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

Maximality of correspondence representations

Boris Bilich
Abstract

In this paper, we fully characterize maximal representations of a C*-correspondence, thereby strengthening several earlier results. We demonstrate the maximality criteria through diverse examples. We also describe the noncommutative Choquet boundary and provide additional counterexamples to Arveson’s hyperrigidity conjecture following the counterexample recently found by the author and Dor-On. Furthermore, we identify several classes of correspondences for which the hyperrigidity conjecture holds.

Maximality of correspondence representations


Boris Bilich111University of Haifa and Georg-August-Universität Göttingen
The author is partially supported by the Bloom PhD scholarship and the DFG Middle-Eastern collaboration project no. 529300231
Keywords: C*-correspondence, Cuntz-Pimsner algebra, hyperrigidity, dilation, unique extension property, Choquet boundary. 2020 Mathematics Subject Classification: 46L07, 46L08, 47A20, 47L55.

1 Introduction

A dilation of an operator TT on a Hilbert space HH is an operator SS on a larger Hilbert space KHK\supset H such that TT is a corner of SS, often represented in block-diagonal form as:

S=[T].S=\begin{bmatrix}T&\ast\\ \ast&\ast\end{bmatrix}.

The general paradigm of dilation theory is to study properties of operators by finding a dilation that exhibits features that are easier to work with. This concept is best illustrated by classical results such as Sz.-Nagy’s unitary dilation of a contraction [50], Andô’s dilation theorem [3], which shows that a pair of commuting contractions can be dilated to a pair of commuting unitaries, and Stinespring’s dilation [49] of a unital completely positive map on a C*-algebra to a *-homomorphism. The theory is now flourishing [8, 12, 18, 48, 22, 19, 53, 44] and has many applications. For a comprehensive introduction and overview, see [47].

This paper studies dilation theory for operator systems associated with C*-correspondences. We first define what operator systems and dilations mean in this context to set the stage.

An operator system 𝒮\mathcal{S} is a self-adjoint unital subspace of a C*-algebra BB. We additionally require that 𝒮\mathcal{S} generates BB as a C*-algebra. A unital linear map φ:𝒮(H)\varphi\colon\mathcal{S}\to\mathcal{B}(H) is called unital completely positive (u.c.p. map) if it sends positive nn by nn matrices with entries in 𝒮\mathcal{S} to positive operators on HnH^{\oplus n}. A dilation of φ\varphi is a u.c.p. map ψ:𝒮(K)\psi\colon\mathcal{S}\to\mathcal{B}(K) together with an isometric embedding V:HKV\colon H\hookrightarrow K such that

Vψ(s)V=φ(s) for all s𝒮.V^{*}\psi(s)V=\varphi(s)\text{ for all }s\in\mathcal{S}.

The dilation is called trivial if ψ=φψ\psi=\varphi\oplus\psi^{\prime} under the identification of HH with VHVH. A u.c.p. map φ\varphi is said to be maximal if it admits no non-trivial dilations.

Building on the ideas from Agler [1] and Muhly and Solel [37], Dritschel and McCullough demonstrated that maximal u.c.p. maps can be characterized by the unique extension property (UEP) defined by Arveson [7, 4]. By definition, a u.c.p. map φ:𝒮(H)\varphi\colon\mathcal{S}\to\mathcal{B}(H) satisfies the UEP if there is a *-representation ρ:B(H)\rho\colon B\to\mathcal{B}(H) such that φ=ρ|𝒮\varphi=\rho|_{\mathcal{S}} and ρ\rho is the unique u.c.p. extension of φ\varphi. We say that a representation ρ\rho of BB is maximal (with respect to 𝒮\mathcal{S}) if ρ|𝒮\rho|_{\mathcal{S}} is a maximal u.c.p. map. By the above, there is a bijective correspondence between maximal u.c.p. maps and maximal BB-representations.

According to the theorem of Dritschel and McCullough [24, Theorem 2.1], every u.c.p. map has a maximal dilation. Dritschel and McCullough used maximality to provide the first dilation-theoretic proof of the existence of the C*-envelope Cenv(𝒮)C^{*}_{env}(\mathcal{S}), which is the smallest quotient of BB such that the quotient map is completely isometric on 𝒮\mathcal{S}. As they showed in [24, Theorem 4.1], maximal representations factor through Cenv(𝒮)C^{*}_{env}(\mathcal{S}), and there is a maximal representation that is faithful on Cenv(𝒮)C^{*}_{env}(\mathcal{S}).

However, not every representation of Cenv(𝒮)C^{*}_{env}(\mathcal{S}) is maximal. To better understand this phenomenon, Arveson [5] defined the noncommutative Choquet boundary C𝒮Cenv(𝒮)^\partial_{C}\mathcal{S}\subset\widehat{C^{*}_{env}(\mathcal{S})} as the subset of unitary equivalence classes of maximal irreducible representations. Elements of C𝒮\partial_{C}\mathcal{S} are called boundary representations. Arveson [5] in the separable case and Davidson-Kennedy [21] in general proved that every operator system has sufficiently many boundary representations.

In [6] Arveson introduced the notion of hyperrigidity of operator systems and gave several equivalent definitions. The most relevant definition for us is that an operator system is hyperrigid if all representations of Cenv(𝒮)C^{*}_{env}(\mathcal{S}) are maximal. In an attempt to connect this notion with the noncommutative Choquet boundary, Arveson conjectured [6, Conjecture 4.3] that if C𝒮=Cenv(𝒮)^\partial_{C}\mathcal{S}=\widehat{C^{*}_{env}(\mathcal{S})}, i.e., all irreducible representations of the C*-envelope are boundary representations, then the operator system is hyperrigid. This has come to be known as Arveson’s hyperrigidity conjecture, and it has received significant attention in the literature [21, 33, 14, 13, 15, 17, 20, 27, 34, 30, 51, 42, 28]. Recently, an elementary Type I counterexample was found by the author and Dor-On [9]. However, it is still interesting to determine for which classes of operator systems the conjecture holds. For example, the conjecture remains open for function systems in commutative C*-algebras.

In the current paper, we study the topics described above in the context of a C*-correspondence XX over a C*-algebra AA. There are two natural operator systems associated with XX in the Cuntz-Toeplitz algebra 𝒯X\mathcal{T}_{X}. The first one is 𝒯X++(𝒯X+)\mathcal{T}_{X}^{+}+(\mathcal{T}_{X}^{+})^{*} for the non-self-adjoint tensor algebra 𝒯X+\mathcal{T}_{X}^{+} generated by AA and XX. This operator algebra and the associated dilation theory were studied by Muhly and Solel [39] and by Katsoulis and Ramsey [30, 31]. Another operator system is 𝒮X=X+A+X𝒯X\mathcal{S}_{X}=X^{*}+A+X\subset\mathcal{T}_{X} which was studied by Kim [34]. Moreover, following Muhly and Solel [39], we give a definition of dilations and maximality for representations of XX without any reference to operator systems in Definition 2.6. We thus have, a priori, three notions of maximality for isometric representations of XX. Fortunately, they all coincide (Proposition 2.8).

We study dilation theory for correspondences in two steps. First, we analyze the case when XX is a W*-correspondence over a von Neumann algebra 𝔐\mathfrak{M}. It turns out that the dilations in this case have a much more amenable structure. Maximality is then detected by a spatial condition involving the support projection PX𝔐P_{X}\in\mathfrak{M} of XX as follows: a representation t:X(H)t\colon X\to\mathcal{B}(H) is maximal if and only if t(X)H¯=ρ(PX)H\overline{t(X)H}=\rho(P_{X})H. We call this property the full Cuntz-Pimsner covariance since it generalizes the notion of full Cuntz-Krieger families by Dor-On and Salomon [23] in the context of graph operator algebras.

After dealing with W*-correspondences, we reduce the problem for C*-correspondences to the W*-case by using second dual techniques (see Section 3). If XX is a C*-correspondence over a C*-algebra AA, then A{A}^{**} is the universal von Neumann algebra of AA, and X{X}^{**} is a W*-correspondence over it (see [11, Section 8.5]). We prove that the representations of XX are in canonical bijection with the normal representations of X{X}^{**} and that this bijection respects dilations (Proposition 3.3). Therefore, the notions of maximality for XX and X{X}^{**} coincide and thus the maximality criterion directly follows from the one for W*-correspondences. We define full Cuntz-Pimsner covariance for W*- and C*-correspondences simultaneously in Definition 4.3 and prove that it is equivalent to maximality in Theorem 4.4, which is the main result of the paper. A similar idea to detect maximality by a projection in the second dual algebra was investigated by Clouâtre and Saikia [16]. This shows that the reduction to von Neumann algebras via second duals is a powerful tool for studying non-commutative Choquet boundary theory.

Of course, the second dual A{A}^{**} is a very big algebra which is hard to describe even for the most basic C*-algebras. Therefore, we also prove several maximality criteria which do not refer to the second dual. We describe maximality for an almost proper correspondence in terms of ideals in Proposition 4.10. Furthermore, we interpret full Cuntz-Pimsner covariance for topological graphs in terms of receiving vertices in Theorem 6.5.

Using the maximality criterion, we are able to recover several earlier results. The characterization of maximal representations of graph correspondences by Dor-On and Salomon [23] is an immediate consequence, and we describe it in Example 4.5 (and Theorem 6.5 is a generalization to topological graphs). In Corollary 4.8, we prove the Katsoulis-Kribs theorem, which states that the Cuntz-Pimsner algebra 𝒪X\mathcal{O}_{X} is the C*-envelope of 𝒯X+\mathcal{T}^{+}_{X}.

Kim [34] established a general criterion for the hyperrigidity of a correspondence, refining the result of Katsoulis and Ramsey [30], who provided a sufficient condition for hyperrigidity when the correspondence is countably generated. In Corollary 4.9, we derive Kim’s hyperrigidity criterion: a correspondence XX is hyperrigid if and only if Katsura’s ideal acts non-degenerately on XX.

In Section 5, we study the noncommutative Choquet boundary and Arveson’s hyperrigidity conjecture. We provide an abstract characterization of the Choquet boundary in Theorem 5.3 via the so-called atomic spectrum σa(X)A^\sigma_{a}(X)\subset\widehat{A} (see Definition 5.2). For Type I and abelian C*-algebras, we give a more concrete description of the atomic spectrum, and thus of the Choquet boundary, in Proposition 5.9.

We analyze the hyperrigidity conjecture for correspondences. A whole class of counterexamples to the hyperrigidity conjecture is presented in Proposition 5.4. We also derive a measure-theoretic criterion for a correspondence associated with a topological quiver to be a counterexample in Corollary 6.2. On the other hand, we show that the conjecture holds for proper correspondences (Proposition 5.7) and for topological graphs (Corollary 6.2). This demonstrates that correspondences provide a natural framework for studying the connections between the Choquet boundary and hyperrigidity.

Acknowledgements.

The author is deeply grateful to his PhD advisor, Adam Dor-On, for formulating the original problem and providing guidance throughout this work. Special thanks are also due to Baruch Solel for a productive conversation about W*-correspondences. The author also thanks Orr Shalit for his insightful comments and corrections on the first version of the preprint.

2 C*- and W*-correspondences

Throughout the section, AA, BB will denote C*-algebras and 𝔐\mathfrak{M}, 𝔑\mathfrak{N} will denote von Neumann algebras.

Definition 2.1.

An AA-BB-correspondence is a right Hilbert BB-module XX together with a non-degenerate left AA-action φX:A(X)\varphi_{X}\colon A\to\mathcal{L}(X). The correspondence XX is called proper if φX(A)𝒦(X)\varphi_{X}(A)\subset\mathcal{K}(X) and injective if φX\varphi_{X} is injective. When A=BA=B, we refer to it as an AA-correspondence instead of an AA-AA-correspondence. We write φ\varphi instead of φX\varphi_{X} when XX is clear from the context.

An 𝔐\mathfrak{M}-𝔑\mathfrak{N}-correspondence XX is a W*-correspondence if XX has a Banach predual and the left action φX\varphi_{X} is normal (see [11, Section 8.5] for equivalent definitions). Given von Neumann algebras 𝔐\mathfrak{M}, 𝔑\mathfrak{N}, and a correspondence XX between them, we will always assume that XX is a W*-correspondence.

If XX is an AA-BB-correspondence and YY is a BB-CC-correspondence, then there is an AA-CC-correspondence XBYX\otimes_{B}Y which is a completion of the algebraic tensor product with respect to the tensor inner product defined by

ξη,ξηXBY=η,φY(ξ,ξX)ηY\langle\xi\otimes\eta,\xi^{\prime}\otimes\eta^{\prime}\rangle_{X\otimes_{B}Y}=\langle\eta,\varphi_{Y}(\langle\xi,\xi^{\prime}\rangle_{X})\eta^{\prime}\rangle_{Y}

for all ξ,ξX\xi,\xi^{\prime}\in X and η,ηY\eta,\eta^{\prime}\in Y. In particular, if HH is a BB-representation, then we can view it as a BB-\mathbb{C}-correspondence, so that XBHX\otimes_{B}H is an AA-representation. The tensor product construction is defined analogously for W*-correspondences. If X is a 𝔐\mathfrak{M}-𝔑\mathfrak{N}-W*-correspondence, then X𝔑HX\otimes_{\mathfrak{N}}H is a normal representation of 𝔐\mathfrak{M} if HH is a normal representation of 𝔑\mathfrak{N}.

Definition 2.2.

An (isometric) representation (H,ρH,t)(H,\rho_{H},t) of an AA-correspondence (resp. W*-correspondence over 𝔐\mathfrak{M}) XX on a Hilbert space HH consists of a non-degenerate AA-representation (resp. normal 𝔐\mathfrak{M}-representation) ρH:A(H)\rho_{H}\colon A\to\mathcal{B}(H) and a linear map t:X(H)t\colon X\to\mathcal{B}(H) such that

  • ρH(a)t(ξ)ρH(a)=t(φX(a)ξa)\rho_{H}(a)t(\xi)\rho_{H}(a^{\prime})=t(\varphi_{X}(a)\xi\cdot a^{\prime}),

  • t(ξ)t(η)=ρH(ξ,η)t(\xi)^{*}t(\eta)=\rho_{H}(\langle\xi,\eta\rangle)

for all a,aAa,a^{\prime}\in A and ξ,ηX\xi,\eta\in X. We often suppress the map ρH\rho_{H} from the notation and refer to XX-representations as pairs (H,t)(H,t).

Lemma 2.3.

Let (H,t)(H,t) be a representation of a correspondence XX. The map tt is automatically completely contractive. If XX is a 𝔐\mathfrak{M}-W*-correspondence, then tt is also w*-continuous.

Proof.

We have

t(ξ)2=t(ξ)t(ξ)=ρ(ξ,ξ)ξ,ξ=ξ2{\|t(\xi)\|}^{2}=\|t(\xi)^{*}t(\xi)\|=\|\rho(\langle\xi,\xi\rangle)\|\leq\|\langle\xi,\xi\rangle\|={\|\xi\|}^{2}

and the same computation works for matrices with entries in XX, so tt is completely contractive.

We now show the w*-continuity. The w*-closure of t(X)t(X) in (H)\mathcal{B}(H) is a Hilbert W*-module over ρ(𝔐)\rho(\mathfrak{M}). By viewing ρ(𝔐)\rho(\mathfrak{M}) as an ideal in 𝔐\mathfrak{M}, we can regard t(X)¯w\overline{t(X)}^{w*} as a Hilbert 𝔐\mathfrak{M}-module. Then, the bounded Hilbert 𝔐\mathfrak{M}-module map t:Xt(X)¯wt\colon X\to\overline{t(X)}^{w*} is w*-continuous by [11, Corollary 8.5.8]. ∎

For a C*-correspondence, the Cuntz-Toeplitz algebra 𝒯X\mathcal{T}_{X} was defined by Pimsner [43] as the universal C*-algebra generated by images of AA and XX in representations of XX. There are a *-homomorphism 𝝆:A𝒯X\boldsymbol{\rho}\colon A\to\mathcal{T}_{X} and a linear map 𝒕:X𝒯X\boldsymbol{t}\colon X\to\mathcal{T}_{X} with the following universal property. For all representations (H,t)(H,t) of XX, there is a unique representation ρHt:𝒯X(H)\rho_{H}\rtimes t\colon\mathcal{T}_{X}\to\mathcal{B}(H) such that ρH=ρHt𝝆\rho_{H}=\rho_{H}\rtimes t\circ\boldsymbol{\rho} and t=ρHt𝒕t=\rho_{H}\rtimes t\circ\boldsymbol{t}. We usually identify AA and XX with their images via 𝝆\boldsymbol{\rho} and 𝒕\boldsymbol{t}. We also identify 𝒦(X)\mathcal{K}(X) with its copy inside 𝒯X\mathcal{T}_{X} given by 𝒕(X)𝒕(X)¯\overline{\boldsymbol{t}(X)\boldsymbol{t}(X)^{*}}.

There is a gauge action γ\gamma of the unit circle 𝕋={z:|z|=1}\mathbb{T}=\{z\in\mathbb{C}\colon|z|=1\} on 𝒯X\mathcal{T}_{X}. The action leaves AA invariant and acts on XX by γz(ξ)=zξ\gamma_{z}(\xi)=z\xi for all ξX\xi\in X and z𝕋z\in\mathbb{T}

Let IφX1(𝒦(X))AI\subset\varphi_{X}^{-1}(\mathcal{K}(X))\subset A be an ideal. Consider the ideal 𝒥I𝒯X\mathcal{J}_{I}\subset\mathcal{T}_{X} generated by elements φX(a)𝝆(a)\varphi_{X}(a)-\boldsymbol{\rho}(a) for aIa\in I. The quotient 𝒪X,I=𝒯X/𝒥I\mathcal{O}_{X,I}=\mathcal{T}_{X}/\mathcal{J}_{I} is called the II-relative Cuntz-Pimsner algebra. An XX-representation (H,t)(H,t) for which ρHt\rho_{H}\rtimes t factors through 𝒪X,I\mathcal{O}_{X,I} is called II-relative Cuntz-Pimsner covariant.

The ideal X=φX1(𝒦(X))(kerφX)\mathcal{I}_{X}=\varphi_{X}^{-1}(\mathcal{K}(X))\cap{(\ker\varphi_{X})}^{\perp} is called Katsura’s ideal. Katsura [32] showed that the ideal 𝒥X\mathcal{J}_{\mathcal{I}_{X}} is the largest gauge-invariant ideal of 𝒯X\mathcal{T}_{X} trivially intersecting 𝝆(A)\boldsymbol{\rho}(A) and the algebra 𝒪X=𝒪X,X\mathcal{O}_{X}=\mathcal{O}_{X,\mathcal{I}_{X}} is called just the Cuntz-Pimsner algebra of XX. Consequently, XX-representations for which ρHt\rho_{H}\rtimes t factors through 𝒪X\mathcal{O}_{X} are called Cuntz-Pimsner covariant.

A representation (H,t)(H,t) of XX induces an AA-linear isometry t:XAHHt_{*}\colon X\otimes_{A}H\to H by the formula t(ξv)=t(ξ)vt_{*}(\xi\otimes v)=t(\xi)v for all ξX\xi\in X and vHv\in H. Conversely, given an AA-linear isometry t:XAHHt_{*}\colon X\otimes_{A}H\to H, we can define a representation t:X(H)t\colon X\to\mathcal{B}(H) by the formula t(ξ)v=t(ξv)t(\xi)v=t(\xi\otimes v) for all ξX\xi\in X and vHv\in H.

Proposition 2.4 ([38, Lemma 2.5]).

Let ρ:A(H)\rho\colon A\to\mathcal{B}(H) be a representation. The assignments ttt\leftrightarrow t_{*} above define mutually inverse bijections between the sets of XX-representations t:X(H)t\colon X\to\mathcal{B}(H) and AA-linear isometries t:XAHHt_{*}\colon X\otimes_{A}H\to H. The same applies to a W*-correspondence XX over 𝔐\mathfrak{M} and a normal representation ρ:𝔐(H)\rho\colon\mathfrak{M}\to\mathcal{B}(H).

Construction 2.5.

Let MM be an AA-representation. Consider the XX-Fock Hilbert space MX=k0XkAMM^{X}=\bigoplus_{k\geq 0}X^{\otimes k}\otimes_{A}M where we set X0=AX^{\otimes 0}=A. Observe that XAMX=k1XkAMX\otimes_{A}M^{X}=\bigoplus_{k\geq 1}X^{\otimes k}\otimes_{A}M as an AA-representation and denote by (sM):XAMXMX(s_{M})_{*}\colon X\otimes_{A}M^{X}\to M^{X} the tautological inclusion isometry. The isometry defines an XX-representation (MX,sM)(M^{X},s_{M}) by Proposition 2.4. A representation which is unitarily equivalent to one of the form (MX,sM)(M^{X},s_{M}) is called induced.

This construction applies also when XX is a W*-correspondence over a von Neumann algebra 𝔐\mathfrak{M} and MM is a normal 𝔐\mathfrak{M}-representation.

Definition 2.6.

An (isometric) dilation of an XX-representation (H,t)(H,t) is a triple (K,s,V)(K,s,V), where (K,s)(K,s) is an XX-representation and V:HKV\colon H\to K is an AA-linear isometry such that

Vs(ξ)V=t(ξ)V^{*}s(\xi)V=t(\xi)

for all ξX\xi\in X. The dilation (K,s,V)(K,s,V) is called trivial if VHKVH\subset K is XX-reducing which means that VHVH and (VH)(VH)^{\perp} are invariant under s(ξ)s(\xi) for all ξX\xi\in X.

The representation (H,t)(H,t) is called maximal if it does not admit non-trivial dilations. The correspondence XX is called hyperrigid if all its Cuntz-Pimsner covariant representations are maximal.

The definitions above also apply to the case when XX is a W*-correspondence without any modifications.

Remark.

In this work, we assume all representations and dilations to be isometric. This assumption does not result in any loss of generality, as Muhly and Solel proved that every completely contractive representation admits an isometric dilation [39]. Therefore, the notion of maximality remains the same when considering only isometric dilations.

Lemma 2.7.

If (K,s,V)(K,s,V) is a dilation of (H,t)(H,t), then

s(ξ)V=Vt(ξ)s(\xi)V=Vt(\xi)

for all ξX\xi\in X.

Proof.

We have s(ξ)V=VVs(ξ)V+(1VV)s(ξ)V=Vt(ξ)+(1VV)s(ξ)Vs(\xi)V=VV^{*}s(\xi)V+(1-VV^{*})s(\xi)V=Vt(\xi)+(1-VV^{*})s(\xi)V so it is enough to show that T=(1VV)s(ξ)V=0T=(1-VV^{*})s(\xi)V=0. We compute

TT=Vs(ξ)(1VV)s(ξ)V=VρK(ξ,ξ)VρH(ξ,ξ)=0,T^{*}T=V^{*}s(\xi)^{*}(1-VV^{*})s(\xi)V=V^{*}\rho_{K}(\langle\xi,\xi\rangle)V-\rho_{H}(\langle\xi,\xi\rangle)=0,

where the last equality follows by the AA-linearity of VV. ∎

Proposition 2.8.

Let XX be a C*-correspondence over AA. Consider the operator algebra 𝒯X+𝒯X\mathcal{T}_{X}^{+}\subset\mathcal{T}_{X} generated by A,X𝒯XA,X\subset\mathcal{T}_{X} and the operator system 𝒮X=X+A+X𝒯X\mathcal{S}_{X}=X^{*}+A+X\subset\mathcal{T}_{X}. The following are equivalent for a representation (H,t)(H,t) of XX:

  1. (1)

    (H,t)(H,t) is maximal;

  2. (2)

    ρHt|𝒮X\rho_{H}\rtimes t|_{\mathcal{S}_{X}} satisfies the UEP;

  3. (3)

    ρHt|𝒯X++(𝒯X+)\rho_{H}\rtimes t|_{\mathcal{T}_{X}^{+}+(\mathcal{T}_{X}^{+})^{*}} satisfies the UEP.

Proof.

If AA is non-unital, then so is 𝒯X\mathcal{T}_{X} and by [46, Proposition 2.4] π(H,t)\pi_{(H,t)} has the UEP with respect to 𝒯X+\mathcal{T}_{X}^{+} or 𝒮X\mathcal{S}_{X} if and only if its C*-unitization has the UEP. Therefore, we may assume that AA and 𝒯X\mathcal{T}_{X} are unital. By [4, Proposition 2.2], a representation has the UEP if and only if it is maximal.

Suppose that (1) holds and (H,t)(H,t) is maximal. Since A,X𝒮XA,X\subset\mathcal{S}_{X} and A,X𝒯X+A,X\subset\mathcal{T}_{X}^{+}, then any 𝒯X+\mathcal{T}_{X}^{+}-dilation or 𝒮X\mathcal{S}_{X}-dilation of (H,t)(H,t) is also an XX-dilation and thus trivial. Therefore, (1) implies (3) and (2).

Analogously, 𝒮X𝒯X++(𝒯X+)\mathcal{S}_{X}\subset\mathcal{T}_{X}^{+}+{(\mathcal{T}_{X}^{+})}^{*} and thus if a representation has UEP with respect to 𝒮X\mathcal{S}_{X}, then it also has the UEP with respect to 𝒯X+\mathcal{T}_{X}^{+}. This proves (2)\Rightarrow(3).

Finally, suppose that (3) holds. Let (K,s,V)(K,s,V) be a dilation of (H,t)(H,t). By an inductive application of Lemma 2.7, we have

Vs(ξ1)s(ξ2)s(ξn)V=t(ξ1)t(ξ2)t(ξn)V^{*}s(\xi_{1})s(\xi_{2})\cdots s(\xi_{n})V=t(\xi_{1})t(\xi_{2})\cdots t(\xi_{n})

for all ξ1,ξ2,,ξnX\xi_{1},\xi_{2},\dots,\xi_{n}\in X. Since 𝒯X+\mathcal{T}_{X}^{+} is linearly spanned by AA and elements of the form ξ1ξn\xi_{1}\cdots\xi_{n}, it follows that (ρKs,V)(\rho_{K}\ltimes s,V) is a 𝒯X+\mathcal{T}_{X}^{+}-dilation of ρHt\rho_{H}\ltimes t and therefore it is trivial. We conclude that (3) implies (1). ∎

3 Second dual correspondence

For a C*-algebra AA, its second dual A{A}^{**} is naturally endowed with the structure of a von Neumann algebra (see [10, III.5.2.8]). It is universal in the following sense, if ψ:A𝔐\psi\colon A\to\mathfrak{M} is a *-homomorphism into a von Neumann algebra 𝔐\mathfrak{M}, then ψ\psi extends uniquely to a normal homomorphism ψ~:A𝔐\tilde{\psi}\colon{A}^{**}\to\mathfrak{M}. For example, if ρ:A(H)\rho\colon A\to\mathcal{B}(H) is a representation, then ρ~:A(H)\tilde{\rho}\colon{A}^{**}\to\mathcal{B}(H) is a normal representation. Therefore, there is a one-to-one correspondence between representations of AA and normal representations of A{A}^{**}.

The multiplier algebra (A)\mathcal{M}(A) can be naturally embedded into A{A}^{**} (see [2]). Consequently, if BB is another C*-algebra and ϕ:A(B)B\phi\colon A\to\mathcal{M}(B)\subset{B}^{**} is a *-homomorphism, then there is a unique normal homomorphism ϕ~:AB\tilde{\phi}\colon{A}^{**}\to{B}^{**}, which is unital if and only if ϕ(A)B=B\phi(A)B=B.

If ψ:𝔐𝔑\psi\colon\mathfrak{M}\to\mathfrak{N} is a normal *-homomorphism, then there is a unique central projection PψZ(𝔐)P_{\psi}\in Z(\mathfrak{M}) called the support projection of ψ\psi such that kerψ=𝔐(1Pψ)\ker\psi=\mathfrak{M}(1-P_{\psi}) and ψ(𝔐)𝔐Pψ\psi(\mathfrak{M})\cong\mathfrak{M}P_{\psi}. For a *-homomorphism ϕ:A(B)\phi\colon A\to\mathcal{M}(B), we define the support projection of ϕ\phi to be Pϕ=Pϕ~AP_{\phi}=P_{\tilde{\phi}}\in{A}^{**}. For an ideal ιI:IB\iota_{I}\colon I\hookrightarrow B, we use the notation PI=ι~I(1)BP_{I}=\tilde{\iota}_{I}(1)\in{B}^{**}. It is the support projection of the associated map B(I)B\to\mathcal{M}(I).

If XX is an AA-BB-correspondence, then by [11, Proposition 8.5.17], X{X}^{**} is a WW^{*}-module over B{B}^{**} such that (X)=𝒦(X)\mathcal{L}({X}^{**})=\mathcal{K}(X)^{**}. The map φX:A(X)=(𝒦(X))\varphi_{X}\colon A\to\mathcal{L}(X)=\mathcal{M}(\mathcal{K}(X)) lifts uniquely to the normal map φX=φ~X:A(X)\varphi_{{X}^{**}}=\tilde{\varphi}_{X}\colon{A}^{**}\to\mathcal{L}({X}^{**}) which endows X{X}^{**} with a structure of a W*-correspondence. We define the support projection of XX to be PX=PφXAP_{X}=P_{\varphi_{X}}\in{A}^{**}.

Lemma 3.1.

Let XX be an AA-BB-correspondence and II be an ideal. Then, (φ(I)X)=φ~(PI)XX{(\varphi(I)X)}^{**}=\tilde{\varphi}(P_{I}){X}^{**}\subset{X}^{**}. In particular, if HH is an AA-representation, then ρH(I)H=ρ~H(PI)H\rho_{H}(I)H=\tilde{\rho}_{H}(P_{I})H.

Proof.

The subspace (φ(I)X){(\varphi(I)X)}^{**} is a w*-closure of φ(I)X\varphi(I)X in X{X}^{**}. Since PII=IP_{I}I=I in A{A}^{**}, it follows that (φ(I)X)φ~(PI)X{(\varphi(I)X)}^{**}\subset\tilde{\varphi}(P_{I}){X}^{**}. On the other hand, PIP_{I} is in the w*-closure of II in A{A}^{**}, which implies (φ(I)X)φ~(PI)X{(\varphi(I)X)}^{**}\subset\tilde{\varphi}(P_{I}){X}^{**}. This proves the first claim. The second claim follows from the first one by considering HH as an AA-\mathbb{C}-correspondence and from the fact that Hilbert spaces are reflexive. ∎

Thanks to the universal property of the second dual, there is a bijection between *-representations ρ:AB(H)\rho\colon A\to B(H) and normal *-representations ρ~:A(H)\tilde{\rho}\colon{A}^{**}\to\mathcal{B}(H). This can be extended to correspondences as follows. Since XX is w*-dense in X{X}^{**}, every linear map t:X(H)t\colon X\to\mathcal{B}(H) extends uniquely to a w*-continuous map t~:X(H)\tilde{t}\colon{X}^{**}\to\mathcal{B}(H). Because the left and right actions of A{A}^{**} on X{X}^{**} and the inner product are w*-continuous, the maps ρ~H\tilde{\rho}_{H} and t~\tilde{t} define a representation of X{X}^{**} on HH. On the other hand, we can recover ρH\rho_{H} and tt by restricting to w*-dense subspaces AA and XX. Thus, we have proved the following.

Lemma 3.2.

The assignments (H,ρH,t)(H,ρ~h,t~)(H,\rho_{H},t)\leftrightarrow(H,\tilde{\rho}_{h},\tilde{t}) define a bijection between the sets of XX-representations and XX^{**}-representations on HH.

The following result reduces the study of dilations for C*-correspondences to W*-correspondences. This is the key result to prove Theorem 4.4, which is the main theorem.

Proposition 3.3.

A triple (K,s,V)(K,s,V) is a dilation of (H,t)(H,t) if and only if (K,s~,V)(K,\tilde{s},V) is a dilation of (H,t~)(H,\tilde{t}). In particular, (H,t)(H,t) is maximal if and only if (H,t~)(H,\tilde{t}) is maximal.

Proof.

By construction, we have s~(ξ)=s(ξ)\tilde{s}(\xi)=s(\xi) and t~(ξ)=t(ξ)\tilde{t}(\xi)=t(\xi) for all ξX\xi\in X. Therefore, if (K,s~,V)(K,\tilde{s},V) is a dilation of (H,t~)(H,\tilde{t}), then (K,s,V)(K,s,V) is a dilation of (H,t)(H,t).

On the other hand, suppose that (K,s,V)(K,s,V) is a dilation of (H,t)(H,t). The multiplication of bounded operators is separately w*-continuous. Therefore, if ξX\xi\in{X}^{**} is an arbitrary element and {ξλ}λΛ\{\xi_{\lambda}\}_{\lambda\in\Lambda} is a net in XX which converges to ξ\xi in the w*-topology, then

Vs~(ξ)V=wlimλΛVs(ξλ)V=wlimλΛt(ξλ)=t~(ξ).V^{*}\tilde{s}(\xi)V=\operatorname*{w*-lim}_{\lambda\in\Lambda}V^{*}s(\xi_{\lambda})V=\operatorname*{w*-lim}_{\lambda\in\Lambda}t(\xi_{\lambda})=\tilde{t}(\xi).

It follows that (K,s~,V)(K,\tilde{s},V) is a dilation of (H,t~)(H,\tilde{t}).

We conclude that dilations of (H,t)(H,t) are in bijection with dilations of (H,t~)(H,\tilde{t}). In particular, (H,t)(H,t) is maximal if and only if (H,t~)(H,\tilde{t}) is maximal. ∎

4 Maximality criteria

In this section we develop the general theory of dilations for correspondences. We first deal with W*-correspondences and then proceed to the case of C*-correspondences using the second dual technique.

Consider a W*-correspondence XX over a W*-algebra 𝔐\mathfrak{M}. We assume all representations of 𝔐\mathfrak{M} to be normal. Let (H,t)(H,t) be a representation of a correspondence XX. Denote H0=Ht(X)H¯H_{0}=H\ominus\overline{t(X)H} which is an 𝔐\mathfrak{M}-reducing subspace. The map t:X𝔐Ht(X)H¯H0=Ht_{*}\colon X\otimes_{\mathfrak{M}}H\to\overline{t(X)H}\oplus H_{0}=H can be now written as a column block-matrix

t=[Qt0]:X𝔐Ht(X)H¯H0,t_{*}=\begin{bmatrix}Qt_{*}\\ 0\end{bmatrix}\colon X\otimes_{\mathfrak{M}}H\to\overline{t(X)H}\oplus H_{0},

where QQ is the orthogonal projection Ht(X)H¯H\twoheadrightarrow\overline{t(X)H}. Dilations of the representation (H,t)(H,t) correspond to dilations of the map tt_{*} and the above matrix form helps to classify those dilations.

Construction 4.1.

Consider an 𝔐\mathfrak{M}-representation LL together with a pair of bounded 𝔐\mathfrak{M}-linear operators s0:X𝔐LH0s_{0}\colon X\otimes_{\mathfrak{M}}L\to H_{0} and s1:X𝔐LLs_{1}\colon X\otimes_{\mathfrak{M}}L\to L satisfying s0s0+s1s1=1X𝔐Ls_{0}^{*}s_{0}+s_{1}^{*}s_{1}=1_{X\otimes_{\mathfrak{M}}L}. For K=HLK=H\oplus L, we can form an 𝔐\mathfrak{M}-linear map s:X𝔐K=X𝔐HX𝔐LK=t(X)H¯H0Ls_{*}\colon X\otimes_{\mathfrak{M}}K=X\otimes_{\mathfrak{M}}H\oplus X\otimes_{\mathfrak{M}}L\to K=\overline{t(X)H}\oplus H_{0}\oplus L by the block matrix

s=[Qt00s00s1]:(X𝔐H)(X𝔐L)t(X)H¯H0H.s_{*}=\begin{bmatrix}Qt_{*}&0\\ 0&s_{0}\\ 0&s_{1}\end{bmatrix}\colon(X\otimes_{\mathfrak{M}}H)\oplus(X\otimes_{\mathfrak{M}}L)\to\overline{t(X)H}\oplus H_{0}\oplus H.

Due to the relation on s0s_{0} and s1s_{1}, the map ss_{*} is an isometry and thus defines an isometric XX-representation (K,s)(K,s). Together with the tautological inclusion V:HKV\colon H\hookrightarrow K, the triple (K,s,V)(K,s,V) is a dilation of (H,t)(H,t), which is trivial if and only if s0=0s_{0}=0.

Lemma 4.2.

Every dilation of (H,t)(H,t) is equivalent to (K,s,V)(K,s,V) from Construction 4.1 for some L,s0,s1L,s_{0},s_{1}.

Proof.

Let (K,s,V)(K,s,V) be a dilation of (H,t)(H,t). Identify HH with VHVH and set L=KHL=K\ominus H. Since it is a dilation, the map ss_{*} has the block form

s=[Qts20s0s3s1]:(X𝔐H)(X𝔐L)t(X)H¯H0L,s_{*}=\begin{bmatrix}Qt_{*}&s_{2}\\ 0&s_{0}\\ s_{3}&s_{1}\end{bmatrix}\colon(X\otimes_{\mathfrak{M}}H)\oplus(X\otimes_{\mathfrak{M}}L)\to\overline{t(X)H}\oplus H_{0}\oplus L,

where sis_{i} are 𝔐\mathfrak{M}-linear operators between corresponding Hilbert spaces.

To prove the lemma, we must show that s2=0s_{2}=0, s3=0s_{3}=0, and s0s0+s1s1=1X𝔐Ls_{0}^{*}s_{0}+s_{1}^{*}s_{1}=1_{X\otimes_{\mathfrak{M}}L}. Everything follows from the condition that ss_{*} is an isometry. Indeed, we have

[1X𝔐H001X𝔐L]=ss=[tQQt+s3s3tQs2+s3s1s2Qt+s1s3s0s0+s1s1+s2s2].\begin{bmatrix}1_{X\otimes_{\mathfrak{M}}H}&0\\ 0&1_{X\otimes_{\mathfrak{M}}L}\end{bmatrix}=s_{*}^{*}s_{*}=\begin{bmatrix}t_{*}^{*}Q^{*}Qt_{*}+s_{3}^{*}s_{3}&t_{*}^{*}Q^{*}s_{2}+s_{3}^{*}s_{1}\\ s_{2}^{*}Qt_{*}+s_{1}^{*}s_{3}&s_{0}^{*}s_{0}+s_{1}^{*}s_{1}+s_{2}^{*}s_{2}\end{bmatrix}.

Since QtQt_{*} is a unitary X𝔐Ht(X)H¯X\otimes_{\mathfrak{M}}H\to\overline{t(X)H}, we have s3s3=1X𝔐HtQQt=0s_{3}^{*}s_{3}=1_{X\otimes_{\mathfrak{M}}H}-t^{*}_{*}Q^{*}Qt_{*}=0 and thus s3=0s_{3}=0. Consequently, in the lower left corner we have 0=s2Qt0=s_{2}^{*}Qt_{*}, which implies s2=0s_{2}=0. Finally, the condition on (2,2)-entry is s0s0+s1s1=1X𝔐Ls_{0}^{*}s_{0}+s_{1}^{*}s_{1}=1_{X\otimes_{\mathfrak{M}}L} and the dilation (K,t,V)(K,t,V) is of the form of Construction 4.1 for the data L,s0,s1L,s_{0},s_{1}. ∎

Recall that for a W*-correspondence XX over 𝔐\mathfrak{M} we define the support projection PXZ(𝔐)P_{X}\in Z(\mathfrak{M}) as the unique central projection satisfying kerφX=𝔐(1PX)\ker\varphi_{X}=\mathfrak{M}(1-P_{X}).

Definition 4.3.

Let XX be a W*-correspondence over 𝔐\mathfrak{M} and let PZ(𝔐)P\in Z(\mathfrak{M}) be a central projection. A representation (H,t)(H,t) is called PP-relative Cuntz-Pimsner covariant if ρH(P)Ht(X)H¯\rho_{H}(P)H\subset\overline{t(X)H}. A PXP_{X}-relative Cuntz-Pimsner covariant representation is called fully Cuntz-Pimsner covariant.

If XX is a C*-correspondence over AA, then a representation (H,t)(H,t) is called PP-relative Cuntz-Pimsner covariant for PZ(A)P\in Z(A^{**}) or fully Cuntz-Pimsner covariant if the associated representation (H,t~)(H,\tilde{t}) of XX^{**} is respectively PP-relative or fully Cuntz-Pimsner covariant.

Since φ(PX)X=X\varphi(P_{X})X=X for a W*-correspondence XX, we have ρ(PX)t(X)=t(X)\rho(P_{X})t(X)=t(X) for any representation (H,t)(H,t) and thus t(X)H¯ρ(PX)H\overline{t(X)H}\subset\rho(P_{X})H. Therefore, the XX-representation is fully Cuntz-Pimsner covariant if and only if t(X)H¯=ρ(PX)H\overline{t(X)H}=\rho(P_{X})H.

If XX is a C*-correspondence and (H,t)(H,t) is an XX-representation, then by w*-continuity t(X)H¯=t~(X)H¯\overline{t(X)H}=\overline{\tilde{t}({X}^{**})H}. In particular, the representation is fully Cuntz-Pimsner covariant if and only if t(X)H¯=ρ~(PX)H\overline{t(X)H}=\tilde{\rho}(P_{X})H.

Theorem 4.4.

A representation of a C*- or W*-correspondence is maximal if and only if it is fully Cuntz-Pimsner covariant.

Proof.

By Proposition 3.3, it is enough to prove the theorem for a W*-correspondence XX over 𝔐\mathfrak{M}. Let H0=Ht(X)H¯H_{0}=H\ominus\overline{t(X)H}, which is an 𝔐\mathfrak{M}-reducing subspace. Then, the representation is fully Cuntz-Pimsner covariant if and only if ρ(PX)H0=0\rho(P_{X})H_{0}=0.

Suppose that (H,t)(H,t) is not maximal. Then, by Lemma 4.2, there exists an 𝔐\mathfrak{M}-representation LL together with 𝔐\mathfrak{M}-linear maps s0s_{0} and s1s_{1} such that s0:X𝔐LH0s_{0}\colon X\otimes_{\mathfrak{M}}L\to H_{0} is nonzero. The projection PXP_{X} acts by identity on XX and thus on X𝔐LX\otimes_{\mathfrak{M}}L. Therefore, since s0s_{0} is 𝔐\mathfrak{M}-linear, the projection PXP_{X} also acts by identity on Ims0¯H0\overline{\operatorname{Im}s_{0}}\subset H_{0}. Consequently, ρ(PX)H00\rho(P_{X})H_{0}\neq 0 and the representation is not fully Cuntz-Pimsner covariant.

Suppose now that ρ(PX)H00\rho(P_{X})H_{0}\neq 0. Let ρM:𝔐(M)\rho_{M}\colon\mathfrak{M}\hookrightarrow\mathcal{B}(M) be a faithful normal representation on a Hilbert space MM. Then, by the Morita-Rieffel theory for von Neumann algebras (see [11, 8.5.12]), X𝔐MX\otimes_{\mathfrak{M}}M is a faithful representation of (X)\mathcal{L}(X) and thus of 𝔐PXφ(𝔐)\mathfrak{M}P_{X}\cong\varphi(\mathfrak{M}). Since ρ(PX)H0\rho(P_{X})H_{0} is a non-trivial representation of 𝔐PX\mathfrak{M}P_{X}, from the representation theory of von Neumann algebras it follows that ρ(PX)H0\rho(P_{X})H_{0} is quasi-equivalent to a subrepresentation of X𝔐MX\otimes_{\mathfrak{M}}M. Therefore, there exists a non-zero 𝔐\mathfrak{M}-linear contraction s0:X𝔐MH0s_{0}^{\prime}\colon X\otimes_{\mathfrak{M}}M\to H_{0}.

Let L=MXL=M^{X} be the XX-Fock space of MM and (sM):X𝔐LL{(s_{M})}_{*}\colon X\otimes_{\mathfrak{M}}L\to L be the corresponding 𝔐\mathfrak{M}-linear isometry (see Construction 2.5). Define s0:X𝔐LH0s_{0}\colon X\otimes_{\mathfrak{M}}L\to H_{0} to be equal to s0s_{0}^{\prime} on X𝔐MX𝔐LX\otimes_{\mathfrak{M}}M\subset X\otimes_{\mathfrak{M}}L and zero on the orthogonal complement. Also, set s1=(sM)1s0s0:X𝔐LLs_{1}={(s_{M})}_{*}\sqrt{1-s_{0}^{*}s_{0}}\colon X\otimes_{\mathfrak{M}}L\to L. We have s0s0+s1s1=1X𝔐Ls_{0}^{*}s_{0}+s_{1}^{*}s_{1}=1_{X\otimes_{\mathfrak{M}}L} and Construction 4.1 gives a dilation (K,s,V)(K,s,V) with K=HLK=H\oplus L and

s=[Qt00s00s1]:(X𝔐H)(X𝔐L)t(X)H¯H0L.s_{*}=\begin{bmatrix}Qt_{*}&0\\ 0&s_{0}\\ 0&s_{1}\end{bmatrix}\colon(X\otimes_{\mathfrak{M}}H)\oplus(X\otimes_{\mathfrak{M}}L)\to\overline{t(X)H}\oplus H_{0}\oplus L.

The dilation is non-trivial since s0s_{0} is non-zero by construction. We conclude that full Cuntz-Pimsner covariance is equivalent to maximality. ∎

Example 4.5.

We analyze topological graphs and quivers in Section 6. However, we already have enough tools to describe maximality for discrete graphs.

Let (E0,E1,r,s)(E^{0},E^{1},r,s) be a countable discrete graph and X=X(E)X=X(E) be its graph correspondence over c0(E0)c_{0}(E^{0}) (see [45, Chapter 8]). Representations (H,t)(H,t) of XX are in bijection with Cuntz-Krieger families {pv,se}vE0,eE1\{p_{v},s_{e}\}_{v\in E^{0},e\in E^{1}}. Moreover, we have t(X)H¯=eE1seseH\overline{t(X)H}=\bigoplus_{e\in E^{1}}s_{e}s_{e}^{*}H.

We have c0(E0)=(E0){c_{0}(E^{0})}^{**}=\ell^{\infty}(E^{0}) and kerφ~X=(E0r(E0))\ker\tilde{\varphi}_{X}=\ell^{\infty}(E^{0}\setminus r(E^{0})). Therefore, the support projection PXP_{X} equals to the sum of projections corresponding to non-sink vertices r(E0)r(E^{0}). Consequently, the representation is fully Cuntz-Pimsner covariant if and only if eE1sese=vr(E0)pv\sum_{e\in E^{1}}s_{e}s_{e}^{*}=\sum_{v\in r(E^{0})}p_{v}, where the sum is taken in the w*-topology. This coincides with the notion of full Cuntz-Krieger families by Dor-On and Salomon [23]. Therefore, Theorem 4.4 is a generalization of [23, Theorem 3.5].

Corollary 4.6.

Let MM be a representation of AA. The induced representation (MX,sM)(M^{X},s_{M}) is maximal if and only if ρ~M(PX)=0\tilde{\rho}_{M}(P_{X})=0.

Proof.

By construction, M=MXsM(X)MX¯M=M^{X}\ominus\overline{s_{M}(X)M^{X}} as an AA-representation. Therefore, ρ~MX(PX)MXsM(X)MX¯\tilde{\rho}_{M^{X}}(P_{X})M^{X}\subset\overline{s_{M}(X)M^{X}} if and only if ρ~M(PX)=0\tilde{\rho}_{M}(P_{X})=0. ∎

Lemma 4.7.

A representation (H,t)(H,t) of a C*-correspondence XX is PIP_{I}-relative Cuntz-Pimsner covariant if and only if it is II-relative covariant in the sense of Katsura. In particular, the representation is Cuntz-Pimsner covariant if and only if it is PXP_{\mathcal{I}_{X}}-relative Cuntz-Pimsner covariant, where X\mathcal{I}_{X} is Katsura’s ideal.

Proof.

If HH is an AA-representation, then we have ρ(X)H=ρ~(PX)H\rho(\mathcal{I}_{X})H=\tilde{\rho}(P_{\mathcal{I}_{X}})H by Lemma 3.1. The statement follows from [36, Proposition 2.16] for B=(H)B=\mathcal{B}(H). ∎

Let 𝒜\mathcal{A} be a non-self-adjoint subalgebra of a C*-algebra BB. In [7], Arveson defined the Shilov ideal JJ of 𝒜\mathcal{A} to be the largest ideal of BB such that the quotient BB/JB\twoheadrightarrow B/J is completely isometric on 𝒜\mathcal{A}. The quotient Cenv(𝒜)B/JC^{*}_{env}(\mathcal{A})\coloneqq B/J is called the C*-envelope of 𝒜\mathcal{A}. The existence of the Shilov ideal and of the C*-envelope was established by Hamana [26].

Corollary 4.8 ([29, Theorem 3.7]).

Maximal representations of a C*-correspondence XX are Cuntz-Pimsner covariant in the sense of Katsura. In particular, Cenv(𝒯X+)=𝒪XC^{*}_{env}(\mathcal{T}_{X}^{+})=\mathcal{O}_{X}.

Proof.

The Cuntz-Pimsner covariance is the same as the PXP_{\mathcal{I}_{X}}-relative Cuntz-Pimsner covariance by Lemma 4.7. The restriction φX|X:X𝒦(X)\varphi_{X}|_{\mathcal{I}_{X}}\colon\mathcal{I}_{X}\to\mathcal{K}(X) is injective and thus φ~X|X:X=APX(X)\tilde{\varphi}_{X}|_{{\mathcal{I}_{X}}^{**}}\colon{\mathcal{I}_{X}}^{**}={A}^{**}P_{\mathcal{I}_{X}}\to\mathcal{L}({X}^{**}) is also injective. Therefore, we have PXPXP_{\mathcal{I}_{X}}\leq P_{X} and, hence, fully Cuntz-Pimsner covariant representations are Cuntz-Pimsner covariant. By the proof of [24, Theorem 4.1], there is a completely isometric maximal representation π\pi of 𝒯X+\mathcal{T}_{X}^{+} such that the C*-algebra generated by its image is the C*-envelope Cenv(𝒯X+)C^{*}_{env}(\mathcal{T}_{X}^{+}) (see also [4, Section 3]). Since maximal representations are Cuntz-Pimsner covariant, the representation π\pi factors through 𝒪X\mathcal{O}_{X} and thus Cenv(𝒯X+)C^{*}_{env}(\mathcal{T}_{X}^{+}) is a quotient of 𝒪X\mathcal{O}_{X}.

Moreover, if J𝒪XJ\subset\mathcal{O}_{X} is the Shilov ideal and z𝕋z\in\mathbb{T}, then the quotient 𝒪X/γz(J)\mathcal{O}_{X}/\gamma_{z}(J) is completely isometric on 𝒯X+\mathcal{T}_{X}^{+} and thus γz(J)J\gamma_{z}(J)\subset J. Therefore, the Shilov ideal is gauge-invariant. However, any non-zero gauge-invariant ideal of 𝒪X\mathcal{O}_{X} intersects non-trivially with A𝒯X+A\subset\mathcal{T}_{X}^{+} by [32, Proposition 7.14], and thus the quotient is not completely isometric on 𝒯X+\mathcal{T}_{X}^{+}. Therefore, 𝒪X\mathcal{O}_{X} is the C*-envelope of 𝒯X+\mathcal{T}_{X}^{+}. ∎

Corollary 4.9 ([34, Theorem 1.2]).

A C*-correspondence XX is hyperrigid if and only if φ(X)X=X\varphi(\mathcal{I}_{X})X=X.

Proof.

We have seen in the proof of Corollary 4.8 that PXPXP_{\mathcal{I}_{X}}\leq P_{X}. By Lemma 3.1, we have φ~(PX)X=(φ(X)X)X\tilde{\varphi}(P_{\mathcal{I}_{X}}){X}^{**}={(\varphi(\mathcal{I}_{X})X)}^{**}\subset{X}^{**}. Therefore, PX=PXP_{\mathcal{I}_{X}}=P_{X} if and only if φ(X)X=X\varphi(\mathcal{I}_{X})X=X. In particular, if the latter holds, then all Cuntz-Pimsner covariant representations are automatically fully Cuntz-Pimsner covariant and XX is hyperrigid.

Conversely, if PX<PXP_{\mathcal{I}_{X}}<P_{X}, then there is an AA-representation MM such that ρ~M(PX)=0\tilde{\rho}_{M}(P_{\mathcal{I}_{X}})=0 but ρ~M(PX)=1\tilde{\rho}_{M}(P_{X})=1. The first condition ensures that the induced representation (MX,sM)(M^{X},s_{M}) is Cuntz-Pimsner covariant and the second condition implies that it is not maximal by Corollary 4.6. Therefore, XX is not hyperrigid if φ(X)XX\varphi(\mathcal{I}_{X})X\neq X. ∎

We say that a C*-correspondence XX is almost proper, if the ideal φ1(𝒦(X))A\varphi^{-1}(\mathcal{K}(X))\subset A acts non-degenerately on XX. For an almost proper correspondence, we give an equivalent characterization of the full Cuntz-Pimsner covariance and thus of maximality without any reference to the second dual.

Proposition 4.10.

A representation (H,t)(H,t) of an almost proper C*-correspondence XX is fully Cuntz-Pimsner covariant if and only if the identity ρ(φ1(𝒦(X)))H=t(X)H¯ρ(kerφ)H\rho\left(\varphi^{-1}(\mathcal{K}(X))\right)H=\overline{t(X)H}\oplus\rho(\ker\varphi)H holds.

Proof.

Let I=φ1(𝒦(X))I=\varphi^{-1}(\mathcal{K}(X)) and J=kerφIJ=\ker\varphi\subset I. The inclusion I/J𝒦(X)I/J\hookrightarrow\mathcal{K}(X) induces the inclusion (I/J)=A(PIPJ)(X){(I/J)}^{**}={A}^{**}(P_{I}-P_{J})\hookrightarrow\mathcal{L}({X}^{**}). Therefore, we have PIPJPXP_{I}-P_{J}\leq P_{X}.

By Lemma 3.1 and the assumption that XX is almost proper, we have φ~(PI)X=X\tilde{\varphi}(P_{I}){X}^{**}={X}^{**} and φ~(PJ)X=0\tilde{\varphi}(P_{J}){X}^{**}=0. Therefore, PIPJP_{I}-P_{J} acts by the identity on XX which implies PXPIPJP_{X}\leq P_{I}-P_{J} and thus PX=PIPJP_{X}=P_{I}-P_{J}. We conclude that a representation (H,t)(H,t) is fully Cuntz-Pimsner covariant if and only if t(X)H¯=ρ~(PIPJ)H=ρ(I)Hρ(J)H\overline{t(X)H}=\tilde{\rho}(P_{I}-P_{J})H=\rho(I)H\ominus\rho(J)H. ∎

5 Boundary representations

Let XX be a C*-correspondence over AA. An irreducible representation of 𝒯X\mathcal{T}_{X} is called boundary if its restriction to the operator system 𝒯X++(𝒯X+)\mathcal{T}_{X}^{+}+(\mathcal{T}_{X}^{+})^{*} satisfies the UEP. The noncommutative Choquet boundary C𝒯X+\partial_{C}\mathcal{T}_{X}^{+} is the set of unitary equivalence classes of boundary representations. Since any boundary representation factors through the C*-envelope, C𝒯X+\partial_{C}\mathcal{T}_{X}^{+} is naturally a subset of 𝒪^X\widehat{\mathcal{O}}_{X}. The goal of this section is to compute CXC𝒯X+\partial_{C}X\coloneqq\partial_{C}\mathcal{T}_{X}^{+}.

We say that a representation (H,t)(H,t) of a C*-correspondence XX is irreducible if the induced representation ρHt\rho_{H}\rtimes t of 𝒯X\mathcal{T}_{X} is irreducible. By Proposition 2.8, any boundary representation of 𝒯X\mathcal{T}_{X} is induced, up to unitary equivalence, by an irreducible maximal representation of XX. Therefore, we also refer to irreducible maximal representations of XX as boundary representations.

A representation (H,t)(H,t) is called fully coisometric if t(X)H¯=H\overline{t(X)H}=H. Fully coisometric representations are automatically fully Cuntz-Pimsner covariant and thus they are maximal. By the Wold decomposition [40] for correspondences, every XX-representation is a direct sum of fully coisometric and induced parts. Therefore, irreducible representations of 𝒯X\mathcal{T}_{X} and 𝒪X\mathcal{O}_{X} are either fully co-isometric or induced: 𝒯^X=(𝒯^X)fc(𝒯^X)ind\widehat{\mathcal{T}}_{X}={(\widehat{\mathcal{T}}_{X})}_{\mathrm{fc}}\sqcup{(\widehat{\mathcal{T}}_{X})}_{\mathrm{ind}} and 𝒪^X=(𝒪^X)fc(𝒪^X)ind\widehat{\mathcal{O}}_{X}={(\widehat{\mathcal{O}}_{X})}_{\mathrm{fc}}\sqcup{(\widehat{\mathcal{O}}_{X})}_{\mathrm{ind}}.

Since fully coisometric representations are Cuntz-Pimsner covariant, we have (𝒯^X)fc=(𝒪^X)fc{(\widehat{\mathcal{T}}_{X})}_{\mathrm{fc}}={(\widehat{\mathcal{O}}_{X})}_{\mathrm{fc}} and we denote this set by X^fc\widehat{X}_{\mathrm{fc}}. By the maximality of fully coisometric representations, we have X^fcCX\widehat{X}_{\mathrm{fc}}\subset\partial_{C}X. To describe CX\partial_{C}X we thus only need to classify the induced boundary representations.

Lemma 5.1.

An induced representation (MX,sM)(M^{X},s_{M}) is irreducible if and only if MM is an irreducible representation of AA. Therefore, there is a bijection ind:A^(𝒯^X)ind\operatorname{ind}\colon\widehat{A}\to{(\widehat{\mathcal{T}}_{X})}_{\mathrm{ind}} such that indA/X^=(𝒪^X)ind\operatorname{ind}\widehat{A/\mathcal{I}_{X}}={(\widehat{\mathcal{O}}_{X})}_{\mathrm{ind}}.

Proof.

Suppose that M=MM′′M=M^{\prime}\oplus M^{\prime\prime} is not irreducible. Then, (M)XMX(M^{\prime})^{X}\subset M^{X} is a proper 𝒯X\mathcal{T}_{X}-reducing subspace and the induced representation on MXM^{X} is not irreducible. Conversely, suppose that MXM^{X} is not irreducible and consider an 𝒯X\mathcal{T}_{X}-reducing subspace HMXH\subset M^{X} having orthogonal projection PP. We claim that PMMPM\subset M. Indeed, M=MXsM(X)MXM=M^{X}\ominus s_{M}(X)M^{X} can be characterized as a subspace of vectors vv having v,sM(ξ)w=0\langle v,s_{M}(\xi)w\rangle=0 for all ξX\xi\in X and wMXw\in M^{X}. By 𝒯X\mathcal{T}_{X}-reducibility, we have Pv,sM(ξ)w=v,sM(ξ)Pw=v,sM(ξ)w=0\langle Pv,s_{M}(\xi)w\rangle=\langle v,s_{M}(\xi)Pw\rangle=\langle v,s_{M}(\xi)w\rangle=0 and thus PvMPv\in M.

Since PMPM is an AA-subrepresentation of an irreducible representation MM, we have either PM=0PM=0 or PM=MPM=M. If PM=MHPM=M\subset H, then H=MXH=M^{X} since subspaces t(X)nMt(X)^{n}M for n0n\geq 0 span the whole MXM^{X}. Otherwise, we have PM=0PM=0 which implies M=(1P)MHM=(1-P)M\subset H^{\perp} and H=MXH^{\perp}=M^{X} for the same reason. We conclude that MXM^{X} does not contain any proper 𝒯X\mathcal{T}_{X}-reducing subspace. ∎

Definition 5.2.

The atomic spectrum σa(X)A^\sigma_{a}(X)\subset\widehat{A} of XX consists of those irreducible representations (M,πM)A^(M,\pi_{M})\in\widehat{A} with π~M(PX)=1H\tilde{\pi}_{M}(P_{X})=1_{H} or, equivalently, PπMPXP_{\pi_{M}}\leq P_{X}.

The term atomic spectrum is motivated by Proposition 5.4 which says that the atomic spectrum of the correspondence constructed from a representation coincides with the set of isomorphism classes of irreducible subrepresentations (atoms).

Note that ^Xσa(X)\widehat{\mathcal{I}}_{X}\subset\sigma_{a}(X) since PXPXP_{\mathcal{I}_{X}}\leq P_{X} by the proof of Corollary 4.8. It turns out that the equality holds exactly when the noncommutative Choquet boundary coincides with the noncommutative Shilov boundary.

Theorem 5.3.

Let XX be a C*-correspondence over a C*-algebra AA. The Choquet boundary of XX consists of all irreducible representations of 𝒪X\mathcal{O}_{X} except those induced from elements of the atomic spectrum:

CX=𝒪^Xind(σa(X))=Xfc^ind(A^σa(X)).\partial_{C}X=\widehat{\mathcal{O}}_{X}\setminus\operatorname{ind}\left(\sigma_{a}(X)\right)=\widehat{X_{\mathrm{fc}}}\sqcup\operatorname{ind}\left(\widehat{A}\setminus\sigma_{a}(X)\right).

In particular, CX=𝒪^X\partial_{C}X=\widehat{\mathcal{O}}_{X} if and only if σa(X)=^X\sigma_{a}(X)=\widehat{\mathcal{I}}_{X}.

Proof.

Let (H,t)𝒪^X(H,t)\in\widehat{\mathcal{O}}_{X} be an irreducible representation of XX. It is enough to prove that (H,t)(H,t) is not maximal if and only if it is induced from a representation in the atomic spectrum.

By Wold decomposition, (H,t)(H,t) is either fully coisometric or induced. If it is fully coisometric, then it is maximal. Suppose that it is induced by an AA-representation MM which is irreducible by Lemma 5.1. Since PXP_{X} is a central projection, ρ~M(PX)ρM(A)′′=\tilde{\rho}_{M}(P_{X})\in\rho_{M}(A)^{\prime\prime}=\mathbb{C} and thus either ρ~M(PX)=0\tilde{\rho}_{M}(P_{X})=0 or Mσa(X)M\in\sigma_{a}(X). By Corollary 4.6, (H,t)(H,t) is maximal if and only if ρ~M(PX)=0\tilde{\rho}_{M}(P_{X})=0. We conclude that (H,t)(H,t) is non-maximal exactly when Mσa(X)M\in\sigma_{a}(X). ∎

Arveson [6] conjectured that an operator system is hyperrigid if and only if all irreducible representations of its C*-envelope are boundary. The author together with Dor-On [9] found an elementary type I counterexample to the conjecture. However, it is still interesting to understand for which classes of operator systems the conjecture holds.

We say that a C*-correspondence XX violates the hyperrigidity conjecture if CX=𝒪^X\partial_{C}X=\widehat{\mathcal{O}}_{X} but XX is not hyperrigid. Theorem 5.3 has an immediate consequence that XX violates the hyperrigidity conjecture if and only if σa(X)=^X\sigma_{a}(X)=\widehat{\mathcal{I}}_{X} while φ(X)XX\varphi(\mathcal{I}_{X})X\neq X. We will see that there are plenty of correspondences having this property.

Let LL be a Hilbert space together with an AA-representation ρL:A(L)\rho_{L}\colon A\to\mathcal{B}(L). We can view LL as an AA-\mathbb{C}-correspondence and AA as a \mathbb{C}-AA-correspondence to produce an AA-correspondence LAL\otimes_{\mathbb{C}}A. Kumjian [35] showed that the Cuntz-Pimsner algebra 𝒪LA\mathcal{O}_{L\otimes_{\mathbb{C}}A} is purely infinite and simple provided that ρL(A)𝒦(L)=0\rho_{L}(A)\cap\mathcal{K}(L)=0.

Proposition 5.4.

The atomic spectrum σa(LA)\sigma_{a}(L\otimes_{\mathbb{C}}A) coincides with the atomic spectrum of LL, that is, the set of isomorphism classes of irreducible subrepresentations of LL. Consequently, if LL does not contain any irreducible subrepresentations, then LAL\otimes_{\mathbb{C}}A violates the hyperrigidity conjecture.

Proof.

Let X=LAX=L\otimes_{\mathbb{C}}A. We have kerφ~X=kerρ~L\ker\tilde{\varphi}_{X}=\ker\tilde{\rho}_{L}. Therefore, we also have ρ~L(A)=ρL(A)′′APX\tilde{\rho}_{L}(A^{**})=\rho_{L}(A)^{\prime\prime}\cong A^{**}P_{X}. By the elementary theory of von Neumann algebras, PXP_{X} acts non-degenerately on an irreducible representation of AA if and only if the representation is contained in LL.

If LL does not contain any irreducible subrepresentations, then σa(X)=\sigma_{a}(X)=\emptyset and CX=𝒪^X\partial_{C}X=\widehat{\mathcal{O}}_{X} by Theorem 5.3. On the other hand, the induced representation LXL^{X} of XX is not maximal by Corollary 4.6 and thus XX is not hyperrigid. Alternatively, one could show that X=0\mathcal{I}_{X}=0 and apply Kim’s hyperrigidity criterion. ∎

Example 5.5.

Let A=C()A=C^{*}(\mathbb{Z}) and X=2(,C())X=\ell^{2}(\mathbb{Z},C^{*}(\mathbb{Z})) with basis (ek)k{(e_{k})}_{k\in\mathbb{Z}}. The group \mathbb{Z} acts on XX by the automorphism ekek+1e_{k}\mapsto e_{k+1}, which defines the correspondence structure φX:C()(X)\varphi_{X}\colon C^{*}(\mathbb{Z})\to\mathcal{L}(X). An XX-representation on HH is equivalent to isometries Sk=t(ek)S_{k}=t(e_{k}) with pairwise orthogonal ranges and a unitary ρ(δ1)=U\rho(\delta_{1})=U such that USk=Sk+1US_{k}=S_{k+1}. Therefore, the purely infinite simple C*-algebra 𝒪X\mathcal{O}_{X} is obtained by attaching a unitary satisfying the above relation to the infinite Cuntz algebra 𝒪\mathcal{O}_{\infty}.

The correspondence XX is isomorphic to the tensor product 2()C()\ell^{2}(\mathbb{Z})\otimes_{\mathbb{C}}C^{*}(\mathbb{Z}), where the group \mathbb{Z} acts on 2()\ell^{2}(\mathbb{Z}) by the bilateral shift. Since the bilateral shift does not have any eigenvectors, XX violates the hyperrigidity conjecture by Theorem 5.4.

Example 5.6.

Let θ(0,1)\theta\in(0,1) be an irrational number and let Aθ=u,v:u=u1,v=v1,uv=e2πiθvuA_{\theta}=\langle u,v\colon u^{*}=u^{-1},v^{*}=v^{-1},uv=e^{2\pi i\theta}vu\rangle be the irrational rotation algebra. The algebra AθA_{\theta} is a simple C*-algebra with A^θ=𝕋/θ\widehat{A}_{\theta}=\mathbb{T}/\mathbb{Z}\theta. In particular, it is not of Type I.

Let M=2(×)M=\ell^{2}(\mathbb{Z}\times\mathbb{Z}) with basis (ξk,j)k,j{(\xi_{k,j})}_{k,j\in\mathbb{Z}}. Define unitaries U,VU,V on MM by Uξk,j=ξk+1,jU\xi_{k,j}=\xi_{k+1,j} and Vξk,j=e2πiθkξk,j+1V\xi_{k,j}=e^{-2\pi i\theta k}\xi_{k,j+1}. Setting ρ(u)=U\rho(u)=U and ρ(v)=V\rho(v)=V defines a representation AθA_{\theta} on MM. This representation satisfies the assumption of Theorem 5.4 since in any irreducible representation of uu has an eigenvector, whereas UU is a direct sum of bilateral shifts and does not contain any. Therefore, X=MAθX=M\otimes_{\mathbb{C}}A_{\theta} violates the hyperrigidity conjecture.

Proposition 5.7.

If XX is a proper correspondence, then σa(X)=φX(A)^=A^kerφX^\sigma_{a}(X)=\widehat{\varphi_{X}(A)}=\widehat{A}\setminus\widehat{\ker\varphi_{X}}. Consequently, Arveson’s hyperrigidity conjecture holds for proper correspondences.

Proof.

We have PX=1PkerφXP_{X}=1-P_{\ker\varphi_{X}} and thus Mσa(X)M\in\sigma_{a}(X) if and only if ρ~M(PkerφX)=0\tilde{\rho}_{M}(P_{\ker\varphi_{X}})=0, which is equivalent to kerφXkerρ\ker\varphi_{X}\subset\ker\rho. Therefore, σa(X)\sigma_{a}(X) consists of the irreducible representations which factor through AφX(A)A/kerφXA\twoheadrightarrow\varphi_{X}(A)\cong A/\ker\varphi_{X}.

Suppose now that CX=𝒪^X\partial_{C}X=\widehat{\mathcal{O}}_{X}. Therefore, we have φX(A)^=CX=^X\widehat{\varphi_{X}(A)}=\partial_{C}X=\widehat{\mathcal{I}}_{X} by Theorem 5.3. Since XφX(X)\mathcal{I}_{X}\cong\varphi_{X}(\mathcal{I}_{X}) is an ideal in φX(A)\varphi_{X}(A), this implies φX(X)=φX(A)\varphi_{X}(\mathcal{I}_{X})=\varphi_{X}(A). Therefore, the ideal X\mathcal{I}_{X} acts non-degenerately on XX and the correspondence is hyperrigid. We conclude that Arveson’s conjecture holds in this case. ∎

Suppose now that AA is a separable Type I C*-algebra. It is equivalent to the fact that any irreducible representation is uniquely determined up to isomorphism by its kernel by Glimm’s Theorem [25, Theorem 1]. Therefore, there is a canonical bijection between A^\widehat{A} and PrimA\operatorname{Prim}A and we identify those sets. Moreover, if A=C0(Y)A=C_{0}(Y), then we also identify A^\widehat{A} and PrimA\operatorname{Prim}A with YY.

Let 𝔓PrimA\mathfrak{P}\in\operatorname{Prim}A be some primitive ideal and let π𝔓\pi_{\mathfrak{P}} be the corresponding irreducible representation on K𝔓K_{\mathfrak{P}}. Again by Glimm’s Theorem [25, Theorem 1], the irreducible representation on K𝔓K_{\mathfrak{P}} is GCR, meaning that 𝕂(K𝔓)π𝔓(A)\mathbb{K}(K_{\mathfrak{P}})\subset\pi_{\mathfrak{P}}(A). Then, J0(𝔓)=π𝔓1(𝕂(K𝔓))J_{0}(\mathfrak{P})=\pi_{\mathfrak{P}}^{-1}(\mathbb{K}(K_{\mathfrak{P}})) is the smallest ideal of AA properly containing 𝔓=kerπ𝔓\mathfrak{P}=\ker\pi_{\mathfrak{P}}. If AA is commutative, then every primitive ideal is maximal and J0(𝔓)=AJ_{0}(\mathfrak{P})=A.

Lemma 5.8.

The support projection Pπ𝔓P_{\pi_{\mathfrak{P}}} equals to PJ0(𝔓)P𝔓P_{J_{0}(\mathfrak{P})}-P_{\mathfrak{P}}.

Proof.

Suppose that an AA-representation HH satisfies ρ~H(PJ0(𝔓)P𝔓)=1H\tilde{\rho}_{H}(P_{J_{0}(\mathfrak{P})}-P_{\mathfrak{P}})=1_{H}. It is equivalent to ρ~H(PJ0(𝔓))=1H\tilde{\rho}_{H}(P_{J_{0}(\mathfrak{P})})=1_{H} and ρ~H(P𝔓)=0\tilde{\rho}_{H}(P_{\mathfrak{P}})=0. The first equality implies that J0(𝔓)J_{0}(\mathfrak{P}) acts non-degenerately on HH while the second equality implies that 𝔓kerρH\mathfrak{P}\in\ker\rho_{H}.

Consequently, J0(𝔓)/𝔓𝕂(K𝔓)J_{0}(\mathfrak{P})/\mathfrak{P}\cong\mathbb{K}(K_{\mathfrak{P}}) acts non-degenerately on HH. By the representation theory of the C*-algebra of compact operators, HH is isomorphic to a direct sum of copies of K𝔓K_{\mathfrak{P}} and ρ~H(A)=ρH(A)′′π𝔓(A)′′=π~𝔓(A)\tilde{\rho}_{H}({A}^{**})=\rho_{H}(A)^{\prime\prime}\cong\pi_{\mathfrak{P}}(A)^{\prime\prime}=\tilde{\pi}_{\mathfrak{P}}({A}^{**}). Therefore, PJ0(𝔓)P𝔓P_{J_{0}(\mathfrak{P})}-P_{\mathfrak{P}} equals to the support projection of π𝔓\pi_{\mathfrak{P}}. ∎

Proposition 5.9.

The atomic spectrum of a correspondence XX over a Type I C*-algebra AA has the form

σa(X)={𝔓PrimA:φ(𝔓)Xφ(J0(𝔓))X}.\sigma_{a}(X)=\{\mathfrak{P}\in\operatorname{Prim}A\colon\varphi(\mathfrak{P})X\neq\varphi(J_{0}(\mathfrak{P}))X\}.

If A=C0(Y)A=C_{0}(Y) is abelian, then the formula above simplifies to

σa(X)={yY:φ(C0(Y{y}))XX}.\sigma_{a}(X)=\{y\in Y\colon\varphi(C_{0}(Y\setminus\{y\}))X\neq X\}.
Proof.

By Lemma 3.1, we have φ(𝔓)Xφ(J0(𝔓))X\varphi(\mathfrak{P})X\neq\varphi(J_{0}(\mathfrak{P}))X if and only if φ~(P𝔓)Xφ~(PJ0(𝔓))X\tilde{\varphi}(P_{\mathfrak{P}}){X}^{**}\neq\tilde{\varphi}(P_{J_{0}(\mathfrak{P})}){X}^{**}. The latter inequality holds if and only if φ~(Pπ𝔓)=φ~(PJ0(𝔓)P𝔓)0\tilde{\varphi}(P_{\pi_{\mathfrak{P}}})=\tilde{\varphi}(P_{J_{0}(\mathfrak{P})}-P_{\mathfrak{P}})\neq 0. Therefore, we need to prove that φ~(Pπ𝔓)0\tilde{\varphi}(P_{\pi_{\mathfrak{P}}})\neq 0 if and only if Pπ𝔓PXP_{\pi_{\mathfrak{P}}}\leq P_{X}.

Suppose that Pπ𝔓PXP_{\pi_{\mathfrak{P}}}\leq P_{X}. We have APπ𝔓APX{A}^{**}P_{\pi_{\mathfrak{P}}}\subset{A}^{**}P_{X}. Therefore, APπ𝔓{A}^{**}P_{\pi_{\mathfrak{P}}} acts on X{X}^{**} faithfully and φ~(Pπ𝔓)0\tilde{\varphi}(P_{\pi_{\mathfrak{P}}})\neq 0.

On the other hand, if φ~(Pπ𝔓)0\tilde{\varphi}(P_{\pi_{\mathfrak{P}}})\neq 0, then 0PXPπ𝔓Pπ𝔓0\neq P_{X}P_{\pi_{\mathfrak{P}}}\leq P_{\pi_{\mathfrak{P}}}. Since Pπ𝔓P_{\pi_{\mathfrak{P}}} is the support projection of an irreducible representation, it is a minimal central projection. Therefore, we get PXPπ𝔓=Pπ𝔓PXP_{X}P_{\pi_{\mathfrak{P}}}=P_{\pi_{\mathfrak{P}}}\leq P_{X}.

The formula for abelian C*-algebras is a special case of the general one. The simplification comes from the fact that all primitive ideals are maximal and of the form C0(Y{y})C_{0}(Y\setminus\{y\}). ∎

6 Topological quivers

Here we apply the results of previous sections to topological quivers. In our exposition we follow the initial reference [41] with the reversed role of the maps ss and rr. We assume all topological spaces to be second countable, locally compact, and Hausdorff.

Let E0,E1E^{0},E^{1} be two topological spaces together with continuous maps s,r:E1E0s,r\colon E^{1}\rightrightarrows E^{0} called source and range maps. Let λ=(λx)xE0\lambda=(\lambda_{x})_{x\in E^{0}} be the collection of measures on E1E^{1} such that suppλx=s1(x)\operatorname{supp}\lambda_{x}=s^{-1}(x) for all xE0x\in E^{0}. We additionally require that the function

xs1(x)ξ(t)𝑑λx(t)x\mapsto\int_{s^{-1}(x)}\xi(t)d\lambda_{x}(t)

is in Cc(E0)C_{c}(E^{0}) for all ξCc(E1)\xi\in C_{c}(E^{1}). The tuple (E0,E1,r,s,λ)(E^{0},E^{1},r,s,\lambda) satisfying the above conditions is called a topological quiver.

We define a C0(E0)C_{0}(E^{0})-valued inner product on Cc(E1)C_{c}(E^{1}) by

ξ,η(x)=s1(x)ξ(t)¯η(t)𝑑λx(t)\langle\xi,\eta\rangle(x)=\int_{s^{-1}(x)}\overline{\xi(t)}\eta(t)d\lambda_{x}(t)

for all ξ,ηCc(E1)\xi,\eta\in C_{c}(E^{1}). We denote the completion of Cc(E1)C_{c}(E^{1}) with respect to this inner product by X=X(E)X=X(E) which is a Hilbert C0(E0)C_{0}(E^{0})-module. The left action by fC0(E1)f\in C_{0}(E^{1}) on XX is given by (φ(f)ξ)(t)=f(r(t))ξ(t)(\varphi(f)\xi)(t)=f(r(t))\xi(t) for all ξCc(E1)\xi\in C_{c}(E^{1}), tE1t\in E^{1}, and extended to all XX by continuity.

Ideals of C0(E0)C_{0}(E^{0}) are in bijection with open subsets UE0U\subset E^{0} via the assignment UC0(U)C0(E0)U\mapsto C_{0}(U)\subset C_{0}(E^{0}). The open subset corresponding to Katsura’s ideal X\mathcal{I}_{X} is denoted by Ereg0E^{0}_{\mathrm{reg}}. The points of Ereg0E^{0}_{\mathrm{reg}} are called regular vertices. There is a topological characterization of regular vertices in [41, Proposition 3.15].

As in the previous section, we identify C0(E0)^\widehat{C_{0}(E^{0})} with E0E^{0}.

Theorem 6.1.

Let E=(E0,E1,r,s,λ)E=(E^{0},E^{1},r,s,\lambda) be a topological quiver and X=X(E)X=X(E) be the associated correspondence. The atomic spectrum σa(X)\sigma_{a}(X) consists of points xE0x\in E^{0} with λy(r1(x))0\lambda_{y}(r^{-1}(x))\neq 0 for some yE0y\in E^{0}.

Proof.

We will use the description of the atomic spectrum from Proposition 5.9. Suppose that λy(r1(x))=0\lambda_{y}(r^{-1}(x))=0 for all yE0y\in E^{0}. Let {ei}i\{e_{i}\}_{i\in\mathbb{N}} be an increasing approximate identity of C0(E0{x})C_{0}(E^{0}\setminus\{x\}). Then, for every ξ,ηCc(E1)\xi,\eta\in C_{c}(E^{1}) and yE0y\in E^{0}, we have

limiφ(ei)ξ,η(y)=s1(y)ei(r(t))ξ(t)¯η(t)𝑑λy(t)=s1(y)ξ(t)¯η(t)𝑑λy(t)=ξ,η(y)\lim_{i\to\infty}\langle\varphi(e_{i})\xi,\eta\rangle(y)=\int_{s^{-1}(y)}e_{i}(r(t))\overline{\xi(t)}\eta(t)d\lambda_{y}(t)\\ =\int_{s^{-1}(y)}\overline{\xi(t)}\eta(t)d\lambda_{y}(t)=\langle\xi,\eta\rangle(y)

since eirλy-a.e.1e_{i}\circ r\xrightarrow{\lambda_{y}\text{-a.e.}}1. Therefore, φ(C0(E0{x}))X=X\varphi(C_{0}(E^{0}\setminus\{x\}))X=X and xσa(X)x\notin\sigma_{a}(X).

Conversely, suppose that λy(r1(x))>0\lambda_{y}(r^{-1}(x))>0 for some yE0y\in E^{0}. Then, we have φ(C0(E0{x}))X{ξC0(E0):ξ|r1(x)=0}¯X\varphi(C_{0}(E^{0}\setminus\{x\}))X\subset\overline{\{\xi\in C_{0}(E^{0})\colon\xi|_{r^{-1}(x)}=0\}}\subsetneq X. We conclude that xσa(X)x\in\sigma_{a}(X). ∎

A topological quiver is called a topological graph if the map ss is a local homeomorphism. This implies that s1(x)=suppλxs^{-1}(x)=\operatorname{supp}\lambda_{x} is discrete for all xE0x\in E^{0}.

Corollary 6.2.

A topological quiver violates the hyperrigidity conjecture if and only if λy(r1(x))=0\lambda_{y}(r^{-1}(x))=0 for all xE0Ereg0x\in E^{0}\setminus E^{0}_{\mathrm{reg}} and yE0y\in E^{0} while λy(r1(E0Ereg0))0\lambda_{y}(r^{-1}(E^{0}\setminus E^{0}_{\mathrm{reg}}))\neq 0 for some yE0y\in E^{0}. In particular, correspondences of topological graphs satisfy the hyperrigidity conjecture.

Proof.

By definition, we have ^X(E)=Ereg0\widehat{\mathcal{I}}_{X(E)}=E^{0}_{\mathrm{reg}}. By Theorem 5.3 and Theorem 6.1, we have C(X)=𝒪^X\partial_{C}(X)=\widehat{\mathcal{O}}_{X} if and only if σa(X)=^X\sigma_{a}(X)=\widehat{\mathcal{I}}_{X} if and only if λy(r1(x))=0\lambda_{y}(r^{-1}(x))=0 for all xE0Ereg0x\in E^{0}\setminus E^{0}_{\mathrm{reg}}. On the other hand, we have φ(X)X=X\varphi(\mathcal{I}_{X})X=X if and only if all the measures λx\lambda_{x} are concentrated on r1(Ereg0)r^{-1}(E^{0}_{\mathrm{reg}}). By Kim’s Theorem (Corollary 4.9), it follows that XX is hyperrigid if and only if λy(r1(E0Ereg0))=0\lambda_{y}(r^{-1}(E^{0}\setminus E^{0}_{\mathrm{reg}}))=0 for all yE0y\in E^{0}.

When EE is a topological quiver, the measure λx\lambda_{x} has the discrete support s1(x)s^{-1}(x) and thus it is a discrete measure for all xXx\in X. Assume that λy(r1(x))=0\lambda_{y}(r^{-1}(x))=0 for all xE0Ereg0x\in E^{0}\setminus E^{0}_{\mathrm{reg}}. We have λy(r1(E0Ereg0))=xE0Ereg0λy(r1(x))=0\lambda_{y}(r^{-1}(E^{0}\setminus E^{0}_{\mathrm{reg}}))=\sum_{x\in E^{0}\setminus E^{0}_{\mathrm{reg}}}\lambda_{y}(r^{-1}(x))=0 for all yE0y\in E^{0}. Therefore, the two conditions for the violation of the hyperrigidity conjecture cannot both hold for topological graphs. ∎

Example 6.3.

Let YY be a connected topological space which is not a single point and μ\mu be a probability measure on YY with full support. Then, we may define a topological quiver with E0=Y{}E^{0}=Y\sqcup\{\bullet\}, E1=YE^{1}=Y, s(Y)=s(Y)=\bullet, r=idYr=\operatorname{id}_{Y}, and λ=μ\lambda_{\bullet}=\mu. Since ss is not a homeomorphism on any open set, Ereg0=E^{0}_{\mathrm{reg}}=\emptyset. The atomic spectrum of EE is exactly the set of atoms of μ\mu. Therefore, if μ\mu is atomless, then EE violates the hyperrigidity conjecture. One may check manually that the representation induced by L2(Y,μ)L^{2}(Y,\mu) is not maximal.

Let YY be a topological space. We identify C0(Y)C_{0}(Y)^{*} with the set of bounded complex Radon measures on YY. Let B(Y)B(Y) be the *-algebra of bounded Borel functions on YY. Then, integration defines an injective *-homomorphism B(Y)C0(Y)B(Y)\hookrightarrow C_{0}(Y)^{**} and we identify B(Y)B(Y) with its image. In particular, if SYS\subset Y is a Borel subset, then the characteristic function 𝟏S\mathbf{1}_{S} is a projection in the W*-algebra C0(Y)C_{0}(Y)^{**}.

For a representation ρH:C0(Y)(H)\rho_{H}\colon C_{0}(Y)\to\mathcal{B}(H), we say that the projection ρ~H(𝟏S)\tilde{\rho}_{H}(\mathbf{1}_{S}) is the spectral projection corresponding to SS. If Y=σ(T)Y=\sigma(T) for a normal operator T(H)T\in\mathcal{B}(H) and the representation ρH\rho_{H} is the tautological one, then the notion of spectral projection coincides with the classical one for normal operators.

If ZZ is another topological space, then by Gelfand–Naimark duality, a *-homomorphism ψ:C0(Y)(C0(Z))=Cb(Z)\psi\colon C_{0}(Y)\to\mathcal{M}(C_{0}(Z))=C_{b}(Z) is equivalent to a continuous map ψˇ:ZY\check{\psi}\colon Z\to Y. The dual of ψ\psi restricted to C0(Z){C_{0}(Z)}^{*} acts as the pushforward of measures ψˇ:C0(Z)C0(Y)\check{\psi}_{*}\colon{C_{0}(Z)}^{*}\to{C_{0}(Y)}^{*}. Finally, the dual of ψˇ\check{\psi}_{*} is the normal homomorphism ψ~:C0(Y)C0(Z)\tilde{\psi}\colon{C_{0}(Y)}^{**}\to{C_{0}(Z)}^{**}. If fB(Y)C0(Y)f\in B(Y)\subset{C_{0}(Y)}^{**}, then ψ~(f)=fψB(Z)\tilde{\psi}(f)=f\circ\psi\in B(Z).

Lemma 6.4.

The support projection of ψ:C0(Y)(C0(Z))\psi\colon C_{0}(Y)\to\mathcal{M}(C_{0}(Z)) is the characteristic function 𝟏ψˇ(Z)B(Y)C0(Y)\mathbf{1}_{\check{\psi}(Z)}\in B(Y)\subset C_{0}(Y)^{**} of the Borel set ψˇ(Z)Y\check{\psi}(Z)\subset Y.

Proof.

Denote C=ψˇ(Z)C=\check{\psi}(Z). We have ψ~(𝟏C)=ψ𝟏C=1\tilde{\psi}(\mathbf{1}_{C})=\psi\circ\mathbf{1}_{C}=1 so that 𝟏CPψ\mathbf{1}_{C}\geq P_{\psi}. Suppose that 𝟏C>Pψ\mathbf{1}_{C}>P_{\psi}. Hence, there exists a probability measure μC(Y)\mu\in{C(Y)}^{*} such that 𝟏C(μ)=μ(C)>Pψ(μ)\mathbf{1}_{C}(\mu)=\mu(C)>P_{\psi}(\mu). By replacing μ\mu with 1μ(C)μ|C\frac{1}{\mu(C)}\mu|_{C} we may assume that μ(C)=1\mu(C)=1 and μ(YC)=1\mu(Y\setminus C)=1. Since we assume YY to be second countable LCH and thus separable and metrizable, we may apply [52, Lemma 2.2] to find a probability measure ν\nu on ZZ such that μ=ψˇν\mu=\check{\psi}_{*}\nu. We arrive at a contradiction by noting that

1>Pψ(μ)=Pψ(ψˇν)=ψ~(Pψ)(ν)=1.1>P_{\psi}(\mu)=P_{\psi}(\check{\psi}_{*}\nu)=\tilde{\psi}(P_{\psi})(\nu)=1.

We conclude that 𝟏C=Pψ\mathbf{1}_{C}=P_{\psi}. ∎

Theorem 6.5.

A representation (H,t)(H,t) of the correspondence X=X(E)X=X(E) of a topological graph E=(E0,E1,r,s,λ)E=(E^{0},E^{1},r,s,\lambda) is fully Cuntz-Pimsner covariant if and only if t(X)H¯=Hr(E1)\overline{t(X)H}=H_{r(E^{1})}, where Hr(E1)H_{r(E^{1})} is the range of the spectral projection on HH corresponding to the Borel subset r(E1)E0r(E^{1})\subset E^{0}.

Proof.

Since ss is a local homeomorphism, from [41, Theorem 3.11] it follows that pointwise multiplication defines a natural embedding ι:C0(E1)𝒦(X(E))\iota\colon C_{0}(E^{1})\hookrightarrow\mathcal{K}(X(E)). Consequently, the induced map ι~:C0(E1)(X(E))\tilde{\iota}\colon{C_{0}(E^{1})}^{**}\to\mathcal{L}({X(E)}^{**}) is injective. The left action is given by the composition

φ:C0(E0)r^Cb(E1)=(C0(E1))𝜄(𝒦(X))=(X(E)).\varphi\colon C_{0}(E^{0})\xrightarrow{\hat{r}}C_{b}(E^{1})=\mathcal{M}(C_{0}(E^{1}))\xrightarrow{\iota}\mathcal{M}(\mathcal{K}(X))=\mathcal{L}(X(E)).

Therefore, the map φ~\tilde{\varphi} is given by the composition C0(E0)r~C0(E1)(X(E)){C_{0}(E^{0})}^{**}\xrightarrow{\tilde{r}}{C_{0}(E^{1})}^{**}\hookrightarrow\mathcal{L}({X(E)}^{**}) and thus the kernel of φ~\tilde{\varphi} equals the kernel of r~\tilde{r}. Finally, the support projection PXP_{X} of φ\varphi equals the support projection of rr which is 𝟏r(E1)\mathbf{1}_{r(E^{1})} by Lemma 6.4. We conclude that the representation is fully Cuntz-Pimsner covariant if and only if t(X(E))H=Hr(E1)=ρ~(𝟏r(E1))Ht(X(E))H=H_{r(E^{1})}=\tilde{\rho}(\mathbf{1}_{r(E^{1})})H. ∎

The above theorem generalizes the analogous result for discrete graphs by Dor-On and Salomon which was explained earlier in Example 4.5.

References

  • [1] Jim Agler “An abstract approach to model theory” In Surveys of some recent results in operator theory, Vol. II 192, Pitman Res. Notes Math. Ser. Longman Sci. Tech., Harlow, 1988, pp. 1–23
  • [2] Charles A. Akemann, Gert K. Pedersen and Jun Tomiyama “Multipliers of CC^{*}-algebras” In J. Functional Analysis 13, 1973, pp. 277–301 DOI: 10.1016/0022-1236(73)90036-0
  • [3] T. Andô “On a pair of commutative contractions” In Acta Sci. Math. (Szeged) 24, 1963, pp. 88–90
  • [4] William Arveson “Notes on the unique extension property” In Unpublished note, availible at http://users.uoa.gr/~akatavol/newtexfil/arveson/unExt.pdf, 2003
  • [5] William Arveson “The noncommutative Choquet boundary” In J. Amer. Math. Soc. 21.4, 2008, pp. 1065–1084 DOI: 10.1090/S0894-0347-07-00570-X
  • [6] William Arveson “The noncommutative Choquet boundary II: hyperrigidity” In Israel J. Math. 184, 2011, pp. 349–385 DOI: 10.1007/s11856-011-0071-z
  • [7] William B. Arveson “Subalgebras of CC^{\ast}-algebras” In Acta Math. 123, 1969, pp. 141–224 DOI: 10.1007/BF02392388
  • [8] Tirthankar Bhattacharyya, Sourav Pal and Subrata Shyam Roy “Dilations of Γ\Gamma-contractions by solving operator equations” In Adv. Math. 230.2, 2012, pp. 577–606 DOI: 10.1016/j.aim.2012.02.016
  • [9] Boris Bilich and Adam Dor-On “Arveson’s hyperrigidity conjecture is false”, 2024 arXiv:2404.05018 [math.OA]
  • [10] B. Blackadar “Operator algebras” Theory of CC^{*}-algebras and von Neumann algebras, Operator Algebras and Non-commutative Geometry, III 122, Encyclopaedia of Mathematical Sciences Springer-Verlag, Berlin, 2006, pp. xx+517 DOI: 10.1007/3-540-28517-2
  • [11] David P. Blecher and Christian Le Merdy “Operator algebras and their modules—an operator space approach” Oxford Science Publications 30, London Mathematical Society Monographs. New Series The Clarendon Press, Oxford University Press, Oxford, 2004, pp. x+387 DOI: 10.1093/acprof:oso/9780198526599.001.0001
  • [12] Man-Duen Choi and Kenneth R. Davidson “A 3×33\times 3 dilation counterexample” In Bull. Lond. Math. Soc. 45.3, 2013, pp. 511–519 DOI: 10.1112/blms/bds109
  • [13] Raphaël Clouâtre “Non-commutative peaking phenomena and a local version of the hyperrigidity conjecture” In Proc. Lond. Math. Soc. (3) 117.2, 2018, pp. 221–245 DOI: 10.1112/plms.12133
  • [14] Raphaël Clouâtre “Unperforated pairs of operator spaces and hyperrigidity of operator systems” In Canad. J. Math. 70.6, 2018, pp. 1236–1260 DOI: 10.4153/CJM-2018-008-1
  • [15] Raphaël Clouâtre and Michael Hartz “Multiplier algebras of complete Nevanlinna-Pick spaces: dilations, boundary representations and hyperrigidity” In J. Funct. Anal. 274.6, 2018, pp. 1690–1738 DOI: 10.1016/j.jfa.2017.10.008
  • [16] Raphaël Clouâtre and Hridoyananda Saikia “A boundary projection for the dilation order”, 2023 arXiv:2310.17601 [math.OA]
  • [17] Raphaël Clouâtre and Ian Thompson “Rigidity of operator systems: tight extensions and noncommutative measurable structures”, 2024 arXiv: https://arxiv.org/abs/2406.16806
  • [18] Kenneth R. Davidson, Adam Dor-On, Orr Moshe Shalit and Baruch Solel “Dilations, inclusions of matrix convex sets, and completely positive maps” In Int. Math. Res. Not. IMRN, 2017, pp. 4069–4130 DOI: 10.1093/imrn/rnw140
  • [19] Kenneth R. Davidson, Adam H. Fuller and Evgenios T.. Kakariadis “Semicrossed products of operator algebras by semigroups” In Mem. Amer. Math. Soc. 247.1168, 2017, pp. v+97 DOI: 10.1090/memo/1168
  • [20] Kenneth R. Davidson and Matthew Kennedy “Choquet order and hyperrigidity for function systems” In Adv. Math. 385, 2021, pp. Paper No. 107774\bibrangessep30 DOI: 10.1016/j.aim.2021.107774
  • [21] Kenneth R. Davidson and Matthew Kennedy “The Choquet boundary of an operator system” In Duke Math. J. 164.15, 2015, pp. 2989–3004 DOI: 10.1215/00127094-3165004
  • [22] Kenneth R. Davidson, Stephen C. Power and Dilian Yang “Dilation theory for rank 2 graph algebras” In J. Operator Theory 63.2, 2010, pp. 245–270
  • [23] Adam Dor-On and Guy Salomon “Full Cuntz-Krieger dilations via non-commutative boundaries” In J. Lond. Math. Soc. (2) 98.2, 2018, pp. 416–438 DOI: 10.1112/jlms.12140
  • [24] Michael A. Dritschel and Scott A. McCullough “Boundary representations for families of representations of operator algebras and spaces” In J. Operator Theory 53.1, 2005, pp. 159–167
  • [25] James Glimm “Type I CC^{\ast}-algebras” In Ann. of Math. (2) 73, 1961, pp. 572–612 DOI: 10.2307/1970319
  • [26] Masamichi Hamana “Injective envelopes of operator systems” In Publ. Res. Inst. Math. Sci. 15.3, 1979, pp. 773–785 DOI: 10.2977/prims/1195187876
  • [27] Samuel J. Harris and Se-Jin Kim “Crossed products of operator systems” In J. Funct. Anal. 276.7, 2019, pp. 2156–2193 DOI: 10.1016/j.jfa.2018.11.017
  • [28] Evgenios T.. Kakariadis and Orr Moshe Shalit “Operator algebras of monomial ideals in noncommuting variables” In J. Math. Anal. Appl. 472.1, 2019, pp. 738–813 DOI: 10.1016/j.jmaa.2018.11.050
  • [29] Elias G. Katsoulis and David W. Kribs “Tensor algebras of CC^{*}-correspondences and their CC^{*}-envelopes” In J. Funct. Anal. 234.1, 2006, pp. 226–233 DOI: 10.1016/j.jfa.2005.12.013
  • [30] Elias G. Katsoulis and Christopher Ramsey “The hyperrigidity of tensor algebras of C\rm C^{\ast}-correspondences” In J. Math. Anal. Appl. 483.1, 2020, pp. 123611\bibrangessep10 DOI: 10.1016/j.jmaa.2019.123611
  • [31] Elias G. Katsoulis and Christopher Ramsey “The non-selfadjoint approach to the Hao-Ng isomorphism” In Int. Math. Res. Not. IMRN, 2021, pp. 1160–1197 DOI: 10.1093/imrn/rnz271
  • [32] Takeshi Katsura “On CC^{*}-algebras associated with CC^{*}-correspondences” In J. Funct. Anal. 217.2, 2004, pp. 366–401 DOI: 10.1016/j.jfa.2004.03.010
  • [33] Matthew Kennedy and Orr Moshe Shalit “Essential normality, essential norms and hyperrigidity” In J. Funct. Anal. 268.10, 2015, pp. 2990–3016 DOI: 10.1016/j.jfa.2015.03.014
  • [34] Se-Jin Kim “Hyperrigidity of C*-correspondences” In Integral Equations Operator Theory 93.4, 2021, pp. Paper No. 47\bibrangessep17 DOI: 10.1007/s00020-021-02663-3
  • [35] Alex Kumjian “On certain Cuntz-Pimsner algebras” In Pacific J. Math. 217.2, 2004, pp. 275–289 DOI: 10.2140/pjm.2004.217.275
  • [36] Ralf Meyer and Camila F. Sehnem “A bicategorical interpretation for relative Cuntz-Pimsner algebras” In Math. Scand. 125.1, 2019, pp. 84–112 DOI: 10.7146/math.scand.a-112630
  • [37] Paul S. Muhly and Baruch Solel “An algebraic characterization of boundary representations” In Nonselfadjoint operator algebras, operator theory, and related topics 104, Oper. Theory Adv. Appl. Birkhäuser, Basel, 1998, pp. 189–196
  • [38] Paul S. Muhly and Baruch Solel “Hardy algebras, WW^{\ast}-correspondences and interpolation theory” In Math. Ann. 330.2, 2004, pp. 353–415 DOI: 10.1007/s00208-004-0554-x
  • [39] Paul S. Muhly and Baruch Solel “Tensor algebras over CC^{*}-correspondences: representations, dilations, and CC^{*}-envelopes” In J. Funct. Anal. 158.2, 1998, pp. 389–457 DOI: 10.1006/jfan.1998.3294
  • [40] Paul S. Muhly and Baruch Solel “Tensor algebras, induced representations, and the Wold decomposition” In Canad. J. Math. 51.4, 1999, pp. 850–880 DOI: 10.4153/CJM-1999-037-8
  • [41] Paul S. Muhly and Mark Tomforde “Topological quivers” In Internat. J. Math. 16.7, 2005, pp. 693–755 DOI: 10.1142/S0129167X05003077
  • [42] Paweł Pietrzycki and Jan Stochel “Hyperrigidity I: operator moments and convergence of subnormal operators”, 2024 arXiv:2405.20814 [math.OA]
  • [43] Michael V. Pimsner “A class of CC^{*}-algebras generalizing both Cuntz-Krieger algebras and crossed products by 𝐙{\bf Z} In Free probability theory (Waterloo, ON, 1995) 12, Fields Inst. Commun. Amer. Math. Soc., Providence, RI, 1997, pp. 189–212
  • [44] Gelu Popescu “Doubly Λ\Lambda-commuting row isometries, universal models, and classification” In J. Funct. Anal. 279.12, 2020, pp. 108798\bibrangessep69 DOI: 10.1016/j.jfa.2020.108798
  • [45] Iain Raeburn “Graph algebras” 103, CBMS Reg. Conf. Ser. Math. Providence, RI: American Mathematical Society (AMS), 2005
  • [46] Guy Salomon “Hyperrigid subsets of Cuntz-Krieger algebras and the property of rigidity at zero” In J. Operator Theory 81.1, 2019, pp. 61–79 DOI: 10.7900/jot
  • [47] Orr Moshe Shalit “Dilation theory: a guided tour” In Operator theory, functional analysis and applications 282, Oper. Theory Adv. Appl. Birkhäuser/Springer, Cham, 2021, pp. 551–623
  • [48] Orr Moshe Shalit “Representing a product system representation as a contractive semigroup and applications to regular isometric dilations” In Canad. Math. Bull. 53.3, 2010, pp. 550–563 DOI: 10.4153/CMB-2010-060-8
  • [49] W. Stinespring “Positive functions on CC^{*}-algebras” In Proc. Amer. Math. Soc. 6, 1955, pp. 211–216 DOI: 10.2307/2032342
  • [50] Béla Sz.-Nagy “Sur les contractions de l’espace de Hilbert” In Acta Sci. Math. 15, 1953, pp. 87–92
  • [51] Ian Thompson “An approximate unique extension property for completely positive maps” In J. Funct. Anal. 286.1, 2024, pp. Paper No. 110193\bibrangessep32 DOI: 10.1016/j.jfa.2023.110193
  • [52] V.. Varadarajan “Groups of automorphisms of Borel spaces” In Trans. Amer. Math. Soc. 109, 1963, pp. 191–220 DOI: 10.2307/1993903
  • [53] Alexander Vernik “Dilations of CP-maps commuting according to a graph” In Houston J. Math. 42.4, 2016, pp. 1291–1329