\mathcal{B}}. Thus e=Ie=I_{\mathcal{B}}.

As (𝒜,)({\mathcal{A}},{\mathcal{B}}) is a Cartan inclusion, it has the AUP. Therefore II_{\mathcal{B}} is the unit for 𝒜{\mathcal{A}}. The equality (3.19) now follows because uu_{\mathcal{B}} is the identity map on {\mathcal{B}}, α~=α\tilde{\alpha}=\alpha and 𝒟~=𝒟\tilde{\mathcal{D}}={\mathcal{D}}.

Case 2: 𝒟{\mathcal{D}} is not unital. Let yu(J)α~(𝒟~)y\in u_{\mathcal{B}}(J)\cap\tilde{\alpha}(\tilde{\mathcal{D}}). Then for some (d,λ)𝒟~(d,\lambda)\in\tilde{\mathcal{D}} and jJj\in J,

y=α(d)+λI~=u(j).y=\alpha(d)+\lambda I_{\tilde{\mathcal{B}}}=u_{\mathcal{B}}(j).

For any h𝒟h\in{\mathcal{D}}, multiplying each side of the second equality by α~(h,0)\tilde{\alpha}(h,0) gives

α(dh)+λα(h)=u(j)α~(h,0).\alpha(dh)+\lambda\alpha(h)=u_{\mathcal{B}}(j)\tilde{\alpha}(h,0).

Since the left side belongs to u(α(𝒟))u_{\mathcal{B}}(\alpha({\mathcal{D}})) and the right belongs to u(J)u_{\mathcal{B}}(J), the hypothesis that α(𝒟)J={0}\alpha({\mathcal{D}})\cap J=\{0\} gives α(dh)+λα(h)=0\alpha(dh)+\lambda\alpha(h)=0. Thus for every h𝒟h\in{\mathcal{D}},

dh=λh.dh=-\lambda h.

Let (eλ)(e_{\lambda}) be an approximate unit for 𝒟{\mathcal{D}}. Choosing h=eλh=e_{\lambda} and taking the limit, we obtain limλeλ\lim\lambda e_{\lambda} exists in 𝒟{\mathcal{D}} and d=limλeλd=-\lim\lambda e_{\lambda}. As 𝒟{\mathcal{D}} is non-unital, this forces y=0y=0. This completes the proof of (3.19).

Equality 3.19 and the fact that kerγ~\ker\tilde{\gamma} is the maximal ideal of ~\tilde{\mathcal{B}} disjoint from α~(𝒟~)\tilde{\alpha}(\tilde{\mathcal{D}}) gives u(J)kerγ~u_{\mathcal{B}}(J)\subseteq\ker\tilde{\gamma}. Therefore u(J)=u(J)u()kerγ~u()=u(kerγ)u_{\mathcal{B}}(J)=u_{\mathcal{B}}(J)\cap u_{\mathcal{B}}({\mathcal{B}})\subseteq\ker\tilde{\gamma}\cap u_{\mathcal{B}}({\mathcal{B}})=u_{\mathcal{B}}(\ker\gamma), whence JkerγJ\subseteq\ker\gamma. Thus kerγ\ker\gamma is the unique maximal α(𝒟)\alpha({\mathcal{D}})-disjoint ideal of {\mathcal{B}}.

We are now prepared to show kerγ\ker\gamma is a regular ideal of {\mathcal{B}}. Suppose d𝒟d\in{\mathcal{D}} satisfies α(d)(kerγ)\alpha(d)\in(\ker\gamma)^{\perp\perp}. Let d{\mathcal{I}}_{d} be the closed ideal of {\mathcal{B}} generated by α(d)\alpha(d). Then d{\mathcal{I}}_{d} is contained in (kerγ)(\ker\gamma)^{\perp\perp}, so α(d)d(kerγ)\alpha(d)\in{\mathcal{I}}_{d}\cap(\ker\gamma)^{\perp\perp}. But dkerγα(𝒟)kerγ={0}{\mathcal{I}}_{d}\cap\ker\gamma\subseteq\alpha({\mathcal{D}})\cap\ker\gamma=\{0\}. Fact 3 shows ((kerγ),kerγ)((\ker\gamma)^{\perp\perp},\ker\gamma) has the ideal intersection property, so d={0}{\mathcal{I}}_{d}=\{0\}. That d=0d=0 follows. Therefore (kerγ)(\ker\gamma)^{\perp\perp} is an ideal of {\mathcal{B}} having trivial intersection with α(𝒟)\alpha({\mathcal{D}}). As kerγ(kerγ)\ker\gamma\subseteq(\ker\gamma)^{\perp\perp}, the maximality property of kerγ\ker\gamma gives kerγ=(kerγ)\ker\gamma=(\ker\gamma)^{\perp\perp}, so kerγ\ker\gamma is a regular ideal of {\mathcal{B}}.

Turning now to (b), we will use the notation from (3.9). We claim that for v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}),

(3.20) γθα(v)=θ¯vγ|R(α(v)).\gamma\circ\theta_{\alpha(v)}=\overline{\theta}_{v}\circ\gamma|_{R(\alpha(v))}.

To see this, let v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). Regularity of the map α\alpha and the fact that Cartan inclusions have the AUP, gives α(vv)\alpha(vv^{*})\in{\mathcal{B}}.

Let 0{\mathbb{P}}_{0} be the collection of all polynomials pp with complex coefficients such that p(0)=0p(0)=0. Given p0p\in{\mathbb{P}}_{0}, factor p(t)=tq(t)p(t)=tq(t), where qq is another polynomial. For x𝒞x\in{\mathcal{C}}, let

y:=α(p(vv))Δ(α(x))=α(vv)α(q(vv))Δ(α(x)).y:=\alpha(p(vv^{*}))\Delta(\alpha(x))=\alpha(vv^{*})\alpha(q(vv^{*}))\Delta(\alpha(x)).

Using Lemma 3, (2.1.10), and the fact that Δ\Delta is a conditional expectation,

γ(θα(v)(y))\displaystyle\gamma(\theta_{\alpha(v)}(y)) =γ(α(v)Δ(α(q(vv)x))α(v))=γ(Δ(α(vq(vv)xv)))\displaystyle=\gamma(\alpha(v)^{*}\Delta(\alpha(q(vv^{*})x))\alpha(v))=\gamma(\Delta(\alpha(v^{*}q(vv^{*})xv)))
=(3.17)E(vq(vv)xv)=θ¯v(E(vvq(vv)x))=θ¯v(γ(Δ(α(vvq(vv)x))))\displaystyle\stackrel{{\scriptstyle\eqref{ush1}}}{{=}}E(v^{*}q(vv^{*})xv)=\overline{\theta}_{v}(E(vv^{*}q(vv^{*})x))=\overline{\theta}_{v}(\gamma(\Delta(\alpha(vv^{*}q(vv^{*})x))))
=θ¯v(γ(y)).\displaystyle=\overline{\theta}_{v}(\gamma(y)).

The maps γ\gamma, θα(v)\theta_{\alpha(v)}, and θ¯v\overline{\theta}_{v} are multiplicative on their domains. Let

v:={p(vv):p0}.{\mathbb{P}}_{v}:=\{p(vv^{*}):p\in{\mathbb{P}}_{0}\}.

Since =C(Δ(α(𝒞)){\mathcal{B}}=C^{*}(\Delta(\alpha({\mathcal{C}})), we find that for bα(v)b\in\alpha({\mathbb{P}}_{v}){\mathcal{B}},

γ(θα(v)(b))=θ¯v(γ(b)).\gamma(\theta_{\alpha(v)}(b))=\overline{\theta}_{v}(\gamma(b)).

Since α(v)\alpha({\mathbb{P}}_{v}){\mathcal{B}} is dense in R(α(v))R(\alpha(v)), continuity gives (3.20).

With (3.20) in hand, we now complete the proof of the invariance of kerγ\ker\gamma. Suppose tkerγt\in\ker\gamma, and let bb\in{\mathcal{B}}. Then

γ(α(v)tbα(v))=γ(θα(v)(tbα(vv)))=(3.20)θ¯v(γ(tbα(vv)))=0.\gamma(\alpha(v)^{*}tb\alpha(v))=\gamma(\theta_{\alpha(v)}(tb\alpha(vv^{*})))\stackrel{{\scriptstyle\eqref{fJreg2a}}}{{=}}\overline{\theta}_{v}(\gamma(tb\alpha(vv^{*})))=0.

By replacing bb with elements from an approximate unit for {\mathcal{B}}, we find α(v)tα(v)kerγ\alpha(v^{*})t\alpha(v)\in\ker\gamma, so (3.18) holds. This completes the proof of (b).

Let NN be the *-semigroup generated by α(𝒩(𝒞,𝒟)){\mathcal{B}}\cup\alpha({\mathcal{N}}({\mathcal{C}},{\mathcal{D}})). For every wNw\in N, using (b), we find that w(kerγ)wkerγw(\ker\gamma)w^{*}\subseteq\ker\gamma, that is kerγ\ker\gamma is an NN-invariant regular ideal in {\mathcal{B}}. While the statement of [BrownFullerPittsReznikoffReIdIdInQu, Proposition 3.19(ii)], concerns a regular invariant ideals, the proof applies for NN-invariant regular ideals (in fact, the statement should have been for NN-invariant regular ideals). Thus, [BrownFullerPittsReznikoffReIdIdInQu, Proposition 3.19(ii)] shows 𝔍{\mathfrak{J}} is a regular ideal in 𝒜{\mathcal{A}} such that 𝔍=kerγ{\mathfrak{J}}\cap{\mathcal{B}}=\ker\gamma. The proof of Claim 3 is now complete. \diamondsuit

Let q:𝒜𝒜/𝔍q:{\mathcal{A}}\rightarrow{\mathcal{A}}/{\mathfrak{J}} be the quotient map. We must show (𝒜/𝔍,/(𝔍) :: qα)({\mathcal{A}}/{\mathfrak{J}},{\mathcal{B}}/({\mathfrak{J}}\cap{\mathcal{B}})\hbox{\,:\hskip-1.0pt:\,}q\circ\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

First we show qα:(𝒞,𝒟)(𝒜/𝔍,/(𝒥))q\circ\alpha:({\mathcal{C}},{\mathcal{D}})\rightarrow({\mathcal{A}}/{\mathfrak{J}},{\mathcal{B}}/({\mathcal{J}}\cap{\mathcal{B}})) is a regular *-monomorphism. To show qαq\circ\alpha is one-to-one, suppose x𝔍α(𝒞)x\in{\mathfrak{J}}\cap\alpha({\mathcal{C}}). Then there exists c𝒞c\in{\mathcal{C}} so that x=α(c)x=\alpha(c). Since xx𝔍α(𝒞)x^{*}x\in{\mathfrak{J}}\cap\alpha({\mathcal{C}}), (3.17) shows

0=γ(Δ(α(cc)))=E(cc).0=\gamma(\Delta(\alpha(c^{*}c)))=E(c^{*}c).

Since the pseudo-expectation for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is faithful, we find c=0c=0, hence x=0x=0. Therefore qαq\circ\alpha is faithful. For v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) and hh\in{\mathcal{B}}, (3.18) gives

α(v)(h+kerγ)α(v)α(v)hα(v)+kerγ/(𝔍),\alpha(v)^{*}(h+\ker\gamma)\alpha(v)\subseteq\alpha(v^{*})h\alpha(v)+\ker\gamma\in{\mathcal{B}}/({\mathcal{B}}\cap{\mathfrak{J}}),

whence qαq\circ\alpha is a regular *-monomorphism.

By [BrownFullerPittsReznikoffReIdIdInQu, Theorem 4.8], (𝒜/𝔍,/(𝔍))({\mathcal{A}}/{\mathfrak{J}},{\mathcal{B}}/({\mathcal{B}}\cap{\mathfrak{J}})) is a Cartan inclusion. Clearly (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) generates (𝒜/𝔍,/(𝔍))({\mathcal{A}}/{\mathfrak{J}},{\mathcal{B}}/({\mathfrak{J}}\cap{\mathcal{B}})) in the sense of part (ii) of Definition 3(a).

It remains to establish that (/(𝔍),𝒟,qα)({\mathcal{B}}/({\mathfrak{J}}\cap{\mathcal{B}}),{\mathcal{D}},q\circ\alpha) has the ideal intersection property. Suppose K/(𝔍)K\unlhd{\mathcal{B}}/({\mathcal{B}}\cap{\mathfrak{J}}) satisfies K(q(α(𝒟)))={0}K\cap(q(\alpha({\mathcal{D}})))=\{0\}. Since qαq\circ\alpha is faithful, q1(K){\mathcal{B}}\cap q^{-1}(K) is an ideal of {\mathcal{B}} having trivial intersection with α(𝒟)\alpha({\mathcal{D}}). By the maximality property of kerγ\ker\gamma, q1(K)kerγ{\mathcal{B}}\cap q^{-1}(K)\subseteq\ker\gamma, so K={0}K=\{0\}. Therefore (/(𝔍),𝒟,qα)({\mathcal{B}}/({\mathcal{B}}\cap{\mathfrak{J}}),{\mathcal{D}},q\circ\alpha) has the ideal intersection property. Thus (𝒜/𝔍,/(𝔍) :: qα)({\mathcal{A}}/{\mathfrak{J}},{\mathcal{B}}/({\mathcal{B}}\cap{\mathfrak{J}})\hbox{\,:\hskip-1.0pt:\,}q\circ\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). This finishes the proof of the minimality statement and also concludes the proof of Theorem 3. ∎

4. Pseudo-Cartan Inclusions and their Cartan Envelopes

In this section  we give the formal definition of a pseudo-Cartan inclusion. The key results of this section  are: Theorem 4.2, which shows that in the presence of the AUP, the unitization process commutes with taking Cartan envelopes, and Theorem 4.3 which gives some properties shared by a pseudo-Cartan inclusion and its Cartan envelope. Proposition 4.4 gives a construction which can be used to produce a family of pseudo-Cartan inclusions starting with a Cartan inclusion.

4.1. Definition of Pseudo-Cartan Inclusions and Examples

We are now prepared to define the notion of pseudo-Cartan inclusion. Standing Assumption 3 remains in force throughout, that is, we assume that 𝒟{\mathcal{D}} is abelian for every inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

Definition \the\numberby.

A regular inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion if it satisfies the following:

  1. (a)

    𝒟c{\mathcal{D}}^{c} is abelian; and

  2. (b)

    both (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) and (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) have the ideal intersection property.

We will frequently use the alternate descriptions of a pseudo-Cartan inclusion found in Theorem 3, often without comment.

The following is immediate from Definition 4.1.

Observation \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be an inclusion with 𝒞{\mathcal{C}} abelian. Then 𝒟c=𝒞{\mathcal{D}}^{c}={\mathcal{C}}, so (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion if and only if (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is regular and has the ideal intersection property.

Let us compare the notion of pseudo-Cartan inclusion with other related classes of inclusions.

Cartan Inclusions:

Proposition 2.3 shows that every Cartan inclusion is a pseudo-Cartan inclusion. Furthermore, Proposition 2.3 may be interpreted as stating that a pseudo-Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a Cartan inclusion if and only if the pseudo-expectation E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) “is” a conditional expectation in the sense that E(𝒞)=ι(𝒟)E({\mathcal{C}})=\iota({\mathcal{D}}). This characterization is the reason we chose the term “pseudo-Cartan inclusion” in Definition 4.1.

Virtual Cartan Inclusions:

We will use the term virtual Cartan inclusion for a regular MASA inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) having the ideal intersection property. Theorem 3 shows that every virtual Cartan inclusion is a pseudo-Cartan inclusion; furthermore, the pseudo-Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a virtual Cartan inclusion if and only if 𝒟{\mathcal{D}} is a MASA in 𝒞{\mathcal{C}}. By [PittsNoApUnInC*Al, Theorem 2.6], every virtual Cartan inclusion has the AUP.

Weak Cartan Inclusions:

Weak Cartan inclusions were defined in [ExelPittsChGrC*AlNoHaEtGr, Definition 2.11.5]. We shall show in Proposition 4.1 below that that every weak Cartan inclusion is a pseudo-Cartan inclusion. Since the axioms for weak Cartan inclusions include the AUP, the class of pseudo-Cartan inclusions properly contains the class of weak Cartan inclusions.

The example given in [ExelPittsChGrC*AlNoHaEtGr, Remark 2.11.6] shows that there exists a Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) (as in Definition 2.1(h) above) which is not a weak Cartan inclusion, so the classes of pseudo-Cartan inclusions with the AUP and weak Cartan inclusions are also distinct. Also, an example of a unital pseudo-Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) with 𝒞{\mathcal{C}} abelian which is not a weak Cartan inclusion is given in the paragraph just prior to [ExelPittsChGrC*AlNoHaEtGr, 2.11.16]. However, in both these examples, the algebra 𝒞{\mathcal{C}} is not separable. We do not know whether every pseudo-Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) with the AUP and such that 𝒞{\mathcal{C}} is separable is a weak Cartan inclusion.222During his lecture at the 2025 Canadian Operator Theory Symposium, Dan Ursu stated that he found an example of a separable pseudo-Cartan inclusion with the AUP which is not a weak Cartan inclusion.

Pseudo-Diagonals and Abelian Cores:

These are classes of regular inclusions which are not assumed to be regular. Nagy and Reznikoff define the notion of a pseudo-diagonal, see [NagyReznikoffPsDiUnTh, p. 268]. One of the requirements for the inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) to be a pseudo-diagonal is that there exists a faithful conditional expectation E:𝒞𝒟E:{\mathcal{C}}\rightarrow{\mathcal{D}}, however regularity of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is not required. Nagy and Reznikoff observe that when (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-diagonal, 𝒟{\mathcal{D}} is necessarily a MASA in 𝒞{\mathcal{C}}[NagyReznikoffPsDiUnTh, Corollary 3.2]. Thus, when (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular pseudo-diagonal, (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a Cartan inclusion and so is a pseudo-Cartan inclusion. Likewise, every regular abelian core (see [NagyReznikoffPsDiUnTh, p. 272]) is a Cartan inclusion and hence a pseudo-Cartan inclusion.

For a unital pseudo-Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), the relative commutant 𝒟c{\mathcal{D}}^{c} is a MASA in 𝒞{\mathcal{C}} and [PittsStReInI, Lemma 2.10] shows (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) is a regular inclusion. As (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) has the ideal intersection property (Definition 4.1(b)), it is a virtual Cartan inclusion. We now observe the same is true for possibly non-unital pseudo-Cartan inclusions.

Observation \the\numberby.

If (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion, then (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) is a virtual Cartan inclusion.

Proof.

Lemma 3(a) shows that (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) is a regular inclusion, so it is a regular MASA inclusion with the ideal intersection property. ∎

Our next goal is to show that every weak Cartan inclusion is a pseudo-Cartan inclusion. Before proceeding, we recall some terminology from [ExelPittsChGrC*AlNoHaEtGr].

Definition \the\numberby ([ExelPittsChGrC*AlNoHaEtGr, Definitions 2.7.2, 2.8.1 and 2.11.5]).

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be an inclusion and σ𝒟^\sigma\in\hat{\mathcal{D}}.

  1. (a)

    For x𝒞x\in{\mathcal{C}}, we say σ\sigma is free relative to xx if whenever f1,f2f_{1},f_{2} are states on 𝒞{\mathcal{C}} such that f1|𝒟=σ=f2|𝒟f_{1}|_{\mathcal{D}}=\sigma=f_{2}|_{\mathcal{D}}, we have f1(x)=f2(x)f_{1}(x)=f_{2}(x).

  2. (b)

    We say σ𝒟^\sigma\in\hat{\mathcal{D}} is free for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), or a free point when the context is clear, if σ\sigma uniquely extends to a state ff on 𝒞{\mathcal{C}}.

  3. (c)

    The inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is called topologically free if the set of free points for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is dense in 𝒟^\hat{\mathcal{D}}. Other terminology for topologically free inclusions may be found in the literature, for example, the almost extension property is used in [KwasniewskiMeyerApAlExPrUnPsEx, NagyReznikoffPsDiUnTh, ZarikianPuExPrDiCrPr] instead of ‘topologically free inclusion’.

Proposition \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a weak Cartan inclusion. Then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion.

Proof.

First note that the weak Cartan inclusions considered in [ExelPittsChGrC*AlNoHaEtGr] have the AUP and are regular (see [ExelPittsChGrC*AlNoHaEtGr, Definition 2.3.1, Hypothesis 2.3.2, and Definition 2.11.5]). Next, [ExelPittsChGrC*AlNoHaEtGr, Proposition 2.11.7] shows that (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is topologically free. Combining [KwasniewskiMeyerApAlExPrUnPsEx, Theorems 5.5 and 3.6], we find that (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the unique pseudo-expectation property; let E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be its pseudo-expectation. We must show EE is faithful.

Since (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP, (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a regular inclusion by Lemma 3(e). Corollary 2.3 shows E~\tilde{E} is the unique pseudo-expectation for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}). Let (𝒞~,𝒟~){\mathcal{L}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) be the left kernel of E~\tilde{E}. It follows from [PittsStReInII, Theorem 6.5] that (𝒞~,𝒟~){\mathcal{L}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is the largest ideal of 𝒞~\tilde{\mathcal{C}} having trivial intersection with 𝒟~\tilde{\mathcal{D}}.

If J𝒞J\unlhd{\mathcal{C}} has trivial intersection with 𝒟{\mathcal{D}}, then JJ is also an ideal of 𝒞~\tilde{\mathcal{C}} which has trivial intersection with 𝒟~\tilde{\mathcal{D}}. Thus J(𝒞~,𝒟~)J\subseteq{\mathcal{L}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}). Hence 1:=𝒞(𝒞~,𝒟~){\mathcal{L}}_{1}:={\mathcal{C}}\cap{\mathcal{L}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is the largest ideal of 𝒞{\mathcal{C}} having trivial intersection with 𝒟{\mathcal{D}}. Therefore, [ExelPittsChGrC*AlNoHaEtGr, Proposition 2.11.3] shows that 1{\mathcal{L}}_{1} is the grey ideal for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). The grey ideal in a weak Cartan inclusion is {0}\{0\} (see [ExelPittsChGrC*AlNoHaEtGr, Definition 2.11.5]), so 1={0}{\mathcal{L}}_{1}=\{0\}. But 1{\mathcal{L}}_{1} is the left kernel of EE, whence (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property. Theorem 3 then shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion. ∎

4.2. The AUP and Unitization of Cartan Envelopes

Given a weakly non-degenerate regular inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a unital inclusion by Lemma 2.2(e), but we have been unable to establish whether (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is regular. Two conditions which ensure that regularity of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) implies (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is also regular are: (i) (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP (Corollary 3); and (ii) 𝒞{\mathcal{C}} is abelian (because every unitary in 𝒞~\tilde{\mathcal{C}} normalizes 𝒟~\tilde{\mathcal{D}} and 𝒞~\tilde{\mathcal{C}} is spanned by its unitaries). On the other hand, [PittsNoApUnInC*Al, Example 3.1] shows that regularity of (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) does not imply regularity of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Theorem 2.3 (and Theorem 3) give the following.

Observation \the\numberby.

Suppose both (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) and (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) are regular inclusions. Then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion if and only if (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a pseudo-Cartan inclusion.

A pseudo-Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) may not have the AUP. Our next two results show that it is possible to enlarge 𝒟{\mathcal{D}} to obtain a pseudo-Cartan inclusion (𝒞,)({\mathcal{C}},{\mathcal{E}}) having the AUP such that the Cartan envelopes of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) and (𝒞,)({\mathcal{C}},{\mathcal{E}}) are the same. The significance of these results is that it is often possible to replace (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) with (𝒞,)({\mathcal{C}},{\mathcal{E}}) when the AUP is required. The first of these results shows that {\mathcal{E}} may be taken to be 𝒟c{\mathcal{D}}^{c}; it extends [PittsStReInII, Proposition 5.29(b)] to include not-necessarily unital inclusions.

Proposition \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion and consider the virtual Cartan inclusion (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}). Then (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) if and only if (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is Cartan envelope for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}).

Proof.

Proposition 3 shows that if (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}), then (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

Suppose (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Then we may find a Cartan envelope (𝒜, :: α)({\mathcal{A}}^{\prime},{\mathcal{B}}^{\prime}\hbox{\,:\hskip-1.0pt:\,}\alpha^{\prime}) for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}), which by Proposition 3, is also a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). By uniqueness of Cartan envelopes for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), there exists a unique regular *-isomorphism ψ:(𝒜,)(𝒜,)\psi:({\mathcal{A}}^{\prime},{\mathcal{B}}^{\prime})\rightarrow({\mathcal{A}},{\mathcal{B}}) such that ψα=α\psi\circ\alpha^{\prime}=\alpha. Since α:(𝒞,𝒟c)(𝒜,)\alpha^{\prime}:({\mathcal{C}},{\mathcal{D}}^{c})\rightarrow({\mathcal{A}}^{\prime},{\mathcal{B}}^{\prime}) is regular, we conclude that α:(𝒞,𝒟c)(𝒜,)\alpha:({\mathcal{C}},{\mathcal{D}}^{c})\rightarrow({\mathcal{A}},{\mathcal{B}}) is regular. Therefore, (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}). ∎

Given the pseudo-Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), examples show that 𝒟c{\mathcal{D}}^{c} need not be contained in 𝒩(𝒞,𝒟){\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). Nevertheless, by suitably enlarging 𝒟{\mathcal{D}}, we can use (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) to produce pseudo-Cartan inclusions (𝒞,)({\mathcal{C}},{\mathcal{E}}) having the AUP such that 𝒩(𝒞,𝒟){\mathcal{E}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) and which have the same Cartan envelope as (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). However, unlike 𝒟c{\mathcal{D}}^{c}, {\mathcal{E}} may not be a MASA in 𝒞{\mathcal{C}}.

Proposition \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a pseudo-Cartan inclusion. Suppose 𝒫𝒩(𝒞,𝒟){\mathcal{P}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) is a *-semigroup such that 𝒟𝒫{\mathcal{D}}\subseteq{\mathcal{P}}, and span𝒫\operatorname{span}{\mathcal{P}} is dense in 𝒞{\mathcal{C}}. Let

:=C({p𝒫:p0}).{\mathcal{E}}:=C^{*}(\{p\in{\mathcal{P}}:p\geq 0\}).

Then (𝒞,)({\mathcal{C}},{\mathcal{E}}) is a pseudo-Cartan inclusion having the AUP. Furthermore,

  1. (a)

    𝒟𝒟c{\mathcal{D}}\subseteq{\mathcal{E}}\subseteq{\mathcal{D}}^{c};

  2. (b)

    𝒩(𝒞,𝒟){\mathcal{E}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}); and

  3. (c)

    (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) if and only if (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,)({\mathcal{C}},{\mathcal{E}}).

Proof.

We first establish parts (a) and (b), then show (𝒞,)({\mathcal{C}},{\mathcal{E}}) is a pseudo-Cartan inclusion with the AUP, and finally show part (c).

The first inclusion in (a) follows from the fact that 𝒟𝒫{\mathcal{D}}\subseteq{\mathcal{P}} and 𝒟{\mathcal{D}} is the span of its positive elements. For the second, [PittsNoApUnInC*Al, Proposition 2.1] shows that for 0p𝒫0\leq p\in{\mathcal{P}}, p2𝒟cp^{2}\in{\mathcal{D}}^{c}, so p𝒟cp\in{\mathcal{D}}^{c}; thus 𝒟c{\mathcal{E}}\subseteq{\mathcal{D}}^{c}. This gives (a).

Before moving to part (b), we introduce some notation. Let us write,

𝒫+:={p𝒫:p0}.{\mathcal{P}}^{+}:=\{p\in{\mathcal{P}}:p\geq 0\}.

Since 𝒟c{\mathcal{D}}^{c} is abelian, 𝒫+{\mathcal{P}}^{+} is a *-semigroup, whence =span¯𝒫+{\mathcal{E}}=\overline{\operatorname{span}}\,{\mathcal{P}}^{+}.

We now turn to part (b). For p𝒫+p\in{\mathcal{P}}^{+} and and d𝒟d\in{\mathcal{D}}, [PittsNoApUnInC*Al, Proposition 2.1] gives p2dd=ddp2𝒟p^{2}d^{*}d=d^{*}dp^{2}\in{\mathcal{D}}. Taking square roots, we obtain p|d|=|d|p𝒟p|d|=|d|p\in{\mathcal{D}}. Since 𝒟{\mathcal{D}} is the span of its positive elements, we see that for every d𝒟d\in{\mathcal{D}} and p𝒫+p\in{\mathcal{P}}^{+},

pd=dp𝒟.pd=dp\in{\mathcal{D}}.

Thus, if xspan𝒫+x\in\operatorname{span}{\mathcal{P}}^{+}, xd𝒟xd\in{\mathcal{D}}, so xdx=xdx=xxd𝒟x^{*}dx=xdx^{*}=x^{*}xd\in{\mathcal{D}} also. Therefore, span𝒫+𝒩(𝒞,𝒟)\operatorname{span}{\mathcal{P}}^{+}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). As 𝒩(𝒞,𝒟){\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) is closed, part (b) holds.

We next show (𝒞,)({\mathcal{C}},{\mathcal{E}}) is a regular inclusion. For any v𝒫v\in{\mathcal{P}}, v𝒫+vv𝒫+v𝒫+v^{*}{\mathcal{P}}^{+}v\cup v{\mathcal{P}}^{+}v^{*}\subseteq{\mathcal{P}}^{+}, whence vvvvv{\mathcal{E}}v^{*}\cup v^{*}{\mathcal{E}}v\subseteq{\mathcal{E}}. Therefore, 𝒫𝒩(𝒞,){\mathcal{P}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{E}}), and because 𝒞=span¯𝒫{\mathcal{C}}=\overline{\operatorname{span}}\,{\mathcal{P}}, (𝒞,)({\mathcal{C}},{\mathcal{E}}) is a regular inclusion.

Note that c=𝒟c{\mathcal{E}}^{c}={\mathcal{D}}^{c}. Thus, part (a) and Lemma 2.4 show (c,)({\mathcal{E}}^{c},{\mathcal{E}}) has the ideal intersection property. As (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion, (𝒞,c)({\mathcal{C}},{\mathcal{E}}^{c}) has the ideal intersection property, so (𝒞,)({\mathcal{C}},{\mathcal{E}}) is a pseudo-Cartan inclusion.

To show (𝒞,)({\mathcal{C}},{\mathcal{E}}) has the AUP, let (uλ)(u_{\lambda}) be an approximate unit for {\mathcal{E}}. Let ε>0\varepsilon>0. For v𝒫v\in{\mathcal{P}}, we may choose nn\in{\mathbb{N}} so that v(vv)1/nv<ε/3\left\lVert v(v^{*}v)^{1/n}-v\right\rVert<\varepsilon/3. Since (vv)1/n(v^{*}v)^{1/n}\in{\mathcal{E}}, there exists λ0\lambda_{0} so that for λλ0\lambda\geq\lambda_{0}, v(vv)1/nuλ(vv)1/n<ε/3.\left\lVert v\right\rVert\left\lVert(v^{*}v)^{1/n}u_{\lambda}-(v^{*}v)^{1/n}\right\rVert<\varepsilon/3. For λλ0\lambda\geq\lambda_{0},

vuλv\displaystyle\left\lVert vu_{\lambda}-v\right\rVert vuλv(vv)1/nuλ+v(vv)1/nuλv(vv)1/n+v(vv)1/nv\displaystyle\leq\left\lVert vu_{\lambda}-v(v^{*}v)^{1/n}u_{\lambda}\right\rVert+\left\lVert v(v^{*}v)^{1/n}u_{\lambda}-v(vv^{*})^{1/n}\right\rVert+\left\lVert v(v^{*}v)^{1/n}-v\right\rVert
2vv(vv)1/n+v(vv)1/nuλ(vv)1/n<ε.\displaystyle\leq 2\left\lVert v-v(v^{*}v)^{1/n}\right\rVert+\left\lVert v\right\rVert\left\lVert(v^{*}v)^{1/n}u_{\lambda}-(v^{*}v)^{1/n}\right\rVert<\varepsilon.

Therefore vuλvvu_{\lambda}\rightarrow v; similarly uλvvu_{\lambda}v\rightarrow v. Since span𝒫\operatorname{span}{\mathcal{P}} is dense in 𝒞{\mathcal{C}}, it follows that (𝒞,)({\mathcal{C}},{\mathcal{E}}) has the AUP.

Finally, using Proposition 4.2 and the equality c=𝒟c{\mathcal{E}}^{c}={\mathcal{D}}^{c},

(𝒜, :: α) is a Cartan envelope for (𝒞,𝒟)\displaystyle({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha)\text{ is a Cartan envelope for }({\mathcal{C}},{\mathcal{D}})\iff (𝒜, :: α) is a Cartan envelope for\displaystyle({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha)\text{ is a Cartan envelope for }
(𝒞,𝒟c)=(𝒞,c)\displaystyle({\mathcal{C}},{\mathcal{D}}^{c})=({\mathcal{C}},{\mathcal{E}}^{c})
\displaystyle\iff (𝒜, :: α) is a Cartan envelope for\displaystyle({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha)\text{ is a Cartan envelope for }
(𝒞,).\displaystyle({\mathcal{C}},{\mathcal{E}}).\hfill\qed

As noted in Corollary 3, for the weakly non-degenerate and non-unital inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), the inclusion mapping u𝒞u_{\mathcal{C}} of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) into (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is regular if and only if (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP. We now study the relationships between the Cartan envelopes of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) and (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) when (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP. We begin with a lemma concerning the image under α\alpha of an approximate unit for 𝒟{\mathcal{D}} in a Cartan package (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha).

Lemma \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is an inclusion with the AUP and (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). If (uλ)𝒟(u_{\lambda})\subseteq{\mathcal{D}} is an approximate unit for 𝒞{\mathcal{C}}, then (α(uλ))(\alpha(u_{\lambda}))\subseteq{\mathcal{B}} is an approximate unit for 𝒜{\mathcal{A}}.

Proof.

Let Δ:𝒜\Delta:{\mathcal{A}}\rightarrow{\mathcal{B}} be the conditional expectation. For x𝒞x\in{\mathcal{C}}, we have

limλΔ(α(x))α(uλ)=limλΔ(α(xuλ))=Δ(α(x)).\lim_{\lambda}\Delta(\alpha(x))\alpha(u_{\lambda})=\lim_{\lambda}\Delta(\alpha(xu_{\lambda}))=\Delta(\alpha(x)).

Since {\mathcal{B}} is generated by Δ(α(𝒞))\Delta(\alpha({\mathcal{C}})), it follows that (α(uλ))(\alpha(u_{\lambda})) is an approximate unit for {\mathcal{B}}. As (𝒜,)({\mathcal{A}},{\mathcal{B}}) is a regular MASA inclusion, an application of [PittsNoApUnInC*Al, Theorem 2.6] shows (α(uλ))(\alpha(u_{\lambda})) is an approximate unit for 𝒜{\mathcal{A}}. ∎

Lemma \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular inclusion with the AUP such that 𝒞{\mathcal{C}} is not unital. Suppose (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan package (resp. Cartan envelope) for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Then (𝒜~,~ :: α~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\tilde{\alpha}) is a Cartan package (resp. Cartan envelope) for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}).

Proof.

We will use Δ:𝒜\Delta:{\mathcal{A}}\rightarrow{\mathcal{B}} for the conditional expectation. Since (𝒜,)({\mathcal{A}},{\mathcal{B}}) is a Cartan pair, so is (𝒜~,~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}}); also, Δ~:𝒜~~\tilde{\Delta}:\tilde{\mathcal{A}}\rightarrow\tilde{\mathcal{B}} is the conditional expectation of 𝒜~\tilde{\mathcal{A}} onto ~\tilde{\mathcal{B}}.

That 𝒜~\tilde{\mathcal{A}} is generated by α~(𝒞~)Δ~(α~(𝒞~))\tilde{\alpha}(\tilde{\mathcal{C}})\cup\tilde{\Delta}(\tilde{\alpha}(\tilde{\mathcal{C}})) follows from the fact that 𝒜{\mathcal{A}} is generated by α(𝒞)Δ(α(𝒞))\alpha({\mathcal{C}})\cup\Delta(\alpha({\mathcal{C}})). Similarly, ~\tilde{\mathcal{B}} is generated by Δ~(π(𝒞~))\tilde{\Delta}(\pi(\tilde{\mathcal{C}})). It remains to show that α~\tilde{\alpha} is a regular *-homomorphism, that is, α~(𝒩(𝒞~,𝒟~))𝒩(𝒜~,~)\tilde{\alpha}({\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}))\subseteq{\mathcal{N}}(\tilde{\mathcal{A}},\tilde{\mathcal{B}}). For this, we use the relative strict topology on 𝒜~\tilde{\mathcal{A}} , which we now describe.

For each x𝒜x\in{\mathcal{A}}, the map 𝒜~axa+ax\tilde{\mathcal{A}}\ni a\mapsto\left\lVert xa\right\rVert+\left\lVert ax\right\rVert is a seminorm on 𝒜~\tilde{\mathcal{A}}, and the relative strict topology on 𝒜~\tilde{\mathcal{A}} is the smallest topology on 𝒜~\tilde{\mathcal{A}} making each of these seminorms continuous. Thus, a net (aλ)(a_{\lambda}) in 𝒜~\tilde{\mathcal{A}} converges to a𝒜~a\in\tilde{\mathcal{A}} in the relative strict topology if and only if for every x𝒜x\in{\mathcal{A}},

limλ((aλa)x+(x(aλa))=0.\lim_{\lambda}(\left\lVert(a_{\lambda}-a)x\right\rVert+\left\lVert(x(a_{\lambda}-a)\right\rVert)=0.

We write a=rslimλaλa=^{rs}\lim_{\lambda}a_{\lambda} when the net (aλ)(a_{\lambda}) in 𝒜~\tilde{\mathcal{A}} converges in the relative strict topology to a𝒜~a\in\tilde{\mathcal{A}}. The relative strict topology on 𝒜~\tilde{\mathcal{A}} is locally convex, and since 𝒜{\mathcal{A}} is an essential ideal in 𝒜~\tilde{\mathcal{A}}, it is Hausdorff. Routine arguments show that when 𝒜~\tilde{\mathcal{A}} is equipped with the relative strict topology, the adjoint operation is continuous and multiplication is jointly continuous on norm-bounded subsets of 𝒜~\tilde{\mathcal{A}}.

We claim that if (wλ)(w_{\lambda}) is a bounded net in 𝒩(𝒜~,~){\mathcal{N}}(\tilde{\mathcal{A}},\tilde{\mathcal{B}}) which converges relative strictly to w𝒜~w\in\tilde{\mathcal{A}}, then w𝒩(𝒜~,~)w\in{\mathcal{N}}(\tilde{\mathcal{A}},\tilde{\mathcal{B}}). Indeed, for h,k~h,k\in\tilde{\mathcal{B}}, as wλhwλ~w_{\lambda}^{*}hw_{\lambda}\in\tilde{\mathcal{B}},

whwk=limλrswλhwλk=limλrskwλhwλ=kwhw.w^{*}hwk={}^{rs}\lim_{\lambda}w_{\lambda}^{*}hw_{\lambda}k={}^{rs}\lim_{\lambda}kw_{\lambda}^{*}hw_{\lambda}=kw^{*}hw.

As ~\tilde{\mathcal{B}} is a MASA in 𝒜~\tilde{\mathcal{A}}, whw~w^{*}hw\in\tilde{\mathcal{B}}. Likewise whw~whw^{*}\in\tilde{\mathcal{B}}, so our claim holds.

Now let (uλ)(u_{\lambda}) be a net in 𝒟{\mathcal{D}} which is an approximate unit for 𝒞{\mathcal{C}}. Choose v𝒩(𝒞~,𝒟~)v\in{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) and write v=(x,ξ)v=(x,\xi) for some x𝒞x\in{\mathcal{C}} and ξ\xi\in{\mathbb{C}}. For each λ\lambda, uλ𝒩(𝒞,𝒟)u_{\lambda}\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), so (since 𝒩(𝒞~,𝒟~){\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is closed under multiplication) vuλ=(xuλ+ξuλ,0)𝒩(𝒞,𝒟)vu_{\lambda}=(xu_{\lambda}+\xi u_{\lambda},0)\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). The regularity of the map α\alpha and the fact that 𝒩(𝒜,)𝒩(𝒜~,~){\mathcal{N}}({\mathcal{A}},{\mathcal{B}})\subseteq{\mathcal{N}}(\tilde{\mathcal{A}},\tilde{\mathcal{B}}) (Lemma 3(c)) gives

𝒩(𝒜,)α(vuλ)=α~(vuλ)𝒩(𝒜~,~).{\mathcal{N}}({\mathcal{A}},{\mathcal{B}})\ni\alpha(vu_{\lambda})=\tilde{\alpha}(vu_{\lambda})\in{\mathcal{N}}(\tilde{\mathcal{A}},\tilde{\mathcal{B}}).

By Lemma 4.2, α(uλ)\alpha(u_{\lambda}) is an approximate unit for 𝒜{\mathcal{A}}. Therefore, α(uλ)\alpha(u_{\lambda}) converges in the relative strict topology to I𝒜~I_{\tilde{\mathcal{A}}}. As α~(vuλ)\tilde{\alpha}(vu_{\lambda}) is a bounded net in 𝒩(𝒜~,~){\mathcal{N}}(\tilde{\mathcal{A}},\tilde{\mathcal{B}}), the claim gives

α~(v)=α~(v)limλrsα(uλ)=limλrsα~(vuλ)𝒩(𝒜~,~).\tilde{\alpha}(v)=\tilde{\alpha}(v)\,{}^{rs}\lim_{\lambda}\alpha(u_{\lambda})={}^{rs}\lim_{\lambda}\tilde{\alpha}(vu_{\lambda})\in{\mathcal{N}}(\tilde{\mathcal{A}},\tilde{\mathcal{B}}).

Thus α~\tilde{\alpha} is a regular *-monomorphism. This completes the proof that (𝒜~,~ :: α~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\tilde{\alpha}) is a Cartan package for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}).

Now suppose that (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). It remains to establish part (b) of Definition 3. By hypothesis, (,α(𝒟))({\mathcal{B}},\alpha({\mathcal{D}})) has the ideal intersection property, so (~,α~(𝒟~))=(~,(α(𝒟)))(\tilde{\mathcal{B}},\tilde{\alpha}(\tilde{\mathcal{D}}))=(\tilde{\mathcal{B}},(\alpha({\mathcal{D}}))^{\sim}) has the ideal intersection property by Lemma 2.4. ∎

The following result shows that in some settings, a Cartan envelope for a non-unital inclusion may be constructed from a Cartan envelope for its unitization.

Lemma \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular inclusion having the AUP with 𝒞{\mathcal{C}} not unital. Suppose (𝒜, :: π)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\pi) is a Cartan envelope for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) and write Δ\Delta for the conditional expectation of 𝒜{\mathcal{A}} onto {\mathcal{B}}. Put

α:=π|𝒞,𝒟1:=C(Δ(α(𝒞))),and𝒞1:=C(α(𝒞)𝒟1).\alpha:=\pi|_{\mathcal{C}},\quad{\mathcal{D}}_{1}:=C^{*}(\Delta(\alpha({\mathcal{C}}))),\quad\text{and}\quad{\mathcal{C}}_{1}:=C^{*}(\alpha({\mathcal{C}})\cup{\mathcal{D}}_{1}).

Then (𝒞1,𝒟1 :: α)({\mathcal{C}}_{1},{\mathcal{D}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}); further, 𝒞1𝒜{\mathcal{C}}_{1}\unlhd{\mathcal{A}} and 𝒟1{\mathcal{D}}_{1}\unlhd{\mathcal{B}} are ideals having codimension one.

Proof.

Note that since (,𝒟~,π)({\mathcal{B}},\tilde{\mathcal{D}},\pi) and (𝒟~,𝒟)(\tilde{\mathcal{D}},{\mathcal{D}}) have the ideal intersection property, (,𝒟,π)({\mathcal{B}},{\mathcal{D}},\pi) also has the ideal intersection property. Proposition 3 shows (𝒞1,𝒟1 :: α)({\mathcal{C}}_{1},{\mathcal{D}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

Let I𝒜I_{\mathcal{A}} denote the unit of 𝒜{\mathcal{A}} and consider the CC^{*}-algebras

𝒟1+=𝒟1+I𝒜and𝒞1+=𝒞1+I𝒜.{\mathcal{D}}_{1}^{+}={\mathcal{D}}_{1}+{\mathbb{C}}I_{\mathcal{A}}\quad\text{and}\quad{\mathcal{C}}_{1}^{+}={\mathcal{C}}_{1}+{\mathbb{C}}I_{\mathcal{A}}.

Obviously, 𝒟1+{\mathcal{D}}_{1}^{+}\subseteq{\mathcal{B}}, and 𝒞1+𝒜{\mathcal{C}}_{1}^{+}\subseteq{\mathcal{A}}; Lemma 3 gives π(0,1)=I𝒜\pi(0,1)=I_{\mathcal{A}}. Since (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}), {\mathcal{B}} is generated by π(Δ(𝒞~))=π(Δ(𝒞))+I𝒜\pi(\Delta(\tilde{\mathcal{C}}))=\pi(\Delta({\mathcal{C}}))+{\mathbb{C}}I_{\mathcal{A}} and 𝒜{\mathcal{A}} is generated by π(𝒞~)=(π(𝒞)+I𝒜){\mathcal{B}}\cup\pi(\tilde{\mathcal{C}})={\mathcal{B}}\cup(\pi({\mathcal{C}})+{\mathbb{C}}I_{\mathcal{A}}). 𝒞1+=𝒜{\mathcal{C}}_{1}^{+}={\mathcal{A}}. Thus,

(4.2.1) (𝒞~1,𝒟~1)(𝒞1+,𝒟1+)=(𝒜,).(\tilde{\mathcal{C}}_{1},\tilde{\mathcal{D}}_{1})\simeq({\mathcal{C}}_{1}^{+},{\mathcal{D}}_{1}^{+})=({\mathcal{A}},{\mathcal{B}}).

That 𝒞1𝒜{\mathcal{C}}_{1}\unlhd{\mathcal{A}} and 𝒟1{\mathcal{D}}_{1}\unlhd{\mathcal{B}} are ideals of codimension one follows from (4.2.1). ∎

Theorem \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular inclusion with the AUP such that 𝒞{\mathcal{C}} is not unital. Then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion if and only if (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a pseudo-Cartan inclusion. Moreover, the following statements hold.

  1. (a)

    If (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), then (𝒜~,~ :: α~)(\tilde{\mathcal{A}},\tilde{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\tilde{\alpha}) is a Cartan envelope for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}).

  2. (b)

    Suppose (𝒜, :: π)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\pi) is a Cartan envelope for (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) and Δ:𝒜\Delta:{\mathcal{A}}\rightarrow{\mathcal{B}} is the conditional expectation. Let

    α:=π|𝒞,𝒟1=C(Δ(π(𝒞))),and𝒞1=C(π(𝒞)𝒟1).\alpha:=\pi|_{{\mathcal{C}}},\quad{\mathcal{D}}_{1}=C^{*}(\Delta(\pi({\mathcal{C}}))),\quad\text{and}\quad{\mathcal{C}}_{1}=C^{*}(\pi({\mathcal{C}})\cup{\mathcal{D}}_{1}).

    Then 𝒟1{\mathcal{D}}_{1}\unlhd{\mathcal{B}} and 𝒞1𝒜{\mathcal{C}}_{1}\unlhd{\mathcal{A}} are ideals with codimension 1 and (𝒞1,𝒟1 :: α)({\mathcal{C}}_{1},{\mathcal{D}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

Proof.

That (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion if and only if (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a pseudo-Cartan inclusion follows by combining Lemma 3(e) and Observation 4.2. Lemma 4.2 gives part (a), and Lemma 4.2 gives part (b). ∎

4.3. Some Properties Shared by a Pseudo-Cartan Inclusion and its Cartan Envelope

This subsection is devoted to establishing Theorem 4.3, which shows that certain desirable properties are common to both a pseudo-Cartan inclusion and its Cartan envelope.

Recall that Definition 3(a) states that (𝒞1,𝒟1 :: α)({\mathcal{C}}_{1},{\mathcal{D}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an essential expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) if the inclusion (𝒟1,𝒟,α|𝒟)({\mathcal{D}}_{1},{\mathcal{D}},\alpha|_{{\mathcal{D}}}) has the ideal intersection property. We next observe that more can be said for essential expansions in the context of pseudo-Cartan inclusions.

Observation \the\numberby.

For i=1,2i=1,2, suppose (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are pseudo-Cartan inclusions, and (𝒞2,𝒟2 :: α)({\mathcal{C}}_{2},{\mathcal{D}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an essential expansion of (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}). Then the inclusion (𝒞2,𝒞1,α)({\mathcal{C}}_{2},{\mathcal{C}}_{1},\alpha) has the ideal intersection property.

Proof.

Suppose J𝒞2J\unlhd{\mathcal{C}}_{2} and Jα(𝒞1)={0}J\cap\alpha({\mathcal{C}}_{1})=\{0\}. Then Jα(𝒟1)={0}J\cap\alpha({\mathcal{D}}_{1})=\{0\}. By hypothesis, (𝒟2,α(𝒟1))({\mathcal{D}}_{2},\alpha({\mathcal{D}}_{1})) has the ideal intersection property, so J𝒟2={0}J\cap{\mathcal{D}}_{2}=\{0\}. Since (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) is a pseudo-Cartan inclusion, both (𝒟2c,𝒟2)({\mathcal{D}}_{2}^{c},{\mathcal{D}}_{2}) and (𝒞2,𝒟2c)({\mathcal{C}}_{2},{\mathcal{D}}_{2}^{c}) have the ideal intersection property. As (J𝒟2c)𝒟2J𝒟2={0}(J\cap{\mathcal{D}}_{2}^{c})\cap{\mathcal{D}}_{2}\subseteq J\cap{\mathcal{D}}_{2}=\{0\}, we find J𝒟2c={0}J\cap{\mathcal{D}}_{2}^{c}=\{0\}, whence J={0}J=\{0\}. ∎

Lemma \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a pseudo-Cartan inclusion with Cartan envelope (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha). Let J𝒞J\unlhd{\mathcal{C}}, and define J1𝒜J_{1}\subseteq{\mathcal{A}} to be the norm-closed {\mathcal{B}}-bimodule generated by α(J)\alpha(J). Then J1𝒜J_{1}\unlhd{\mathcal{A}}.

Proof.

Suppose xJx\in J, hh\in{\mathcal{B}} and v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). We first show that

(4.3.1) α(v)hα(x)J1.\alpha(v)h\alpha(x)\in J_{1}.

Since α\alpha is a regular map, α(v)𝒩(𝒜,)\alpha(v)\in{\mathcal{N}}({\mathcal{A}},{\mathcal{B}}). By Lemma 3, given uα(vv)¯u\in\overline{\alpha(v^{*}v){\mathcal{B}}},

α(v)uhα(x)=θα(v)1(uh)α(v)α(x).\alpha(v)uh\alpha(x)=\theta_{\alpha(v)}^{-1}(uh)\alpha(v)\alpha(x).

As θα(v)1(uh)α(vv)¯\theta_{\alpha(v)}^{-1}(uh)\in\overline{\alpha(vv^{*}){\mathcal{B}}} and α(vx)α(J)\alpha(vx)\in\alpha(J), we see that

(4.3.2) α(v)uhα(x)J1.\alpha(v)uh\alpha(x)\in J_{1}.

Since (𝒜,)({\mathcal{A}},{\mathcal{B}}) is Cartan and α\alpha is a regular map, Lemma 3(b) gives α(vv)c=\alpha(v^{*}v)\in{\mathcal{B}}^{c}={\mathcal{B}}. By considering an approximate unit for {\mathcal{B}}, we find α(vv)α(vv)¯\alpha(v^{*}v)\in\overline{\alpha(v^{*}v){\mathcal{B}}}. Since α(v)α(vv)1/nα(v)\alpha(v)\alpha(v^{*}v)^{1/n}\rightarrow\alpha(v), taking u=α(vv)1/nu=\alpha(v^{*}v)^{1/n} in (4.3.2), we obtain (4.3.1).

Next using (4.3.1), we see that if nn\in{\mathbb{N}} and xiJx_{i}\in J, hi,kih_{i},k_{i}\in{\mathcal{B}} for 1in1\leq i\leq n and v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}),

α(v)(i=1nhiα(xi)ki)J1,\alpha(v)\left(\sum_{i=1}^{n}h_{i}\alpha(x_{i})k_{i}\right)\in J_{1},

from which it follows that

(4.3.3) α(v)J1J1.\alpha(v)J_{1}\subseteq J_{1}.

Let MM be the set of all finite products of elements of α(𝒩(𝒞,𝒟)){\mathcal{B}}\cup\alpha({\mathcal{N}}({\mathcal{C}},{\mathcal{D}})). An induction argument using (4.3.3) shows that for wMw\in M and x1J1x_{1}\in J_{1}, wx1J1wx_{1}\in J_{1}. As spanM\operatorname{span}M is dense in 𝒜{\mathcal{A}}, we conclude that J1J_{1} is a left ideal of 𝒜{\mathcal{A}}. Similar arguments show J1J_{1} is a right ideal, so J1J_{1}\unlhd{\mathcal{B}}. ∎

Theorem \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a pseudo-Cartan inclusion with Cartan envelope (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha). Then

  1. (a)

    𝒞{\mathcal{C}} is simple if and only if 𝒜{\mathcal{A}} is simple; and

  2. (b)

    𝒞{\mathcal{C}} is unital if and only if 𝒜{\mathcal{A}} is unital.

  3. (c)

    𝒞{\mathcal{C}} is separable if and only if 𝒜{\mathcal{A}} is separable.

Proof.

Proposition 4.2 shows that (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) is also a Cartan envelope for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}). Thus, by replacing 𝒟{\mathcal{D}} with 𝒟c{\mathcal{D}}^{c} if necessary, without loss of generality we may assume (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion with the AUP.

Let Δ:𝒜\Delta:{\mathcal{A}}\rightarrow{\mathcal{B}} be the conditional expectation and let E:𝒞E:{\mathcal{C}}\rightarrow{\mathcal{B}} be the map

E:=Δα.E:=\Delta\circ\alpha.

(a) Suppose 𝒞{\mathcal{C}} is simple and let J𝒜J\unlhd{\mathcal{A}}. Then Jα(𝒞){{0},α(𝒞)}J\cap\alpha({\mathcal{C}})\in\{\{0\},\alpha({\mathcal{C}})\}. If Jα(𝒞)={0}J\cap\alpha({\mathcal{C}})=\{0\}, Observation 4.3 shows J={0}J=\{0\}. On the other hand, suppose Jα(𝒞)=α(𝒞)J\cap\alpha({\mathcal{C}})=\alpha({\mathcal{C}}). Since we are assuming (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP, we may choose a net (eλ)(e_{\lambda}) in 𝒟{\mathcal{D}} which is an approximate unit for 𝒞{\mathcal{C}}. As =C(E(𝒞)){\mathcal{B}}=C^{*}(E({\mathcal{C}})), we find (α(eλ))(\alpha(e_{\lambda})) is an approximate unit for {\mathcal{B}}. But (𝒜,)({\mathcal{A}},{\mathcal{B}}) is a Cartan inclusion, so by [PittsNoApUnInC*Al, Theorem 2.6], every approximate unit for {\mathcal{B}} is an approximate unit for 𝒜{\mathcal{A}}. As α(𝒟)α(𝒞)J\alpha({\mathcal{D}})\subseteq\alpha({\mathcal{C}})\subseteq J, we conclude that JJ contains an approximate unit for 𝒜{\mathcal{A}}, whence J=𝒜J={\mathcal{A}}. Thus 𝒜{\mathcal{A}} is simple.

For the converse, suppose 𝒜{\mathcal{A}} is simple. Let J𝒞J\unlhd{\mathcal{C}} be a non-zero ideal, let KK be the ideal in {\mathcal{B}} generated by α(J𝒟)\alpha(J\cap{\mathcal{D}}) and put

J:={x𝒞:E(xx)K}.J^{\prime}:=\{x\in{\mathcal{C}}:E(x^{*}x)\in K\}.

We aim to show JJ^{\prime} is a non-zero ideal in 𝒞{\mathcal{C}}. By construction, JJ^{\prime} is a closed subset of 𝒞{\mathcal{C}}.

Next we show JJ^{\prime} is a non-zero subspace of 𝒞{\mathcal{C}}. For τ^\tau\in\hat{\mathcal{B}}, τE\tau\circ E is a state on 𝒞{\mathcal{C}}, so the map,

𝒞x(τ(E(xx)))1/2{\mathcal{C}}\ni x\mapsto(\tau(E(x^{*}x)))^{1/2}

is a semi-norm on 𝒞{\mathcal{C}}. Thus for x,y𝒞x,y\in{\mathcal{C}}, τ(E((x+y)(x+y)))1/2τ(E(xx))1/2+τ(E(yy))1/2\tau(E((x+y)^{*}(x+y)))^{1/2}\leq\tau(E(x^{*}x))^{1/2}+\tau(E(y^{*}y))^{1/2}. Allowing τ\tau to vary throughout ^\hat{\mathcal{B}}, we conclude

E((x+y)(x+y))(E(xx)1/2+E(yy)1/2)2.E((x+y)^{*}(x+y))\leq(E(x^{*}x)^{1/2}+E(y^{*}y)^{1/2})^{2}.

Thus when x,yJx,y\in J^{\prime} we obtain x+yJx+y\in J^{\prime}. That JJ^{\prime} is invariant under scalar multiplication is obvious, so JJ^{\prime} is a closed linear subspace of 𝒞{\mathcal{C}}. Clearly J𝒟JJ\cap{\mathcal{D}}\subseteq J^{\prime}. Since (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) and (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) both have the ideal intersection property, J𝒟{0}J\cap{\mathcal{D}}\neq\{0\}, and hence J{0}J^{\prime}\neq\{0\}.

We are now ready to show JJ^{\prime} is an ideal. For v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), v(J𝒟)vJ𝒟v^{*}(J\cap{\mathcal{D}})v\subseteq J\cap{\mathcal{D}}, and regularity of α\alpha gives α(v)𝒩(𝒜,)\alpha(v)\in{\mathcal{N}}({\mathcal{A}},{\mathcal{B}}). So for bb\in{\mathcal{B}} and dJDd\in J\cap D, we have

α(v)α(d)bα(v)\displaystyle\alpha(v^{*})\alpha(d)b\alpha(v) =limα(v)α(d)bα(vv)1/nα(v)\displaystyle=\lim\alpha(v^{*})\alpha(d)b\alpha(vv^{*})^{1/n}\alpha(v)
=lim(α(v)α(d)α(v))θα(v)(α(vv)1/nb)K.\displaystyle=\lim(\alpha(v)^{*}\alpha(d)\alpha(v))\theta_{\alpha(v)}(\alpha(vv^{*})^{1/n}b)\in K.

It follows that

α(v)Kα(v)K,v𝒩(𝒞,𝒟).\alpha(v)^{*}K\alpha(v)\subseteq K,\quad v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}).

Therefore, when v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) and xJx\in J^{\prime}, xvJxv\in J^{\prime} because

E(vxxv)\displaystyle E(v^{*}x^{*}xv) =Δ(α(v)α(xx)α(v))=α(v)Δ(α(xx))α(v)\displaystyle=\Delta(\alpha(v)^{*}\alpha(x^{*}x)\alpha(v))=\alpha(v)^{*}\Delta(\alpha(x^{*}x))\alpha(v)
=α(v)E(xx)α(v)K.\displaystyle=\alpha(v)^{*}E(x^{*}x)\alpha(v)\in K.

Since span𝒩(𝒞,𝒟)\operatorname{span}{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) is dense in 𝒞{\mathcal{C}}, we conclude that JJ^{\prime} is a right ideal in 𝒞{\mathcal{C}}. For xJx\in J^{\prime} and y𝒞y\in{\mathcal{C}}, E(xyyx)y2E(xx)E(x^{*}y^{*}yx)\leq\left\lVert y\right\rVert^{2}E(x^{*}x), whence JJ^{\prime} is a left ideal in 𝒞{\mathcal{C}}. Thus {0}J𝒞\{0\}\neq J^{\prime}\unlhd{\mathcal{C}}.

Next we show that J𝒟=𝒟J\cap{\mathcal{D}}={\mathcal{D}}. For this, let σ𝒟^\sigma\in\hat{\mathcal{D}} and let τ^\tau\in\hat{\mathcal{B}} satisfy τα=σ\tau\circ\alpha=\sigma. Let MM be the norm-closed {\mathcal{B}}-bimodule generated by α(J)\alpha(J^{\prime}). By Lemma 4.3 and the hypothesis that 𝒜{\mathcal{A}} is simple,

M=𝒜.M={\mathcal{A}}.

Let 0b0\leq b\in{\mathcal{B}} be chosen so that b=1\left\lVert b\right\rVert=1 and τ(b)=1\tau(b)=1. Since bMb\in M there is nn\in{\mathbb{N}}, xiJx_{i}\in J^{\prime}, and hi,kih_{i},k_{i}\in{\mathcal{B}} (1in1\leq i\leq n) so that

bi=1nhiα(xi)ki<1/2.\left\lVert b-\sum_{i=1}^{n}h_{i}\alpha(x_{i})k_{i}\right\rVert<1/2.

Then

|τ(Δ(bi=1nhiα(xi)ki))|=|1i=1nτ(hi)τ(Δ(α(xi)))τ(ki)|<1/2.\left|\tau\left(\Delta\left(b-\sum_{i=1}^{n}h_{i}\alpha(x_{i})k_{i}\right)\right)\right|=\left|1-\sum_{i=1}^{n}\tau(h_{i})\tau(\Delta(\alpha(x_{i})))\tau(k_{i})\right|<1/2.

It follows that there exists xJx\in J^{\prime} such that τ(Δ(α(x)))0\tau(\Delta(\alpha(x)))\neq 0. Hence

0|τ(Δ(α(x)))|2=τ(E(x)E(x))τ(E(xx)).0\neq|\tau(\Delta(\alpha(x)))|^{2}=\tau(E(x^{*})E(x))\leq\tau(E(x^{*}x)).

Since xJx\in J^{\prime}, E(xx)KE(x^{*}x)\in K, and thus τ\tau does not annihilate KK. But KK is the ideal of {\mathcal{B}} generated by α(J𝒟)\alpha(J\cap{\mathcal{D}}), so τ\tau does not annihilate α(J𝒟)\alpha(J\cap{\mathcal{D}}). Therefore, σ|J𝒟0\sigma|_{J\cap{\mathcal{D}}}\neq 0. Since this holds for all σ𝒟^\sigma\in\hat{\mathcal{D}}, we conclude that J𝒟=𝒟J\cap{\mathcal{D}}={\mathcal{D}}.

Finally, the fact that 𝒟{\mathcal{D}} contains an approximate unit for 𝒞{\mathcal{C}} implies that J=𝒞J={\mathcal{C}}. Thus 𝒞{\mathcal{C}} is simple.

(b) Suppose 𝒞{\mathcal{C}} is unital. Lemma 3 shows 𝒜{\mathcal{A}} is unital and α(I𝒞)=I𝒜\alpha(I_{\mathcal{C}})=I_{\mathcal{A}}.

Now suppose 𝒜{\mathcal{A}} is unital. Let (uλ)𝒟(u_{\lambda})\subseteq{\mathcal{D}} be an approximate unit for 𝒞{\mathcal{C}}. Choose x1,,xn𝒞x_{1},\dots,x_{n}\in{\mathcal{C}} and let

z=k=1nΔ(α(xk)).z=\prod_{k=1}^{n}\Delta(\alpha(x_{k})).

Since Δ(α(xn))α(uλ)=Δ(α(xnuλ))\Delta(\alpha(x_{n}))\,\alpha(u_{\lambda})=\Delta(\alpha(x_{n}u_{\lambda})), zα(uλ)zz\alpha(u_{\lambda})\rightarrow z; likewise, α(uλ)zz\alpha(u_{\lambda})z\rightarrow z. As the collection of all finite products of Δ(α(𝒞))\Delta(\alpha({\mathcal{C}})) has dense span in {\mathcal{B}}, it follows that α(uλ)\alpha(u_{\lambda}) is an approximate unit for {\mathcal{B}}, and hence also for 𝒜{\mathcal{A}} (because (𝒜,)({\mathcal{A}},{\mathcal{B}}) has the AUP). Therefore,

α(uλ)I𝒜I𝒜=α(uλ)I𝒜0.\left\lVert\alpha(u_{\lambda})I_{\mathcal{A}}-I_{\mathcal{A}}\right\rVert=\left\lVert\alpha(u_{\lambda})-I_{\mathcal{A}}\right\rVert\rightarrow 0.

Hence (uλ)(u_{\lambda}) converges to an element e𝒞e\in{\mathcal{C}}. Since α(e)=I𝒜\alpha(e)=I_{\mathcal{A}}, we find e=I𝒞e=I_{\mathcal{C}}.

(c) Since every non-empty subset of a separable metric space is a separable metric space, separability of 𝒜{\mathcal{A}} implies α(𝒞)\alpha({\mathcal{C}}) is separable. So 𝒞{\mathcal{C}} is separable.

Now suppose 𝒞{\mathcal{C}} is separable and let Q𝒞Q\subseteq{\mathcal{C}} be countable and dense. Then 𝒜{\mathcal{A}} is generated by α(Q)Δ(α(Q))\alpha(Q)\cup\Delta(\alpha(Q)), so 𝒜{\mathcal{A}} is separable. ∎

It would be desirable to include nuclearity in the list of properties shared by 𝒞{\mathcal{C}} and 𝒜{\mathcal{A}}.

Conjecture \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a pseudo-Cartan inclusion and let (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) be a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Then 𝒞{\mathcal{C}} is nuclear if and only if 𝒜{\mathcal{A}} is nuclear.

4.4. Constructing Pseudo-Cartan Inclusions from Cartan Inclusions

We have seen that every pseudo-Cartan inclusion has a Cartan envelope, and we now consider the reverse process, that of constructing pseudo-Cartan inclusions from a given Cartan inclusion. We will work more generally, starting instead with a given pseudo-Cartan inclusion and constructing new pseudo-Cartan inclusions from it. Our next result, Proposition 4.4, extends parts of [PittsStReInII, Lemma 5.26 and Proposition 5.31] to the setting of pseudo-Cartan inclusions.

Proposition \the\numberby.

Suppose (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) is a pseudo-Cartan inclusion and 𝒟{\mathcal{D}} is a CC^{*}-subalgebra of 𝒟1{\mathcal{D}}_{1} such that 𝒟𝒟1{\mathcal{D}}\subseteq{\mathcal{D}}_{1} has the ideal intersection property. Let 𝒩(𝒞1,𝒟){\mathcal{M}}\subseteq{\mathcal{N}}({\mathcal{C}}_{1},{\mathcal{D}}) be a *-semigroup such that 𝒟span¯{\mathcal{D}}\subseteq\overline{\operatorname{span}}\,{\mathcal{M}} and set 𝒞=span¯{\mathcal{C}}=\overline{\operatorname{span}}\,{\mathcal{M}}. The following statements hold.

  1. (a)

    (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion.

  2. (b)

    Let (𝒜1,1 :: α)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha) be a Cartan envelope for (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and let Δ1:𝒜11\Delta_{1}:{\mathcal{A}}_{1}\rightarrow{\mathcal{B}}_{1} be the conditional expectation. Set

    :=C(Δ1(α(𝒞)))and𝒜:=C(α(𝒞)).{\mathcal{B}}:=C^{*}(\Delta_{1}(\alpha({\mathcal{C}})))\quad\text{and}\quad{\mathcal{A}}:=C^{*}(\alpha({\mathcal{C}})\cup{\mathcal{B}}).

    Then (𝒜, :: α|𝒞)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha|_{\mathcal{C}}) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

Proof.

(a) Since {\mathcal{M}} is a *-semigroup, 𝒞{\mathcal{C}} is a CC^{*}-algebra. Therefore, (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular inclusion. Since (𝒞1,𝒟1 :: )({\mathcal{C}}_{1},{\mathcal{D}}_{1}\hbox{\,:\hskip-1.0pt:\,}\subseteq) is an essential expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), Lemma 3(a) shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property. Hence (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion.

(b) As (𝒟1,𝒟)({\mathcal{D}}_{1},{\mathcal{D}}) and (1,α(𝒟1))({\mathcal{B}}_{1},\alpha({\mathcal{D}}_{1})) have the ideal intersection property, so does (1,α(𝒟))({\mathcal{B}}_{1},\alpha({\mathcal{D}})). Therefore (𝒜1,1 :: α|𝒟)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha|_{{\mathcal{D}}}) is an essential and Cartan expansion of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Applying Proposition 3, we obtain (b). ∎

While Proposition 4.4 applies when (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) is a Cartan inclusion, in general, (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) need not be the Cartan envelope of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). In fact, (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) can itself be Cartan. A trivial example of this occurs when {\mathcal{M}} is taken to be 𝒟{\mathcal{D}}, in which case, (𝒞,𝒟)=(𝒟,𝒟)({\mathcal{C}},{\mathcal{D}})=({\mathcal{D}},{\mathcal{D}}) is a Cartan inclusion. Here is a more interesting example where (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a Cartan inclusion.

Example \the\numberby. Let the amenable discrete group Γ\Gamma act topologically freely on the compact Hausdorff space XX. Suppose ZXZ\subseteq X is a non-empty closed invariant set which is nowhere dense in XX. Then

𝒟:={fC(X):f|Z is constant}{\mathcal{D}}:=\{f\in C(X):f|_{Z}\text{ is constant}\}

has the ideal intersection property in C(X)C(X). Let Y=𝒟^Y=\hat{\mathcal{D}} and fix x0Zx_{0}\in Z. Since every f𝒟f\in{\mathcal{D}} is constant on ZZ, for every f𝒟f\in{\mathcal{D}}, ff(x0)C0(XZ)f-f(x_{0})\in C_{0}(X\setminus Z), and it follows that YY is homeomorphic to the one point compactification of XZX\setminus Z. Let π:C(Y)𝒟\pi:C(Y)\rightarrow{\mathcal{D}} be the inverse of the Gelfand transformation and let {Us:sΓ}\{U_{s}:s\in\Gamma\} be the canonical copy of Γ\Gamma in C(X)rΓC(X)\rtimes_{r}\Gamma. The restriction of the action of Γ\Gamma to XZX\setminus Z is topologically free, and hence determines a topologically free action of Γ\Gamma on YY.

Letting (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) be the Cartan inclusion (C(X)rΓ,C(X))(C(X)\rtimes_{r}\Gamma,C(X)), apply Proposition 4.4 to (𝒟1,𝒟)({\mathcal{D}}_{1},{\mathcal{D}}) to produce a pseudo-Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Since ZZ is a Γ\Gamma invariant set, 𝒟{\mathcal{D}} is invariant under {Us}sΓ\{U_{s}\}_{s\in\Gamma}. Therefore, the pair (π,U)(\pi,U) is a covariant representation for the action of Γ\Gamma on YY. Since 𝒞{\mathcal{C}} is generated by 𝒟{\mathcal{D}} and {Us}sΓ\{U_{s}\}_{s\in\Gamma}, we obtain a *-epimorphism (π×U):C(Y)fΓ=C(Y)rΓ𝒞(\pi\times U):C(Y)\rtimes_{f}\Gamma=C(Y)\rtimes_{r}\Gamma\rightarrow{\mathcal{C}}. The topological freeness of Γ\Gamma acting on YY and Proposition 4.4 implies that (C(Y)rΓ,C(Y))(C(Y)\rtimes_{r}\Gamma,C(Y)) is a Cartan pair, and in particular, has the ideal intersection property. As π\pi is faithful, we conclude that (C(Y)Γ,C(Y))(C(Y)\rtimes\Gamma,C(Y)) is isomorphic to (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Thus (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is already Cartan, so is its own Cartan envelope. This shows (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) is not the Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

It is natural to wonder when a reduced crossed product is a pseudo-Cartan inclusion. This question is answered by the following result.

Proposition \the\numberby.

Let XX be a locally compact Hausdorff space and suppose Γ\Gamma is a discrete group acting as homeomorphisms on XX. With the corresponding action of Γ\Gamma on C0(X)C_{0}(X), the following statements are equivalent:

  1. (a)

    (C0(X)rΓ,C0(X))(C_{0}(X)\rtimes_{r}\Gamma,C_{0}(X)) is a pseudo-Cartan inclusion;

  2. (b)

    (C0(X)rΓ,C0(X))(C_{0}(X)\rtimes_{r}\Gamma,C_{0}(X)) is a Cartan inclusion; and

  3. (c)

    the action of Γ\Gamma on XX is topologically free.

Proof.

For each of (a)(b)(a)\Leftrightarrow(b) and (b)(c)(b)\Leftrightarrow(c), we use the fact that (C0(X)rΓ,C0(X))(C_{0}(X)\rtimes_{r}\Gamma,C_{0}(X)) is a regular inclusion and there is a faithful conditional expectation Δ:C0(X)ΓC0(X)\Delta:C_{0}(X)\rtimes\Gamma\rightarrow C_{0}(X).

(a)(b)(a)\Leftrightarrow(b). Apply Proposition 2.3.

(b)(c)(b)\Leftrightarrow(c). By [Zeller-MeierPrCrCstAlGrAu, Proposition 4.14], (C0(X)rΓ,C0(X))(C_{0}(X)\rtimes_{r}\Gamma,C_{0}(X)) is a MASA inclusion if and only if the action of Γ\Gamma on XX is topologically free. ∎

Remark \the\numberby. Despite the equivalence of parts (a) and (b) in Proposition 4.4, reduced crossed products can be used to construct examples of pseudo-Cartan inclusions which are not Cartan inclusions. Indeed, [PittsStReInI, Theorem 6.15] shows that reduced crossed products can be used to construct unital virtual Cartan inclusions. Thus by combining Proposition 4.4 with [PittsStReInI, Theorem 6.15], we obtain a wide variety of pseudo-Cartan inclusions.

5. The Twisted Groupoid of the Cartan Envelope

In [PittsStReInII, Section 7], we described the twist for the Cartan envelope of a unital regular inclusion. Combining Theorem 4.2 with results of [PittsStReInII] allows us to describe the twist associated to the Cartan envelope for any pseudo-Cartan inclusion regardless of whether it is unital. This description is found in Theorem 5.2 below.

5.1. Reprising the Unital Case

We begin with reprising some definitions and results from [PittsStReInII, Section 7]. Because we shall also require these notions in Section 6.3, we shall first consider unital regular inclusions with the unique (but not necessarily faithful) pseudo-expectation property. We then turn our attention to unital pseudo-Cartan inclusions.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular and unital inclusion with the unique pseudo-expectation property and let (I(𝒟),ι)(I({\mathcal{D}}),\iota) be an injective envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Let E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be the pseudo-expectation.

  1. (a)

    An eigenfunctional (see [DonsigPittsCoSyBoIs, Definition 2.1]) on 𝒞{\mathcal{C}} is a non-zero bounded linear functional ϕ\phi on 𝒞{\mathcal{C}} which is an eigenvector for the left and right actions of 𝒟{\mathcal{D}} on the dual space of 𝒞{\mathcal{C}}. When ϕ\phi is an eigenfunctional, there exist unique r(ϕ)𝒟^r(\phi)\in\hat{\mathcal{D}} and s(ϕ)𝒟^s(\phi)\in\hat{\mathcal{D}} such that for every d1,d2𝒟d_{1},d_{2}\in{\mathcal{D}} and x𝒞x\in{\mathcal{C}},

    ϕ(d1xd2)=r(ϕ)(d1)ϕ(x)s(ϕ)(d2).\phi(d_{1}xd_{2})=r(\phi)(d_{1})\,\phi(x)\,s(\phi)(d_{2}).

    If for every v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}),

    |ϕ(v)|2{0,s(ϕ)(vv)},|\phi(v)|^{2}\in\{0,s(\phi)(v^{*}v)\},

    ϕ\phi is a compatible eigenfunctional ([PittsStReInII, Definition 7.6]). A compatible state is a state on 𝒞{\mathcal{C}} which is also a compatible eigenfunctional. The collection of all compatible eigenfunctionals of unit norm is denoted c1(𝒞,𝒟){\mathcal{E}}^{1}_{c}({\mathcal{C}},{\mathcal{D}}) and 𝔖(𝒞,𝒟){\mathfrak{S}}({\mathcal{C}},{\mathcal{D}}) denotes the set of compatible states.

  2. (b)

    For ϕc1(𝒞,𝒟)\phi\in{\mathcal{E}}^{1}_{c}({\mathcal{C}},{\mathcal{D}}), [PittsStReInII, Theorem 7.9] shows that there are unique states 𝔰(ϕ),𝔯(ϕ){\mathfrak{s}}(\phi),{\mathfrak{r}}(\phi) on 𝒞{\mathcal{C}} with the following properties:

    • s(ϕ)=𝔰(ϕ)|𝒟s(\phi)={\mathfrak{s}}(\phi)|_{\mathcal{D}}, r(ϕ)=𝔯(ϕ)|𝒟r(\phi)={\mathfrak{r}}(\phi)|_{\mathcal{D}};

    • when v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) satisfies ϕ(v)0\phi(v)\neq 0 and x𝒞x\in{\mathcal{C}},

      ϕ(vx)=ϕ(v)𝔰(ϕ)(x)andϕ(xv)=𝔯(ϕ)(x)ϕ(v).\phi(vx)=\phi(v)\,{\mathfrak{s}}(\phi)(x)\quad\text{and}\quad\phi(xv)={\mathfrak{r}}(\phi)(x)\,\phi(v).

    For later use, here are formulas for 𝔰(ϕ){\mathfrak{s}}(\phi) and 𝔯(ϕ){\mathfrak{r}}(\phi): for x𝒞x\in{\mathcal{C}} and v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) with ϕ(v)>0\phi(v)>0,

    (5.1.1) 𝔰(ϕ)(x)=ϕ(vx)ϕ(v)and𝔯(ϕ)(x)=ϕ(xv)ϕ(v).{\mathfrak{s}}(\phi)(x)=\frac{\phi(vx)}{\phi(v)}\quad\text{and}\quad{\mathfrak{r}}(\phi)(x)=\frac{\phi(xv)}{\phi(v)}.

    In addition, note that (still assuming v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) satisfies ϕ(v)>0\phi(v)>0) 𝔰(ϕ)(vv)0{\mathfrak{s}}(\phi)(v^{*}v)\neq 0 and for every x𝒞x\in{\mathcal{C}},

    ϕ(x)=𝔰(ϕ)(vx)𝔰(ϕ)(vv)1/2.\phi(x)=\frac{{\mathfrak{s}}(\phi)(v^{*}x)}{{\mathfrak{s}}(\phi)(v^{*}v)^{1/2}}.
  3. (c)

    For ρ𝔖(𝒞,𝒟)\rho\in{\mathfrak{S}}({\mathcal{C}},{\mathcal{D}}) and v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) with ρ(vv)0\rho(v^{*}v)\neq 0, we use the notation [v,ρ][v,\rho] for the linear functional

    [v,ρ](x):=ρ(vx)ρ(vv)1/2.[v,\rho](x):=\frac{\rho(v^{*}x)}{\rho(v^{*}v)^{1/2}}.

    By [PittsStReInII, Corollary 7.11],

    c1(𝒞,𝒟)={[v,ρ]:ρ𝔖(𝒞,𝒟) and ρ(vv)0}.{\mathcal{E}}^{1}_{c}({\mathcal{C}},{\mathcal{D}})=\{[v,\rho]:\rho\in{\mathfrak{S}}({\mathcal{C}},{\mathcal{D}})\text{ and }\rho(v^{*}v)\neq 0\}.
  4. (d)

    A strongly compatible state on 𝒞{\mathcal{C}} is a state ρ\rho on 𝒞{\mathcal{C}} for which there exists σI(𝒟)^\sigma\in\widehat{I({\mathcal{D}})} such that

    ρ=σE.\rho=\sigma\circ E.

    We denote the family of all strongly compatible states on 𝒞{\mathcal{C}} by 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}). Since

    𝒟^={ρ|𝒟:ρ𝔖s(𝒞,𝒟)},\hat{\mathcal{D}}=\{\rho|_{\mathcal{D}}:\rho\in{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}})\},

    there is a rich supply of strongly compatible states.

    By [PittsStReInII, Theorem 6.9] (whose statement is reproduced in Theorem A) every strongly compatible state is a compatible state, that is,

    𝔖s(𝒞,𝒟)𝔖(𝒞,𝒟).{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}})\subseteq{\mathfrak{S}}({\mathcal{C}},{\mathcal{D}}).
  5. (e)

    A compatible eigenfunctional ϕc1(𝒞,𝒟)\phi\in{\mathcal{E}}^{1}_{c}({\mathcal{C}},{\mathcal{D}}) is called a strongly compatible eigenfunctional if 𝔰(ϕ)𝔖s(𝒞,𝒟){\mathfrak{s}}(\phi)\in{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}). When this occurs, 𝔯(ϕ){\mathfrak{r}}(\phi) also belongs to 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}).

  6. (f)

    Let Σ(𝒞,𝒟)\Sigma({\mathcal{C}},{\mathcal{D}}) be the collection of all norm-one strongly compatible eigenfunctionals and put

    G(𝒞,𝒟):={|ϕ|:ϕΣ(𝒞,𝒟)},G({\mathcal{C}},{\mathcal{D}}):=\{|\phi|:\phi\in\Sigma({\mathcal{C}},{\mathcal{D}})\},

    where |ϕ||\phi| denotes the composition of the absolute value function on {\mathbb{C}} with ϕ\phi.

  7. (g)

    The representation of ϕΣ(𝒞,𝒟)\phi\in\Sigma({\mathcal{C}},{\mathcal{D}}) as ϕ=[v,𝔰(ϕ)]\phi=[v,{\mathfrak{s}}(\phi)], allows the sets Σ(𝒞,𝒟)\Sigma({\mathcal{C}},{\mathcal{D}}) and G(𝒞,𝒟)G({\mathcal{C}},{\mathcal{D}}) to be equipped with groupoid operations [PittsStReInII, Definition 7.17]. Upon doing so, the unit space, G(𝒞,𝒟)(0)G({\mathcal{C}},{\mathcal{D}})^{(0)}, may be identified with 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}). With the topology of pointwise convergence, Σ(𝒞,𝒟)\Sigma({\mathcal{C}},{\mathcal{D}}) and G(𝒞,𝒟)G({\mathcal{C}},{\mathcal{D}}) become Hausdorff topological groupoids (take F=𝔖s(𝒞,𝒟)F={\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) in [PittsStReInII, Theorem 7.18]) and we obtain a twist

    G(𝒞,𝒟)(0)×𝕋Σ(𝒞,𝒟)G(𝒞,𝒟)G({\mathcal{C}},{\mathcal{D}})^{(0)}\times{\mathbb{T}}\hookrightarrow\Sigma({\mathcal{C}},{\mathcal{D}})\twoheadrightarrow G({\mathcal{C}},{\mathcal{D}})

    associated with (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

  8. (h)

    For each a𝒞a\in{\mathcal{C}}, let 𝔤(a):Σ(𝒞,𝒟)\mathfrak{g}(a):\Sigma({\mathcal{C}},{\mathcal{D}})\rightarrow{\mathbb{C}} be given by

    𝔤(a)(ϕ)=ϕ(a).\mathfrak{g}(a)(\phi)=\phi(a).
  9. (i)

    Set

    (𝒞,𝒟):={a𝒞:ρ(aa)=0 for all ρ𝔖s(𝒞,𝒟)}={a𝒞:E(aa)=0}.{\mathcal{L}}({\mathcal{C}},{\mathcal{D}}):=\{a\in{\mathcal{C}}:\rho(a^{*}a)=0\text{ for all }\rho\in{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}})\}=\{a\in{\mathcal{C}}:E(a^{*}a)=0\}.

    By [PittsStReInII, Theorem 6.5], (𝒞,𝒟){\mathcal{L}}({\mathcal{C}},{\mathcal{D}}) is an ideal of 𝒞{\mathcal{C}} such that (𝒞,𝒟)𝒟={0}{\mathcal{L}}({\mathcal{C}},{\mathcal{D}})\cap{\mathcal{D}}=\{0\}. Also, if J𝒞J\unlhd{\mathcal{C}} satisfies J𝒟={0}J\cap{\mathcal{D}}=\{0\}, then J(𝒞,𝒟)J\subseteq{\mathcal{L}}({\mathcal{C}},{\mathcal{D}}).

Now suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a unital pseudo-Cartan inclusion. Recall from Proposition 2.4 that (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) has the faithful unique pseudo-expectation property and E|𝒟cE|_{{\mathcal{D}}^{c}} is a *-monomorphism; also E|𝒟cE|_{{\mathcal{D}}^{c}} is the pseudo-expectation for (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}). Simplify the notation of 5.1(f) somewhat by writing Σ\Sigma and GG instead of Σ(𝒞,𝒟)\Sigma({\mathcal{C}},{\mathcal{D}}) and G(𝒞,𝒟)G({\mathcal{C}},{\mathcal{D}}). As in 5.1(g), let

G(0)×𝕋ΣG,G^{(0)}\times{\mathbb{T}}\hookrightarrow\Sigma\twoheadrightarrow G,

be the twist for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). We remind the reader that G(0)G^{(0)} is identified with 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}).

By [PittsStReInII, Theorem 7.24], 𝔤\mathfrak{g} induces a regular *-monomorphism θ:𝒞Cr(Σ,G)\theta:{\mathcal{C}}\rightarrow C^{*}_{r}(\Sigma,G) and by [PittsStReInII, Corollary 7.30],

(5.1.2) (Cr(Σ,G),C(G(0)) :: θ)(C^{*}_{r}(\Sigma,G),C(G^{(0)})\hbox{\,:\hskip-1.0pt:\,}\theta)

is a package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). While [PittsStReInII, Corollary 7.31] states (Cr(Σ,G),C(G(0)) :: θ)(C^{*}_{r}(\Sigma,G),C(G^{(0)})\hbox{\,:\hskip-1.0pt:\,}\theta) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), full details were omitted in [PittsStReInII][PittsCoStReInII, Remark 1.2] provides the missing details.

5.2. The Non-Unital Case

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a non-unital pseudo-Cartan inclusion, and fix an injective envelope (I(𝒟),ι)(I({\mathcal{D}}),\iota) for 𝒟{\mathcal{D}}. As (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}) has the ideal intersection property, there is a unique *-monomorphism u:𝒟cI(𝒟)u:{\mathcal{D}}^{c}\rightarrow I({\mathcal{D}}) such that u|𝒟=ι|𝒟u|_{\mathcal{D}}=\iota|_{\mathcal{D}}. Thus (I(𝒟),u~)(I({\mathcal{D}}),\tilde{u}) is an injective envelope for 𝒟c{\mathcal{D}}^{c}. It follows that E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) is a pseudo-expectation for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) relative to (I(𝒟),ι)(I({\mathcal{D}}),\iota) if and only if EE is a pseudo-expectation for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) relative to (I(𝒟),u)(I({\mathcal{D}}),u). Since (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) and (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) have the faithful unique pseudo-expectation property, we may define 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) and 𝔖s(𝒞,𝒟c){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}^{c}) as in 5.1(d). Then

(5.2.1) 𝔖s(𝒞,𝒟)=𝔖s(𝒞,𝒟c).{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}})={\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}^{c}).

Recalling from Proposition 4.2 that (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) and (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) have the same Cartan envelope, we will give a groupoid description for a Cartan envelope of (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}).

Let Σ~\tilde{\Sigma} be the set of norm-one strongly compatible eigenfunctionals for (𝒞~,(𝒟c))(\tilde{\mathcal{C}},({\mathcal{D}}^{c})^{\sim}) and let G~={|ϕ|:ϕΣ~}\tilde{G}=\{|\phi|:\phi\in\tilde{\Sigma}\}. Applying the discussion for the unital case to (𝒞~,(𝒟c))(\tilde{\mathcal{C}},({\mathcal{D}}^{c})^{\sim}) we obtain the twist,

G~(0)×𝕋Σ~G~\tilde{G}^{(0)}\times{\mathbb{T}}\rightarrow\tilde{\Sigma}\twoheadrightarrow\tilde{G}

with unit space G~(0)=𝔖s(𝒞~,(𝒟c))\tilde{G}^{(0)}={\mathfrak{S}}_{s}(\tilde{\mathcal{C}},({\mathcal{D}}^{c})^{\sim}). Taking θ\theta as in (5.1.2), (Cr(Σ~,G~),C(G~(0)) :: θ)(C^{*}_{r}(\tilde{\Sigma},\tilde{G}),C(\tilde{G}^{(0)})\hbox{\,:\hskip-1.0pt:\,}\theta) is a Cartan envelope for (𝒞~,(𝒟c))(\tilde{\mathcal{C}},({\mathcal{D}}^{c})^{\sim}). We use Δ\Delta to denote the conditional expectation of Cr(Σ~,G~)C^{*}_{r}(\tilde{\Sigma},\tilde{G}) onto C(G~(0))C(\tilde{G}^{(0)}); note that Δ\Delta arises from the restriction of an element in Cc(Σ~,G~)C_{c}(\tilde{\Sigma},\tilde{G}) to G~(0)\tilde{G}^{(0)}.

Let qq denote the linear functional, 𝒞~(c,λ)λ\tilde{\mathcal{C}}\ni(c,\lambda)\mapsto\lambda. Since qq is a multiplicative linear functional, it is a compatible state for (𝒞~,(𝒟c))(\tilde{\mathcal{C}},({\mathcal{D}}^{c})^{\sim}), and we claim it is actually a strongly compatible state. Let ψ\psi be a multiplicative linear functional on I(𝒟c)I({\mathcal{D}}^{c}) satisfying ψu~=q|(𝒟c)\psi\circ\tilde{u}=q|_{({\mathcal{D}}^{c})^{\sim}}. Let E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be the pseudo-expectation and E~\tilde{E} its unitization. Given v𝒩(𝒞,𝒟c)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}^{c}), we have

E~(v)E~(v)E~(vv)=u(vv).\tilde{E}(v)^{*}\tilde{E}(v)\leq\tilde{E}(v^{*}v)=u(v^{*}v).

Applying ψ\psi yields,

|ψ(E~(v))|2ψ(u(vv))=q(vv)=0.|\psi(\tilde{E}(v))|^{2}\leq\psi(u(v^{*}v))=q(v^{*}v)=0.

It follows that ψE~\psi\circ\tilde{E} annihilates 𝒩(𝒞,𝒟c){\mathcal{N}}({\mathcal{C}},{\mathcal{D}}^{c}), and hence annihilates 𝒞{\mathcal{C}}. As qq and ψE~\psi\circ\tilde{E} are states on 𝒞~\tilde{\mathcal{C}} which have the same kernel, we obtain

q=ψE~,q=\psi\circ\tilde{E},

showing that qq is a strongly compatible state.

Let

Σ:={ϕΣ~:ϕ|𝒞0}andG:={|ϕ|:ϕΣ~ and ϕ|𝒞0}\Sigma:=\{\phi\in\tilde{\Sigma}:\phi|_{\mathcal{C}}\neq 0\}\quad\text{and}\quad G:=\{|\phi|:\phi\in\tilde{\Sigma}\text{ and }\phi|_{\mathcal{C}}\neq 0\}

be equipped with their relative topologies. Note that if ϕΣ~\phi\in\tilde{\Sigma} satisfies ϕ|𝒞=0\phi|_{\mathcal{C}}=0, then kerϕ=kerq=𝒞\ker\phi=\ker q={\mathcal{C}}, so ϕ\phi is a scalar multiple of qq. It follows that

Σ=Σ~𝕋qandG=G~{|q|}.\Sigma=\tilde{\Sigma}\setminus{\mathbb{T}}q\quad\text{and}\quad G=\tilde{G}\setminus\{|q|\}.

The definition of the topology and groupoid operations on GG show that GG is an open subgroupoid of G~\tilde{G}. Since G~(0)=𝔖s(𝒞~,(𝒟c))\tilde{G}^{(0)}={\mathfrak{S}}_{s}(\tilde{\mathcal{C}},({\mathcal{D}}^{c})^{\sim}),

G(0)={ρ𝔖s(𝒞~,(𝒟c)):ρ|𝒞0}.G^{(0)}=\{\rho\in{\mathfrak{S}}_{s}(\tilde{\mathcal{C}},({\mathcal{D}}^{c})^{\sim}):\rho|_{\mathcal{C}}\neq 0\}.

By [BrownFullerPittsReznikoffGrC*AlTwGpC*Al, Lemma 2.7]. this produces the twist,

(5.2.2) G(0)×𝕋ΣG.G^{(0)}\times{\mathbb{T}}\hookrightarrow\Sigma\twoheadrightarrow G.

Routine modifications to the proof of [PittsStReInII, Theorem 7.24] show that with θ:𝒞~Cr(Σ~,G~)\theta:\tilde{\mathcal{C}}\rightarrow C^{*}_{r}(\tilde{\Sigma},\tilde{G}) arising from 𝔤\mathfrak{g} as described in the unital case above,

C(θ(𝒞)C0(G(0)))=Cr(Σ,G).C^{*}(\theta({\mathcal{C}})\cup C_{0}(G^{(0)}))=C^{*}_{r}(\Sigma,G).

We now show that C(Δ(θ(𝒞)))=C0(G(0))C^{*}(\Delta(\theta({\mathcal{C}})))=C_{0}(G^{(0)}). Let

𝒩0(𝒞,𝒟c):={v𝒩(𝒞,𝒟c):𝔤(v) is compactly supported}{\mathcal{N}}_{0}({\mathcal{C}},{\mathcal{D}}^{c}):=\{v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}^{c}):\mathfrak{g}(v)\text{ is compactly supported}\}

and let 𝒞0=span𝒩0(𝒞,𝒟c){\mathcal{C}}_{0}=\operatorname{span}{\mathcal{N}}_{0}({\mathcal{C}},{\mathcal{D}}^{c}). As in the proof of  [PittsStReInII, Theorem 7.24], 𝒩0(𝒞,𝒟c){\mathcal{N}}_{0}({\mathcal{C}},{\mathcal{D}}^{c}) is a *-semigroup and 𝒞0{\mathcal{C}}_{0} is dense in 𝒞{\mathcal{C}}. Now suppose ρ1\rho_{1} and ρ2\rho_{2} are distinct elements of G(0)G^{(0)}, in other words, ρ1\rho_{1} and ρ2\rho_{2} are distinct strongly compatible states on 𝒞~\tilde{\mathcal{C}}, neither of which vanish on 𝒞{\mathcal{C}}. Then ρ1|𝒞ρ2|𝒞\rho_{1}|_{\mathcal{C}}\neq\rho_{2}|_{\mathcal{C}}, so we may find an element v𝒩0(𝒞,𝒟c)v\in{\mathcal{N}}_{0}({\mathcal{C}},{\mathcal{D}}^{c}) such that ρ1(v)ρ2(v)\rho_{1}(v)\neq\rho_{2}(v). Thus, 𝔤(v)(ρ1)𝔤(v)(ρ2)\mathfrak{g}(v)(\rho_{1})\neq\mathfrak{g}(v)(\rho_{2}). Since Δ\Delta is determined by restriction to G(0)G^{(0)}, we obtain Δ(𝔤(v)(ρ1)Δ(𝔤(v))(ρ2)\Delta(\mathfrak{g}(v)(\rho_{1})\neq\Delta(\mathfrak{g}(v))(\rho_{2}). As θ(v)=𝔤(v)\theta(v)=\mathfrak{g}(v), we conclude that Δ(θ(𝒞))\Delta(\theta({\mathcal{C}})) separates points of G(0)G^{(0)}. Also, recalling that 𝒟c{\mathcal{D}}^{c} contains an approximate unit for 𝒞{\mathcal{C}}, we see that given ρG(0)\rho\in G^{(0)}, there exists h𝒟ch\in{\mathcal{D}}^{c} such that ρ(h)=θ(h)(ρ)0\rho(h)=\theta(h)(\rho)\neq 0. By the Stone-Wierstrauß theorem, C(Δ(θ(𝒞)))=C0(G(0))C^{*}(\Delta(\theta({\mathcal{C}})))=C_{0}(G^{(0)}).

An application of Theorem 4.2 shows that (Cr(Σ,G),C0(G(0)) :: θ|𝒞)(C^{*}_{r}(\Sigma,G),C_{0}(G^{(0)})\hbox{\,:\hskip-1.0pt:\,}\theta|_{\mathcal{C}}) is a Cartan envelope for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}), giving a groupoid description for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) (and (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}})).

The definition of compatible eigenfunctional makes sense regardless of whether the pseudo-Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is unital. Since 𝒩(𝒞,𝒟)𝒩(𝒞,𝒟c){\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}^{c}),

c1(𝒞,𝒟c)c1(𝒞,𝒟).{\mathcal{E}}^{1}_{c}({\mathcal{C}},{\mathcal{D}}^{c})\subseteq{\mathcal{E}}^{1}_{c}({\mathcal{C}},{\mathcal{D}}).

Actually, equality holds. To see this, let ϕc1(𝒞,𝒟)\phi\in{\mathcal{E}}^{1}_{c}({\mathcal{C}},{\mathcal{D}}). Regularity of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) implies there exists v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) such that ϕ(v)>0\phi(v)>0. With 𝔰(ϕ){\mathfrak{s}}(\phi) as defined in (5.1.1), we obtain

ϕ=[v,𝔰(ϕ)]c1(𝒞,𝒟c).\phi=[v,{\mathfrak{s}}(\phi)]\in{\mathcal{E}}^{1}_{c}({\mathcal{C}},{\mathcal{D}}^{c}).

Using (5.2.1), we conclude that the collection of norm-one strongly compatible eigenfunctionals for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is the same as the set of norm-one strongly compatible eigenfunctionals for (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}).

We summarize our discussion with the following result, which extends the groupoid description of the Cartan envelope given in [PittsStReInII, Corollary 7.31] (once again, with F=𝔖s(𝒞,𝒟))F={\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}))) from the unital setting to include both the unital and non-unital cases.

Theorem \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a pseudo-Cartan inclusion, let Σ\Sigma be the collection of all norm-one strongly compatible eigenfunctionals for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), and let G={|ϕ|:ϕΣ}G=\{|\phi|:\phi\in\Sigma\}. Then the following statements hold.

  1. (a)

    Σ\Sigma and GG are Hausdorff topological groupoids with GG effective and étale, G(0)=𝔖s(𝒞,𝒟)G^{(0)}={\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) and

    G(0)×𝕋ΣGG^{(0)}\times{\mathbb{T}}\hookrightarrow\Sigma\twoheadrightarrow G

    is a twist.

  2. (b)

    The map 𝔤\mathfrak{g} extends to a regular *-monomorphism θ:𝒞Cr(Σ,G)\theta:{\mathcal{C}}\rightarrow C^{*}_{r}(\Sigma,G) and

    (Cr(Σ,G),C0(G(0)) :: θ)(C^{*}_{r}(\Sigma,G),C_{0}(G^{(0)})\hbox{\,:\hskip-1.0pt:\,}\theta)

    is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

6. Constructions and Properties of Pseudo-Cartan Inclusions

In this section, we explore the behavior of pseudo-Cartan inclusions and their Cartan envelopes under mappings and also some familiar constructions. Theorem 6.1 shows that under a suitable hypothesis, a regular map between pseudo-Cartan inclusions extends to the Cartan envelopes and Proposition 6.1 shows that a regular automorphism of a pseudo-Cartan inclusion uniquely extends to its Cartan envelope. We examine the behavior of pseudo-Cartan inclusions and their inductive limits for suitable connecting maps in Theorems 6.2 and 6.2, and Theorem 6.3 describes behavior of pseudo-Cartan inclusions and their Cartan envelops under the minimal tensor product. In some of the results, obtaining regularity of maps under these constructions is a delicate and technical issue.

6.1. Mapping Results

The purposes of this subsection are to establish a mapping property, Theorem 6.1, for Cartan envelopes and to show that a regular automorphism of a pseudo-Cartan inclusion extends uniquely to its Cartan envelope, Proposition 6.1. While interesting in its own right, Theorem 6.1 is a key tool for studying inductive limits of pseudo-Cartan inclusions.

We require some preparation, beginning with a simple fact about sums of normalizers. The sum of normalizers is not usually a normalizer. However, when the normalizers are “orthogonal,” their sum is again a normalizer. Here is the precise statement.

Fact \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is an inclusion and w,v𝒩(𝒞,𝒟)w,v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). If ww and vv are orthogonal in the sense that vw=0=wvv^{*}w=0=wv^{*}, then v+w𝒩(𝒞,𝒟)v+w\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}).

Proof.

Let d𝒟d\in{\mathcal{D}}. We have

(v+w)d(v+w)=vdv+wdw+vdw+wdv=vdv+wdw+limvd(ww)1/nw+limwd(vv)1/nv=vdv+wdw+limvwθw(d(ww)1/n)+limwvθv(d(vv)1/n)v=vdv+wdw𝒟.\displaystyle\begin{split}(v+w)^{*}d(v+w)&=v^{*}dv+w^{*}dw+v^{*}dw+w^{*}dv\\ &=v^{*}dv+w^{*}dw+\lim v^{*}d(ww^{*})^{1/n}w+\lim w^{*}d(vv^{*})^{1/n}v\\ &=v^{*}dv+w^{*}dw+\lim v^{*}w\theta_{w}(d(ww^{*})^{1/n})\\ &\qquad+\lim w^{*}v\theta_{v}(d(vv^{*})^{1/n})v\\ &=v^{*}dv+w^{*}dw\in{\mathcal{D}}.\end{split}

Similarly (v+w)d(v+w)𝒟(v+w)d(v+w)^{*}\in{\mathcal{D}}, so v+w𝒩(𝒞,𝒟)v+w\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). ∎

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is an unital inclusion and (I(𝒟),ι)(I({\mathcal{D}}),\iota) is an injective envelope for 𝒟{\mathcal{D}}. Given v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), let P,QP,Q be the support projections in I(𝒟)I({\mathcal{D}}) for the ideals vv𝒟¯\overline{vv^{*}{\mathcal{D}}} and vv𝒟¯\overline{v^{*}v{\mathcal{D}}} respectively. (Given any ideal J𝒟J\unlhd{\mathcal{D}}, the supremum in I(𝒟)s.a.I({\mathcal{D}})_{s.a.} of {ι(x):0xJ and x1}\{\iota(x):0\leq x\in J\text{ and }\left\lVert x\right\rVert\leq 1\} is called the support projection for JJ.) Recall from [PittsStReInI, Proposition 1.11] that the isomorphism θv:vv𝒟¯vv𝒟¯\theta_{v}:\overline{vv^{*}{\mathcal{D}}}\rightarrow\overline{v^{*}v{\mathcal{D}}} uniquely extends to a *-isomorphism θ~v:I(𝒟)PI(𝒟)Q\tilde{\theta}_{v}:I({\mathcal{D}})P\rightarrow I({\mathcal{D}})Q. We then obtain a Frolík decomposition {Rj}j=04\{R_{j}\}_{j=0}^{4} for θ~v\tilde{\theta}_{v}, see [PittsStReInI, Definition 2.12], and also the Frolík ideals {Ki(v)}i=04\{K_{i}(v)\}_{i=0}^{4}[PittsStReInI, Definition 2.13]. Our immediate interest will be with K0(v)K_{0}(v). This is the fixed point ideal for v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}); its definition is

(6.1.1) K0(v):=ι1(R0I(𝒟)).K_{0}(v):=\iota^{-1}(R_{0}I({\mathcal{D}})).

This is a regular ideal in 𝒟{\mathcal{D}}, and it has the following alternate description:

(6.1.2) K0(v)={d(vv𝒟):vd=dv𝒟c}={d(vv𝒟):vd=dv𝒟c}.K_{0}(v)=\{d\in(vv^{*}{\mathcal{D}})^{\perp\perp}:vd=dv\in{\mathcal{D}}^{c}\}=\{d\in(v^{*}v{\mathcal{D}})^{\perp\perp}:vd=dv\in{\mathcal{D}}^{c}\}.

See [PittsStReInI, Definition 2.13 and Lemma 2.15] for further details.

To extend the previous considerations to weakly non-degenerate inclusions, recall that if (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a weakly non-degenerate inclusion, then (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a unital inclusion and the standard embeddings (described in (2.1.2)) satisfy u𝒞|𝒟=u𝒟u_{\mathcal{C}}|_{\mathcal{D}}=u_{\mathcal{D}}. For v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), define

(6.1.3) K0(v):=u𝒟1(K0(u𝒞(v)))={d𝒟:u𝒟(d)K0(u𝒞(v))}.K_{0}(v):=u_{\mathcal{D}}^{-1}(K_{0}(u_{\mathcal{C}}(v)))=\{d\in{\mathcal{D}}:u_{\mathcal{D}}(d)\in K_{0}(u_{\mathcal{C}}(v))\}.

When a normalizer vv for a regular MASA inclusion is not annihilated by the pseudo-expectation, K0(v)K_{0}(v) is non-trivial. In fact, a bit more is true.

Lemma \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular MASA inclusion, let (I(𝒟),ι)(I({\mathcal{D}}),\iota) be an injective envelope for 𝒟{\mathcal{D}} and let E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be the pseudo-expectation. Suppose v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) satisfies E(v)0E(v)\neq 0. Then K0(v){0}K_{0}(v)\neq\{0\} and there exists kK0(v)vv𝒟¯k\in K_{0}(v)\cap\overline{v^{*}v{\mathcal{D}}} such that vk0vk\neq 0.

Proof.

Assume first that in addition to the hypotheses stated, (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a unital inclusion. By [PittsStReInI, Theorem 3.5(iii)], |E(v)|2=R0(ι(vv))|E(v)|^{2}=R_{0}(\iota(v^{*}v)), so R00R_{0}\neq 0. Therefore, K0(v){0}K_{0}(v)\neq\{0\} by (6.1.1) and the fact that (I(𝒟),𝒟,ι)(I({\mathcal{D}}),{\mathcal{D}},\iota) has the ideal intersection property. Next, [PittsStReInII, Lemma 2.15] shows K0(v)=(K0(v)vv𝒟¯)K_{0}(v)=(K_{0}(v)\cap\overline{v^{*}v{\mathcal{D}}})^{\perp\perp}, thus K0(v)vv𝒟¯{0}K_{0}(v)\cap\overline{v^{*}v{\mathcal{D}}}\neq\{0\}. Choose 0kK0(v)vv𝒟¯0\neq k\in K_{0}(v)\cap\overline{v^{*}v{\mathcal{D}}}. Then vk0vk\neq 0. (Otherwise, vvk=0v^{*}vk=0, whence k{vv𝒟}{vv𝒟}k\in\{v^{*}v{\mathcal{D}}\}^{\perp}\cap\{v^{*}v{\mathcal{D}}\}^{\perp\perp}, contrary to k0k\neq 0.) This completes the unital case.

Now assume (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is not unital. First observe that (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a unital regular MASA inclusion. Indeed, [PittsNoApUnInC*Al, Theorem 2.5] shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP, so Lemma 3(e) yields regularity of (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) and a routine argument shows (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a MASA inclusion. Since (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the AUP, the standard embedding u𝒞:𝒞𝒞~u_{\mathcal{C}}:{\mathcal{C}}\rightarrow\tilde{\mathcal{C}} is a regular map by Lemma 3(c). Thus, (v,0)=u𝒞(v)𝒩(𝒞~,𝒟~)(v,0)=u_{\mathcal{C}}(v)\in{\mathcal{N}}(\tilde{\mathcal{C}},\tilde{\mathcal{D}}). By the unital case, there exists (k,λ)K0((v,0))(vv,0)𝒟~¯(k,\lambda)\in K_{0}((v,0))\cap\overline{(v^{*}v,0)\tilde{\mathcal{D}}}, with (v,0)(k,λ)(0,0)(v,0)(k,\lambda)\neq(0,0). Since (k,λ)(vv,0)𝒟~¯(k,\lambda)\in\overline{(v^{*}v,0)\tilde{\mathcal{D}}}, we must have λ=0\lambda=0. In other words, there exists kK0(v)vv𝒟¯k\in K_{0}(v)\cap\overline{v^{*}v{\mathcal{D}}} such that vk0vk\neq 0. ∎

Sometimes it is possible to establish regularity of a *-monomorphism between inclusions with only partial knowledge of the normalizers in the domain of the map. The following useful technical result gives a setting where this can be done.

Proposition \the\numberby.

Let (𝒜2,2 :: α)({\mathcal{A}}_{2},{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) be an essential and Cartan expansion for the pseudo-Cartan inclusion (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}) and suppose N𝒩(𝒜1,1)N\subseteq{\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}) is a *-semigroup such that 1N{\mathcal{B}}_{1}\subseteq N and spanN\operatorname{span}N is dense in 𝒜1{\mathcal{A}}_{1}. If α(w)𝒩(𝒜2,2)\alpha(w)\in{\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2}) for every wNw\in N, then α:(𝒜1,1)(𝒜2,2)\alpha:({\mathcal{A}}_{1},{\mathcal{B}}_{1})\rightarrow({\mathcal{A}}_{2},{\mathcal{B}}_{2}) is a regular *-monomorphism.

Proof.

Throughout the proof, let E2:𝒜22E_{2}:{\mathcal{A}}_{2}\rightarrow{\mathcal{B}}_{2} be the conditional expectation. The first step in the proof is to establish the following lemma.

Lemma \the\numberby.

For v𝒩(𝒜1,1)v\in{\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}), let

𝒵v:={b11:α(vb1)𝒩(𝒜2,2)}.{\mathcal{Z}}_{v}:=\{b_{1}\in{\mathcal{B}}_{1}:\alpha(vb_{1})\in{\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2})\}.

If J1J\unlhd{\mathcal{B}}_{1} is an essential ideal such that J𝒵vJ\subseteq{\mathcal{Z}}_{v}, then for every b22b_{2}\in{\mathcal{B}}_{2}, α(v)b2α(v)2\alpha(v)^{*}b_{2}\alpha(v)\in{\mathcal{B}}_{2}.

Proof.

We first show that for any h2h\in{\mathcal{B}}_{2} and xJx\in J,

(6.1.4) α(v)hα(v)α(x)2.\alpha(v)^{*}h\alpha(v)\alpha(x)\in{\mathcal{B}}_{2}.

Indeed for yn:=α((vx)(vx)))1/nα((vx)(vx))2¯y_{n}:=\alpha((vx)(vx)^{*}))^{1/n}\in\overline{\alpha((vx)(vx)^{*}){\mathcal{B}}_{2}},

α(v)hα(vx)=limα(v)hynα(vx)=limα(vvx)θα(vx)(hyn)2;\alpha(v)^{*}h\alpha(vx)=\lim\alpha(v)^{*}hy_{n}\alpha(vx)=\lim\alpha(v^{*}vx)\theta_{\alpha(vx)}(hy_{n})\in{\mathcal{B}}_{2};

thus (6.1.4) holds.

Fixing b22b_{2}\in{\mathcal{B}}_{2}, put

m:=α(v)b2α(v)E2(α(v)b2α(v)).m:=\alpha(v^{*})b_{2}\alpha(v)-E_{2}(\alpha(v)^{*}b_{2}\alpha(v)).

We shall show m=0m=0. To do this, let

M:={h2:E2(mm)h=0}.M:=\{h\in{\mathcal{B}}_{2}:E_{2}(m^{*}m)h=0\}.

Then M2M\unlhd{\mathcal{B}}_{2}, and because M={E2(mm)}M=\{E_{2}(m^{*}m)\}^{\perp}, M=MM=M^{\perp\perp}. By (6.1.4),

α(J)2¯M.\overline{\alpha(J){\mathcal{B}}_{2}}\subseteq M.

We claim MM is an essential ideal of 2{\mathcal{B}}_{2}. Indeed, suppose K2K\unlhd{\mathcal{B}}_{2} satisfies KM={0}K\cap M=\{0\}. Put

L:=α1(K)1L:=\alpha^{-1}(K)\unlhd{\mathcal{B}}_{1}

and note that LJ={0}L\cap J=\{0\} because

α(LJ)Kα(J)Kα(J)2¯KM={0}.\alpha(L\cap J)\subseteq K\cap\alpha(J)\subseteq K\cap\overline{\alpha(J){\mathcal{B}}_{2}}\subseteq K\cap M=\{0\}.

But JJ is an essential ideal in 1{\mathcal{B}}_{1}, so L={0}L=\{0\}. This gives Kα(1)={0}K\cap\alpha({\mathcal{B}}_{1})=\{0\}. Then K={0}K=\{0\} because (2,1,α|1)({\mathcal{B}}_{2},{\mathcal{B}}_{1},\alpha|_{{\mathcal{B}}_{1}}) has the ideal intersection property. Thus, MM is an essential ideal.

Since MM is an essential ideal, M={0}M^{\perp}=\{0\}. As M=MM=M^{\perp\perp}, M=2M={\mathcal{B}}_{2}. Hence E2(mm)=0E_{2}(m^{*}m)=0, so m=0m=0 by faithfulness of E2E_{2}. This gives α(v)b2α(v)2\alpha(v^{*})b_{2}\alpha(v)\in{\mathcal{B}}_{2}. ∎

We now return to the proof of Proposition 6.1. Suppose first that (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}) is a virtual Cartan inclusion. Given v𝒩(𝒜1,1)v\in{\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}), we shall show that 𝒵v{\mathcal{Z}}_{v} contains an essential ideal. We require two claims. Claim 1. Recall that {vv}={b11:vvb1=0}\{v^{*}v\}^{\perp}=\{b_{1}\in{\mathcal{B}}_{1}:v^{*}vb_{1}=0\}. Then

{vv}={b11:vb1=0}𝒵v.\{v^{*}v\}^{\perp}=\{b_{1}\in{\mathcal{B}}_{1}:vb_{1}=0\}\subseteq{\mathcal{Z}}_{v}.

Proof. To see {vv}{b11:vb1=0}\{v^{*}v\}^{\perp}\subseteq\{b_{1}\in{\mathcal{B}}_{1}:vb_{1}=0\}, note that for any polynomial pp with p(0)=0p(0)=0, 0=vp(vv)b1=p(vv)vb10=vp(v^{*}v)b_{1}=p(vv^{*})vb_{1}; then use

0=inf{p(vv)vv:p is a polynomial with p(0)=0}.0=\inf\{\left\lVert p(vv^{*})v-v\right\rVert:p\text{ is a polynomial with }p(0)=0\}.

The opposite inclusion is trivial, so {vv}={b11:vb1=0}\{v^{*}v\}^{\perp}=\{b_{1}\in{\mathcal{B}}_{1}:vb_{1}=0\}; this description gives {vv}𝒵v\{v^{*}v\}^{\perp}\subseteq{\mathcal{Z}}_{v}. \diamondsuit

Claim 2. Suppose Λ\Lambda is a non-empty set and {Jλ:λΛ}\{J_{\lambda}:\lambda\in\Lambda\} is a collection of ideals in 1{\mathcal{B}}_{1} such that: for every λΛ\lambda\in\Lambda, Jλ𝒵vJ_{\lambda}\subseteq{\mathcal{Z}}_{v}; and for distinct λ,μΛ\lambda,\mu\in\Lambda, JλJμ={0}J_{\lambda}\cap J_{\mu}=\{0\}. Write λΛJλ:=span¯λΛJλ\bigvee_{\lambda\in\Lambda}J_{\lambda}:=\overline{\operatorname{span}}\bigcup_{\lambda\in\Lambda}J_{\lambda}. Then λΛJλ1\bigvee_{\lambda\in\Lambda}J_{\lambda}\unlhd{\mathcal{B}}_{1} and λΛJλ𝒵v\bigvee_{\lambda\in\Lambda}J_{\lambda}\subseteq{\mathcal{Z}}_{v}.Proof. That λΛJλ\bigvee_{\lambda\in\Lambda}J_{\lambda} is an ideal in 1{\mathcal{B}}_{1} is clear. Let hspanλΛJλh\in\operatorname{span}\bigcup_{\lambda\in\Lambda}J_{\lambda}. Then we may find a finite set FΛF\subseteq\Lambda and for each λF\lambda\in F, an element hλJλh_{\lambda}\in J_{\lambda} so that

h=λFhλ.h=\sum_{\lambda\in F}h_{\lambda}.

For distinct λ,μF\lambda,\mu\in F, hλhμJλJμ={0}h_{\lambda}h_{\mu}\in J_{\lambda}\cap J_{\mu}=\{0\}, so hλhμ=0h_{\lambda}h_{\mu}=0. A calculation then shows that {vhλ:λF}\{vh_{\lambda}:\lambda\in F\} is a collection of pairwise orthogonal normalizers in 𝒩(𝒜1,1){\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}). But hλ𝒵vh_{\lambda}\in{\mathcal{Z}}_{v} for every λ\lambda, whence {α(vhλ)}λF\{\alpha(vh_{\lambda})\}_{\lambda\in F} is a family of pairwise orthogonal normalizers in 𝒩(𝒜2,2){\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2}). Fact 6.1 shows that λFα(hλ)𝒩(𝒜2,2)\sum_{\lambda\in F}\alpha(h_{\lambda})\in{\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2}). Since

α(v)(λFα(hλ))𝒩(𝒜2,2),\alpha(v)\left(\sum_{\lambda\in F}\alpha(h_{\lambda})\right)\in{\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2}),

h𝒵vh\in{\mathcal{Z}}_{v}. The proof of Claim 2 is completed by observing that 𝒵v{\mathcal{Z}}_{v} is a closed set. (Indeed, if zi𝒵vz_{i}\in{\mathcal{Z}}_{v} converges to z1z\in{\mathcal{B}}_{1}, then

α(vz)=limα(vzi)𝒩(𝒜2,2)¯=𝒩(𝒜2,2),\alpha(vz)=\lim\alpha(vz_{i})\in\overline{{\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2})}={\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2}),

so z𝒵vz\in{\mathcal{Z}}_{v}.) \diamondsuit We are now prepared to show 𝒵v{\mathcal{Z}}_{v} contains an essential ideal. This fact is trivial when v=0v=0, so assume v0v\neq 0.

Let 𝒬{\mathcal{Q}} be a collection of ideals in 1{\mathcal{B}}_{1}. We shall say 𝒬{\mathcal{Q}} is a nice collection of ideals if:

  • for each J𝒬J\in{\mathcal{Q}}, J𝒵vJ\subseteq{\mathcal{Z}}_{v};

  • If J,K𝒬J,K\in{\mathcal{Q}} and JKJ\neq K, then JK={0}J\cap K=\{0\}; and

  • {vv}𝒬\{v^{*}v\}^{\perp}\in{\mathcal{Q}}.

By Claim 1, the singleton set, {{vv}}\{\{v^{*}v\}^{\perp}\}, is a nice collection, so the family 𝔔{\mathfrak{Q}} consisting of all nice collections of ideals is non-empty. Partially order 𝔔{\mathfrak{Q}} by inclusion. Zorn’s lemma provides a maximal nice collection of ideals; call it 𝔐{\mathfrak{M}}. Consider the ideal,

𝔍:=J𝔐J.{\mathfrak{J}}:=\bigvee_{J\in{\mathfrak{M}}}J.

Claim 2 shows 𝔍𝒵v{\mathfrak{J}}\subseteq{\mathcal{Z}}_{v}. We will show 𝔍{\mathfrak{J}} is an essential ideal in 1{\mathcal{B}}_{1} by showing 𝔍={0}{\mathfrak{J}}^{\perp}=\{0\}.

Suppose to the contrary that 𝔍{0}{\mathfrak{J}}^{\perp}\neq\{0\}. Since {vv}𝔍\{v^{*}v\}^{\perp}\subseteq{\mathfrak{J}},

𝔍{vv}.{\mathfrak{J}}^{\perp}\subseteq\{v^{*}v\}^{\perp\perp}.

Note that {vv}=vv1¯\{v^{*}v\}^{\perp\perp}=\overline{v^{*}v{\mathcal{B}}_{1}}^{\perp\perp}, and, since vv1¯{vv}\overline{v^{*}v{\mathcal{B}}_{1}}\subseteq\{v^{*}v\}^{\perp\perp} has the ideal intersection property, we find

𝔍vv1¯{0}.{\mathfrak{J}}^{\perp}\cap\overline{v^{*}v{\mathcal{B}}_{1}}\neq\{0\}.

Let 0b𝔍vv1¯0\neq b\in{\mathfrak{J}}^{\perp}\cap\overline{v^{*}v{\mathcal{B}}_{1}}. If vvb=0v^{*}vb=0, then b{vv}{vv}={0}b\in\{v^{*}v\}^{\perp}\cap\{v^{*}v\}^{\perp\perp}=\{0\}, which is impossible as b0b\neq 0. Therefore, vb0vb\neq 0. Since (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}) is a pseudo-Cartan inclusion, it has the faithful unique pseudo-expectation property. Thus, letting E:𝒜1I(1)E:{\mathcal{A}}_{1}\rightarrow I({\mathcal{B}}_{1}) be the pseudo-expectation, E(bvvb)0E(b^{*}v^{*}vb)\neq 0. Since spanN\operatorname{span}N is dense in 𝒜1{\mathcal{A}}_{1}, there exists a sequence tspanNt_{\ell}\in\operatorname{span}N such that tvbt_{\ell}\rightarrow vb. Then E(tvb)0E(t_{\ell}^{*}vb)\neq 0 for sufficiently large \ell, so there exists wNw\in N such that

E(wvb)0.E(w^{*}vb)\neq 0.

Because we have assumed (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}) is a virtual Cartan inclusion, we may apply Lemma 6.1 to obtain kK0(wvb)(wvb)(wvb)1¯k\in K_{0}(w^{*}vb)\cap\overline{(w^{*}vb)^{*}(w^{*}vb){\mathcal{B}}_{1}} such that

0wvbk1.0\neq w^{*}vbk\in{\mathcal{B}}_{1}.

If (eλ)(e_{\lambda}) is an approximate unit for 𝔍vv1¯{\mathfrak{J}}^{\perp}\cap\overline{v^{*}v{\mathcal{B}}_{1}}, then

wvbk=limwvk(beλ),whence wvbk𝔍vv1¯.w^{*}vbk=\lim w^{*}vk(be_{\lambda}),\quad\text{whence }\quad w^{*}vbk\in{\mathfrak{J}}^{\perp}\cap\overline{v^{*}v{\mathcal{B}}_{1}}.

Therefore v(wvbk)0v(w^{*}vbk)^{*}\neq 0, for otherwise (wvbk)vv1¯vv1¯(w^{*}vbk)^{*}\in\overline{v^{*}v{\mathcal{B}}_{1}}^{\perp}\cap\overline{v^{*}v{\mathcal{B}}_{1}}. Then

0v(wvbk)=(vkbv)wN.0\neq v(w^{*}vbk)^{*}=(vk^{*}b^{*}v^{*})w\in N.

Since α(N)𝒩(𝒜2,2)\alpha(N)\subseteq{\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2}), we conclude that

h:=(wvbk)h:=(w^{*}vbk)^{*}

is a non-zero element of JvJ_{v}.

Let L:=h1¯L:=\overline{h{\mathcal{B}}_{1}} be the ideal generated by hh. Then L𝒵vL\subseteq{\mathcal{Z}}_{v}, and since we noted above that h𝔍vv1¯h^{*}\in{\mathfrak{J}}^{\perp}\cap\overline{v^{*}v{\mathcal{B}}_{1}}, we see that for every J𝔐J\in{\mathfrak{M}}, LJ={0}L\cap J=\{0\}. Thus 𝔐{L}{\mathfrak{M}}\cup\{L\} is a nice collection of ideals properly containing 𝔐{\mathfrak{M}}. As this contradicts maximality of 𝔐{\mathfrak{M}}, we obtain 𝔍={0}{\mathfrak{J}}^{\perp}=\{0\}. Therefore, 𝔍{\mathfrak{J}} is an essential ideal.

Apply Lemma 6.1 to 𝒵v{\mathcal{Z}}_{v} and 𝒵v{\mathcal{Z}}_{v^{*}} to conclude that α(v)𝒩(𝒜2,2)\alpha(v)\in{\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2}). If follows that α\alpha is a regular *-monomorphism. This completes the proof of the proposition when (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}) is a virtual Cartan inclusion.

Now suppose (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}) is a general pseudo-Cartan inclusion. By Observation 4.1, (𝒜1,1c)({\mathcal{A}}_{1},{\mathcal{B}}_{1}^{c}) is a virtual Cartan inclusion. Lemma 3(a) shows 𝒩(𝒜1,1)𝒩(𝒜1,1c){\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1})\subseteq{\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}^{c}) and Lemma 3(b) gives α(1c)2\alpha({\mathcal{B}}_{1}^{c})\subseteq{\mathcal{B}}_{2}. Let NN^{\prime} be the *-semigroup generated by N1cN\cup{\mathcal{B}}_{1}^{c}. Then 1cN𝒩(𝒜1,1c){\mathcal{B}}_{1}^{c}\subseteq N^{\prime}\subseteq{\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}^{c}). Writing wNw^{\prime}\in N^{\prime} as a finite product with factors belonging to N1cN\cup{\mathcal{B}}_{1}^{c}, we find α(w)𝒩(𝒜2,2)\alpha(w^{\prime})\in{\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2}). Therefore,

α(𝒩(𝒜1,1))α(𝒩(𝒜1,1c))𝒩(𝒜2,2),\alpha({\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}))\subseteq\alpha({\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}^{c}))\subseteq{\mathcal{N}}({\mathcal{A}}_{2},{\mathcal{B}}_{2}),

with the second inclusion obtained from the virtual Cartan case applied to (𝒜1,1c)({\mathcal{A}}_{1},{\mathcal{B}}_{1}^{c}) and NN^{\prime}. This completes the proof. ∎

We need an intertwining property for conditional expectations before stating and proving Theorem 6.1. We will discuss this property further in the introduction to Section 6.2.

Lemma \the\numberby.

Let (𝒜2,2 :: α)({\mathcal{A}}_{2},{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) be an essential and Cartan expansion for the Cartan inclusion (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}), and for i=1,2i=1,2, let Ei:𝒜iiE_{i}:{\mathcal{A}}_{i}\rightarrow{\mathcal{B}}_{i} be the conditional expectations. Then αE1=E2α\alpha\circ E_{1}=E_{2}\circ\alpha.

Proof.

Let (I(1),ι1)(I({\mathcal{B}}_{1}),\iota_{1}) be an injective envelope for 1{\mathcal{B}}_{1}. Since (𝒜2,2 :: α)({\mathcal{A}}_{2},{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) is an essential expansion of (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}), (2,1,α|1)({\mathcal{B}}_{2},{\mathcal{B}}_{1},\alpha|_{{\mathcal{B}}_{1}}) is an inclusion with the ideal intersection property. Proposition 2.4 shows there is a unique *-monomorphism ι2:2I(1)\iota_{2}:{\mathcal{B}}_{2}\rightarrow I({\mathcal{B}}_{1}) such that

(6.1.5) ι1=ι2(α|1).\iota_{1}=\iota_{2}\circ(\alpha|_{{\mathcal{B}}_{1}}).

For b11b_{1}\in{\mathcal{B}}_{1}, we have

(ι2E2α)(b1)=ι2(α(b1))=ι1(b1),(\iota_{2}\circ E_{2}\circ\alpha)(b_{1})=\iota_{2}(\alpha(b_{1}))=\iota_{1}(b_{1}),

so ι2E2α\iota_{2}\circ E_{2}\circ\alpha is a pseudo-expectation for (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}). Since ι1E1\iota_{1}\circ E_{1} is the pseudo-expectation for (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}), we have

ι2αE1=(6.1.5)ι1E1=ι2E2α.\iota_{2}\circ\alpha\circ E_{1}\stackrel{{\scriptstyle\eqref{EWork1}}}{{=}}\iota_{1}\circ E_{1}=\iota_{2}\circ E_{2}\circ\alpha.

Since ι2\iota_{2} is one-to-one, the lemma follows. ∎

We are now ready to establish a key mapping property of Cartan envelopes.

Theorem \the\numberby.

Suppose (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) are pseudo-Cartan inclusions and (𝒞2,𝒟2 :: α)({\mathcal{C}}_{2},{\mathcal{D}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) is a regular and essential expansion of (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}). For i=1,2i=1,2, let (𝒜i,i :: τi)({\mathcal{A}}_{i},{\mathcal{B}}_{i}\hbox{\,:\hskip-1.0pt:\,}\tau_{i}) be Cartan envelopes for (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}). The following statements hold.

  1. (a)

    There is a unique *-monomorphism α˘:𝒜1𝒜2\breve{\alpha}:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2} such that

    (6.1.6) α˘τ1=τ2α.\breve{\alpha}\circ\tau_{1}=\tau_{2}\circ\alpha.
  2. (b)

    (𝒜2,2 :: α˘)({\mathcal{A}}_{2},{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\breve{\alpha}) is a regular, essential expansion of (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}).

  3. (c)

    For i=1,2i=1,2, let Δi:𝒜ii\Delta_{i}:{\mathcal{A}}_{i}\rightarrow{\mathcal{B}}_{i} denote the conditional expectation. Then

    (6.1.7) α˘Δ1=Δ2α˘\breve{\alpha}\circ\Delta_{1}=\Delta_{2}\circ\breve{\alpha}

    and

    (6.1.8) α˘Δ1τ1=Δ2τ2α.\breve{\alpha}\circ\Delta_{1}\circ\tau_{1}=\Delta_{2}\circ\tau_{2}\circ\alpha.

The following commutative diagram illustrates the maps involved; the unlabelled vertical arrows represent inclusion maps.

(6.1.9) 𝒜1\textstyle{{\mathcal{A}}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}!α˘\scriptstyle{\exists!\,\breve{\alpha}}Δ1\scriptstyle{\Delta_{1}}𝒜2\textstyle{{\mathcal{A}}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ2\scriptstyle{\Delta_{2}}𝒞1\textstyle{{\mathcal{C}}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ1\scriptstyle{\tau_{1}}α\scriptstyle{\alpha}𝒞2\textstyle{{\mathcal{C}}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ2\scriptstyle{\tau_{2}}𝒟1\textstyle{{\mathcal{D}}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ1|𝒟1\scriptstyle{\tau_{1}|_{{\mathcal{D}}_{1}}}α|𝒟1\scriptstyle{\alpha|_{{\mathcal{D}}_{1}}}𝒟2\textstyle{{\mathcal{D}}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ2|𝒟2\scriptstyle{\tau_{2}|_{{\mathcal{D}}_{2}}}1\textstyle{{\mathcal{B}}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α˘|1\scriptstyle{\breve{\alpha}|_{{\mathcal{B}}_{1}}}2\textstyle{{\mathcal{B}}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
Proof.

(a) Note that (𝒜2,2 :: τ2α)({\mathcal{A}}_{2},{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\tau_{2}\circ\alpha) is a Cartan expansion of (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}). Also, the inclusions (2,𝒟2,τ2|𝒟2)({\mathcal{B}}_{2},{\mathcal{D}}_{2},\tau_{2}|_{{\mathcal{D}}_{2}}) and (𝒟2,𝒟1,α|𝒟1)({\mathcal{D}}_{2},{\mathcal{D}}_{1},\alpha|_{{\mathcal{D}}_{1}}) have the ideal intersection property, and therefore (2,𝒟1,τ2α|𝒟1)({\mathcal{B}}_{2},{\mathcal{D}}_{1},\tau_{2}\circ\alpha|_{{\mathcal{D}}_{1}}) also has the ideal intersection property. Since τ2α\tau_{2}\circ\alpha is a regular map, (𝒜2,2 :: τ2α)({\mathcal{A}}_{2},{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\tau_{2}\circ\alpha) is an essential and regular Cartan expansion of (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}).

Let

(6.1.10) 1:=C((Δ2τ2α)(𝒞1))and𝒜1:=C(1τ2(α(𝒞1)).{\mathcal{B}}_{1}^{\prime}:=C^{*}((\Delta_{2}\circ\tau_{2}\circ\alpha)({\mathcal{C}}_{1}))\quad\text{and}\quad{\mathcal{A}}_{1}^{\prime}:=C^{*}({\mathcal{B}}_{1}^{\prime}\cup\tau_{2}(\alpha({\mathcal{C}}_{1})).

Proposition 3 shows (𝒜1,1 :: τ2α)({\mathcal{A}}_{1}^{\prime},{\mathcal{B}}_{1}^{\prime}\hbox{\,:\hskip-1.0pt:\,}\tau_{2}\circ\alpha) is a Cartan envelope for (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}).

By the uniqueness of the Cartan envelope, there exists a unique *-isomorphism ψ:𝒜1𝒜1\psi:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{1}^{\prime} such that ψ\psi is regular and ψτ1=τ2α\psi\circ\tau_{1}=\tau_{2}\circ\alpha. Let f:𝒜1𝒜2f:{\mathcal{A}}_{1}^{\prime}\hookrightarrow{\mathcal{A}}_{2} be the inclusion map. Taking

α˘:=fψ,\breve{\alpha}:=f\circ\psi,

we obtain (6.1.6).

(b) By construction, α˘(1)=ψ(1)2\breve{\alpha}({\mathcal{B}}_{1})=\psi({\mathcal{B}}_{1})\subseteq{\mathcal{B}}_{2}. Let us show (2,α˘(1))({\mathcal{B}}_{2},\breve{\alpha}({\mathcal{B}}_{1})) has the ideal intersection property. The proof of part (a) shows that (2,τ2(α(𝒟1)))({\mathcal{B}}_{2},\tau_{2}(\alpha({\mathcal{D}}_{1}))) has the ideal intersection property. Since

τ2(α(𝒟1))ψ(1)=α˘(1)2,\tau_{2}(\alpha({\mathcal{D}}_{1}))\subseteq\psi({\mathcal{B}}_{1})=\breve{\alpha}({\mathcal{B}}_{1})\subseteq{\mathcal{B}}_{2},

(2,α˘(1))({\mathcal{B}}_{2},\breve{\alpha}({\mathcal{B}}_{1})) has the ideal intersection property. Thus (𝒜2,2 :: α˘)({\mathcal{A}}_{2},{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\breve{\alpha}) is an essential expansion of (𝒜1,1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}).

Let NN be the *-semigroup generated by 1τ1(𝒩(𝒞1,𝒟1)){\mathcal{B}}_{1}\cup\tau_{1}({\mathcal{N}}({\mathcal{C}}_{1},{\mathcal{D}}_{1})). Since (𝒜1,1 :: τ1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\tau_{1}) is a Cartan envelope for (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}), N𝒩(𝒜1,1)N\subseteq{\mathcal{N}}({\mathcal{A}}_{1},{\mathcal{B}}_{1}), and spanN\operatorname{span}N is dense in 𝒜1{\mathcal{A}}_{1}. Proposition 6.1 shows α˘\breve{\alpha} is a regular *-monomorphism. Thus the proof of (b) is complete.

(c) Lemma 6.1 (applied to α˘\breve{\alpha}) gives  (6.1.7). Finally, (6.1.7) gives

α˘Δ1τ1=Δ2α˘τ1=Δ2τ2α,\breve{\alpha}\circ\Delta_{1}\circ\tau_{1}=\Delta_{2}\circ\breve{\alpha}\circ\tau_{1}=\Delta_{2}\circ\tau_{2}\circ\alpha,

with the last equality following from (6.1.6). Thus (6.1.8) holds, and the proof is complete. ∎

Example \the\numberby. The conclusions of Theorem 6.1 need not hold without the hypothesis that α\alpha is an essential map. Indeed, Section 2 of [PittsCoStReInII] gives an example of a virtual Cartan inclusion (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and describes a Cartan package (𝒞2,𝒟2 :: α)({\mathcal{C}}_{2},{\mathcal{D}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) for (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) which is not the Cartan envelope for (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}). A Cartan envelope (𝒜1,1 :: τ1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\tau_{1}) for (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) is also described in [PittsCoStReInII, Section 2]. For convenience, we summarize a particular case of the discussion in [PittsCoStReInII, Section 2]. Let S=[0110]S_{-}=\begin{bmatrix}0&1\\ 1&0\end{bmatrix}, S+=[1001]S_{+}=\begin{bmatrix}1&0\\ 0&-1\end{bmatrix}, and put S0:=iSS+S_{0}:=iS_{-}S_{+}. Then S,S+S_{-},S_{+} and S0S_{0} are self-adjoint unitary matrices. Put

𝒞1\displaystyle{\mathcal{C}}_{1} :=C([1,1],M2())\displaystyle:=C([-1,1],M_{2}({\mathbb{C}}))
and
𝒟1\displaystyle{\mathcal{D}}_{1} :={f𝒞1:f(t)C(S) if t<0 and f(t)C(S+) if t>0};\displaystyle:=\{f\in{\mathcal{C}}_{1}:f(t)\in C^{*}(S_{-})\text{ if }t<0\text{ and }f(t)\in C^{*}(S_{+})\text{ if }t>0\};

continuity gives f(0)If(0)\in{\mathbb{C}}I for f𝒟1f\in{\mathcal{D}}_{1}. A Cartan package (𝒞2,𝒟2 :: α)({\mathcal{C}}_{2},{\mathcal{D}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha) for (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) is

𝒞2\displaystyle{\mathcal{C}}_{2} :=C([1,0],M2())M2()C([0,1],M2()),\displaystyle:=C([-1,0],M_{2}({\mathbb{C}}))\oplus M_{2}({\mathbb{C}})\oplus C([0,1],M_{2}({\mathbb{C}})),
𝒟2\displaystyle{\mathcal{D}}_{2} :=C([1,0],C(S))C(S0)C([0,1],C(S+)),and\displaystyle:=C([-1,0],C^{*}(S_{-}))\oplus C^{*}(S_{0})\oplus C([0,1],C^{*}(S_{+})),\quad\text{and}\quad
α(f)\displaystyle\alpha(f) :=f|[1,0]f(0)f|[0,1],f𝒞1.\displaystyle:=f|_{[-1,0]}\oplus f(0)\oplus f|_{[0,1]},\quad f\in{\mathcal{C}}_{1}.

A Cartan envelope for (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) is (𝒜1,1 :: τ1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\tau_{1}), where

𝒜1\displaystyle{\mathcal{A}}_{1} :=C([1,0],M2())C([0,1],M2()),\displaystyle:=C([-1,0],M_{2}({\mathbb{C}}))\oplus C([0,1],M_{2}({\mathbb{C}})),
1\displaystyle{\mathcal{B}}_{1} :=C([1,0],C(S))C([0,1],C(S+)),and\displaystyle:=C([-1,0],C^{*}(S_{-}))\oplus C([0,1],C^{*}(S_{+})),\quad\text{and}\quad
τ1(f)\displaystyle\tau_{1}(f) :=f|[1,0]f|[0,1],f𝒞1.\displaystyle:=f|_{[-1,0]}\oplus f|_{[0,1]},\quad f\in{\mathcal{C}}_{1}.

Since (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) is already a Cartan inclusion, a Cartan envelope for (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) is just (𝒞2,𝒟2 :: Id|𝒞2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}\hbox{\,:\hskip-1.0pt:\,}\operatorname{Id}|_{{\mathcal{C}}_{2}}). However, it is not hard to see that there is no embedding α˘:𝒜1𝒞2\breve{\alpha}:{\mathcal{A}}_{1}\rightarrow{\mathcal{C}}_{2} such that α˘τ1=Id|𝒞2α\breve{\alpha}\circ\tau_{1}=\operatorname{Id}|_{{\mathcal{C}}_{2}}\circ\alpha. Thus the conclusions of Theorem 6.1 do not hold.

The uniqueness and minimality statements from Theorem 3 give the following rigidity result for the Cartan envelopes of a pseudo-Cartan inclusion.

Proposition \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a pseudo-Cartan inclusion, with Cartan envelope (𝒜, :: τ)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\tau). If θ:𝒞𝒞\theta:{\mathcal{C}}\rightarrow{\mathcal{C}} is a *-automorphism such that θ(𝒩(𝒞,𝒟))𝒩(𝒞,𝒟)\theta({\mathcal{N}}({\mathcal{C}},{\mathcal{D}}))\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), then θ\theta uniquely extends to a regular *-automorphism θ˘:𝒜𝒜\breve{\theta}:{\mathcal{A}}\rightarrow{\mathcal{A}} such that

(6.1.12) θ˘τ=τθ.\breve{\theta}\circ\tau=\tau\circ\theta.
Proof.

We first show that (𝒜, :: τθ)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\tau\circ\theta) is a Cartan package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). By definition of the Cartan envelope, τ\tau is a regular map, and as θ\theta is assumed regular, we find τθ\tau\circ\theta is a regular map. Let Δ:𝒜\Delta:{\mathcal{A}}\rightarrow{\mathcal{B}} be the conditional expectation. Since (𝒜, :: τ)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\tau) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}),

\displaystyle{\mathcal{B}} =C(Δ(τ(𝒞)))=C(Δ(τ(θ(𝒞))))\displaystyle=C^{*}(\Delta(\tau({\mathcal{C}})))=C^{*}(\Delta(\tau(\theta({\mathcal{C}}))))
and
𝒜\displaystyle{\mathcal{A}} =C(τ(𝒞))=C(τ(θ(𝒞))).\displaystyle=C^{*}(\tau({\mathcal{C}})\cup{\mathcal{B}})=C^{*}(\tau(\theta({\mathcal{C}}))\cup{\mathcal{B}}).

An examination of Definition 3(a) shows (𝒜, :: τθ)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\tau\circ\theta) is a Cartan package for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

The ideal 𝔍𝒜{\mathfrak{J}}\unlhd{\mathcal{A}} obtained from the minimality statement of Theorem 3 satisfies

{0}=𝔍τ(θ(𝒞))=𝔍τ(𝒞).\{0\}={\mathfrak{J}}\cap\tau(\theta({\mathcal{C}}))={\mathfrak{J}}\cap\tau({\mathcal{C}}).

By Observation 4.3, (𝒜,𝒞,τ)({\mathcal{A}},{\mathcal{C}},\tau) has the ideal intersection property, so 𝔍={0}{\mathfrak{J}}=\{0\}. Therefore, (𝒜, :: τθ)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\tau\circ\theta) is a Cartan envelope for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}).

Apply the uniqueness statement of Theorem 3 to obtain a unique regular *-automorphism θ˘\breve{\theta} of 𝒜{\mathcal{A}} satisfying (6.1.12). ∎

Remark \the\numberby. Let (𝒜,)({\mathcal{A}},{\mathcal{B}}) be a Cartan inclusion, or more generally, a regular MASA inclusion, and suppose θ˘:𝒜𝒜\breve{\theta}:{\mathcal{A}}\rightarrow{\mathcal{A}} is a regular *-automorphism. Then θ˘1\breve{\theta}^{-1} is also regular, because

(6.1.14) θ˘()=.\breve{\theta}({\mathcal{B}})={\mathcal{B}}.

(To see (6.1.14) holds, recall that  [PittsNoApUnInC*Al, Theorem 2.5] shows (𝒜,)({\mathcal{A}},{\mathcal{B}}) has the AUP, so by Lemma 3(b) θ˘()c=\breve{\theta}({\mathcal{B}})\subseteq{\mathcal{B}}^{c}={\mathcal{B}}; as θ˘()\breve{\theta}({\mathcal{B}}) is a MASA in 𝒜{\mathcal{A}}, θ˘()=\breve{\theta}({\mathcal{B}})={\mathcal{B}}.) In particular, the family of regular automorphisms of a Cartan inclusion or of a regular MASA inclusion is a group.

Automorphisms of certain Cartan inclusions and virtual Cartan inclusions satisfying (6.1.14) have been studied, see [Komura*HoBeGrC*Al, Corollary 2.2] and [TaylorEsCoCaSuC*Al, Theorem 6.10]. We expect it is possible to obtain a description of regular *-automorphisms of a pseudo-Cartan inclusion by combining these results with Proposition 6.1.

If the MASA hypothesis in Remark 6.1 is dropped, the inverse of a regular automorphism need not be regular. Here is an example of a pseudo-Cartan inclusion where the family of regular *-automorphisms is not a group.

Example \the\numberby.

Let

𝒞\displaystyle{\mathcal{C}} ={hC():limth(t)=0 and limt+h(t) exists}\displaystyle=\{h\in C({\mathbb{R}}):\lim_{t\rightarrow-\infty}h(t)=0\text{ and }\lim_{t\rightarrow+\infty}h(t)\text{ exists}\}
and
𝒟\displaystyle{\mathcal{D}} ={h𝒞:h| is constant}.\displaystyle=\{h\in{\mathcal{C}}:h|_{\mathbb{N}}\text{ is constant}\}.

Since {\mathbb{N}}\subseteq{\mathbb{R}} has empty interior, (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the ideal intersection property. Also,

(6.1.15) 𝒩(𝒞,𝒟)={h𝒞:|h| is constant on },{\mathcal{N}}({\mathcal{C}},{\mathcal{D}})=\{h\in{\mathcal{C}}:|h|\text{ is constant on }{\mathbb{N}}\},

so

N:={h𝒞:h(n)𝕋 for all n}N:=\{h\in{\mathcal{C}}:h(n)\in{\mathbb{T}}\text{ for all }n\in{\mathbb{N}}\}

is a *-semigroup of normalizers. The Stone-Weiererstrauß theorem shows spanN\operatorname{span}N is a dense subalgebra of 𝒞{\mathcal{C}}. Thus (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is regular, and hence a pseudo-Cartan inclusion.

For f𝒞f\in{\mathcal{C}}, define θ(f)(t)=f(t+1)\theta(f)(t)=f(t+1). Using (6.1.15) we see θ\theta is regular, but θ1\theta^{-1} is not.

6.2. Inductive Limits

In  [KumjianOnC*Di], Kumjian introduced CC^{*}-diagonals, a class of inclusions subsequently broadened by Renault to the class of Cartan inclusions. An inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a CC^{*}-diagonal if it is a Cartan inclusion such that every pure state on 𝒟{\mathcal{D}} extends uniquely to a state on 𝒞{\mathcal{C}}. (This definition is equivalent to Kumjian’s original definition, see [PittsNoApUnInC*Al, Proposition 2.10].) The class of unital CC^{*}-diagonals is closed under inductive limits provided the connecting maps are unital, regular, and one-to-one ([DonsigPittsCoSyBoIs, Theorem 4.23]), and it can be shown that the same is true in the non-unital setting (with connecting maps again being regular *-monomorphisms). The unique extension property implies that the connecting mappings behave well with respect to the conditional expectations: if for i=1,2i=1,2, (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are CC^{*}-diagonals and Ei:𝒞i𝒟iE_{i}:{\mathcal{C}}_{i}\rightarrow{\mathcal{D}}_{i} is the conditional expectation, then for any regular *-monomorphism α21:𝒞1𝒞2\alpha_{21}:{\mathcal{C}}_{1}\rightarrow{\mathcal{C}}_{2},

(6.2.1) α21E1=E2α21.\alpha_{21}\circ E_{1}=E_{2}\circ\alpha_{21}.

Turning to the context of Cartan inclusions, when the connecting maps are regular and satisfy (6.2.1), Li showed the inductive limit of a sequence of Cartan inclusions is again a Cartan inclusion, see [LiXinEvClSiC*AlCaSu, Theorem 1.10]. Li’s result was extended for non-commutative Cartan inclusions by Meyer, Raad and Taylor, see [MeyerRaadTaylorInLiNoCaIn, Theorem 3.9].

For Cartan inclusions, it follows from Lemma 6.1 and Remark 3(b) that (6.2.1) holds provided the inclusion (𝒟2,𝒟1,α21)({\mathcal{D}}_{2},{\mathcal{D}}_{1},\alpha_{21}) has the ideal intersection property and α21\alpha_{21} is a regular map. However, when this assumption on (𝒟2,𝒟1,α21)({\mathcal{D}}_{2},{\mathcal{D}}_{1},\alpha_{21}) is dropped, examples show (6.2.1) need not hold. Such examples suggest that when the connecting maps are regular, but fail to be essential, an inductive limit of Cartan inclusions may not be a Cartan inclusion.

In this subsection, we consider inductive limits of pseudo-Cartan inclusions. We show that when the connecting maps in an inductive system are essential and regular *-monomorphisms, then the inductive limit of pseudo-Cartan inclusions is again a pseudo-Cartan inclusion. Further, we show that the Cartan envelope of the inductive limit of such a system is the inductive limits of the Cartan envelopes.

The following gives the notion of a directed system of pseudo-Cartan inclusions suitable for our purposes.

Definition \the\numberby.

Let Λ\Lambda be a directed set. Suppose

  • for each λΛ\lambda\in\Lambda, (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}) is a pseudo-Cartan inclusion;

  • for every (λ,μ)Λ×Λ(\lambda,\mu)\in\Lambda\times\Lambda with λμ\lambda\geq\mu, (𝒞λ,𝒟λ :: αλμ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\alpha_{\lambda\mu}) is an essential and regular expansion of (𝒞μ,𝒟μ)({\mathcal{C}}_{\mu},{\mathcal{D}}_{\mu}); and

  • whenever λμν\lambda\geq\mu\geq\nu, αλν=αλμαμν\alpha_{\lambda\nu}=\alpha_{\lambda\mu}\circ\alpha_{\mu\nu}.

  1. (a)

    We will call the collection {(𝒞λ,𝒟λ :: αλμ):λμ}\{({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\alpha_{\lambda\mu}):\lambda\geq\mu\} a system of pseudo-Cartan inclusions directed by Λ\Lambda. When the directed set is clear from context or it is not necessary to specify it, we will simplify terminology and say {(𝒞λ,𝒟λ :: αλμ):λμ}\{({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\alpha_{\lambda\mu}):\lambda\geq\mu\} is a system of pseudo-Cartan inclusions.

  2. (b)

    Given a system {(𝒞λ,𝒟λ :: αλμ):λμ}\{({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\alpha_{\lambda\mu}):\lambda\geq\mu\}, recall from the definition of expansion (see Definition 2.1(b)) that αλμ(𝒟μ)𝒟λ\alpha_{\lambda\mu}({\mathcal{D}}_{\mu})\subseteq{\mathcal{D}}_{\lambda}. Letting lim𝒞λ\varinjlim{\mathcal{C}}_{\lambda} and lim𝒟λ\varinjlim{\mathcal{D}}_{\lambda} be the inductive limit CC^{*}-algebras, we obtain the inclusion (lim𝒞λ,lim𝒟λ)(\varinjlim{\mathcal{C}}_{\lambda},\varinjlim{\mathcal{D}}_{\lambda}). We will call this inclusion the inductive limit of the system {(𝒞λ,𝒟λ :: αλμ):λμ}\{({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\alpha_{\lambda\mu}):\lambda\geq\mu\}.

  3. (c)

    Finally, for each λΛ\lambda\in\Lambda, we will let αλ:𝒞λlim𝒞λ\underrightarrow{\alpha_{\lambda}}:{\mathcal{C}}_{\lambda}\rightarrow\varinjlim{\mathcal{C}}_{\lambda} be the canonical embedding: it satisfies

    (6.2.2) αλαλμ=αμfor everyλμ.\underrightarrow{\alpha_{\lambda}}\circ\alpha_{\lambda\mu}=\underrightarrow{\alpha_{\mu}}\quad\text{for every}\quad\lambda\geq\mu.

Remark \the\numberby. By Observation 4.3, if {(𝒞λ,𝒟λ :: αλμ):λμ}\{({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\alpha_{\lambda\mu}):\lambda\geq\mu\} is a system of pseudo-Cartan inclusions, then whenever λμ\lambda\geq\mu, (𝒞λ,𝒞μ,αλμ)({\mathcal{C}}_{\lambda},{\mathcal{C}}_{\mu},\alpha_{\lambda\mu}) has the ideal intersection property.

We now show that the inductive limit of a system of pseudo-Cartan inclusions is again a pseudo-Cartan inclusion, and when the system consists of Cartan inclusions, the inductive limit is Cartan. (In the latter case, we do not apply [LiXinEvClSiC*AlCaSu, Theorem 1.10], preferring instead to use Theorem 6.2(a).)

Theorem \the\numberby.

Suppose Λ\Lambda is a directed set and {(𝒞λ,𝒟λ :: αλμ):λμ}\{({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\alpha_{\lambda\mu}):\lambda\geq\mu\} is a system of pseudo-Cartan inclusions. Let

𝒞:=lim𝒞λand𝒟:=lim𝒟λ.{\mathcal{C}}:=\varinjlim{\mathcal{C}}_{\lambda}\quad\text{and}\quad{\mathcal{D}}:=\varinjlim{\mathcal{D}}_{\lambda}.

The following statements hold.

  1. (a)

    (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion, and for each λΛ\lambda\in\Lambda, (𝒞,𝒟 :: αλ)({\mathcal{C}},{\mathcal{D}}\hbox{\,:\hskip-1.0pt:\,}\underrightarrow{\alpha_{\lambda}}) is a regular and essential expansion for (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}).

  2. (b)

    Suppose each (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}) is a Cartan inclusion and Δλ:𝒞λ𝒟λ\Delta_{\lambda}:{\mathcal{C}}_{\lambda}\rightarrow{\mathcal{D}}_{\lambda} is the conditional expectation. Then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a Cartan inclusion and letting Δ:𝒞𝒟\Delta:{\mathcal{C}}\rightarrow{\mathcal{D}} be the conditional expectation, we have Δαλ=αλΔλ\Delta\circ\underrightarrow{\alpha_{\lambda}}=\underrightarrow{\alpha_{\lambda}}\circ\Delta_{\lambda}.

Proof.

(a) Let us show αλ\underrightarrow{\alpha_{\lambda}} is a regular map of (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}) into (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Let v𝒩(𝒞λ,𝒟λ)v\in{\mathcal{N}}({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}) and d𝒟d\in{\mathcal{D}}. For ε>0\varepsilon>0, we may find μλ\mu\geq\lambda and dμ𝒟μd_{\mu}\in{\mathcal{D}}_{\mu} such that v2dαμ(dμ)<ε\left\lVert v\right\rVert^{2}\left\lVert d-\underrightarrow{\alpha_{\mu}}(d_{\mu})\right\rVert<\varepsilon. Since αμαμλ=αλ\underrightarrow{\alpha_{\mu}}\circ\alpha_{\mu\lambda}=\underrightarrow{\alpha_{\lambda}} and αμλ\alpha_{\mu\lambda} is a regular map,

αλ(v)αμ(dμ)αλ(v)=αμ(αμλ(v)dμαμλ(v))αμ(𝒟μ)𝒟.\underrightarrow{\alpha_{\lambda}}(v)\underrightarrow{\alpha_{\mu}}(d_{\mu})\underrightarrow{\alpha_{\lambda}}(v^{*})=\underrightarrow{\alpha_{\mu}}(\alpha_{\mu\lambda}(v)d_{\mu}\alpha_{\mu\lambda}(v)^{*})\in\underrightarrow{\alpha_{\mu}}({\mathcal{D}}_{\mu})\subseteq{\mathcal{D}}.

But

αλ(v)dαλ(v)αλ(v)αμ(dμ)αλ(v)v2dαμ(dμ)<ε.\left\lVert\underrightarrow{\alpha_{\lambda}}(v)d\underrightarrow{\alpha_{\lambda}}(v^{*})-\underrightarrow{\alpha_{\lambda}}(v)\underrightarrow{\alpha_{\mu}}(d_{\mu})\underrightarrow{\alpha_{\lambda}}(v^{*})\right\rVert\leq\left\lVert v\right\rVert^{2}\left\lVert d-\underrightarrow{\alpha_{\mu}}(d_{\mu})\right\rVert<\varepsilon.

It follows that αλ(v)dαλ(v)𝒟\underrightarrow{\alpha_{\lambda}}(v)d\underrightarrow{\alpha_{\lambda}}(v)^{*}\in{\mathcal{D}}. Likewise, αλ(v)dαλ(v)𝒟\underrightarrow{\alpha_{\lambda}}(v)^{*}d\underrightarrow{\alpha_{\lambda}}(v)\in{\mathcal{D}}. Thus αλ\underrightarrow{\alpha_{\lambda}} is a regular map.

We now show (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular inclusion. Since each (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}) is a regular inclusion we find that

𝒞=λαλ(𝒞λ)¯=λspanαλ(𝒩(𝒞λ,𝒟λ))¯span¯𝒩(𝒞,𝒟).{\mathcal{C}}=\overline{\bigcup_{\lambda}\underrightarrow{\alpha_{\lambda}}({\mathcal{C}}_{\lambda})}=\overline{\bigcup_{\lambda}\operatorname{span}\underrightarrow{\alpha_{\lambda}}({\mathcal{N}}({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}))}\subseteq\overline{\operatorname{span}}{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}).

Therefore (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular inclusion.

To complete the proof that (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion, we show that (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property. Let (I(𝒟),ι)(I({\mathcal{D}}),\iota) be an injective envelope for 𝒟{\mathcal{D}} and for λΛ\lambda\in\Lambda, define ιλ:𝒟λI(𝒟)\iota_{\lambda}:{\mathcal{D}}_{\lambda}\rightarrow I({\mathcal{D}}) by

ιλ:=ιαλ|𝒟λ.\iota_{\lambda}:=\iota\circ\underrightarrow{\alpha_{\lambda}}|_{{\mathcal{D}}_{\lambda}}.

Note that for μλ\mu\geq\lambda,

ιμαμλ|𝒟λ=ιλ.\iota_{\mu}\circ\alpha_{\mu\lambda}|_{{\mathcal{D}}_{\lambda}}=\iota_{\lambda}.

We first claim that for every λΛ\lambda\in\Lambda, (I(𝒟),ιλ)(I({\mathcal{D}}),\iota_{\lambda}) is an injective envelope for 𝒟λ{\mathcal{D}}_{\lambda}. It suffices to show the inclusion (I(𝒟),𝒟λ,ιλ)(I({\mathcal{D}}),{\mathcal{D}}_{\lambda},\iota_{\lambda}) has the ideal intersection property. Let JI(𝒟)J\unlhd I({\mathcal{D}}) satisfy ιλ(𝒟λ)J={0}\iota_{\lambda}({\mathcal{D}}_{\lambda})\cap J=\{0\}. Choose μλ\mu\geq\lambda. Since αμλ\alpha_{\mu\lambda} is an essential map, (𝒟μ,𝒟λ,αμλ|𝒟λ)({\mathcal{D}}_{\mu},{\mathcal{D}}_{\lambda},\alpha_{\mu\lambda}|_{{\mathcal{D}}_{\lambda}}) has the ideal intersection property. Therefore,

(ιμ(𝒟μ),𝒟λ,ιμαμλ|𝒟λ)=(ιμ(𝒟μ),𝒟λ,ιλ)(\iota_{\mu}({\mathcal{D}}_{\mu}),{\mathcal{D}}_{\lambda},\iota_{\mu}\circ\alpha_{\mu\lambda}|_{{\mathcal{D}}_{\lambda}})=(\iota_{\mu}({\mathcal{D}}_{\mu}),{\mathcal{D}}_{\lambda},\iota_{\lambda})

has the ideal intersection property as well. As Jιμ(𝒟μ)ιλ(𝒟λ)=Jιλ(𝒟λ)={0}J\cap\iota_{\mu}({\mathcal{D}}_{\mu})\cap\iota_{\lambda}({\mathcal{D}}_{\lambda})=J\cap\iota_{\lambda}({\mathcal{D}}_{\lambda})=\{0\}, we see Jιμ(𝒟μ)={0}J\cap\iota_{\mu}({\mathcal{D}}_{\mu})=\{0\} for every μλ\mu\geq\lambda. But μλαμ(𝒟μ)\bigcup_{\mu\geq\lambda}\alpha_{\mu}({\mathcal{D}}_{\mu}) is dense in 𝒟{\mathcal{D}}, so μλιμ(𝒟μ)\bigcup_{\mu\geq\lambda}\iota_{\mu}({\mathcal{D}}_{\mu}) is dense in ι(𝒟)\iota({\mathcal{D}}). By [BlackadarOpAl, Proposition II.8.2.4], Jι(𝒟)={0}J\cap\iota({\mathcal{D}})=\{0\}. Since (I(𝒟),𝒟,ι)(I({\mathcal{D}}),{\mathcal{D}},\iota) has the ideal intersection property, we conclude J={0}J=\{0\}. Thus our claim holds.

We are now ready to show (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has a unique pseudo-expectation. For i=1,2i=1,2, let Ei:𝒞I(𝒟)E_{i}:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be a pseudo-expectation. Then EiαλE_{i}\circ\underrightarrow{\alpha_{\lambda}} is a pseudo-expectation for (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}) relative to (I(𝒟),ιλ)(I({\mathcal{D}}),\iota_{\lambda}). Since (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}) has the faithful unique pseudo-expectation property, E1αλ=E2αλE_{1}\circ\underrightarrow{\alpha_{\lambda}}=E_{2}\circ\underrightarrow{\alpha_{\lambda}}. As this holds for every λ\lambda, we conclude that E1=E2E_{1}=E_{2}. Thus, (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has a unique pseudo-expectation E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}).

We now show EE is faithful. Let :={x𝒞:E(xx)=0}{\mathcal{L}}:=\{x\in{\mathcal{C}}:E(x^{*}x)=0\} be the left kernel of EE. It follows from [PittsStReInII, Theorem 6.5] and Lemma 2.3 that {\mathcal{L}} is an ideal in 𝒞{\mathcal{C}}. Then αλ1()\underrightarrow{\alpha_{\lambda}}^{-1}({\mathcal{L}}) is an ideal of 𝒞λ{\mathcal{C}}_{\lambda} contained in λ:={x𝒞λ:E(αλ(xx))=0}{\mathcal{L}}_{\lambda}:=\{x\in{\mathcal{C}}_{\lambda}:E(\underrightarrow{\alpha_{\lambda}}(x^{*}x))=0\}. Since (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}) has the faithful unique pseudo-expectation property and Eλ=EαλE_{\lambda}=E\circ\underrightarrow{\alpha_{\lambda}} is the pseudo-expectation for (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}), λ=0{\mathcal{L}}_{\lambda}=0. We conclude that αλ(𝒞λ)={0}{\mathcal{L}}\cap\underrightarrow{\alpha_{\lambda}}({\mathcal{C}}_{\lambda})=\{0\}. Thus, using [BlackadarOpAl, Proposition II.8.2.4] yet again, we find ={0}{\mathcal{L}}=\{0\}. Therefore, EE is faithful. Since (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property, it is a pseudo-Cartan inclusion.

As (I(𝒟),ιλ)(I({\mathcal{D}}),\iota_{\lambda}) is an injective envelope for 𝒟λ{\mathcal{D}}_{\lambda}, (I(𝒟),𝒟λ,ιλ)(I({\mathcal{D}}),{\mathcal{D}}_{\lambda},\iota_{\lambda}) is an essential inclusion. Since

I(𝒟)ι(𝒟)ιλ(𝒟λ)=ι(αλ(𝒟)),I({\mathcal{D}})\supseteq\iota({\mathcal{D}})\supseteq\iota_{\lambda}({\mathcal{D}}_{\lambda})=\iota(\underrightarrow{\alpha_{\lambda}}({\mathcal{D}})),

Lemma 2.4 shows (𝒟,𝒟λ,αλ)({\mathcal{D}},{\mathcal{D}}_{\lambda},\underrightarrow{\alpha_{\lambda}}) is also an essential inclusion. Since we have already shown that αλ:(𝒞λ,𝒟λ)(𝒞,𝒟)\underrightarrow{\alpha_{\lambda}}:({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda})\rightarrow({\mathcal{C}},{\mathcal{D}}) is a regular map, the proof of (a) is complete.

(b) Part (a) shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion. Clearly, for each λ\lambda, (𝒞λ,𝒟λ :: id|𝒞λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}{\operatorname{id}}|_{{\mathcal{C}}_{\lambda}}) is a Cartan envelope for (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}); let Δλ:𝒞λ𝒟λ\Delta_{\lambda}:{\mathcal{C}}_{\lambda}\rightarrow{\mathcal{D}}_{\lambda} be the conditional expectation. For μ,λΛ\mu,\lambda\in\Lambda with λμ\lambda\geq\mu, Lemma 6.1 gives

αλμΔμ=Δλαλμ.\alpha_{\lambda\mu}\circ\Delta_{\mu}=\Delta_{\lambda}\circ\alpha_{\lambda\mu}.

This means that there exists Δ:λαλ(𝒞λ)λαλ(𝒟λ)\Delta:\bigcup_{\lambda}\underrightarrow{\alpha_{\lambda}}({\mathcal{C}}_{\lambda})\rightarrow\bigcup_{\lambda}\underrightarrow{\alpha_{\lambda}}({\mathcal{D}}_{\lambda}) such that Δαλ=αλΔλ\Delta\circ\underrightarrow{\alpha_{\lambda}}=\underrightarrow{\alpha_{\lambda}}\circ\Delta_{\lambda}. Since each Δλ\Delta_{\lambda} is surjective, contractive and idempotent, Δ\Delta also has these properties. Therefore, Δ\Delta extends to a conditional expectation Δ:𝒞𝒟\Delta:{\mathcal{C}}\rightarrow{\mathcal{D}}. Proposition 2.3 shows (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a Cartan inclusion. Lemma 6.1 gives Δαλ=αλΔλ\Delta\circ\underrightarrow{\alpha_{\lambda}}=\underrightarrow{\alpha_{\lambda}}\circ\Delta_{\lambda}. ∎

We next show that the Cartan envelope of an inductive limit of a system of pseudo-Cartan inclusions is the inductive limit of the Cartan envelopes.

Theorem \the\numberby.

Suppose Λ\Lambda is a directed set and {(𝒞λ,𝒟λ :: αλμ):λμ}\{({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\alpha_{\lambda\mu}):\lambda\geq\mu\} is a system of pseudo-Cartan inclusions. By Theorem 6.2, (lim𝒞λ,lim𝒟λ)(\varinjlim{\mathcal{C}}_{\lambda},\varinjlim{\mathcal{D}}_{\lambda}) is a pseudo-Cartan inclusion.

For each λ\lambda, let (𝒜λ,λ :: τλ)({\mathcal{A}}_{\lambda},{\mathcal{B}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\tau_{\lambda}) be a Cartan envelope for (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}). Theorem 6.1 provides unique regular *-monomorphisms α˘λμ:(𝒜μ,μ)(𝒜λ,λ)\breve{\alpha}_{\lambda\mu}:({\mathcal{A}}_{\mu},{\mathcal{B}}_{\mu})\rightarrow({\mathcal{A}}_{\lambda},{\mathcal{B}}_{\lambda}) (λμ\lambda\geq\mu) satisfying

(6.2.4) α˘λμτμ=τλαλμ.\breve{\alpha}_{\lambda\mu}\circ\tau_{\mu}=\tau_{\lambda}\circ\alpha_{\lambda\mu}.

The following statements hold.

  1. (a)

    {(𝒜λ,λ :: α˘λμ):λμ}\{({\mathcal{A}}_{\lambda},{\mathcal{B}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\breve{\alpha}_{\lambda\mu}):\lambda\geq\mu\} is a system of Cartan inclusions and

    (lim𝒜λ,limλ)(\varinjlim{\mathcal{A}}_{\lambda},\varinjlim{\mathcal{B}}_{\lambda})

    is a Cartan inclusion.

  2. (b)

    There is a regular *-monomorphism τ:lim𝒞λlim𝒜λ\tau:\varinjlim{\mathcal{C}}_{\lambda}\rightarrow\varinjlim{\mathcal{A}}_{\lambda} such that for every λΛ\lambda\in\Lambda,

    (6.2.5) α˘λτλ=ταλ\underrightarrow{\breve{\alpha}_{\lambda}}\circ\tau_{\lambda}=\tau\circ\underrightarrow{\alpha_{\lambda}}

    and (lim𝒜λ,limλ :: τ)(\varinjlim{\mathcal{A}}_{\lambda},\varinjlim{\mathcal{B}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\tau) is a Cartan envelope for (lim𝒞λ,lim𝒟λ)(\varinjlim{\mathcal{C}}_{\lambda},\varinjlim{\mathcal{D}}_{\lambda}).

Proof.

Throughout the proof we will sometimes simplify notation and write,

𝒞\displaystyle{\mathcal{C}} :=lim𝒞λand𝒟:=lim𝒟λ.\displaystyle:=\varinjlim{\mathcal{C}}_{\lambda}\quad\text{and}\quad{\mathcal{D}}:=\varinjlim{\mathcal{D}}_{\lambda}.
Similarly, once part (a) is established, we will sometimes write
𝒜\displaystyle{\mathcal{A}} :=lim𝒜λand:=limλ.\displaystyle:=\varinjlim{\mathcal{A}}_{\lambda}\quad\text{and}\quad{\mathcal{B}}:=\varinjlim{\mathcal{B}}_{\lambda}.

(a) Theorem 6.1(b) shows that for λμ\lambda\geq\mu, (𝒜λ,λ :: α˘λμ)({\mathcal{A}}_{\lambda},{\mathcal{B}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\breve{\alpha}_{\lambda\mu}) is a regular and essential expansion of (𝒜μ,μ)({\mathcal{A}}_{\mu},{\mathcal{B}}_{\mu}). Now suppose λμν\lambda\geq\mu\geq\nu. We have

α˘λντν=τλαλν\breve{\alpha}_{\lambda\nu}\circ\tau_{\nu}=\tau_{\lambda}\circ\alpha_{\lambda\nu}

and

(α˘λμα˘μν)τν=α˘λμτμαμν=τλαλμαμν=τλαλν(\breve{\alpha}_{\lambda\mu}\circ\breve{\alpha}_{\mu\nu})\circ\tau_{\nu}=\breve{\alpha}_{\lambda\mu}\circ\tau_{\mu}\circ\alpha_{\mu\nu}=\tau_{\lambda}\circ\alpha_{\lambda\mu}\circ\alpha_{\mu\nu}=\tau_{\lambda}\circ\alpha_{\lambda\nu}

By the uniqueness part of Theorem 6.1(a), we find α˘λν=α˘λμα˘μν\breve{\alpha}_{\lambda\nu}=\breve{\alpha}_{\lambda\mu}\circ\breve{\alpha}_{\mu\nu}, so

{(𝒜λ,λ :: α˘λμ):λν}\{({\mathcal{A}}_{\lambda},{\mathcal{B}}_{\lambda}\hbox{\,:\hskip-1.0pt:\,}\breve{\alpha}_{\lambda\mu}):\lambda\geq\nu\}

is a system of Cartan inclusions. Theorem 6.2(b) shows (lim𝒜λ,limλ)(\varinjlim{\mathcal{A}}_{\lambda},\varinjlim{\mathcal{B}}_{\lambda}) is a Cartan inclusion.

(b) By properties of inductive limits and (6.2.4), there is a uniquely determined *-monomorphism

τ:𝒞𝒜\tau:{\mathcal{C}}\rightarrow{\mathcal{A}}

satisfying (6.2.5). While ταλ\tau\circ\underrightarrow{\alpha_{\lambda}} is a regular map, we have been unable to directly show that this forces τ\tau to be a regular map, because we lack a description of 𝒩(𝒞,𝒟){\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). (We expect 𝒩(𝒞,𝒟){\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) is the closure of λαλ(𝒩(𝒞λ,𝒟λ))\bigcup_{\lambda}\underrightarrow{\alpha_{\lambda}}({\mathcal{N}}({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda})), but we have not found a proof.)

Therefore, we proceed as follows. Fix a Cartan envelope

(𝒜, :: τ)({\mathcal{A}}^{\prime},{\mathcal{B}}^{\prime}\hbox{\,:\hskip-1.0pt:\,}\tau^{\prime})

for 𝒞,𝒟){\mathcal{C}},{\mathcal{D}}), with conditional expectation Δ:𝒜\Delta^{\prime}:{\mathcal{A}}^{\prime}\rightarrow{\mathcal{B}}^{\prime}. We shall show (𝒜,)({\mathcal{A}}^{\prime},{\mathcal{B}}^{\prime}) is an inductive limit of a system

{(𝒜λ,λ :: τλ):λμ}\{({\mathcal{A}}_{\lambda}^{\prime},{\mathcal{B}}_{\lambda}^{\prime}\hbox{\,:\hskip-1.0pt:\,}\tau_{\lambda}^{\prime}):\lambda\geq\mu\}

of Cartan envelopes for (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}), where each 𝒜λ𝒜{\mathcal{A}}_{\lambda}^{\prime}\subseteq{\mathcal{A}}^{\prime}. We will then use uniqueness of Cartan envelopes to produce an isomorphism ψ:lim𝒜λlim𝒜λ\psi^{\prime}:\varinjlim{\mathcal{A}}_{\lambda}^{\prime}\rightarrow\varinjlim{\mathcal{A}}_{\lambda} which carries limλ\varinjlim{\mathcal{B}}_{\lambda}^{\prime} onto limλ\varinjlim{\mathcal{B}}_{\lambda}. The regularity of τ\tau will then follow from the regularity of τ\tau^{\prime}. As we construct and discuss the various spaces and maps involved, the reader might find it helpful to consult Figure 1.

For λΛ\lambda\in\Lambda, define τλ:𝒞λ𝒜\tau_{\lambda}^{\prime}:{\mathcal{C}}_{\lambda}\rightarrow{\mathcal{A}}^{\prime} by

τλ:=ταλ,\tau_{\lambda}^{\prime}:=\tau^{\prime}\circ\underrightarrow{\alpha_{\lambda}},

and put

λ:=C(Δ(τλ(𝒞λ)),𝒜λ:=C(τλ(𝒞λ)λ),andΔλ:=Δ|𝒜λ.{\mathcal{B}}_{\lambda}^{\prime}:=C^{*}(\Delta^{\prime}(\tau_{\lambda}^{\prime}({\mathcal{C}}_{\lambda})),\quad{\mathcal{A}}_{\lambda}^{\prime}:=C^{*}(\tau_{\lambda}^{\prime}({\mathcal{C}}_{\lambda})\cup{\mathcal{B}}_{\lambda}^{\prime}),\quad\text{and}\quad\Delta_{\lambda}^{\prime}:=\Delta^{\prime}|_{{\mathcal{A}}_{\lambda}^{\prime}}.

For future use, we note that by construction,

(6.2.6) λand𝒜λ𝒜.{\mathcal{B}}_{\lambda}^{\prime}\subseteq{\mathcal{B}}^{\prime}\quad\text{and}\quad{\mathcal{A}}_{\lambda}^{\prime}\subseteq{\mathcal{A}}^{\prime}.

Theorem 6.2(a) shows (𝒞,𝒟 :: αλ)({\mathcal{C}},{\mathcal{D}}\hbox{\,:\hskip-1.0pt:\,}\underrightarrow{\alpha_{\lambda}}) is a regular and essential expansion of (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}) and τ\tau^{\prime} is regular by hypothesis. Therefore τλ\tau_{\lambda}^{\prime} is a regular map. Also,

τλ(𝒟λ)=τ(αλ(𝒟λ))τ(𝒟),\tau_{\lambda}^{\prime}({\mathcal{D}}_{\lambda})=\tau^{\prime}(\underrightarrow{\alpha_{\lambda}}({\mathcal{D}}_{\lambda}))\subseteq\tau^{\prime}({\mathcal{D}})\subseteq{\mathcal{B}}^{\prime},

so (,τ(𝒟λ))({\mathcal{B}}^{\prime},\tau^{\prime}({\mathcal{D}}_{\lambda})) has the ideal intersection property by Lemma 2.4. As τλ(𝒟λ)λ\tau_{\lambda}^{\prime}({\mathcal{D}}_{\lambda})\subseteq{\mathcal{B}}_{\lambda}^{\prime}\subseteq{\mathcal{B}}^{\prime} we see that (𝒜λ,λ :: τλ)({\mathcal{A}}_{\lambda}^{\prime},{\mathcal{B}}_{\lambda}^{\prime}\hbox{\,:\hskip-1.0pt:\,}\tau_{\lambda}^{\prime}) is a regular and essential expansion for (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}). Since (𝒜, :: τλ)({\mathcal{A}}^{\prime},{\mathcal{B}}^{\prime}\hbox{\,:\hskip-1.0pt:\,}\tau_{\lambda}^{\prime}) is an essential, regular and Cartan expansion for (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}), Proposition 3 shows (𝒜λ,λ :: τλ)({\mathcal{A}}_{\lambda}^{\prime},{\mathcal{B}}_{\lambda}^{\prime}\hbox{\,:\hskip-1.0pt:\,}\tau^{\prime}_{\lambda}) is a Cartan envelope for (𝒞λ,𝒟λ)({\mathcal{C}}_{\lambda},{\mathcal{D}}_{\lambda}) and Δλ\Delta_{\lambda}^{\prime} is the conditional expectation of 𝒜λ{\mathcal{A}}_{\lambda}^{\prime} onto λ{\mathcal{B}}_{\lambda}^{\prime}.

For λμ\lambda\geq\mu, let α˘λμ:𝒜μ𝒜λ\breve{\alpha}_{\lambda\mu}^{\prime}:{\mathcal{A}}_{\mu}^{\prime}\rightarrow{\mathcal{A}}_{\lambda}^{\prime} be the *-monomorphism satisfying

(6.2.7) α˘λμτμ=τλαλμ\breve{\alpha}_{\lambda\mu}^{\prime}\circ\tau_{\mu}^{\prime}=\tau_{\lambda}^{\prime}\circ\alpha_{\lambda\mu}

obtained from Theorem 6.1. Since

τμ=ταμ=ταλαλμ=τλαλμ,\tau_{\mu}^{\prime}=\tau^{\prime}\circ\underrightarrow{\alpha_{\mu}}=\tau^{\prime}\circ\underrightarrow{\alpha_{\lambda}}\circ\alpha_{\lambda\mu}=\tau_{\lambda}^{\prime}\circ\alpha_{\lambda\mu},

the uniqueness portion of Theorem 6.1 shows α˘λμ\breve{\alpha}_{\lambda\mu}^{\prime} is the inclusion map. Thus, α˘λ:𝒜λ𝒜\underrightarrow{\breve{\alpha}_{\lambda}}:{\mathcal{A}}_{\lambda}^{\prime}\rightarrow{\mathcal{A}}^{\prime} is also the inclusion map (see (6.2.6)). Theorem 6.1 shows (𝒜λ,λ :: α˘λμ)({\mathcal{A}}_{\lambda}^{\prime},{\mathcal{B}}_{\lambda}^{\prime}\hbox{\,:\hskip-1.0pt:\,}\breve{\alpha}_{\lambda\mu}^{\prime}) is a regular and essential expansion of (𝒜μ,μ)({\mathcal{A}}_{\mu}^{\prime},{\mathcal{B}}_{\mu}^{\prime}).

Since λαλ(𝒞λ)\bigcup_{\lambda}\underrightarrow{\alpha_{\lambda}}({\mathcal{C}}_{\lambda}) is dense in 𝒞{\mathcal{C}}, we find λΔ(τ(αλ(𝒞λ)))=Δ(τ(λαλ(𝒞λ)))\bigcup_{\lambda}\Delta^{\prime}(\tau^{\prime}(\underrightarrow{\alpha_{\lambda}}({\mathcal{C}}_{\lambda})))=\Delta^{\prime}(\tau^{\prime}(\bigcup_{\lambda}\underrightarrow{\alpha_{\lambda}}({\mathcal{C}}_{\lambda}))) is dense in Δ(τ(𝒞))\Delta^{\prime}(\tau^{\prime}({\mathcal{C}})). Therefore since =C(Δ(τ(𝒞))){\mathcal{B}}^{\prime}=C^{*}(\Delta^{\prime}(\tau^{\prime}({\mathcal{C}}))),

λλ=λC(Δ(τ(αλ(𝒞λ))))\bigcup_{\lambda}{\mathcal{B}}_{\lambda}^{\prime}=\bigcup_{\lambda}C^{*}(\Delta^{\prime}(\tau^{\prime}(\underrightarrow{\alpha_{\lambda}}({\mathcal{C}}_{\lambda}))))

is dense in {\mathcal{B}}^{\prime}. Similarly, λ(τ(αλ(𝒞λ))λ)=(λτ(αλ(𝒞λ)))(λλ)\bigcup_{\lambda}(\tau^{\prime}(\underrightarrow{\alpha_{\lambda}}({\mathcal{C}}_{\lambda}))\cup{\mathcal{B}}_{\lambda}^{\prime})=\left(\bigcup_{\lambda}\tau^{\prime}(\underrightarrow{\alpha_{\lambda}}({\mathcal{C}}_{\lambda}))\right)\cup(\bigcup_{\lambda}{\mathcal{B}}_{\lambda}^{\prime}) is dense in τ(𝒞)\tau^{\prime}({\mathcal{C}})\cup{\mathcal{B}}^{\prime}, so λ𝒜λ\bigcup_{\lambda}{\mathcal{A}}_{\lambda}^{\prime} is dense in 𝒜{\mathcal{A}}^{\prime}. We conclude that

𝒜=lim𝒜λand=limλ.{\mathcal{A}}^{\prime}=\varinjlim{\mathcal{A}}_{\lambda}^{\prime}\quad\text{and}\quad{\mathcal{B}}^{\prime}=\varinjlim{\mathcal{B}}_{\lambda}^{\prime}.
𝒜μ\textstyle{{\mathcal{A}}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α˘λμ\scriptstyle{\breve{\alpha}_{\lambda\mu}}𝒜λ\textstyle{{\mathcal{A}}_{\lambda}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α˘λ\scriptstyle{\underrightarrow{\breve{\alpha}_{\lambda}}}𝒜=lim𝒜λ\textstyle{{\mathcal{A}}=\varinjlim{\mathcal{A}}_{\lambda}}𝒞μ\textstyle{{\mathcal{C}}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αλμ\scriptstyle{\alpha_{\lambda\mu}}τμ\scriptstyle{\tau_{\mu}^{\prime}}τμ\scriptstyle{\tau_{\mu}}𝒞λ\textstyle{{\mathcal{C}}_{\lambda}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τλ\scriptstyle{\tau_{\lambda}^{\prime}}τλ\scriptstyle{\tau_{\lambda}}αλ\scriptstyle{\underrightarrow{\alpha_{\lambda}}}𝒞=lim𝒞λ\textstyle{{\mathcal{C}}=\varinjlim{\mathcal{C}}_{\lambda}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ\scriptstyle{\tau^{\prime}}τ\scriptstyle{\tau}𝒜μ\textstyle{{\mathcal{A}}_{\mu}^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α˘λμ\scriptstyle{\breve{\alpha}_{\lambda\mu}^{\prime}}ψμ\scriptstyle{\psi_{\mu}^{\prime}}𝒜λ\textstyle{{\mathcal{A}}_{\lambda}^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψλ\scriptstyle{\psi_{\lambda}^{\prime}}α˘λ\scriptstyle{\underrightarrow{\breve{\alpha}_{\lambda}^{\prime}}}𝒜=lim𝒜λ\textstyle{{\mathcal{A}}^{\prime}=\varinjlim{\mathcal{A}}_{\lambda}^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\scriptstyle{\exists\psi^{\prime}}
Figure 1. The squares on the left side having vertical sides labeled with a subscripted τ\tau or subscripted τ\tau^{\prime} commute by Theorem 6.1. The horizontal maps on the bottom row are inclusion maps.

We now use uniqueness of Cartan envelopes to show lim𝒜λ\varinjlim{\mathcal{A}}_{\lambda}^{\prime} is isomorphic to lim𝒜λ\varinjlim{\mathcal{A}}_{\lambda}. The uniqueness part of Theorem 3 gives the existence of a unique *-isomorphism ψλ:𝒜λ𝒜λ\psi^{\prime}_{\lambda}:{\mathcal{A}}_{\lambda}^{\prime}\rightarrow{\mathcal{A}}_{\lambda} satisfying

ψλτλ=τλ.\psi_{\lambda}^{\prime}\circ\tau_{\lambda}^{\prime}=\tau_{\lambda}.

By properties of inductive limits, to show the existence of an isomorphism ψ:lim𝒜λlim𝒜λ\psi^{\prime}:\varinjlim{\mathcal{A}}_{\lambda}^{\prime}\rightarrow\varinjlim{\mathcal{A}}_{\lambda} it suffices to show

(6.2.8) α˘λμψμ=ψλα˘λμ;\breve{\alpha}_{\lambda\mu}\circ\psi_{\mu}^{\prime}=\psi_{\lambda}^{\prime}\circ\breve{\alpha}_{\lambda\mu}^{\prime};

(that ψ\psi^{\prime} is invertible will follow from rewriting (6.2.8) as (ψλ)1α˘λμ=α˘λμ(ψμ)1(\psi_{\lambda}^{\prime})^{-1}\circ\breve{\alpha}_{\lambda\mu}=\breve{\alpha}^{\prime}_{\lambda\mu}\circ(\psi_{\mu}^{\prime})^{-1}). Recalling that 𝒜λ{\mathcal{A}}_{\lambda}^{\prime} is generated by τλ(𝒞λ)Δλ(τλ(𝒞λ))\tau_{\lambda}^{\prime}({\mathcal{C}}_{\lambda})\cup\Delta_{\lambda}^{\prime}(\tau_{\lambda}^{\prime}({\mathcal{C}}_{\lambda})), it is enough to show

(6.2.9) α˘λμψμτμ\displaystyle\breve{\alpha}_{\lambda\mu}\circ\psi_{\mu}^{\prime}\circ\tau_{\mu}^{\prime} =ψλα˘λμτμ\displaystyle=\psi_{\lambda}^{\prime}\circ\breve{\alpha}_{\lambda\mu}^{\prime}\circ\tau_{\mu}^{\prime}
and
(6.2.10) α˘λμψμΔμτμ\displaystyle\breve{\alpha}_{\lambda\mu}\circ\psi_{\mu}^{\prime}\circ\Delta_{\mu}^{\prime}\circ\tau_{\mu}^{\prime} =ψλα˘λμΔμτμ.\displaystyle=\psi_{\lambda}^{\prime}\circ\breve{\alpha}_{\lambda\mu}^{\prime}\circ\Delta_{\mu}^{\prime}\circ\tau_{\mu}^{\prime}.

For (6.2.9), the uniqueness part of Theorem 3 gives,

α˘λμψμτμ=(3)α˘λμτμ=(6.1)τλαλμ=(3)ψλτλαλμ=(6.1)ψλα˘λμτμ.\breve{\alpha}_{\lambda\mu}\circ\psi_{\mu}^{\prime}\circ\tau_{\mu}^{\prime}\stackrel{{\scriptstyle\eqref{!pschar}}}{{=}}\breve{\alpha}_{\lambda\mu}\circ\tau_{\mu}\stackrel{{\scriptstyle\eqref{cmap}}}{{=}}\tau_{\lambda}\circ\alpha_{\lambda\mu}\stackrel{{\scriptstyle\eqref{!pschar}}}{{=}}\psi_{\lambda}^{\prime}\circ\tau_{\lambda}^{\prime}\circ\alpha_{\lambda\mu}\stackrel{{\scriptstyle\eqref{cmap}}}{{=}}\psi_{\lambda}^{\prime}\circ\breve{\alpha}_{\lambda\mu}^{\prime}\circ\tau_{\mu}^{\prime}.

Turning to (6.2.10), ψμΔμ(ψμ)1\psi_{\mu}^{\prime}\circ\Delta_{\mu}^{\prime}\circ(\psi_{\mu}^{\prime})^{-1} is a conditional expectation from 𝒜μ{\mathcal{A}}_{\mu} onto μ{\mathcal{B}}_{\mu}. By uniqueness of the conditional expectation for Cartan inclusions, we obtain ψμΔμ(ψμ)1=Δμ\psi_{\mu}^{\prime}\circ\Delta_{\mu}^{\prime}\circ(\psi_{\mu}^{\prime})^{-1}=\Delta_{\mu}. Thus

(6.2.11) ψμΔμ=Δμψμ;similarlyψλΔλ=Δλψλ.\psi_{\mu}^{\prime}\circ\Delta_{\mu}^{\prime}=\Delta_{\mu}\circ\psi_{\mu}^{\prime};\quad\text{similarly}\quad\psi_{\lambda}^{\prime}\circ\Delta_{\lambda}^{\prime}=\Delta_{\lambda}\circ\psi_{\lambda}^{\prime}.

Hence,

α˘λμψμΔμτμ\displaystyle\breve{\alpha}_{\lambda\mu}\circ\psi_{\mu}^{\prime}\circ\Delta_{\mu}^{\prime}\circ\tau_{\mu}^{\prime} =(6.2.11)α˘λμΔμψμτμ\displaystyle\stackrel{{\scriptstyle\eqref{indlim3.b}}}{{=}}\breve{\alpha}_{\lambda\mu}\circ\Delta_{\mu}\circ\psi_{\mu}^{\prime}\circ\tau_{\mu}^{\prime}
=(3)α˘λμΔμτμ=(6.1.8)Δλτλαλμ\displaystyle\stackrel{{\scriptstyle\eqref{!pschar}}}{{=}}\breve{\alpha}_{\lambda\mu}\circ\Delta_{\mu}\circ\tau_{\mu}\stackrel{{\scriptstyle\eqref{cmap1.1}}}{{=}}\Delta_{\lambda}\circ\tau_{\lambda}\circ\alpha_{\lambda\mu}
=(3)Δλψλτλαλμ=(6.1.6)Δλψλα˘λμτμ\displaystyle\stackrel{{\scriptstyle\eqref{!pschar}}}{{=}}\Delta_{\lambda}\circ\psi_{\lambda}^{\prime}\circ\tau_{\lambda}^{\prime}\circ\alpha_{\lambda\mu}\stackrel{{\scriptstyle\eqref{cmap1}}}{{=}}\Delta_{\lambda}\circ\psi_{\lambda}^{\prime}\circ\breve{\alpha}_{\lambda\mu}^{\prime}\circ\tau_{\mu}^{\prime}
=(6.2.11)ψλΔλα˘λμτμ=(6.1.7)ψλα˘λμΔμτμ,\displaystyle\stackrel{{\scriptstyle\eqref{indlim3.b}}}{{=}}\psi_{\lambda}^{\prime}\circ\Delta_{\lambda}^{\prime}\circ\breve{\alpha}_{\lambda\mu}^{\prime}\circ\tau_{\mu}^{\prime}\stackrel{{\scriptstyle\eqref{cmap1.05}}}{{=}}\psi_{\lambda}^{\prime}\circ\breve{\alpha}_{\lambda\mu}^{\prime}\circ\Delta_{\mu}^{\prime}\circ\tau_{\mu}^{\prime},

so (6.2.10) holds. This completes the proof of (6.2.8), and the existence of the isomorphism ψ:lim𝒜λlim𝒜λ\psi^{\prime}:\varinjlim{\mathcal{A}}_{\lambda}^{\prime}\rightarrow\varinjlim{\mathcal{A}}_{\lambda}.

For each μΛ\mu\in\Lambda, ψμ(μ)=μ\psi_{\mu}^{\prime}({\mathcal{B}}_{\mu}^{\prime})={\mathcal{B}}_{\mu}, and α˘λμ(ψμ|λ)=ψλ(α˘λμ|μ)\breve{\alpha}_{\lambda\mu}\circ(\psi_{\mu}^{\prime}|_{{\mathcal{B}}_{\lambda}^{\prime}})=\psi_{\lambda}^{\prime}\circ(\breve{\alpha}_{\lambda\mu}^{\prime}|_{{\mathcal{B}}_{\mu}^{\prime}}), so

ψ(limλ)=limλ.\psi^{\prime}(\varinjlim{\mathcal{B}}_{\lambda}^{\prime})=\varinjlim{\mathcal{B}}_{\lambda}.

This implies that ψ\psi^{\prime} is regular. Since τλ=ψλτλ\tau_{\lambda}=\psi_{\lambda}^{\prime}\circ\tau_{\lambda}^{\prime} we obtain

τ=ψτ.\tau=\psi^{\prime}\circ\tau^{\prime}.

As ψ\psi and τ\tau^{\prime} are regular, we conclude τ\tau is regular as well. Hence (lim𝒜λ,limλ,τ)(\varinjlim{\mathcal{A}}_{\lambda},\varinjlim{\mathcal{B}}_{\lambda},\tau) is a Cartan envelope for (lim𝒞λ,lim𝒟λ)(\varinjlim{\mathcal{C}}_{\lambda},\varinjlim{\mathcal{D}}_{\lambda}). This completes the proof. ∎

6.3. Minimal Tensor Products

The purpose of this subsection is to show that the minimal tensor product of two pseudo-Cartan inclusions is again a pseudo-Cartan inclusion. Throughout, for i=1,2i=1,2, let (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) be pseudo-Cartan inclusions. For algebras 𝒜{\mathcal{A}} and {\mathcal{B}}, we use the notation 𝒜{\mathcal{A}}\odot{\mathcal{B}} for the linear span of {ab:a𝒜,b}\{a\otimes b:a\in{\mathcal{A}},b\in{\mathcal{B}}\}.

Regularity of the tensor product inclusion is straightforward, and recorded in the following lemma.

Lemma \the\numberby.

Let 𝒮:={v1v2:vi𝒩(𝒞i,𝒟i){\mathcal{S}}:=\{v_{1}\otimes v_{2}:v_{i}\in{\mathcal{N}}({\mathcal{C}}_{i},{\mathcal{D}}_{i}). Then 𝒮𝒩(𝒞1min𝒞2,𝒟1𝒟2){\mathcal{S}}\subseteq{\mathcal{N}}({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) and span𝒮\operatorname{span}{\mathcal{S}} is dense in 𝒞1min𝒞2{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}. In particular, (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) is a regular inclusion.

Proof.

Let 𝔄:=span𝒮{\mathfrak{A}}:=\operatorname{span}{\mathcal{S}}.

That 𝒮𝒩(𝒞1min𝒞2,𝒟1𝒟2){\mathcal{S}}\subseteq{\mathcal{N}}({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) is evident. For i=1,2i=1,2, suppose xi𝒞ix_{i}\in{\mathcal{C}}_{i}. We wish to show that x1x2𝔄¯x_{1}\otimes x_{2}\in\overline{\mathfrak{A}}. Let ε>0\varepsilon>0. By regularity of (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}), there exists NN\in{\mathbb{N}}, {vj}j=1N𝒩(𝒞1,𝒟1)\{v_{j}\}_{j=1}^{N}\subseteq{\mathcal{N}}({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and {wj}j=1N𝒩(𝒞2,𝒟2)\{w_{j}\}_{j=1}^{N}\subseteq{\mathcal{N}}({\mathcal{C}}_{2},{\mathcal{D}}_{2}) such that

x1j=1Nvj<εandx2j=1Nwj<ε.\left\lVert x_{1}-\sum_{j=1}^{N}v_{j}\right\rVert<\varepsilon\quad\text{and}\quad\left\lVert x_{2}-\sum_{j=1}^{N}w_{j}\right\rVert<\varepsilon.

Then

x1x2(j=1Nvj)(j=1Nwj)\displaystyle\left\lVert x_{1}\otimes x_{2}-\left(\sum_{j=1}^{N}v_{j}\right)\otimes\left(\sum_{j=1}^{N}w_{j}\right)\right\rVert (x1j=1Nvj)x2\displaystyle\leq\left\lVert\left(x_{1}-\sum_{j=1}^{N}v_{j}\right)\otimes x_{2}\right\rVert
+(j=1Nvj)(x2j=1Nwj)\displaystyle+\left\lVert\left(\sum_{j=1}^{N}v_{j}\right)\otimes\left(x_{2}-\sum_{j=1}^{N}w_{j}\right)\right\rVert
<εx2+(x1+ε)ε.\displaystyle<\varepsilon\left\lVert x_{2}\right\rVert+(\left\lVert x_{1}\right\rVert+\varepsilon)\varepsilon.

Therefore, x1x2𝔄¯x_{1}\otimes x_{2}\in\overline{\mathfrak{A}}. As the span of elementary tensors is dense in 𝒞1min𝒞2{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}, the lemma follows. ∎

Our next goal is Proposition 6.3, which shows that when (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) are unital pseudo-Cartan inclusions, then (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) has the unique pseudo-expectation property. On first glance, one might expect that if Ei:𝒞iI(𝒟i)E_{i}:{\mathcal{C}}_{i}\rightarrow I({\mathcal{D}}_{i}) are the unique pseudo-expectations for (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}), then E1E2E_{1}\otimes E_{2} will be the unique pseudo-expectation for (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}). However, the codomain of E1E2E_{1}\otimes E_{2} is I(𝒟1)I(𝒟2)I({\mathcal{D}}_{1})\otimes I({\mathcal{D}}_{2}), rather than I(𝒟1𝒟2)I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) and these CC^{*}-algebras need not be the same [WillardGeTo, Exercises 15G and 19I.2]. Nevertheless, Proposition 6.3 below implies there is an essential embedding gg of I(𝒟1)I(𝒟2)I({\mathcal{D}}_{1})\otimes I({\mathcal{D}}_{2}) into I(𝒟1𝒟2)I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}). Using gg to identify I(𝒟1)I(𝒟2)I({\mathcal{D}}_{1})\otimes I({\mathcal{D}}_{2}) with its image in I(𝒟1𝒟2)I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) will allow us to view E1E2E_{1}\otimes E_{2} as a pseudo-expectation for (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}).

Dual to the notion of an essential embedding of C(X)C(X) into C(Y)C(Y) is the notion of essential surjection of YY onto XX: a continuous surjection π:YX\pi:Y\twoheadrightarrow X between compact Hausdorff spaces is called an essential surjection if whenever FYF\subseteq Y is a closed set and π(F)=X\pi(F)=X, then F=YF=Y. (Essential surjections are also called irreducible maps.)

Lemma \the\numberby.

Suppose for i=1,2i=1,2, YiY_{i} and XiX_{i} are compact Hausdorff spaces, and πi:YiXi\pi_{i}:Y_{i}\twoheadrightarrow X_{i} are continuous and essential surjections. Then π1×π2:Y1×Y2X1×X2\pi_{1}\times\pi_{2}:Y_{1}\times Y_{2}\twoheadrightarrow X_{1}\times X_{2}, given by (y1,y2)(π1(y1),π2(y2))(y_{1},y_{2})\mapsto(\pi_{1}(y_{1}),\pi_{2}(y_{2})), is an essential surjection.

Proof.

We start by proving a special case: assume that Y2=X2Y_{2}=X_{2} and π2\pi_{2} is the identity map.

For typographical ease, let π:=π1×idX2\pi:=\pi_{1}\times{\operatorname{id}}_{X_{2}}. Let FY1×X2F\subseteq Y_{1}\times X_{2} be a closed set such that π(F)=X1×X2\pi(F)=X_{1}\times X_{2}. For tX2t\in X_{2}, let

Ft:={y1Y1:(y1,t)F}.F_{t}:=\{y_{1}\in Y_{1}:(y_{1},t)\in F\}.

Then FtF_{t} is a closed subset of Y1Y_{1}. Given sX1s\in X_{1}, since π(F)=X1×X2\pi(F)=X_{1}\times X_{2}, there is (y1,t)F(y_{1},t)\in F such that π(y1,t)=(s,t)\pi(y_{1},t)=(s,t). Thus π1(Ft)=X1\pi_{1}(F_{t})=X_{1}. Since π1\pi_{1} is essential, we obtain Ft=Y1F_{t}=Y_{1}. Hence for tX2t\in X_{2}, Y1×{t}FY_{1}\times\{t\}\subseteq F. We conclude that F=Y1×X2F=Y_{1}\times X_{2}, showing that the lemma holds in this special case.

For the general case, consider the composition,

Y1×Y2π1×idY2X1×Y2idX1×π2X1×X2.Y_{1}\times Y_{2}\stackrel{{\scriptstyle\pi_{1}\times{\operatorname{id}}_{Y_{2}}}}{{\longrightarrow}}X_{1}\times Y_{2}\stackrel{{\scriptstyle{\operatorname{id}}_{X_{1}}\times\pi_{2}}}{{\longrightarrow}}X_{1}\times X_{2}.

By the special case, each of the maps in the composition are essential surjections. As the composition of essential surjections is an essential surjection, the proof is complete. ∎

Proposition \the\numberby.

For i=1,2i=1,2, let i{\mathcal{B}}_{i} be abelian CC^{*}-algebras (perhaps not unital) and suppose (i,𝒟i,αi)({\mathcal{B}}_{i},{\mathcal{D}}_{i},\alpha_{i}) are essential inclusions. Then (12,𝒟1𝒟2,α1α2)({\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2},\alpha_{1}\otimes\alpha_{2}) is an essential inclusion.

Proof.

First assume the inclusions (i,𝒟i,αi)({\mathcal{B}}_{i},{\mathcal{D}}_{i},\alpha_{i}) are unital. For i=1,2i=1,2, let Yi:=^iY_{i}:=\widehat{\mathcal{B}}_{i} and Xi:=𝒟^iX_{i}:=\widehat{\mathcal{D}}_{i}. The maps αi\alpha_{i} dualize to continuous surjections πi:YiXi\pi_{i}:Y_{i}\rightarrow X_{i} (thus for d𝒟id\in{\mathcal{D}}_{i}, πi(yi)(d)=yi(αi(d))\pi_{i}(y_{i})(d)=y_{i}(\alpha_{i}(d))). The proof in this case now follows from Lemma 6.3 after noting that the inclusion (i,𝒟i,αi)({\mathcal{B}}_{i},{\mathcal{D}}_{i},\alpha_{i}) is essential if and only if πi\pi_{i} is an essential surjection of YiY_{i} onto XiX_{i}, see [PittsIrMaIsBoReOpSeReId, Lemma 4.9].

For the general case, note that (𝒟~1𝒟~2,𝒟1𝒟2,u𝒟1u𝒟2)(\tilde{\mathcal{D}}_{1}\otimes\tilde{\mathcal{D}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2},u_{{\mathcal{D}}_{1}}\otimes u_{{\mathcal{D}}_{2}}) is an essential inclusion because the image of 𝒟1𝒟2{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2} under u𝒟1u𝒟2u_{{\mathcal{D}}_{1}}\otimes u_{{\mathcal{D}}_{2}} is an essential ideal in 𝒟~1𝒟~2\tilde{\mathcal{D}}_{1}\otimes\tilde{\mathcal{D}}_{2}. (This can be shown directly or one can use the facts that 𝒟~1𝒟~2M(𝒟1)M(𝒟2)M(𝒟1𝒟2)\tilde{\mathcal{D}}_{1}\otimes\tilde{\mathcal{D}}_{2}\subseteq M({\mathcal{D}}_{1})\otimes M({\mathcal{D}}_{2})\subseteq M({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) and any CC^{*}-algebra is an essential ideal in its multiplier algebra. A proof of the inclusion of the multiplier algebras can be found at https://math.stackexchange.com/a/4458451.) This fact, together with the unital case show that both of the inclusions,

α1(𝒟1)α2(𝒟2)α~1(𝒟~1)α~2(𝒟~2))~1~2\alpha_{1}({\mathcal{D}}_{1})\otimes\alpha_{2}({\mathcal{D}}_{2})\subseteq\tilde{\alpha}_{1}(\tilde{\mathcal{D}}_{1})\otimes\tilde{\alpha}_{2}(\tilde{\mathcal{D}}_{2}))\subseteq\tilde{\mathcal{B}}_{1}\otimes\tilde{\mathcal{B}}_{2}

have the ideal intersection property. By Lemma 2.4, α1(𝒟1)α2(𝒟2)~1~2\alpha_{1}({\mathcal{D}}_{1})\otimes\alpha_{2}({\mathcal{D}}_{2})\subseteq\tilde{\mathcal{B}}_{1}\otimes\tilde{\mathcal{B}}_{2} has the ideal intersection property. Finally, since α1(𝒟1)α2(𝒟2)12~1~2\alpha_{1}({\mathcal{D}}_{1})\otimes\alpha_{2}({\mathcal{D}}_{2})\subseteq{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2}\subseteq\tilde{\mathcal{B}}_{1}\otimes\tilde{\mathcal{B}}_{2}, another application of Lemma 2.4 shows α1(𝒟1)α2(𝒟2)12\alpha_{1}({\mathcal{D}}_{1})\otimes\alpha_{2}({\mathcal{D}}_{2})\subseteq{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2} has the ideal intersection property, as desired. ∎

Assume now that (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are unital pseudo-Cartan inclusions, let (I(𝒟i),ιi)(I({\mathcal{D}}_{i}),\iota_{i}) be injective envelopes for 𝒟i{\mathcal{D}}_{i} and let (I(𝒟1𝒟2),ι)(I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}),\iota) be an injective envelope for 𝒟1𝒟2{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}. By Proposition 6.3,

(I(𝒟1)I(𝒟2),𝒟1𝒟2,ι1ι2)(I({\mathcal{D}}_{1})\otimes I({\mathcal{D}}_{2}),{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2},\iota_{1}\otimes\iota_{2})

is an essential inclusion. Thus [PittsZarikianUnPsExC*In, Corollary 3.22] gives a unique *-monomorphism g:I(𝒟1)I(𝒟2)I(𝒟1𝒟2)g:I({\mathcal{D}}_{1})\otimes I({\mathcal{D}}_{2})\rightarrow I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) such that

ι=g(ι1ι2).\iota=g\circ(\iota_{1}\otimes\iota_{2}).

Let Ei:𝒞iI(𝒟i)E_{i}:{\mathcal{C}}_{i}\rightarrow I({\mathcal{D}}_{i}) be the pseudo-expectations for (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) (relative to the envelopes (I(𝒟i),ιi)(I({\mathcal{D}}_{i}),\iota_{i})). By [BrownOzawaC*AlFiDiAp, Theorem 3.5.3], there is a unique unital completely positive map E1E2:𝒞1min𝒞2I(𝒟1)I(D2)E_{1}\otimes E_{2}:{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}\rightarrow I({\mathcal{D}}_{1})\otimes I(D_{2}) such that for every elementary tensor x1x2𝒞1𝒞2x_{1}\otimes x_{2}\in{\mathcal{C}}_{1}\odot{\mathcal{C}}_{2}, (E1E2)(x1x2)=E1(x1)E2(x2)(E_{1}\otimes E_{2})(x_{1}\otimes x_{2})=E_{1}(x_{1})\otimes E_{2}(x_{2}). Let

(6.3.1) E:=g(E1E2).E:=g\circ(E_{1}\otimes E_{2}).

Then E:𝒞1min𝒞2I(𝒟1𝒟2)E:{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}\rightarrow I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) is a pseudo-expectation for (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) relative to (I(𝒟1𝒟2),ι)(I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}),\iota). The following commuting diagram illustrates these maps; (the existence of the map labeled “inclusion” follows from [BrownOzawaC*AlFiDiAp, Proposition 3.6.1]).

𝒞1min𝒞2\textstyle{{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E1E2\scriptstyle{E_{1}\otimes E_{2}}I(𝒟1)I(𝒟2)\textstyle{I({\mathcal{D}}_{1})\otimes I({\mathcal{D}}_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}!g\scriptstyle{\exists!g}𝒟1𝒟2\textstyle{{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι1ι2\scriptstyle{\iota_{1}\otimes\iota_{2}}inclusionι\scriptstyle{\iota}I(𝒟1𝒟2)\textstyle{I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2})}

We are now ready to show (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) has the unique pseudo-expectation property.

Proposition \the\numberby.

For i=1,2i=1,2, let (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) be unital pseudo-Cartan inclusions. The map EE defined in (6.3.1) is the unique pseudo-expectation for the inclusion (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) (relative to (I(𝒟1𝒟2),ι)(I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}),\iota)).

Proof.

Since (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are pseudo-Cartan inclusions, 𝒟ic{\mathcal{D}}_{i}^{c} is abelian and (𝒟ic,𝒟i)({\mathcal{D}}_{i}^{c},{\mathcal{D}}_{i}) has the ideal intersection property, so by Proposition 2.4, (𝒟ic,𝒟i)({\mathcal{D}}_{i}^{c},{\mathcal{D}}_{i}) has the faithful unique pseudo-expectation property. In particular, (I(𝒟i),Ei|𝒟i)(I({\mathcal{D}}_{i}),E_{i}|_{{\mathcal{D}}_{i}}) is an injective envelope for 𝒟ic{\mathcal{D}}_{i}^{c} (see [PittsZarikianUnPsExC*In, Corollary 3.22]). Thus, EiE_{i} is also a pseudo-expectation for (𝒞i,𝒟ic)({\mathcal{C}}_{i},{\mathcal{D}}_{i}^{c}). Observation 4.1 shows (𝒞i,𝒟ic)({\mathcal{C}}_{i},{\mathcal{D}}_{i}^{c}) is a virtual Cartan inclusion, and hence (𝒞i,𝒟ic)({\mathcal{C}}_{i},{\mathcal{D}}_{i}^{c}) has the unique pseudo-expectation property. Therefore, EiE_{i} is the unique pseudo-expectation for (𝒞i,𝒟ic)({\mathcal{C}}_{i},{\mathcal{D}}_{i}^{c}).

Now let Δ:𝒞1min𝒞2I(𝒟1𝒟2)\Delta:{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}\rightarrow I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) be a pseudo-expectation. Proposition 6.3 shows (𝒟1c𝒟2c,𝒟1𝒟2)({\mathcal{D}}_{1}^{c}\otimes{\mathcal{D}}_{2}^{c},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) has the ideal intersection property, so another application of [PittsZarikianUnPsExC*In, Corollary 3.22] gives

(6.3.2) Δ|𝒟1c𝒟2c=E|𝒟1c𝒟2c.\Delta|_{{\mathcal{D}}_{1}^{c}\otimes{\mathcal{D}}_{2}^{c}}=E|_{{\mathcal{D}}_{1}^{c}\otimes{\mathcal{D}}_{2}^{c}}.

Clearly 𝒟1c𝒟2c(𝒟1𝒟2)c{\mathcal{D}}_{1}^{c}\otimes{\mathcal{D}}_{2}^{c}\subseteq({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2})^{c}. If equality held, an application of [PittsStReInII, Proposition 6.11] would complete the proof. However, such formulae for tensor products of relative commutants do not hold in general, see [ArchboldCoCoTePrC*Al]. Since we do not know whether 𝒟1c𝒟2c=(𝒟1𝒟2)c{\mathcal{D}}_{1}^{c}\otimes{\mathcal{D}}_{2}^{c}=({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2})^{c}, we adapt the arguments found in the proof of  [PittsStReInII, Proposition 6.11] and  [PittsStReInI, Proposition 3.4] to show Δ=E\Delta=E.

We claim that for any vi𝒩(𝒞i,𝒟i)v_{i}\in{\mathcal{N}}({\mathcal{C}}_{i},{\mathcal{D}}_{i}) (i=1,2i=1,2),

(6.3.3) Δ(v1v2)=E(v1v2).\Delta(v_{1}\otimes v_{2})=E(v_{1}\otimes v_{2}).

Let

J:={d𝒟1𝒟2:(Δ(v1v2)E(v1v2))ι(d)=0},J:=\{d\in{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}:(\Delta(v_{1}\otimes v_{2})-E(v_{1}\otimes v_{2}))\iota(d)=0\},

and let {Kj(vi)}j=04\{K_{j}(v_{i})\}_{j=0}^{4} be a right Frolík family of ideals for viv_{i} (i=1,2i=1,2), see [PittsStReInI, Definition 2.13]. As noted there, for j=1,2,3j=1,2,3 (and i=1,2)i=1,2)),

(6.3.4) θvi(Kj(vi))Kj(vi)=0.\theta_{v_{i}}(K_{j}(v_{i}))K_{j}(v_{i})=0.

Since Ki:=j=04Kj(vi)K_{i}:=\bigvee_{j=0}^{4}K_{j}(v_{i}) is an essential ideal in 𝒟i{\mathcal{D}}_{i} (see [PittsStReInI, Definition 2.13]), it follows that K:=K1K2K:=K_{1}\otimes K_{2} is an essential ideal of 𝒟1𝒟2{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}. Our goal is to show that JJ is an essential ideal of 𝒟1𝒟2{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}, which we do by showing KJK\subseteq J.

Let s,t{0,1,2,3,4}s,t\in\{0,1,2,3,4\}. We shall show that

(6.3.5) Ks(v1)Kt(v2)J.K_{s}(v_{1})\otimes K_{t}(v_{2})\subseteq J.

Starting with the case s=t=0s=t=0, let hiK0(vi)h_{i}\in K_{0}(v_{i}). By [PittsStReInI, Lemma 2.15],

(h1h2)(v1v2)=(v1v2)(h1h2)𝒟1c𝒟2c.(h_{1}\otimes h_{2})(v_{1}\otimes v_{2})=(v_{1}\otimes v_{2})(h_{1}\otimes h_{2})\in{\mathcal{D}}_{1}^{c}\otimes{\mathcal{D}}_{2}^{c}.

Then

E(v1v2)ι(h1h2)\displaystyle E(v_{1}\otimes v_{2})\iota(h_{1}\otimes h_{2}) =E((v1v2)(h1h2))\displaystyle=E((v_{1}\otimes v_{2})(h_{1}\otimes h_{2}))
=(6.3.2)Δ((v1v2)(h1h2))=Δ(v1v2)ι(h1h2),\displaystyle\stackrel{{\scriptstyle\eqref{upse0}}}{{=}}\Delta((v_{1}\otimes v_{2})(h_{1}\otimes h_{2}))=\Delta(v_{1}\otimes v_{2})\iota(h_{1}\otimes h_{2}),

where the first and third equalities follow from [PaulsenCoBoMaOpAl, Corollary 3.19]. Thus h1h2Jh_{1}\otimes h_{2}\in J, so (6.3.5) holds when s=t=0s=t=0.

The remaining cases are similar to those found in the proof of [PittsStReInI, Proposition 3.4]. For example, if h1K1(v1)h_{1}\in K_{1}(v_{1}) and h2Kt(v2)h_{2}\in K_{t}(v_{2}), where 0t40\leq t\leq 4, let xK1(v1)x\in K_{1}(v_{1}). Then using [PittsStReInI, Lemma 2.1],

E(v1v2)ι(h1xh2))\displaystyle E(v_{1}\otimes v_{2})\iota(h_{1}x\otimes h_{2})) =E((v1h1x)(v2h2))=E(θv1(h1)v1xv2h2)\displaystyle=E((v_{1}h_{1}x)\otimes(v_{2}h_{2}))=E(\theta_{v_{1}^{*}}(h_{1})v_{1}x\otimes v_{2}h_{2})
=ι(θv1(h1)I)E(v1v2)ι(xh2)\displaystyle=\iota(\theta_{v_{1}^{*}}(h_{1})\otimes I)E(v_{1}\otimes v_{2})\iota(x\otimes h_{2})
=ι((θv1(h1)I))(xh2))E(v1v2)\displaystyle=\iota((\theta_{v_{1}^{*}}(h_{1})\otimes I))(x\otimes h_{2}))E(v_{1}\otimes v_{2})
=0,\displaystyle=0,

because θv1(h1)x=0\theta_{v_{1}^{*}}(h_{1})\,x=0 by (6.3.4). Taking xx from an approximate unit for K1(v1)K_{1}(v_{1}), we obtain

E(v1v2)ι(h1h2)=0E(v_{1}\otimes v_{2})\iota(h_{1}\otimes h_{2})=0

Similarly, Δ(v1v2)ι(h1h2)=0\Delta(v_{1}\otimes v_{2})\iota(h_{1}\otimes h_{2})=0, so that h1h2Jh_{1}\otimes h_{2}\in J. Thus K1(v1)Kt(v2)JK_{1}(v_{1})\otimes K_{t}(v_{2})\subseteq J. As the other combinations of ss and tt are obtained in the same way, we obtain (6.3.5).

We now have KJK\subseteq J, so JJ is an essential ideal of 𝒟1𝒟2{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}. An application of [PittsStReInI, Lemma 3.3] yields E=ΔE=\Delta. ∎

Our next goal is to show EE is faithful. This requires some preparation, and we first deal with some generalities involving unital regular inclusions having the unique pseudo-expectation property.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular inclusion having the unique (but not necessarily faithful) pseudo-expectation property and let Φ\Phi be the pseudo-expectation. We will use notation and terminology from [PittsStReInII] as reprised in Section 5.1(a)–(i).

Lemma \the\numberby.

Assume (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a unital regular inclusion having the unique pseudo-expectation Φ\Phi. The following statements hold.

  1. (a)

    (𝒞,𝒟)={x𝒞:ϕ(x)=0 for all ϕΣ(𝒞,𝒟)}{\mathcal{L}}({\mathcal{C}},{\mathcal{D}})=\{x\in{\mathcal{C}}:\phi(x)=0\text{ for all }\phi\in\Sigma({\mathcal{C}},{\mathcal{D}})\}.

  2. (b)

    Suppose in addition that Φ\Phi is faithful. Then spanΣ(𝒞,𝒟)\operatorname{span}\Sigma({\mathcal{C}},{\mathcal{D}}) is weak-* dense in the dual space, 𝒞#{\mathcal{C}}^{\#}, of 𝒞{\mathcal{C}}.

Proof.

(a) Suppose x𝒞x\in{\mathcal{C}} and Φ(xx)=0\Phi(x^{*}x)=0. Given ϕΣ(𝒞,𝒟)\phi\in\Sigma({\mathcal{C}},{\mathcal{D}}) we may find v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) and ρ𝔖s(𝒞,𝒟)\rho\in{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) so that for every y𝒞y\in{\mathcal{C}},

ϕ(y)=[v,ρ](y)=ρ(vy)ρ(vv)1/2.\phi(y)=[v,\rho](y)=\frac{\rho(v^{*}y)}{\rho(v^{*}v)^{1/2}}.

As ρ=σΦ\rho=\sigma\circ\Phi for some σI(𝒟)^\sigma\in\widehat{I({\mathcal{D}})}, the Cauchy-Schwartz inequality shows |ϕ(x)|2ρ(xx)=σ(Φ(xx))=0|\phi(x)|^{2}\leq\rho(x^{*}x)=\sigma(\Phi(x^{*}x))=0.

On the other hand, suppose x𝒞x\in{\mathcal{C}} satisfies ϕ(x)=0\phi(x)=0 for every ϕΣ(𝒞,𝒟)\phi\in\Sigma({\mathcal{C}},{\mathcal{D}}). Fix ρ𝔖s(𝒞,𝒟)\rho\in{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) and let v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). If ρ(vv)0\rho(v^{*}v)\neq 0, then [v,ρ](x)=0[v,\rho](x)=0, so ρ(vx)=0\rho(v^{*}x)=0. On the other hand, if ρ(vv)=0\rho(v^{*}v)=0, the Cauchy-Schwartz inequality gives ρ(vx)=0\rho(v^{*}x)=0. Thus, for every v𝒩(C,𝒟)v\in{\mathcal{N}}(C,{\mathcal{D}}),

ρ(vx)=0.\rho(v^{*}x)=0.

By regularity of (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), we find ρ(xx)=0\rho(x^{*}x)=0.

Allowing ρ\rho to vary throughout 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}), we conclude that σ(Φ(xx))=0\sigma(\Phi(x^{*}x))=0 for every σI(𝒟)^\sigma\in\widehat{I({\mathcal{D}})}, that is, Φ(xx)=0\Phi(x^{*}x)=0.

(b) Let 𝒮:=spanΣ(𝒞,𝒟){\mathcal{S}}:=\operatorname{span}\Sigma({\mathcal{C}},{\mathcal{D}}). Suppose u:𝒞#u:{\mathcal{C}}^{\#}\rightarrow{\mathbb{C}} is a weak *-continuous linear functional which annihilates 𝒮{\mathcal{S}}. As the weak *-continuous linear functionals on 𝒞#{\mathcal{C}}^{\#} may be identified with 𝒞{\mathcal{C}}, there exists x𝒞x\in{\mathcal{C}} such that u(f)=f(x)u(f)=f(x) for every f𝒞#f\in{\mathcal{C}}^{\#}. Since ϕ(x)=u(ϕ)=0\phi(x)=u(\phi)=0 for every ϕΣ(𝒞,𝒟)\phi\in\Sigma({\mathcal{C}},{\mathcal{D}}), part (a) shows x(𝒞,𝒟)x\in{\mathcal{L}}({\mathcal{C}},{\mathcal{D}}). Since Φ\Phi is faithful by assumption, x=0x=0. This gives u=0u=0, whence 𝒮{\mathcal{S}} is weak-* dense in 𝒞#{\mathcal{C}}^{\#}. ∎

Now that these preliminaries have been completed, we return to the context of Proposition 6.3 and the task of showing EE is faithful. Since the map gg appearing in (6.3.1) is a *-monomorphism, we may use gg to identify I(𝒟1)I(𝒟2)I({\mathcal{D}}_{1})\otimes I({\mathcal{D}}_{2}) with a subalgebra of I(𝒟1𝒟2)I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}); upon doing so, we have

E=E1E2.E=E_{1}\otimes E_{2}.

This enables us to describe Σ(𝒞1min𝒞2,𝒟1𝒟2)\Sigma({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}). The notation in the next lemma is discussed in Section 5.

Lemma \the\numberby.

For unital pseudo-Cartan inclusions (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}), the following statements hold.

  1. (a)

    𝔖s(𝒞1min𝒞2,𝒟1𝒟2)={ρ1ρ2:ρi𝔖s(𝒞i,𝒟i)}{\mathfrak{S}}_{s}({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2})=\{\rho_{1}\otimes\rho_{2}:\rho_{i}\in{\mathfrak{S}}_{s}({\mathcal{C}}_{i},{\mathcal{D}}_{i})\}; and

  2. (b)

    Σ(𝒞1min𝒞2,𝒟1𝒟2)={ϕ1ϕ2:ϕiΣ(𝒞i,𝒟i)}\Sigma({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2})=\{\phi_{1}\otimes\phi_{2}:\phi_{i}\in\Sigma({\mathcal{C}}_{i},{\mathcal{D}}_{i})\}.

Proof.

(a) For i=1,2i=1,2, let ρi𝔖s(𝒞i,𝒟i)\rho_{i}\in{\mathfrak{S}}_{s}({\mathcal{C}}_{i},{\mathcal{D}}_{i}) and find σiI(𝒟i)^\sigma_{i}\in\widehat{I({\mathcal{D}}_{i})} so that ρi=σiEi\rho_{i}=\sigma_{i}\circ E_{i}. As E=E1E2E=E_{1}\otimes E_{2} and σ1σ2\sigma_{1}\otimes\sigma_{2} extends from a multiplicative linear functional on I(𝒟1)I(𝒟2)I({\mathcal{D}}_{1})\otimes I({\mathcal{D}}_{2}) to a multiplicative linear functional σ\sigma on I(𝒟1𝒟2)I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}). we see that

ρ1ρ2=(σ1E1)(σ2E2)=(σ1σ2)(E1E2)=σE𝔖s(𝒞1min𝒞2,𝒟1𝒟2).\rho_{1}\otimes\rho_{2}=(\sigma_{1}\circ E_{1})\otimes(\sigma_{2}\circ E_{2})=(\sigma_{1}\otimes\sigma_{2})\circ(E_{1}\otimes E_{2})=\sigma\circ E\in{\mathfrak{S}}_{s}({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}).

On the other hand, if ρ𝔖s(𝒞1min𝒞2,𝒟1𝒟2)\rho\in{\mathfrak{S}}_{s}({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}), there exists a multiplicative linear functional σ\sigma on I(𝒟1𝒟2)I({\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) so that ρ=σE=σ(E1E2)\rho=\sigma\circ E=\sigma\circ(E_{1}\otimes E_{2}). Since the maximal ideal space of I(𝒟1)I(𝒟2)I({\mathcal{D}}_{1})\otimes I({\mathcal{D}}_{2}) is homeomorphic to I(𝒟1)^×I(𝒟2)^\widehat{I({\mathcal{D}}_{1})}\times\widehat{I({\mathcal{D}}_{2})}, the restriction of σ\sigma to I(𝒟1)I(𝒟2)I({\mathcal{D}}_{1})\otimes I({\mathcal{D}}_{2}) has the form σ1σ2\sigma_{1}\otimes\sigma_{2} where σiI(𝒟i)^\sigma_{i}\in\widehat{I({\mathcal{D}}_{i})}. Therefore ρ=(σ1E1)(σ2E2)\rho=(\sigma_{1}\circ E_{1})\otimes(\sigma_{2}\circ E_{2}), as desired.

(b) Let ϕiΣ(𝒞i,𝒟i)\phi_{i}\in\Sigma({\mathcal{C}}_{i},{\mathcal{D}}_{i}). Then there are vi𝒩(𝒞i,𝒟i)v_{i}\in{\mathcal{N}}({\mathcal{C}}_{i},{\mathcal{D}}_{i}) and ρi𝔖s(𝒞i,𝒟i)\rho_{i}\in{\mathfrak{S}}_{s}({\mathcal{C}}_{i},{\mathcal{D}}_{i}) so that ϕi=[vi,ρi]\phi_{i}=[v_{i},\rho_{i}]. Since v1v2𝒩(𝒞1min𝒞2,𝒟1𝒟2)v_{1}\otimes v_{2}\in{\mathcal{N}}({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}), we get ϕ1ϕ2=[v1v2,ρ1ρ2]Σ(𝒞1min𝒞2,𝒟1𝒟2)\phi_{1}\otimes\phi_{2}=[v_{1}\otimes v_{2},\rho_{1}\otimes\rho_{2}]\in\Sigma({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}).

For the reverse inclusion, suppose ϕΣ(𝒞1min𝒞2,𝒟1𝒟2)\phi\in\Sigma({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}), and let ρ=𝔰(ϕ)\rho={\mathfrak{s}}(\phi). By part (a), ρ=ρ1ρ2\rho=\rho_{1}\otimes\rho_{2}, where ρi𝔖s(𝒞i,𝒟i)\rho_{i}\in{\mathfrak{S}}_{s}({\mathcal{C}}_{i},{\mathcal{D}}_{i}). Since span{v1v2:vi𝒩(𝒞i,𝒟i)}\operatorname{span}\{v_{1}\otimes v_{2}:v_{i}\in{\mathcal{N}}({\mathcal{C}}_{i},{\mathcal{D}}_{i})\} is dense in 𝒞1min𝒞2{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}, we may choose vi𝒩(𝒞i,𝒟i)v_{i}\in{\mathcal{N}}({\mathcal{C}}_{i},{\mathcal{D}}_{i}) such that ϕ(v1v2)>0\phi(v_{1}\otimes v_{2})>0. Then

ϕ=[v1v2,ρ]=[v1v2,ρ1ρ2]=[v1,ρ1][v2,ρ2],\phi=[v_{1}\otimes v_{2},\rho]=[v_{1}\otimes v_{2},\rho_{1}\otimes\rho_{2}]=[v_{1},\rho_{1}]\otimes[v_{2},\rho_{2}],

with the first equality following from [PittsStReInII, Theorem 7.9(f)]. ∎

Proposition \the\numberby.

For i=1,2i=1,2, let (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) be unital pseudo-Cartan inclusions. The following statements hold.

  1. (a)

    (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) is a pseudo-Cartan inclusion.

  2. (b)

    If (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are Cartan inclusions, so is (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}).

Before giving the proof, we remark that part (b) is probably known, but we do not have a reference.

Proof.

(a) We already know (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) is a regular inclusion having the unique pseudo-expectation property. It remains to show the pseudo-expectation EE (see (6.3.1)) is faithful.

For typographical reasons, let us write :={z𝒞1min𝒞2:E(zz)=0}{\mathcal{L}}:=\{z\in{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}:E(z^{*}z)=0\} for the left kernel of EE. Recall that {\mathcal{L}} is an ideal in 𝒞1min𝒞2{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2} by [PittsStReInII, Theorem 6.5].

We claim

(6.3.6) (𝒞1𝒞2)={0}.{\mathcal{L}}\cap({\mathcal{C}}_{1}\odot{\mathcal{C}}_{2})=\{0\}.

Suppose z(𝒞1𝒞2)z\in{\mathcal{L}}\cap({\mathcal{C}}_{1}\odot{\mathcal{C}}_{2}). Write

z=j=1nxjyj,z=\sum_{j=1}^{n}x_{j}\otimes y_{j},

where xj𝒞1x_{j}\in{\mathcal{C}}_{1}, yj𝒞2y_{j}\in{\mathcal{C}}_{2} and {y1,,yn}\{y_{1},\dots,y_{n}\} is a linearly independent set. For ϕ1ϕ2Σ(𝒞1min𝒞2,𝒟1𝒟2)\phi_{1}\otimes\phi_{2}\in\Sigma({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}),

0=(ϕ1ϕ2)(z)=j=1nϕ1(xj)ϕ2(yj),0=(\phi_{1}\otimes\phi_{2})(z)=\sum_{j=1}^{n}\phi_{1}(x_{j})\phi_{2}(y_{j}),

with Lemma 6.3(a) giving the first equality. Holding ϕ1\phi_{1} fixed, this equality persists if ϕ2\phi_{2} is replaced by ψspanΣ(𝒞2,𝒟2)\psi\in\operatorname{span}\Sigma({\mathcal{C}}_{2},{\mathcal{D}}_{2}), and hence also for any ψ𝒞2#\psi\in{\mathcal{C}}_{2}^{\#} by Lemma 6.3(b) applied to (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) and E2E_{2}. Therefore, setting

s:=j=1nϕ1(xj)yj,s:=\sum_{j=1}^{n}\phi_{1}(x_{j})y_{j},

we find ψ(s)=0\psi(s)=0 for every ψ𝒞2#\psi\in{\mathcal{C}}_{2}^{\#}, whence s=0s=0. By the linear independence of {y1,,yn}\{y_{1},\dots,y_{n}\}, we conclude that ϕ1(xj)=0\phi_{1}(x_{j})=0 for each jj. Varying ϕ1\phi_{1} and applying Lemma 6.3(b) to (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and E1E_{1}, we see that xj=0x_{j}=0 for every jj, so z=0z=0. Thus (6.3.6) holds.

For z𝒞1𝒞2z\in{\mathcal{C}}_{1}\odot{\mathcal{C}}_{2}, define

η(z):=z+(𝒞1min𝒞2)/.\eta(z):=\left\lVert z+{\mathcal{L}}\right\rVert_{({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2})/{\mathcal{L}}}.

By the claim, η\eta is a CC^{*}-norm on 𝒞1𝒞2{\mathcal{C}}_{1}\odot{\mathcal{C}}_{2}. But 𝒞1min𝒞2\left\lVert\cdot\right\rVert_{{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}} is the smallest CC^{*}-norm on 𝒞1𝒞2{\mathcal{C}}_{1}\odot{\mathcal{C}}_{2}, so for any z𝒞1C2z\in{\mathcal{C}}_{1}\odot C_{2},

η(z)z𝒞1min𝒞2η(z).\eta(z)\leq\left\lVert z\right\rVert_{{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}}\leq\eta(z).

Thus for x𝒞1min𝒞2x\in{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},

x+(𝒞1min𝒞2)/=x𝒞1min𝒞2.\left\lVert x+{\mathcal{L}}\right\rVert_{({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2})/{\mathcal{L}}}=\left\lVert x\right\rVert_{{\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}}.

It follows that ={0}{\mathcal{L}}=\{0\}, completing the proof of (a).

(b) When (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are Cartan inclusions, ιi1Ei\iota_{i}^{-1}\circ E_{i} and (ι1ι2)1(E1E2)(\iota_{1}\otimes\iota_{2})^{-1}\circ(E_{1}\otimes E_{2}) are faithful conditional expectations. Thus part (b) follows from part (a) and the characterization of Cartan inclusions given following Observation 4.1. ∎

Proposition \the\numberby.

Suppose (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) are unital pseudo-Cartan inclusions with Cartan envelopes (𝒜1,1 :: α1)({\mathcal{A}}_{1},{\mathcal{B}}_{1}\hbox{\,:\hskip-1.0pt:\,}\alpha_{1}) and (𝒜2,2 :: α2)({\mathcal{A}}_{2},{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha_{2}) respectively. Then (𝒜1min𝒜2,12 :: α1α2)({\mathcal{A}}_{1}\otimes_{\min}{\mathcal{A}}_{2},{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha_{1}\otimes\alpha_{2}) is a Cartan envelope for (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}).

Proof.

Proposition 6.3(b) shows that (𝒜1min𝒜2,12)({\mathcal{A}}_{1}\otimes_{\min}{\mathcal{A}}_{2},{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2}) is a Cartan inclusion. Let Δi:𝒜ii\Delta_{i}:{\mathcal{A}}_{i}\rightarrow{\mathcal{B}}_{i} be the conditional expectations and put

Δ:=Δ1Δ2andα:=α1α2.\Delta:=\Delta_{1}\otimes\Delta_{2}\quad\text{and}\quad\alpha:=\alpha_{1}\otimes\alpha_{2}.

Then Δ:𝒜1min𝒜212\Delta:{\mathcal{A}}_{1}\otimes_{\min}{\mathcal{A}}_{2}\rightarrow{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2} is the (faithful) conditional expectation. Also, since (𝒜1,1,α)({\mathcal{A}}_{1},{\mathcal{B}}_{1},\alpha) is a Cartan envelope for (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}),

1I=C(Δ(α(𝒞1I)))and𝒜1I=C((α(𝒞1I)(1I)).{\mathcal{B}}_{1}\otimes I=C^{*}(\Delta(\alpha({\mathcal{C}}_{1}\otimes I)))\quad\text{and}\quad{\mathcal{A}}_{1}\otimes I=C^{*}((\alpha({\mathcal{C}}_{1}\otimes I)\cup({\mathcal{B}}_{1}\otimes I)).

Likewise, I𝒜2I\otimes{\mathcal{A}}_{2} is generated by α(I𝒞2)Δ(α(I𝒞2))\alpha(I\otimes{\mathcal{C}}_{2})\cup\Delta(\alpha(I\otimes{\mathcal{C}}_{2})). It follows that (𝒜1min𝒜2,12 :: α1α2)({\mathcal{A}}_{1}\otimes_{\min}{\mathcal{A}}_{2},{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha_{1}\otimes\alpha_{2}) is a Cartan package for (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}). Proposition 6.3 shows that (12,𝒟1𝒟2,(α1α2)|𝒟1𝒟2)({\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2},(\alpha_{1}\otimes\alpha_{2})|_{{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}}) is an essential inclusion. Therefore (𝒜1min𝒜2,12 :: α1α2)({\mathcal{A}}_{1}\otimes_{\min}{\mathcal{A}}_{2},{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha_{1}\otimes\alpha_{2}) is a Cartan envelope for (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}). ∎

Having completed the unital case, we are ready for the main theorem of this section.

Theorem \the\numberby.

Suppose for i=1,2i=1,2 that (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are pseudo-Cartan inclusions (not assumed unital) and let (𝒜1,i :: αi)({\mathcal{A}}_{1},{\mathcal{B}}_{i}\hbox{\,:\hskip-1.0pt:\,}\alpha_{i}) be Cartan envelopes for (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}). Then (𝒞1min𝒟2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{D}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) is a pseudo-Cartan inclusion, and (𝒜1min𝒜2,12 :: α1α2)({\mathcal{A}}_{1}\otimes_{\min}{\mathcal{A}}_{2},{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha_{1}\otimes\alpha_{2}) is a Cartan envelope for (𝒞1min𝒟2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{D}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}).

Proof.

We begin by assuming that (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) have the AUP. Observation 4.2 shows (𝒞~i,𝒟~i)(\tilde{\mathcal{C}}_{i},\tilde{\mathcal{D}}_{i}) is a unital pseudo-Cartan inclusion, and Proposition 6.3 shows that (𝒞~1min𝒞~2,𝒟~1𝒟~2)(\tilde{\mathcal{C}}_{1}\otimes_{\min}\tilde{\mathcal{C}}_{2},\tilde{\mathcal{D}}_{1}\otimes\tilde{\mathcal{D}}_{2}) is a unital pseudo-Cartan inclusion.

Since (𝒟~1,𝒟1)(\tilde{\mathcal{D}}_{1},{\mathcal{D}}_{1}) and (𝒟~2,𝒟2)(\tilde{\mathcal{D}}_{2},{\mathcal{D}}_{2}) have the ideal intersection property, Proposition 6.3 shows (𝒟~1𝒟~2,𝒟1𝒟2)(\tilde{\mathcal{D}}_{1}\otimes\tilde{\mathcal{D}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) has the ideal intersection property. Let

:={v1v2:vi𝒩(𝒞i,𝒟i)}𝒩(𝒞~1min𝒞~2,𝒟1𝒟2).{\mathcal{M}}:=\{v_{1}\otimes v_{2}:v_{i}\in{\mathcal{N}}({\mathcal{C}}_{i},{\mathcal{D}}_{i})\}\subseteq{\mathcal{N}}(\tilde{\mathcal{C}}_{1}\otimes_{\min}\tilde{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}).

Then 𝒟1𝒟2span¯()=𝒞1min𝒞2{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}\subseteq\overline{\operatorname{span}}({\mathcal{M}})={\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}. Applying Proposition 4.4(a) to (𝒞~1min𝒞~2,𝒟~1𝒟~2)(\tilde{\mathcal{C}}_{1}\otimes_{\min}\tilde{\mathcal{C}}_{2},\tilde{\mathcal{D}}_{1}\otimes\tilde{\mathcal{D}}_{2}) with this choice for {\mathcal{M}}, we conclude (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) is a pseudo-Cartan inclusion.

Next, Proposition 6.3 and Theorem 4.2(a) show (𝒜~1min𝒜~2,~1~2 :: α~1α~2)(\tilde{\mathcal{A}}_{1}\otimes_{\min}\tilde{\mathcal{A}}_{2},\tilde{\mathcal{B}}_{1}\otimes\tilde{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\tilde{\alpha}_{1}\otimes\tilde{\alpha}_{2}) is a Cartan envelope for (𝒞~1min𝒞~2,𝒟~1𝒟~2)(\tilde{\mathcal{C}}_{1}\otimes_{\min}\tilde{\mathcal{C}}_{2},\tilde{\mathcal{D}}_{1}\otimes\tilde{\mathcal{D}}_{2}). Let Δi:𝒜ii\Delta_{i}:{\mathcal{A}}_{i}\rightarrow{\mathcal{B}}_{i} be the conditional expectations (for the Cartan inclusions (𝒜i,i)({\mathcal{A}}_{i},{\mathcal{B}}_{i})). Write Δ:=Δ1Δ2\Delta:=\Delta_{1}\otimes\Delta_{2} and α:=α1α2\alpha:=\alpha_{1}\otimes\alpha_{2}. Notice that

12=C(Δ~(α~(𝒞1min𝒞2)))and𝒜1min𝒜2=C((12)α~(𝒞1min𝒞2)){\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2}=C^{*}(\tilde{\Delta}(\tilde{\alpha}({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2})))\quad\text{and}\quad{\mathcal{A}}_{1}\otimes_{\min}{\mathcal{A}}_{2}=C^{*}(({\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2})\cup\tilde{\alpha}({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2}))

That (𝒜1min𝒜2,12 :: α1α2)({\mathcal{A}}_{1}\otimes_{\min}{\mathcal{A}}_{2},{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha_{1}\otimes\alpha_{2}) is a Cartan envelope for (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) now follows from Proposition 4.4(b).

Now suppose (𝒞1,𝒟1)({\mathcal{C}}_{1},{\mathcal{D}}_{1}) and (𝒞2,𝒟2)({\mathcal{C}}_{2},{\mathcal{D}}_{2}) are pseudo-Cartan inclusions, where neither is assumed to have the AUP. Then (𝒞i,𝒟ic)({\mathcal{C}}_{i},{\mathcal{D}}_{i}^{c}) are pseudo-Cartan inclusions with the AUP, whence (𝒞1min𝒞2,𝒟1c𝒟2c)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}^{c}\otimes{\mathcal{D}}_{2}^{c}) is a pseudo-Cartan inclusion. Proposition 6.3 shows (𝒟1c𝒟2c,𝒟1𝒟2)({\mathcal{D}}_{1}^{c}\otimes{\mathcal{D}}_{2}^{c},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) has the ideal intersection property, so an application of Proposition 4.4(a) (with ={v1v2:vi𝒩(𝒞i,𝒟i)}{\mathcal{M}}=\{v_{1}\otimes v_{2}:v_{i}\in{\mathcal{N}}({\mathcal{C}}_{i},{\mathcal{D}}_{i})\}) shows (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}) is a pseudo-Cartan inclusion. Let (𝒜i,i :: αi)({\mathcal{A}}_{i},{\mathcal{B}}_{i}\hbox{\,:\hskip-1.0pt:\,}\alpha_{i}) be a Cartan envelope for (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}). By Proposition 4.2, (𝒜i,i :: αi)({\mathcal{A}}_{i},{\mathcal{B}}_{i}\hbox{\,:\hskip-1.0pt:\,}\alpha_{i}) is a Cartan envelope for (𝒞i,𝒟ic)({\mathcal{C}}_{i},{\mathcal{D}}_{i}^{c}), whence (𝒜1min𝒜2,12 :: α1α2)({\mathcal{A}}_{1}\otimes_{\min}{\mathcal{A}}_{2},{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha_{1}\otimes\alpha_{2}) is a Cartan envelope for (𝒞1min𝒞2,𝒟1c𝒟2c)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}^{c}\otimes{\mathcal{D}}_{2}^{c}). Finally, Proposition 4.4(b) shows (𝒜1min𝒜2,12 :: α1α2)({\mathcal{A}}_{1}\otimes_{\min}{\mathcal{A}}_{2},{\mathcal{B}}_{1}\otimes{\mathcal{B}}_{2}\hbox{\,:\hskip-1.0pt:\,}\alpha_{1}\otimes\alpha_{2}) is also a Cartan envelope for (𝒞1min𝒞2,𝒟1𝒟2)({\mathcal{C}}_{1}\otimes_{\min}{\mathcal{C}}_{2},{\mathcal{D}}_{1}\otimes{\mathcal{D}}_{2}). ∎

7. Applications

In this section we give a few applications of the results presented so far. We begin with an application of Theorem 3 which describes the CC^{*}-envelope for an intermediate Banach algebras. Several of the results in this section apply to unital inclusions, see Definition 2.1(d).

Theorem \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a unital pseudo-Cartan inclusion and 𝒜𝒞{\mathcal{A}}\subseteq{\mathcal{C}} is a closed subalgebra satisfying 𝒟𝒜𝒞{\mathcal{D}}\subseteq{\mathcal{A}}\subseteq{\mathcal{C}} (we do not assume 𝒜=𝒜){\mathcal{A}}={\mathcal{A}}^{*}). Let C(𝒜)C^{*}({\mathcal{A}}) be the CC^{*}-subalgebra of 𝒞{\mathcal{C}} generated by 𝒜{\mathcal{A}} and let Ce(𝒜)C^{*}_{e}({\mathcal{A}}) be the CC^{*}-envelope for 𝒜{\mathcal{A}}. Then

C(𝒜)=Ce(𝒜).C^{*}({\mathcal{A}})=C^{*}_{e}({\mathcal{A}}).
Proof.

Since (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) has the faithful unique pseudo-expectation property, the result follows from [PittsStReInI, Theorem 8.3]. ∎

For unital CC^{*}-algebras 𝒜{\mathcal{A}} and {\mathcal{B}} with 𝒜{\mathcal{B}}\subseteq{\mathcal{A}}, when the assumption that {\mathcal{B}} is abelian and regular is dropped,  [PittsZarikianUnPsExC*In, Example 6.9] shows that the faithful unique pseudo-expectation property for 𝒜{\mathcal{B}}\subseteq{\mathcal{A}} need not imply that {\mathcal{B}} norms 𝒜{\mathcal{A}} in the sense of [PopSinclairSmithNoC*Al]. Nevertheless, [PittsZarikianUnPsExC*In, Section 6.2] argues that the faithful unique pseudo-expectation property is “conducive” to norming. Further evidence for this statement is Theorem 7 below: for unital regular inclusions satisfying Standing Assumption 3, the faithful unique pseudo-expectation property implies norming.

The proof of Theorem 7 requires two preparatory results which we now present.

Lemma \the\numberby.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a unital pseudo-Cartan inclusion. Suppose σ𝒟^\sigma\in\hat{\mathcal{D}} is free and let ff be the unique state extension of σ\sigma to 𝒞{\mathcal{C}}. If (πf,f,ξf)(\pi_{f},{\mathcal{H}}_{f},\xi_{f}) is the GNS triple associated to ff, then πf(𝒟)′′\pi_{f}({\mathcal{D}})^{\prime\prime} is an atomic MASA in (f){\mathcal{B}}({\mathcal{H}}_{f}).

Furthermore, if

𝒪:={τ𝒟^:v𝒩(𝒞,𝒟) such that τ=βv(σ)}{\mathcal{O}}:=\{\tau\in\hat{\mathcal{D}}:\exists\,v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}})\text{ such that }\tau=\beta_{v}(\sigma)\}

is the orbit of σ\sigma under 𝒩(𝒞,𝒟){\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), then

(7.1) ker(πf|𝒟)=τ𝒪kerτ.\ker(\pi_{f}|_{\mathcal{D}})=\bigcap_{\tau\in{\mathcal{O}}}\ker\tau.
Proof.

Our first task is to show that ff is a compatible state on 𝒞{\mathcal{C}} ([PittsStReInI, Definition 4.1]). Let (I(𝒟),ι)(I({\mathcal{D}}),\iota) be an injective envelope for 𝒟{\mathcal{D}} and let E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be the pseudo-expectation for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Choose σI(𝒟)^\sigma^{\prime}\in\widehat{I({\mathcal{D}})} such that σι=σ\sigma^{\prime}\circ\iota=\sigma. Note that σE|𝒟=σι=σ\sigma^{\prime}\circ E|_{\mathcal{D}}=\sigma^{\prime}\circ\iota=\sigma. Since σ\sigma is free, we conclude that f=σEf=\sigma^{\prime}\circ E. By [PittsStReInII, Theorem 6.9] (see Theorem A below) ff is a compatible state.

The Cauchy-Schwartz inequality implies that for any d𝒟d\in{\mathcal{D}} and x𝒞x\in{\mathcal{C}}, f(dx)=σ(d)f(x)=f(xd)f(dx)=\sigma(d)f(x)=f(xd) ([PittsNoApUnInC*Al, Fact 1.2] has the details), so fMod(𝒞,𝒟)f\in\text{Mod}({\mathcal{C}},{\mathcal{D}}) (see [PittsStReInI, Definition 2.4]).

We claim that if v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) satisfies σ(vv)=1\sigma(v^{*}v)=1 and σ(vdv)=σ(d)\sigma(v^{*}dv)=\sigma(d) for every d𝒟d\in{\mathcal{D}}, then

f(vxv)=f(x)for allx𝒞.f(v^{*}xv)=f(x)\quad\text{for all}\quad x\in{\mathcal{C}}.

Define a positive linear functional fvf_{v} on 𝒞{\mathcal{C}} by fv(x)=f(vxv)f_{v}(x)=f(v^{*}xv). Note that 1fv1\leq\left\lVert f_{v}\right\rVert because fv|𝒟=σf_{v}|_{\mathcal{D}}=\sigma. Also for x𝒞x\in{\mathcal{C}} and d𝒟d\in{\mathcal{D}} with σ(d)=1\sigma(d)=1,

|f(vxv)|=|f(dvxvd)|vd2xf=dvvdx.|f(v^{*}xv)|=|f(d^{*}v^{*}xvd)|\leq\left\lVert vd\right\rVert^{2}\left\lVert x\right\rVert\left\lVert f\right\rVert=\left\lVert d^{*}v^{*}vd\right\rVert\left\lVert x\right\rVert.

As inf{dvvd:d𝒟 and σ(d)=1}=σ(vv)=1\inf\{\left\lVert d^{*}v^{*}vd\right\rVert:d\in{\mathcal{D}}\text{ and }\sigma(d)=1\}=\sigma(v^{*}v)=1, we conclude fv=1\left\lVert f_{v}\right\rVert=1. Thus, fvf_{v} is also an extension of σ\sigma to a state on 𝒞{\mathcal{C}}. We conclude that fv=ff_{v}=f, so the claim holds.

It follows that ff is a pure and 𝒟{\mathcal{D}}-rigid state (see [PittsStReInI, Definition 4.11]). By first applying [PittsStReInI, Proposition 4.12] and then [PittsStReInI, Proposition 4.8(iv)], we find πf(𝒟)′′\pi_{f}({\mathcal{D}})^{\prime\prime} is an atomic MASA in (f){\mathcal{B}}({\mathcal{H}}_{f}).

Let Nf={x𝒞:f(xx)=0}N_{f}=\{x\in{\mathcal{C}}:f(x^{*}x)=0\} be the left kernel of ff. Suppose d𝒟d\in{\mathcal{D}} and πf(d)0\pi_{f}(d)\neq 0. Since (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular inclusion, there exists v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) such that πf(d)(v+Nf)0\pi_{f}(d)(v+N_{f})\neq 0. Then f(vv)0f(v^{*}v)\neq 0 because 0f(vddv)d2f(vv)0\neq f(v^{*}d^{*}dv)\leq\left\lVert d\right\rVert^{2}f(v^{*}v). As f|𝒟=σf|_{\mathcal{D}}=\sigma[PittsStReInI, Proposition 4.4(v)], shows βv(σ)(d)0\beta_{v}(\sigma)(d)\neq 0. As βv(σ)𝒪\beta_{v}(\sigma)\in{\mathcal{O}}, we find

τ𝒪kerτker(πf|𝒟).\bigcap_{\tau\in{\mathcal{O}}}\ker\tau\subseteq\ker(\pi_{f}|_{\mathcal{D}}).

On the other hand, if dker(πf|𝒟)d\in\ker(\pi_{f}|_{\mathcal{D}}), then for every v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) with σ(vv)0\sigma(v^{*}v)\neq 0, we have 0=πf(d)(v+N)0=\pi_{f}(d)(v+N), so applying [PittsStReInI, Proposition 4.4(v)] again, βv(σ)(d)=0\beta_{v}(\sigma)(d)=0. Therefore, dτ𝒪kerτd\in\bigcap_{\tau\in{\mathcal{O}}}\ker\tau. Thus (7.1) holds and the proof is complete. ∎

Our second preparatory result, of independent interest, shows that a natural class of pseudo-Cartan inclusions is topologically free.

Proposition \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a unital pseudo-Cartan inclusion such that there exists a countable subset 𝒩(𝒞,𝒟){\mathfrak{C}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) satisfying 𝒞=C(𝒟){\mathcal{C}}=C^{*}({\mathfrak{C}}\cup{\mathcal{D}}). Then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a topologically free inclusion.

Proof.

Let 𝔘:={σ𝒟^:σ is free for (𝒞,𝒟)}{\mathfrak{U}}:=\{\sigma\in\hat{\mathcal{D}}:\sigma\text{ is free for $({\mathcal{C}},{\mathcal{D}})$}\}. Our task is to show 𝔘{\mathfrak{U}} is dense in 𝒟^\hat{\mathcal{D}}. For convenience, set

:=𝒟c{\mathcal{B}}:={\mathcal{D}}^{c}

and let π:^𝒟^\pi:\hat{\mathcal{B}}\rightarrow\hat{\mathcal{D}} be the restriction map,

π(σ1)=σ1|𝒟(σ1^).\pi(\sigma_{1})=\sigma_{1}|_{\mathcal{D}}\qquad(\sigma_{1}\in\hat{\mathcal{B}}).

By [ExelPittsChGrC*AlNoHaEtGr, Theorem 2.11.16], the set of free points for the inclusion (,𝒟)({\mathcal{B}},{\mathcal{D}}) contains a dense GδG_{\delta} set. We may therefore choose dense open sets Un𝒟^U_{n}\subseteq\hat{\mathcal{D}} such that

nUn\bigcap_{n}U_{n}

is contained in the set of free points for (,𝒟)({\mathcal{B}},{\mathcal{D}}). As 𝒟{\mathcal{D}}\subseteq{\mathcal{B}} has the ideal intersection property, π\pi is an essential surjection, so for each nn, π1(Un)\pi^{-1}(U_{n}) is a dense open subset of ^\hat{\mathcal{B}}.

Let 𝒮{\mathcal{S}} be the *-semigroup generated by {I}{\mathfrak{C}}\cup\{I\}, that is, 𝒮{\mathcal{S}} is the collection of all finite products of elements of {I}{\mathfrak{C}}\cup{\mathfrak{C}}^{*}\cup\{I\}. Then 𝒮{\mathcal{S}} is countable, {wd:w𝒮 and d𝒟}𝒩(𝒞,𝒟)\{wd:w\in{\mathcal{S}}\text{ and }d\in{\mathcal{D}}\}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), and the linear span,

𝔖:=span{wd:w𝒮 and d𝒟},{\mathfrak{S}}:=\operatorname{span}\{wd:w\in{\mathcal{S}}\text{ and }d\in{\mathcal{D}}\},

is dense in 𝒞{\mathcal{C}}.

Let v𝒮v\in{\mathcal{S}}. By Lemma 3(a), v𝒩(𝒞,)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{B}}). Moreover, as (𝒞,)({\mathcal{C}},{\mathcal{B}}) is a regular MASA inclusion, the proof of [PittsStReInI, Theorem 3.10] shows there is a dense open set Xv^X_{v}\subseteq\hat{\mathcal{B}} such that each σ1Xv\sigma_{1}\in X_{v} is free relative to vv. Let

𝒴:=(v𝒮Xv)(nπ1(Un)).{\mathcal{Y}}:=\left(\bigcap_{v\in{\mathcal{S}}}X_{v}\right)\cap\left(\bigcap_{n}\pi^{-1}(U_{n})\right).

As {\mathcal{B}} is unital, ^\hat{\mathcal{B}} is compact, so Baire’s theorem implies 𝒴^{\mathcal{Y}}\subseteq\hat{\mathcal{B}} is a dense GδG_{\delta} set.

We claim that if y𝒴y\in{\mathcal{Y}}, then π(y)\pi(y) is a free point for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). To see this, suppose ρ1\rho_{1} and ρ2\rho_{2} are states on 𝒞{\mathcal{C}} such that ρ1|𝒟=ρ2|𝒟=π(y)\rho_{1}|_{\mathcal{D}}=\rho_{2}|_{\mathcal{D}}=\pi(y). Since ynπ1(Un)y\in\bigcap_{n}\pi^{-1}(U_{n}), we find π(y)nUn\pi(y)\in\bigcap_{n}U_{n}. Thus π(y)\pi(y) is a free point for the inclusion (,𝒟)({\mathcal{B}},{\mathcal{D}}). Next, since yv𝒮Xvy\in\bigcap_{v\in{\mathcal{S}}}X_{v}, ρ1(v)=ρ2(v)\rho_{1}(v)=\rho_{2}(v) for every v𝒮v\in{\mathcal{S}}. Using [PittsNoApUnInC*Al, Fact 1.2], for every v𝒮v\in{\mathcal{S}} and d𝒟d\in{\mathcal{D}}, ρ1(vd)=ρ2(vd)\rho_{1}(vd)=\rho_{2}(vd). Therefore, for every x𝔖x\in{\mathfrak{S}}, ρ1(x)=ρ2(x)\rho_{1}(x)=\rho_{2}(x). As 𝔖{\mathfrak{S}} is dense in 𝒞{\mathcal{C}}, we conclude ρ1=ρ2\rho_{1}=\rho_{2}. Thus π(y)\pi(y) is a free point for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}), so the claim holds.

Finally we show that π(𝒴)\pi({\mathcal{Y}}) is dense in 𝒟^\hat{\mathcal{D}}. Let G𝒟^G\subseteq\hat{\mathcal{D}} be an open set. Then π1(G)\pi^{-1}(G) is open in ^\hat{\mathcal{B}}, so there exists y𝒴π1(G)y\in{\mathcal{Y}}\cap\pi^{-1}(G), whence π(y)G\pi(y)\in G, as desired. As π(𝒴)𝔘\pi({\mathcal{Y}})\subseteq{\mathfrak{U}}, the proof is complete. ∎

We now come to our norming result.

Theorem \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a unital pseudo-Cartan inclusion. Then 𝒟{\mathcal{D}} norms 𝒞{\mathcal{C}}.

Proof.

The argument is mostly the same as the proof of  [PittsStReInI, Theorem 8.2], except we use Proposition 7 instead of [PittsStReInI, Theorem 3.10] and Lemma 7 instead of [PittsStReInI, Proposition 4.12]. Therefore, we shall only outline the proof, leaving the reader to consult the proof of [PittsStReInI, Theorem 8.2] for additional details as desired. (The notation here differs somewhat from that used in [PittsStReInI, Theorem 8.2], but this will present no difficulty.)

Arguing as in the last paragraph of the proof of  [PittsStReInI, Theorem 8.2], it suffices to show 𝒟{\mathcal{D}} norms 𝒞{\mathcal{C}} under the additional assumption that there is a countable set 𝒩(𝒞,𝒟){\mathfrak{C}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) such that 𝒞=C(𝒟){\mathcal{C}}=C^{*}({\mathfrak{C}}\cup{\mathcal{D}}). For the remainder of the proof, we assume this.

Let 𝒩(𝒞,𝒟){\mathcal{F}}\subseteq{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) be the *-semigroup generated by 𝒟{\mathfrak{C}}\cup{\mathcal{D}}. By the additional assumption, span\operatorname{span}{\mathcal{F}} is a dense *-subalgebra of 𝒞{\mathcal{C}}.

Let Y𝒟^Y\subseteq\hat{\mathcal{D}} be the set of free points for (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}). Proposition 7 shows YY is dense in 𝒟^\hat{\mathcal{D}}. For yYy\in Y, denote by fyf_{y} the unique state extension of yy to 𝒞{\mathcal{C}}, and let (πy,y,ξy)(\pi_{y},{\mathcal{H}}_{y},\xi_{y}) be the GNS triple for fyf_{y}. Since fyf_{y} is a pure state, πy\pi_{y} is an irreducible representation, and Lemma 7 shows that πy(𝒟)′′\pi_{y}({\mathcal{D}})^{\prime\prime} is a MASA in (y){\mathcal{B}}({\mathcal{H}}_{y}).

For y1,y2Yy_{1},y_{2}\in Y, define y1y2y_{1}\sim y_{2} if and only if there exists vv\in{\mathcal{F}} such that βv(y1)=y2\beta_{v}(y_{1})=y_{2}. As in the proof of [PittsStReInI, Theorem 8.2], \sim is an equivalence relation, and if πy1\pi_{y_{1}} is unitarily equivalent to πy2\pi_{y_{2}}, then y1y2y_{1}\sim y_{2}. Thus, if y1≁y2y_{1}\not\sim y_{2}, πy1\pi_{y_{1}} and πy2\pi_{y_{2}} are disjoint representations.

Let 𝒴Y{\mathcal{Y}}\subseteq Y be a set containing exactly one element of each \sim equivalence class. Consider the representation

π:=y𝒴πy\pi:=\bigoplus_{y\in{\mathcal{Y}}}\pi_{y}

of 𝒞{\mathcal{C}} on π:=y𝒴y{\mathcal{H}}_{\pi}:=\bigoplus_{y\in{\mathcal{Y}}}{\mathcal{H}}_{y}. Since each πy(𝒟)′′\pi_{y}({\mathcal{D}})^{\prime\prime} is an atomic MASA and the representations {πy:y𝒴}\{\pi_{y}:y\in{\mathcal{Y}}\} are pairwise disjoint, π(𝒟)′′\pi({\mathcal{D}})^{\prime\prime} is also an atomic MASA in (π){\mathcal{B}}({\mathcal{H}}_{\pi}).

We now show π\pi is faithful. As YY is dense in 𝒟^\hat{\mathcal{D}} and is the union of the 𝒩(𝒞,𝒟){\mathcal{N}}({\mathcal{C}},{\mathcal{D}})-orbits of the elements of 𝒴{\mathcal{Y}}, (7.1) shows that the restriction of π\pi to 𝒟{\mathcal{D}} is faithful. By the ideal intersection property for (𝒟c,𝒟)({\mathcal{D}}^{c},{\mathcal{D}}), kerπ𝒟c={0}\ker\pi\cap{\mathcal{D}}^{c}=\{0\}. Then kerπ={0}\ker\pi=\{0\} because (𝒞,𝒟c)({\mathcal{C}},{\mathcal{D}}^{c}) also has the ideal intersection property. Therefore π\pi is a faithful representation of 𝒞{\mathcal{C}}.

A direct argument or the argument found on [CameronPittsZarikianBiCaMASAvNAlNoAlMeTh, Page 466] shows π(𝒟)′′\pi({\mathcal{D}})^{\prime\prime} is locally cyclic (as defined in [PopSinclairSmithNoC*Al, Page 173]) for (π){\mathcal{B}}({\mathcal{H}}_{\pi}). Using [PopSinclairSmithNoC*Al, Lemma 2.3 and Theorem 2.7], π(𝒟)\pi({\mathcal{D}}) norms (π){\mathcal{B}}({\mathcal{H}}_{\pi}). As π\pi is faithful, it follows that 𝒟{\mathcal{D}} norms 𝒞{\mathcal{C}}. ∎

We now extend [PittsStReInII, Theorem 8.4] from the setting of virtual Cartan inclusions considered there to pseudo-Cartan inclusions. This is a significant generalization of [PittsStReInII, Theorem 8.4] because it weakens the hypothesis that 𝒞{\mathcal{C}} is unital and relaxes the condition that 𝒟{\mathcal{D}} be a MASA. We remark that [BrownExelFullerPittsReznikoffInC*AlCaEm, Example 5.1] gives an example of a Cartan inclusion (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) and an intermediate CC^{*}-subalgebra 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{B}}\subseteq{\mathcal{C}} such that (,𝒟)({\mathcal{B}},{\mathcal{D}}) is not a regular inclusion. Thus, in the context of Theorem 7, it is possible that the inclusions (C(𝒜i),𝒟i)(C^{*}({\mathcal{A}}_{i}),{\mathcal{D}}_{i}) are not regular.

Theorem \the\numberby.

Suppose for i=1,2i=1,2, (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are pseudo-Cartan inclusions such that (𝒞~i,𝒟~i)(\tilde{\mathcal{C}}_{i},\tilde{\mathcal{D}}_{i}) are pseudo-Cartan inclusions, and 𝒜i{\mathcal{A}}_{i} are Banach algebras satisfying 𝒟i𝒜i𝒞i{\mathcal{D}}_{i}\subseteq{\mathcal{A}}_{i}\subseteq{\mathcal{C}}_{i}. Let C(𝒜i)C^{*}({\mathcal{A}}_{i}) be the CC^{*}-subalgebra of 𝒞i{\mathcal{C}}_{i} generated by 𝒜i{\mathcal{A}}_{i}. If u:𝒜1𝒜2u:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2} is an isometric isomorphism, then uu uniquely extends to a *-isomorphism of C(𝒜1)C^{*}({\mathcal{A}}_{1}) onto C(𝒜2)C^{*}({\mathcal{A}}_{2}).

Remark \the\numberby. We have included the hypothesis that (𝒞~i,𝒟~i)(\tilde{\mathcal{C}}_{i},\tilde{\mathcal{D}}_{i}) are pseudo-Cartan inclusions because we do not know whether regularity is preserved when units are adjoined. When (𝒞~i,𝒟~i)(\tilde{\mathcal{C}}_{i},\tilde{\mathcal{D}}_{i}) are regular, Observation 4.2 shows (𝒞~i,𝒟~i)(\tilde{\mathcal{C}}_{i},\tilde{\mathcal{D}}_{i}) are pseudo-Cartan inclusions. As noted earlier, if (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) have the AUP or if 𝒞i{\mathcal{C}}_{i} are abelian, (𝒞~i,𝒟~i)(\tilde{\mathcal{C}}_{i},\tilde{\mathcal{D}}_{i}) are pseudo-Cartan inclusions provided that (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are pseudo-Cartan inclusions.

Proof of Theorem 7.

Suppose first that (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are unital pseudo-Cartan inclusions. Theorem 7 shows that C(𝒜i)=Ce(𝒜i)C^{*}({\mathcal{A}}_{i})=C^{*}_{e}({\mathcal{A}}_{i}). By Theorem 7, 𝒟i{\mathcal{D}}_{i} norms 𝒞i{\mathcal{C}}_{i}; in particular, 𝒟i{\mathcal{D}}_{i} norms 𝒜i{\mathcal{A}}_{i}. Thus [PittsNoAlAuCoBoIsOpAl, Corollary 1.5] shows uu uniquely extends to a *-isomorphism of C(𝒜1)C^{*}({\mathcal{A}}_{1}) onto C(𝒜2)C^{*}({\mathcal{A}}_{2}).

Now suppose (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are not assumed unital and u:𝒜1𝒜2u:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2} is an isometric isomorphism. Since (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is weakly non-degenerate, (𝒞~i,𝒟~i)(\tilde{\mathcal{C}}_{i},\tilde{\mathcal{D}}_{i}) is a unital pseudo-Cartan inclusion, so I𝒟~i=I𝒞~iI_{\tilde{\mathcal{D}}_{i}}=I_{\tilde{\mathcal{C}}_{i}}. Write 𝒜~i:=𝒜i+I𝒞~i\tilde{\mathcal{A}}_{i}:={\mathcal{A}}_{i}+{\mathbb{C}}I_{\tilde{\mathcal{C}}_{i}}, so that 𝒟~i𝒜~i𝒞~i\tilde{\mathcal{D}}_{i}\subseteq\tilde{\mathcal{A}}_{i}\subseteq\tilde{\mathcal{C}}_{i}. Notice that

C(𝒜~i)=C(𝒜i)+I𝒞~i.C^{*}(\tilde{\mathcal{A}}_{i})=C^{*}({\mathcal{A}}_{i})+{\mathbb{C}}I_{\tilde{\mathcal{C}}_{i}}.

Applying [MeyerAdUnOpAl, Corollary 3.3] with n=1n=1 shows u~:𝒜~1𝒜~2\tilde{u}:\tilde{\mathcal{A}}_{1}\rightarrow\tilde{\mathcal{A}}_{2} is an isometric isomorphism. Therefore, u~\tilde{u} uniquely extends to a *-isomorphism θ\theta of C(𝒜~1)C^{*}(\tilde{\mathcal{A}}_{1}) onto C(𝒜~2)C^{*}(\tilde{\mathcal{A}}_{2}). Thus θ|C(𝒜1)\theta|_{C^{*}({\mathcal{A}}_{1})} is a *-isomorphism of C(𝒜1)C^{*}({\mathcal{A}}_{1}) onto C(𝒜2)C^{*}({\mathcal{A}}_{2}) extending uu.

If π:C(𝒜1)C(𝒜2)\pi:C^{*}({\mathcal{A}}_{1})\rightarrow C^{*}({\mathcal{A}}_{2}) is another *-isomorphism such that π|𝒜1=u\pi|_{{\mathcal{A}}_{1}}=u, then π~|𝒜~1=u~\tilde{\pi}|_{\tilde{\mathcal{A}}_{1}}=\tilde{u}. As θ\theta is the unique extension of u~\tilde{u} to C(𝒜~1)C^{*}(\tilde{\mathcal{A}}_{1}), π~=θ\tilde{\pi}=\theta. Therefore, π=θ|C(𝒜1)\pi=\theta|_{C^{*}({\mathcal{A}}_{1})}. This gives uniqueness of the extension of uu to C(𝒜1)C^{*}({\mathcal{A}}_{1}) and completes the proof.∎

If the hypothesis that (C(𝒜1),𝒟1)(C({\mathcal{A}}_{1}),{\mathcal{D}}_{1}) is a regular inclusion is added to the hypotheses of Theorem 7, more can be said.

Corollary \the\numberby.

Suppose for i=1,2i=1,2, (𝒞i,𝒟i)({\mathcal{C}}_{i},{\mathcal{D}}_{i}) are pseudo-Cartan inclusions such that (𝒞~i,𝒟~i)(\tilde{\mathcal{C}}_{i},\tilde{\mathcal{D}}_{i}) are pseudo-Cartan inclusions, that 𝒜i{\mathcal{A}}_{i} are Banach algebras satisfying 𝒟i𝒜i𝒞i{\mathcal{D}}_{i}\subseteq{\mathcal{A}}_{i}\subseteq{\mathcal{C}}_{i}, and that u:𝒜1𝒜2u:{\mathcal{A}}_{1}\rightarrow{\mathcal{A}}_{2} is an isometric isomorphism.

If (C(𝒜1),𝒟1)(C^{*}({\mathcal{A}}_{1}),{\mathcal{D}}_{1}) is regular, then (C(𝒜1),𝒟1)(C^{*}({\mathcal{A}}_{1}),{\mathcal{D}}_{1}) and (C(𝒜2),u(𝒟1))(C^{*}({\mathcal{A}}_{2}),u({\mathcal{D}}_{1})) are pseudo-Cartan inclusions; let (1,𝒩1 :: τ1)({\mathcal{M}}_{1},{\mathcal{N}}_{1}\hbox{\,:\hskip-1.0pt:\,}\tau_{1}) and (2,𝒩2 :: τ2)({\mathcal{M}}_{2},{\mathcal{N}}_{2}\hbox{\,:\hskip-1.0pt:\,}\tau_{2}) be their Cartan envelopes. Furthermore, there is a unique extension of uu to a *-isomorphism u˘:12\breve{u}:{\mathcal{M}}_{1}\rightarrow{\mathcal{M}}_{2} such that

u˘τ1=τ2u.\breve{u}\circ\tau_{1}=\tau_{2}\circ u.
Proof.

Let u:C(𝒜1)C(𝒜2)u^{\prime}:C^{*}({\mathcal{A}}_{1})\rightarrow C^{*}({\mathcal{A}}_{2}) be the unique *-isomorphism extending uu provided by Theorem 7. Since u|𝒟1=u|𝒟1u^{\prime}|_{{\mathcal{D}}_{1}}=u|_{{\mathcal{D}}_{1}}, u(𝒟1)u({\mathcal{D}}_{1}) is a CC^{*}-algebra. Since (C(𝒜1),𝒟1)(C^{*}({\mathcal{A}}_{1}),{\mathcal{D}}_{1}) is a regular inclusion, so is (C(𝒜2),u(𝒟1))(C^{*}({\mathcal{A}}_{2}),u({\mathcal{D}}_{1})). By Corollary 2.3, both (C(𝒜1),𝒟1)(C^{*}({\mathcal{A}}_{1}),{\mathcal{D}}_{1}) and (C(𝒜2),u(𝒟1)(C^{*}({\mathcal{A}}_{2}),u({\mathcal{D}}_{1})) have the faithful unique pseudo-expectation property, so they are pseudo-Cartan inclusions. Also for any v𝒩(C(𝒜1),𝒟1)v\in{\mathcal{N}}(C^{*}({\mathcal{A}}_{1}),{\mathcal{D}}_{1}) and h𝒟1h\in{\mathcal{D}}_{1},

u(v)u(h)u(v)=u(vhv)u(𝒟1)=u(𝒟1).u^{\prime}(v)^{*}u(h)u^{\prime}(v)=u^{\prime}(v^{*}hv)\in u^{\prime}({\mathcal{D}}_{1})=u({\mathcal{D}}_{1}).

Thus uu^{\prime} is a regular *-isomorphism. An application of Theorem 6.1 completes the proof. ∎

The following is immediate from Theorem 7.

Corollary \the\numberby.

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion such that (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is a pseudo-Cartan inclusion. If 𝒜{\mathcal{A}} is an algebra such that 𝒟𝒜𝒞{\mathcal{D}}\subseteq{\mathcal{A}}\subseteq{\mathcal{C}} and C(𝒜)=𝒞C^{*}({\mathcal{A}})={\mathcal{C}}, then the group of isometric automorphisms of 𝒜{\mathcal{A}} is isomorphic to the group of all *-automorphisms of 𝒞{\mathcal{C}} which leave 𝒜{\mathcal{A}} invariant.

Theorems 77 seem interesting even in the commutative case, for they apply to certain function algebras.

Example \the\numberby.

Let 𝒞{\mathcal{C}} be a unital and abelian CC^{*}-algebra, and suppose J𝒞J\unlhd{\mathcal{C}} is an essential ideal. Let 𝒜𝒞{\mathcal{A}}\subseteq{\mathcal{C}} be a Banach algebra such that: 𝒜{\mathcal{A}} separates points of 𝒞^\hat{\mathcal{C}}, I𝒞𝒜{\mathbb{C}}I_{\mathcal{C}}\subseteq{\mathcal{A}}, and 𝒜J={0}{\mathcal{A}}\cap J=\{0\}. Then 𝒜+J{\mathcal{A}}+J is a Banach algebra satisfying JJ+𝒜𝒞J\subseteq J+{\mathcal{A}}\subseteq{\mathcal{C}}. By the Stone-Weierstrauß Theorem, C(𝒜)=𝒞C^{*}({\mathcal{A}})={\mathcal{C}}. Since (𝒞,J)({\mathcal{C}},J) is a pseudo-Cartan inclusion, Cenv(𝒜+J)=𝒞C^{*}_{env}({\mathcal{A}}+J)={\mathcal{C}} and any isometric automorphism of 𝒜+J{\mathcal{A}}+J uniquely extends to a *-isomorphism of 𝒞{\mathcal{C}}.

8. Questions

In addition to Conjecture 4.3, we now present a few open questions.

Recall that Barlak and Li [BarlakLiCaSuUCTPr, Corollary 1.2] showed that if a separable, nuclear, CC^{*}-algebra contains a Cartan MASA, then it satisfies the Universal Coefficient Theorem (UCT). Separable, nuclear CC^{*}-algebras which satisfy the UCP are of significant interest in the classification program. This is in part the motivation for the following two questions.

Question \the\numberby. Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion. Must 𝒞{\mathcal{C}} contain a Cartan MASA?

Here is an example which motivates our next question. Let XX be the Cantor set and let Γ\Gamma be a countable discrete group having property (T)(T). It follows from a result of Elek [ElekFrMiAcCoGrInPrMe, Theorem 1], that there exists a free and minimal action of Γ\Gamma on XX which admits an ergodic (regular, non-atomic) invariant Borel probability measure μ\mu. Let tUtt\mapsto U_{t} be the unitary representation of Γ\Gamma on L2:=L2(X,μ)L^{2}:=L^{2}(X,\mu): (Utξ)(s)=ξ(t1s)(U_{t}\xi)(s)=\xi(t^{-1}s) and MM be the representation of C(X)C(X) by multiplication operators on L2L^{2}.

Define 𝒞\mathcal{C} to be the CC^{*}-algebra generated by the images of UU and MM, and put 𝒟=M(C(X))\mathcal{D}=M(C(X)). (It turns out 𝒞\mathcal{C} is an exotic crossed product.) The freeness of the action implies (𝒞,𝒟)(\mathcal{C},\mathcal{D}) is a regular MASA inclusion with the unique state extension property, so there exists a (unique) conditional expectation E:𝒞𝒟E:\mathcal{C}\rightarrow\mathcal{D}.

Let J:={x𝒞:E(xx)=0}J:=\{x\in\mathcal{C}:E(x^{*}x)=0\} be the left kernel of EE. By [PittsStReInI, Theorem 3.15], JJ is a ideal of 𝒞\mathcal{C} having trivial intersection with 𝒟\mathcal{D} and [ExelPittsZarikianExIdFrTrGrC*Al, Theorem 3.7] shows J{0}J\neq\{0\}. Then (𝒞/J,𝒟)(\mathcal{C}/J,\mathcal{D}) is a regular inclusion having the unique state extension property by [ArchboldBunceGregsonExStC*AlII, Lemma 3.1]. It follows (𝒞/J,𝒟)(\mathcal{C}/J,\mathcal{D}) is a regular inclusion with the extension property which has a faithful conditional expectation; thus (𝒞/J,𝒟)(\mathcal{C}/J,\mathcal{D}) is a CC^{*}-diagonal.

Question \the\numberby. Suppose 𝒞{\mathcal{C}} is a CC^{*}-algebra and J𝒞J\unlhd{\mathcal{C}}. If 𝒞/J{\mathcal{C}}/J admits a subalgebra {\mathcal{E}} such that (𝒞/J,)({\mathcal{C}}/J,{\mathcal{E}}) is a pseudo-Cartan inclusion (or a Cartan inclusion), must there exist 𝒟𝒞{\mathcal{D}}\subseteq{\mathcal{C}} such that (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion (or Cartan inclusion)?

Next, the uniqueness statement of Theorem 3 suggests the possibility of finding a canonical method of replacing certain non-Hausdorff twists with Hausdorff ones.

Question \the\numberby. Suppose (Σ,G)(\Sigma,G) is a twist over the second countable, étale and topologically free groupoid GG (see [ExelPittsChGrC*AlNoHaEtGr, Definition 3.4.6] for the definition of a topologically free groupoid). We assume G(0)G^{(0)} is Hausdorff, but we do not assume GG is Hausdorff. Then (Cess(Σ,G),C0(G(0)))(C^{*}_{ess}(\Sigma,G),C_{0}(G^{(0)})) is a weak-Cartan inclusion (see [ExelPittsChGrC*AlNoHaEtGr, Corollary 3.9.5]) and hence a pseudo-Cartan inclusion. Let (𝒜, :: α)({\mathcal{A}},{\mathcal{B}}\hbox{\,:\hskip-1.0pt:\,}\alpha) be its Cartan envelope, and let (Σ𝒜,G𝒜)(\Sigma_{{\mathcal{A}}},G_{{\mathcal{A}}}) be the twist associated to (𝒜,)({\mathcal{A}},{\mathcal{B}}). Then G𝒜G_{{\mathcal{A}}} is Hausdorff. What is the relationship between (Σ𝒜,G𝒜)(\Sigma_{{\mathcal{A}}},G_{{\mathcal{A}}}) and (Σ,G)(\Sigma,G)?

We close with a technical question. Several of the results in Section 7 assume that the unitization of a pseudo-Cartan inclusion is again a pseudo-Cartan inclusion and it would be interesting to know if that hypothesis can be removed. The issue is whether (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) is regular.

Question \the\numberby. If (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a pseudo-Cartan inclusion, must (𝒞~,𝒟~)(\tilde{\mathcal{C}},\tilde{\mathcal{D}}) be a pseudo-Cartan inclusion?

Appendix A

Changes in Notation and Terminology. Throughout Appendix A, we depart from the notation and terminology used in Sections 17 above; instead we use the notation and terminology found in [PittsStReInII]. In particular, all inclusions below are unital, and for a unital CC^{*}-algebra BB, an essential extension for {\mathcal{B}} will mean a pair (𝒜,α)({\mathcal{A}},\alpha) consisting of a unital CC^{*}-algebra 𝒜{\mathcal{A}} and a unital *-monomorphism α:𝒜\alpha:{\mathcal{B}}\rightarrow{\mathcal{A}} such that α()𝒜\alpha({\mathcal{B}})\subseteq{\mathcal{A}} has the ideal intersection property.

Lemma 2.3 of [PittsStReInII] is presented without proof, and there is an error in its statement. The error is the assertion that the map Ropen(𝒜^)Gr1(G)\operatorname{\textsc{Ropen}}(\hat{\mathcal{A}})\ni G\mapsto r^{-1}(G) is a Boolean algebra isomorphism of Ropen(𝒜^)\operatorname{\textsc{Ropen}}(\hat{\mathcal{A}}) onto Ropen(^)\operatorname{\textsc{Ropen}}(\hat{\mathcal{B}}): it should have said Ropen(𝒜^)G(r1(G)¯)Ropen(^)\operatorname{\textsc{Ropen}}(\hat{\mathcal{A}})\ni G\mapsto\left(\overline{r^{-1}(G)}\right)^{\circ}\in\operatorname{\textsc{Ropen}}(\hat{\mathcal{B}}) is a Boolean algebra isomorphism. Here is the full and corrected statement.

Lemma \the\numberby (Corrected [PittsStReInII, Lemma 2.3]).

Suppose 𝒜{\mathcal{A}} and {\mathcal{B}} are abelian, unital CC^{*}-algebras, (,α)({\mathcal{B}},\alpha) is an essential extension of 𝒜{\mathcal{A}}, and r:^𝒜^r:\hat{\mathcal{B}}\rightarrow\hat{\mathcal{A}} is the continuous surjection, ρ^ρα\rho\in\hat{\mathcal{B}}\mapsto\rho\circ\alpha. Then the maps

(A.1) Rideal()Jα1(J)\displaystyle\operatorname{\textsc{Rideal}}({\mathcal{B}})\ni J\mapsto\alpha^{-1}(J) andRopen(𝒜^)G(r1(G)¯)\displaystyle\quad\text{and}\quad\operatorname{\textsc{Ropen}}(\hat{\mathcal{A}})\ni G\mapsto\left(\overline{r^{-1}(G)}\right)^{\circ}
are Boolean algebra isomorphisms of Rideal()\operatorname{\textsc{Rideal}}({\mathcal{B}}) onto Rideal(𝒜)\operatorname{\textsc{Rideal}}({\mathcal{A}}) and Ropen(𝒜^)\operatorname{\textsc{Ropen}}(\hat{\mathcal{A}}) onto Ropen(^)\operatorname{\textsc{Ropen}}(\hat{\mathcal{B}}) respectively. The inverses of these maps are
(A.2) Rideal(𝒜)Kα(K)\displaystyle\operatorname{\textsc{Rideal}}({\mathcal{A}})\ni K\mapsto\alpha(K)^{\perp\perp} andRopen(^)H(r(H¯))\displaystyle\quad\text{and}\quad\operatorname{\textsc{Ropen}}(\hat{\mathcal{B}})\ni H\mapsto(r(\overline{H}))^{\circ}
respectively. Furthermore, for JRideal()J\in\operatorname{\textsc{Rideal}}({\mathcal{B}}) and GRopen(𝒜^)G\in\operatorname{\textsc{Ropen}}(\hat{\mathcal{A}}),
(A.3) supp(α1(J))=(r(supp(J)¯)\displaystyle\operatorname{supp}(\alpha^{-1}(J))=(r(\overline{\operatorname{supp}(J)})^{\circ} andideal((r1(G)¯))=α(ideal(G)).\displaystyle\quad\text{and}\quad\operatorname{ideal}(\left(\overline{r^{-1}(G)}\right)^{\circ})=\alpha(\operatorname{ideal}(G))^{\perp\perp}.

A complete proof of Lemma A may be found in [PittsIrMaIsBoReOpSeReId]. The proofs of (A.1) and (A.2) are found in [PittsIrMaIsBoReOpSeReId, Propositions 3.2 and 4.17] and (A.3) is [PittsIrMaIsBoReOpSeReId, Lemma 4.12] combined with [PittsIrMaIsBoReOpSeReId, Lemma 4.7]. To make the notational transition from [PittsIrMaIsBoReOpSeReId] to the notation of Lemma A, take Y:=^Y:=\hat{\mathcal{B}}, X=A^X=\hat{A} and π:=r\pi:=r.

Unfortunately, the misstatement in [PittsStReInII, Lemma 2.3] leads to a gap in the proof of [PittsStReInII, Theorem 6.9]. The statement of [PittsStReInII, Theorem 6.9] is correct, but its proof is insufficient, as we now describe.

Let (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) be a regular inclusion with the unique pseudo-expectation property, let (I(𝒟),ι)(I({\mathcal{D}}),\iota) be an injective envelope for 𝒟{\mathcal{D}} and let E:𝒞I(𝒟)E:{\mathcal{C}}\rightarrow I({\mathcal{D}}) be the pseudo-expectation. In the proof of [PittsStReInII, Theorem 6.9] we claimed that if v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) and ρI(𝒟)^\rho\in\widehat{I({\mathcal{D}})} satisfies ρ(E(v))0\rho(E(v))\neq 0, then ρι(fixβv)\rho\circ\iota\in(\operatorname{fix}\beta_{v})^{\circ}. That claim is then used show that ρE\rho\circ E is a compatible state. Our proof of this claim is insufficient because it uses the incorrectly stated portion of [PittsStReInII, Lemma 2.3].

We now give a proof of [PittsStReInII, Theorem 6.9], whose statement we have reproduced in Theorem A. The proof uses the correction to [PittsStReInII, Lemma 2.3] given as Lemma A above to show that ρE\rho\circ E is a compatible state.

Theorem \the\numberby ([PittsStReInII, Theorem 6.9]).

Suppose (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a regular inclusion with the unique pseudo-expectation property. Then (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a covering inclusion and 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) is a compatible cover for 𝒟^\hat{\mathcal{D}}. Furthermore, if FF is any closed subset of Mod(𝒞,𝒟)\text{Mod}({\mathcal{C}},{\mathcal{D}}) which covers 𝒟^\hat{\mathcal{D}}, then 𝔖s(𝒞,𝒟)F{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}})\subseteq F.

Proof.

Denote by rr the “restriction” map, I(𝒟)^ρρι𝒟^\widehat{I({\mathcal{D}})}\ni\rho\mapsto\rho\circ\iota\in\hat{\mathcal{D}}. Since (I(𝒟),ι)(I({\mathcal{D}}),\iota) is an injective envelope for 𝒟{\mathcal{D}}, it is in particular an essential extension of 𝒟{\mathcal{D}}.

For v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}), let

Uv:={ρI(𝒟)^:ρ(E(v))0}.U_{v}:=\{\rho\in\widehat{I({\mathcal{D}})}:\rho(E(v))\neq 0\}.

If ρUv\rho\in U_{v}, the Cauchy-Schwartz inequality gives,

|ρ(E(v))|2ρ(E(vv))=(ρι)(vv),|\rho(E(v))|^{2}\leq\rho(E(v^{*}v))=(\rho\circ\iota)(v^{*}v),

so r(ρ)domβvr(\rho)\in\operatorname{dom}\beta_{v}. By [PittsStReInI, Lemma 2.5], ρE|𝒟=r(ρ)fixβv\rho\circ E|_{\mathcal{D}}=r(\rho)\in\operatorname{fix}\beta_{v}. Thus,

(A.4) r(Uv)fix(βv).r(U_{v})\subseteq\operatorname{fix}(\beta_{v}).

Let us show that every τ𝔖s(𝒞,𝒟)\tau\in{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) is a compatible state. Fix τ𝔖s(𝒞,𝒟)\tau\in{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) and suppose τ(v)0\tau(v)\neq 0 for some v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}). Our task is to show that |τ(v)|2=τ(vv)|\tau(v)|^{2}=\tau(v^{*}v). Write τ=ρE\tau=\rho\circ E for some ρI(𝒟)^\rho\in\widehat{I({\mathcal{D}})}. Let

X:={ρI(𝒟)^:ρ(E(v))>ρ(E(v))/2}.X:=\{\rho^{\prime}\in\widehat{I({\mathcal{D}})}:\rho^{\prime}(E(v))>\rho(E(v))/2\}.

By construction, XX is an open subset of I(𝒟)^\widehat{I({\mathcal{D}})} and ρX\rho\in X. Since X¯Uv\overline{X}\subseteq U_{v},  (A.4) gives,

r(X¯)(fixβv).r(\overline{X})^{\circ}\subseteq(\operatorname{fix}\beta_{v})^{\circ}.

Put G:=r(X¯)G:=r(\overline{X})^{\circ}. Since I(𝒟)^\widehat{I({\mathcal{D}})} is a Stonean space, X¯\overline{X} and r1(G)¯\overline{r^{-1}(G)} are clopen sets and hence are regular open subsets of I(𝒟)^\widehat{I({\mathcal{D}})}. Then

r1(G)¯=(r1(G)¯)=(r1(r(X¯))¯)=X¯,\overline{r^{-1}(G)}=(\overline{r^{-1}(G)})^{\circ}=\left(\overline{r^{-1}(r(\overline{X})^{\circ})}\right)^{\circ}=\overline{X},

with the last equality following from the portions of Lemma A concerning regular open sets. Therefore, we may find a net (ρλ)(\rho_{\lambda}) in r1(G)r^{-1}(G) such that ρλρ\rho_{\lambda}\rightarrow\rho. Since r(ρλ)=ρλι(fixβv)r(\rho_{\lambda})=\rho_{\lambda}\circ\iota\in(\operatorname{fix}\beta_{v})^{\circ}, we may find dλideal((fixβv))d_{\lambda}\in\operatorname{ideal}((\operatorname{fix}\beta_{v})^{\circ}) with ρλ(ι(dλ))=1\rho_{\lambda}(\iota(d_{\lambda}))=1. Since vdλ𝒟cvd_{\lambda}\in{\mathcal{D}}^{c} (see [PittsStReInII, Lemmas 2.14 and 2.15]), and E|𝒟cE|_{{\mathcal{D}}^{c}} is a homomorphism (by [PittsStReInII, Lemma 6.8])

|ρλ(E(v))|2=|ρλ(E(v)ι(dλ))|2=|ρλ(E(vdλ))|2=|ρλ(E(dλvvdλ))|=ρλ(E(vv)).|\rho_{\lambda}(E(v))|^{2}=|\rho_{\lambda}(E(v)\iota(d_{\lambda}))|^{2}=|\rho_{\lambda}(E(vd_{\lambda}))|^{2}=|\rho_{\lambda}(E(d_{\lambda}^{*}v^{*}vd_{\lambda}))|=\rho_{\lambda}(E(v^{*}v)).

As τ=ρE\tau=\rho\circ E,

|τ(v)|2=limλ|ρλ(E(v))|2=limλρλ(E(vv))=τ(vv).|\tau(v)|^{2}=\lim_{\lambda}|\rho_{\lambda}(E(v))|^{2}=\lim_{\lambda}\rho_{\lambda}(E(v^{*}v))=\tau(v^{*}v).

It follows that τ\tau is a compatible state.

The remainder of the proof now follows exactly as in the proof of [PittsStReInII, Theorem 6.9]; we include the details for convenience.

Let us show the invariance of 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}). Choose τ𝔖s(𝒞,𝒟)\tau\in{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) and write τ=ρE\tau=\rho\circ E for some ρI(𝒟)^\rho\in\widehat{I({\mathcal{D}})}. Suppose v𝒩(𝒞,𝒟)v\in{\mathcal{N}}({\mathcal{C}},{\mathcal{D}}) is such that ρ(ι(vv))0\rho(\iota(v^{*}v))\neq 0. Let PP and QQ be the support projections in I(𝒟)I({\mathcal{D}}) for the ideals vv𝒟¯\overline{vv^{*}{\mathcal{D}}} and vv𝒟¯\overline{v^{*}v{\mathcal{D}}} respectively. Then θ~v\tilde{\theta}_{v} is a partial automorphism with domain PI(𝒟)PI({\mathcal{D}}) and range QI(𝒟)QI({\mathcal{D}}). Define τI(𝒟)^\tau^{\prime}\in\widehat{I({\mathcal{D}})} by

τ(h)=ρ(θ~v(Ph)),hI(𝒟).\tau^{\prime}(h)=\rho(\tilde{\theta}_{v}(Ph)),\qquad h\in I({\mathcal{D}}).

For x𝒞x\in{\mathcal{C}}, [PittsStReInII, Proposition 6.2] gives,

ρ(E(vxv))\displaystyle\rho(E(v^{*}xv)) =ρ(θ~v(E(vvx)))=ρ(θ~v(ι(vv)PE(x)))\displaystyle=\rho(\tilde{\theta}_{v}(E(vv^{*}x)))=\rho(\tilde{\theta}_{v}(\iota(vv^{*})PE(x)))
=ρ(ι(vv))ρ(θ~v(PE(x)))=ρ(ι(vv))(τE)(x).\displaystyle=\rho(\iota(v^{*}v))\rho(\tilde{\theta}_{v}(PE(x)))=\rho(\iota(v^{*}v))(\tau^{\prime}\circ E)(x).

Thus 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) is invariant.

If σ𝒟^\sigma\in\hat{\mathcal{D}}, choose any ρI(𝒟)^\rho\in\widehat{I({\mathcal{D}})} such that ρι=σ\rho\circ\iota=\sigma. Then σ=(ρE)|𝒟\sigma=(\rho\circ E)|_{\mathcal{D}}, so 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) covers 𝒟^\hat{\mathcal{D}}. Thus, 𝔖s(𝒞,𝒟){\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}}) is a compatible cover for 𝒟^\hat{\mathcal{D}} and (𝒞,𝒟)({\mathcal{C}},{\mathcal{D}}) is a covering inclusion.

Finally, if FMod(𝒞,𝒟)F\subseteq\text{Mod}({\mathcal{C}},{\mathcal{D}}) is closed and covers 𝒟^\hat{\mathcal{D}}, then 𝔖s(𝒞,𝒟)F{\mathfrak{S}}_{s}({\mathcal{C}},{\mathcal{D}})\subseteq F by [PittsStReInII, Theorem 6.1(b)].

References