Quick viewing(Text Mode)

Arxiv:2101.09776V1 [Math.OA]

Arxiv:2101.09776V1 [Math.OA]

arXiv:2101.09776v1 [math.OA] 24 Jan 2021 rtragba,acmo hm steaayi fterfinite-dimen their of analysis [ the instance is for theme see common resentations; a , erator hr r osc ol vial o eea prtragba.T algebras. operator general for Ho available provide tools notion. this such this no and of are RFD, characterizations there are op alternative them of developing of classes for concrete which tus for characterize Even to RFD. challenging wh be is undertaken, to was shown algebras were operator tor RFD of study systematic a eairo rirr ersnain yapoiain hsi ca is finite-dimensional. This residually are approximation. that by wo sufficient algebras representations to are arbitrary has representations of one finite-dimensional behavior Indeed, extent, algebra. what operator situation given to representat these a finite-dimensional of in extent structure what approach the to effective clear and entirely natural not is perfectly a is this a atal upre yNFgatDS1096adb h E the by and 839412. DMS-1900916 No grant Sk grant NSF lodowska-Curie Marie by supported partially was semigrou coaction, semigroup approximation, Exel–Loring 47B32. iae.Yt h eea oinhsrcie oprtvl itea little comparatively received has notion [ general the Yet, tigated. disacmltl omn e frpeettoso finite-dimens on representations of C For norming spaces. completely a admits 48 IIEDMNINLAPOIAIN N SEMIGROUP AND APPROXIMATIONS FINITE-DIMENSIONAL ntecasfiainpormfrcrancasso oceenon- concrete of classes certain for program classification the In 2020 noeao ler ssi obe to said is algebra operator An h rtato a atal upre ya SR Discover NSERC an by supported partially was author first The e od n phrases. and words Key n [ and ] ahmtc ujc Classification. Subject Mathematics nmn ae finterest. of cases many in fa prtragbai neie yismxmlC maximal its fin by inherited residual is whether algebra of operator question an semigroups. the of of to related algebras intimately operator approximat is and finite-dimensional functions construct of an to algebras on us semigroup allows a which of coaction finite-dimensional exh such we residually One generality, greater approximations. finite-dimensional this constructing in operator emerge non-self-adjoint) subtleties merous therei (possibly representa general dense of to be space paradigm its should of representations property finite-dimensional topological a in encoded Abstract. 3 ,bt fwihfcso oceeagba ffntos However functions. of algebras concrete on focus which of both ], OCIN O PRTRALGEBRAS OPERATOR FOR COACTIONS ∗ agba,ti sacasclnto hthsbe hruhyinves- thoroughly been has that notion classical a is this -algebras, h eiulfiiedmninlt faC a of finite-dimensionality residual The RAPHA iiedmninlapoiain eiulfinite-dimen residual approximation, Finite-dimensional LCLOU EL ¨ 31 1. , 33 Introduction TEADAA DOR-ON ADAM AND ATRE ˆ , 40 eiulyfinite-dimensional residually rmr 75,4L7 eodr 72,20Mxx, 47A20, Secondary 46L07; 47L55, Primary , 54 1 n h eeecsteen Although therein. references the and ] lers lerso functions. of algebras algebras, p ∗ ∗ cvr hc eresolve we which -, agbai nw obe to known is -algebra rpa no’ oio 2020 Horizon Union’s uropean lers hl nu- While algebras. .W xedthis extend We n. oli oino a of notion a is tool btnvltosfor tools novel ibit rn.Tescn author second The grant. y in,saigthat stating tions, ite-dimensionality u investigation Our prtralgebra, operator osfroperator for ions i tnsi sharp in stands his drwehr and whether, nder ostuycapture truly ions r eea opera- several ere odtriethe determine to ee,a present at wever, rtragba it algebras erator tninbeyond ttention efroperator for se ,i eea it general in s, o F)i it if RFD) (or efdon op- selfadjoint oa Hilbert ional h impe- the s inlrep- sional sionality, n[ in , 20 ] 2 R. CLOUATREˆ AND A. DOR-ON contrast with the more classical C∗-algebraic setting where various conditions have been shown to be equivalent to residual finite-dimensionality [6, 24, 38, 39]. In this paper, we provide new characterizations that apply to general operator algebras. Our approach is inspired by the work of Exel and Loring [38], which we discuss next. Let A be a C∗-algebra, and recall that the spectrum of A is the set A of unitary equivalence classes of irreducible ∗-representations, together with the point strong operator topology. In [38] it is shown that A is RFD if and only if for everyb ∗-representation there exists a net (πλ) of finite dimensional ∗-representations of A that converges to π in the point strong operator topology. Following [38], it was shown by Archbold [6] that a C∗-algebra A is RFD if and only if the set of unitary equivalence classes of finite-dimensional irreducible ∗-representations is dense in A. Hence, the two topological characterizations of residual finite-dimensionality, one in terms of irreducible ∗-representations, and another in terms of arbitrary ∗b-representations, provides a fair amount of additional flexibility. The main driving force behind the present paper is the investigation of those representations of general operator algebras that admit a finite-dimensional ap- proximation of the type described in the previous paragraph. As we will see, there are subtleties intrinsic to working with general operator algebras (as opposed to C∗-algebras), wherein there are several different reasonable interpretations of what a finite-dimensional approximation should be. Interestingly, this flexibility allows one to connect the existence of finite-dimensional approximations to the residual finite-dimensionality of various C∗-covers for the operator algebra. This is emi- nently desirable, as RFD C∗-algebras are typically much better understood. We also note that the properties of the maximal and minimal C∗-covers being RFD have been explored previously in [20, 21], and the findings of these papers aptly illustrate the depth of the problem. Therefore, our analysis based on the existence of finite-dimensional approximations sheds light on some unresolved problems. Next, we describe the organization of the paper. In Section 2, we introduce various notions of finite-dimensional approximations, which we briefly mention here. In what follows, all representations of operator algebras are completely contractive. Let A be an operator algebra and let π : A → B(H) be a representation. Let

πλ : A → B(H), λ ∈ Λ

∗ be a net of representations with the property that the space C (πλ(A))H is finite- dimensional for every λ ∈ Λ. We say that the net (πλ) is a finite-dimensional approximation for π if (πλ(a)) converges to π(a) in the SOT for every a ∈ A. If, in ∗ ∗ , we have that (πλ(a) ) converges to π(a) in the SOT for every a ∈ A, then we say that (πλ) is an Exel–Loring approximation for π. These two notions coincide when A is a C∗-algebra. Finite-dimensional approximations do not necessarily arise in geometrically transparent ways (see Example 3.4) and thus, constructing such approximations can be highly non-trivial. Consequently, it is beneficial to identify operations that preserve the existence of these approximations, and this is the focus of the rest of Section 2. In Section 3, we utilize finite-dimensional approximations to analyze the residual finite-dimensionality of general operator algebras. Recall that the C∗-envelope of ∗ ∗ A is the smallest C -algebra generated by a copy A and it is denoted by Ce(A). As explained in Subsection 2.1, it can be constructed with the aid of so-called extremal representations of A. Using these ideas, we can now state one of our FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 3 main results, which is a version of [38, Theorem 2.4] that characterizes residual finite-dimensionality for general operator algebras (see Theorems 3.1 and 3.2). Theorem 1.1. Let A be a separable operator algebra. Then, the following state- ments are equivalent. (i) The algebra A is RFD. (ii) Every extremal representation of A has a finite-dimensional approximation. (iii) Every extremal representation of A has an Exel–Loring approximation. ∗ (iv) If π is a ∗-representation of Ce(A), then π|A admits an Exel–Loring approx- imation. The curious reader may wonder how the statements appearing above relate to having RFD C∗-envelope. It is easy to construct examples where A is RFD while ∗ ∗ Ce(A) is not, but a precise condition on A which is equivalent to Ce(A) being RFD is unknown; see [21] and [3, Proposition 4.1] for recent related results. At the other extreme in the scale of C∗-algebras generated by a copy of A is ∗ ∗ the maximal C -cover, denoted by Cmax(A). Notably, the existence of Exel–Loring approximations can be used to characterize the residual finite-dimensionality of ∗ Cmax(A), as shown in Theorem 3.3. ∗ Theorem 1.2. Let A be an operator algebra. Then Cmax(A) is RFD if and only if every representation of A admits an Exel–Loring approximation. Motivated by these results, let us say that the operator algebra A has property (F ) if all of its representations admit a finite-dimensional approximation. The question of which operator algebras enjoy property (F ) is an interesting one, and Theorems 1.1 and 1.2 imply that ∗ F Cmax(A) is RFD ⇒ A has property ( ) ⇒ A is RFD. In trying to understand property (F ), a natural question emerges.

∗ Question 1. Let A be an RFD operator algebra. Is Cmax(A) necessarily RFD? This question was already raised in [20, Section 5], where some partial results supported an affirmative answer. Our results show that Question 1 is equivalent to the question of whether or not the existence of a representation admitting a finite- dimensional approximation implies that all representations admit an Exel–Loring approximation. Our approximation techniques and the resulting property (F ) thus unearth an intermediate state which may prove useful in answering this question. The rest of the paper is motivated by Question 1, in that we find concrete operator algebras whose maximal C∗-cover is RFD. While C∗-algebras are analogous to groups, general operator algebras are anal- ogous to semigroups. Hence, non-self-adjoint operator algebras arising from semi- groups form a natural class of examples for our study. Semigroup C∗-algebras have been studied by many authors over the years [4,5,16,25–27,43,50], where a unified approach was obtained by Li for the class of independent semigroups in [44–46]. ∗ An operator algebra analogue of the reduced C -algebra is Ar(P ), which is the operator algebra generated by the left regular representation of P on ℓ2(P ). Suppose P is a (both left and right). For p ∈ P we define its set of right divisors

Rp = {r ∈ P : p = qr for some q ∈ P }. 4 R. CLOUATREˆ AND A. DOR-ON

We say that P has the finite divisor property (or FDP) if Rp is finite for every p ∈ P . It is relatively straightforward to show that Ar(P ) is RFD when P has FDP (see Proposition 4.1). With the aim of answering Question 1 in concrete examples, we introduce in Section 4 a notion of an RFD semigroup coaction of P on an operator algebra A, and prove the following (see Theorem 4.6). Theorem 1.3. Let P a countable cancellative semigroup with FDP such that there is a character χ : Ar(P ) → C with χ(λp)=1 for every p ∈ P . Let A be a separable ∗ operator algebra that admits an RFD coaction of P . Then, Cmax(A) is RFD. Coactions of groups on general operator algebras were introduced in [34] for the purpose of showing the existence of C∗-algebras satisfying co-universality with respect to representations of product systems over right LCM . In the C∗-algebra literature, coactions by a discrete group G are interpreted as actions of the quantum dual group of G on the C∗-algebra, and are useful in many instances [17, 37, 53, ?Seh19]. Since our semigroups are assumed to be FDP, they will not contain infinite groups as subsemigroups, and hence our development of coactions by semigroups differs from the one for group coactions in [34]. The class of independent semigroups includes the majority of examples of semi- groups previously studied through their C∗-algebras in the literature. In Section 5 we use Theorem 1.3 to show that the maximal C*-cover is RFD for semigroup alge- bras of many independent semigroups, as well as for operator algebras of functions with circular symmetry (see Theorem 5.7 and Theorem 5.12 respectively). Theorem 1.4. Let P and Q be countable semigroups embedded in some groups G and H respectively. Assume that P and Q are independent, and that Q is left- amenable and has FDP. Assume further that there is a ϕ : G → H such that ϕ(P ) ⊂ Q and ϕ|P is finite-to-one. Then, there is an RFD coaction of Q ∗ on Ar(P ) and Cmax(Ar(P )) is RFD. This theorem allows us to show that for many concrete examples of indepen- ∗ dent semigroups, the operator algebra Ar(P ) has an RFD maximal C -cover. For instance, this works for Nd for any d, left-angled Artin monoids (Example 5.9) and Braid monoids (Example 5.10). Theorem 1.5. Let Ω ⊂ Cn be a balanced and let A be a T-symmetric operator algebra of functions on Ω. Assume that there is a collection of homogeneous polynomials spanning a dense subset of A. Then, A admits an RFD coaction by N. ∗ In particular, Cmax(A) is RFD. This theorem is applied to some uniform algebras (Corollary 5.13), to the famous Schur–Agler class of functions (Corollary 5.14) and to some algebras of multipli- ers on well-behaved reproducing Hilbert spaces (Corollary 5.15). Taking into account Theorem 1.2, these results shed considerable light on the structure of representations of operator algebras that play an important role in multivariate operator theory; see [12, 18, 19, 22] and the references therein. In particular, it follows from Corollary 5.13 that the maximal C∗-cover of the polydisc algebra A(Dn) is RFD. This result is particularly striking in view of the following observation. It is a consequence of Ando’s theorem [51, Theorem 5.5] ∗ D2 ∗ that Cmax(A( )) is the universal C -algebra generated by a pair of commuting contractions. Hence, this universal C*-algebra is RFD. On the other hand, it follows from [23, Theorem 6.11] and the refutation of Connes conjecture [42] FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 5 that the universal C∗-algebra generated by a pair of doubly commuting contractions is not RFD.

2. Finite-dimensional approximations In this section, we examine various notions of finite-dimensional approximations for representations of operator algebras. Before proceeding however, we need to recall some operator algebraic background.

2.1. C∗-covers and extremal representations. Let A be an operator algebra. If B is a C∗-algebra and ι : A → B is a completely isometric homomorphism with B = C∗(ι(A)), then we say that the pair (B, ι) is a C∗-cover of A. Throughout the paper, we will be dealing with two special C∗-covers, which we introduce next. ∗ ∗ The maximal C -cover of A, denoted by (Cmax(A),υ), is the essentially unique C∗-cover satisfying the following : for any C∗-algebra C and any ∗ representation π : A → C, there is a ∗-representation π : Cmax(A) → C such that π ◦υ = π. (Bya representation of an operator algebra, here and elsewhere we mean a completely contractive homomorphism.) For a proofb of existence of the maximal ∗ Cb -cover, the reader should consult [13, Proposition 2.4.2]. The other C∗-cover that is relevant for our purposes is the “minimal” one. It ∗ ∗ is usually referred to as the C -envelope of A, denoted by (Ce (A),ε). Its defining universal property is the following: for any C∗-cover (ι, B) of A, there is a ∗- ∗ representation π : B → Ce(A) such that π ◦ ι = ε. The existence of this object is highly non-trivial, see [13, Theorem 4.3.1 and Proposition 4.3.5]. Due to important work of Dritschel–McCullough [36] (inspired by previous in- sight of Muhly–Solel [49]), it is known that certain special representations of A are intimately related to its C∗-envelope. We recall these important concepts. Let π : A → B(Hπ) and θ : A → B(Hθ) be representations. We say that θ is a dilation of π if Hπ ⊂ Hθ and

π(a)= PHπ θ(a)|Hπ , a ∈ A.

If in addition the space Hπ is invariant for θ(A), then the dilation θ is said to be ∗ an extension of π. Similarly, if Hπ is invariant for θ(A) , then the dilation θ is said to be a coextension of π. If θ is a dilation of π, then it is said to be a trivial dilation if the space Hπ is reducing for θ(A). In other words, a trivial dilation θ of π can be written as

⊥ θ(a)= π(a) ⊕ θ(a)|Hπ , a ∈ A. A representation is said to be extension-extremal  if all its extensions are trivial. Likewise, it is said to be coextension-extremal if all its coextensions are trivial. Moreover, a representation is simply said to be extremal if all its dilations are trivial. Clearly, an extremal representation is necessarily extension-extremal and coextension-extremal. The reader should consult [30] for a thorough study of these notions. We now collect some minor variations on known facts that we require later. We provide a detailed argument, as some care must be taken when dealing with the (or lack thereof). Theorem 2.1. Let (B, ι) be a C∗-cover of some operator algebra A. Let π : A → B(Hπ) be a representation. The following statements hold. 6 R. CLOUATREˆ AND A. DOR-ON

(i) If π is extremal, then there is a ∗-representation σ : B → B(Hπ) such that π = σ ◦ ι. (ii) If π is extension-extremal, then there is an extremal representation σ of A which is a coextension of π. (iii) If π is coextension-extremal, then there is an extremal representation σ of A which is an extension of π. Proof. (i) If A is unital, then necessarily ι and B are both unital and the desired statement is an immediate consequence of Stinespring’s dilation theorem [51, The- orem 4.1] applied to the representation π ◦ ι−1. If A is not unital, then [47, Section 3] implies the existence of a unital represen- tation π1 of the unitization A1 of A that extends π. It is easy to see that π1 is still extremal (see for instance [35, Proposition 2.5]). Hence, it suffices to apply the previous argument to π1. (ii) Assume first that A is unital. Then, π(I) is a self-adjoint projection with range M commuting with π(A). Thus, the map π′ : A → B(M) defined as ′ π (a)= π(a)|M, a ∈ A is a unital representation. It is readily verified that π′ is extension-extremal since π is assumed to be. Invoking [30, Corollary 3.11], we see that there is an extremal ′ ′ unital representation σ : A → B(Hσ) which is a coextension of π . Define now a representation σ : A → B(Hπ) as σ(a)= σ′(a) ⊕ 0, a ∈ A. Clearly, this is a coextension of π. To see that σ is extremal, we argue as follows. First, we note that the zero map on ι(A) has a unique completely positive extension to B, namely the zero map itself. Invoking [35, Proposition 2.4], we conclude that 0 is an extremal representation of A. Once we know this, we may argue as in the proof of [10, Proposition 4.4] (using the version of the Schwarz inequality found in [15, Proposition 1.5.7]) to see that σ′ being extremal implies that σ is also extremal, by another application of [35, Proposition 2.4]. If A is not unital, then as above we may apply [47, Section 3] to find a unital representation π1 of the unitization A1 of A. A trivial modification of the argument in [35, Proposition 2.5] shows that π1 is still extension-extremal. An application of the previous argument to π1 completes the proof. (iii) This is completely analogous to (ii).

2.2. Finite-dimensional approximations. Let A be an operator algebra and let π : A → B(H) be a representation. Let

πλ : A → B(Hπ), λ ∈ Λ ∗ be a net of representations with the property that the space C (πλ(A))H is finite- dimensional for every λ ∈ Λ. We say that the net (πλ) is a finite-dimensional approximation for π if (πλ(a)) converges to π(a) in the SOT for every a ∈ A. If, ∗ ∗ instead, we have that (πλ(a) ) converges to π(a) in the SOT for every a ∈ A, then we say that (πλ) is a finite-dimensional ∗-approximation for π. Finally, we say that (πλ) is an Exel–Loring approximation for π if it is both a finite-dimensional approximation and a finite-dimensional ∗-approximation. Our definitions above are directly inspired by work on Exel–Loring [38, Theorem 2.4], according to which all FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 7

∗-representations of an RFD C∗-algebra admit what we call Exel–Loring approxi- mations. The rest of this section is devoted to establishing basic facts about finite-dimen- sional approximations that we require throughout the paper. First, we examine direct sums. Lemma 2.2. Let A be an operator algebra and let Ω be a set. For each ω ∈ Ω, let πω : A → B(Hω) be a representation that admits an Exel–Loring approximation. Let π : A → B(⊕ω∈ΩHω) be the representation defined as

π(a)= πω(a), a ∈ A. ∈ ωMΩ Then, π admits an Exel–Loring approximation.

Proof. For each ω ∈ Ω, there is a directed set Λω and a net

πω,λ : A → B(Hω), λ ∈ Λω which is an Exel–Loring approximation for πω. Let F denote the directed set of finite of Ω. Given F = {ω1,ω2,...,ωn}∈F, it is readily verified that n

lim lim ··· lim πωi,λi (a) ⊕ 0 = πω(a) ⊕ 0 λ1∈Λω1 λ2∈Λω2 λn∈Λω n i=1 ! ∈ M ωM∈F / ωMF ωM∈F / and n ∗ ∗ lim lim ··· lim πωi,λi (a) ⊕ 0 = πω(a) ⊕ 0 λ1∈Λω1 λ2∈Λω2 λn∈Λω n i=1 ! ∈ M ωM∈F / ωMF ωM∈F / in the SOT for every a ∈ A. On the other hand, we see that

∗ ∗ lim πω(a) ⊕ 0 = π(a) and lim πω(a) ⊕ 0 = π(a) F ∈F F ∈F ∈ ! ∈ ! ωMF ωM∈F / ωMF ωM∈F / in the SOT for every a ∈ A. Thus, there is an Exel–Loring approximation for π contained in the set n π ⊕ 0  ωi,λi  i=1 6 1 M ω=ωM,...,ωn  N where n ∈ , ω1,ω2,...,ωn ∈ Ω, λ1 ∈ Λω1 , λ2 ∈ Λω2 ,...,λn ∈ Λω .   n Two representations π : A → B(Hπ) and σ : A → B(Hσ) are said to be approxi- mately unitarily equivalent if there is a sequence of unitary operators

Un : Hπ → Hσ, n ∈ N with the property that ∗ lim kUnπ(a)Un − σ(a)k =0, a ∈ A. n→∞ Another elementary fact we need is that the existence of Exel–Loring approxima- tions is preserved under approximate unitary equivalence.

Lemma 2.3. Let A be an operator algebra and let π : A → B(Hπ) be a represen- tation that admits an Exel–Loring approximation. Let σ : A → B(Hσ) be another representation which is approximately unitarily equivalent to π. Then, σ admits an Exel–Loring approximation. 8 R. CLOUATREˆ AND A. DOR-ON

Proof. By assumption, there is a sequence of unitary operators

Un : Hπ → Hσ, n ∈ N with the property that ∗ (2.1) lim kUnπ(a)Un − σ(a)k =0, a ∈ A. n→∞ Obviously, we also have that ∗ ∗ ∗ (2.2) lim kUnπ(a) Un − σ(a) k =0, a ∈ A. n→∞ Let

πλ : A → B(H), λ ∈ Λ be an Exel–Loring approximation for π. Now, let Ω be the directed set consisting of pairs (n, λ) where n ∈ N and λ ∈ Λ. For ω = (n, λ) ∈ Ω, define a representation σω : A → B(Hσ) as ∗ σω(a)= Unπλ(a)Un, a ∈ A.

It then follows easily from Equations (2.1) and (2.2) that the set {σω : ω ∈ Ω} contains an Exel–Loring approximation for σ.

Next, we show that restrictions to invariant subspaces preserve the existence of finite-dimensional approximations.

Lemma 2.4. Let A be an operator algebra and let π : A → B(Hπ) be a represen- tation. Let H ⊂ Hπ be a closed subspace. The following statements hold. (i) Assume π admits a finite-dimensional approximation and that H is invariant for π(A). Then, the representation

a 7→ π(a)|H, a ∈ A admits a finite-dimensional approximation. (ii) Assume π admits an Exel–Loring approximation and that H is reducing for π(A). Then, the representation

a 7→ π(a)|H, a ∈ A admits an Exel–Loring approximation. Proof. Clearly, we may assume without loss of generality that H is infinite-dimensional. Throughout the proof, we let Λ denote the directed set of triples (F, V,ε) where ε> 0, F is a finite subset of the unit ball of A, and V is a finite subset of H. (i) By assumption, there is a net

πω : A → B(Hπ), ω ∈ Ω ∗ of representations such that C (πω(A))Hπ is finite-dimensional and (πω(a)) con- verges in the SOT to π(a) for every a ∈ A. Thus,

hπ(a)ξ, π(b)ηi = limhπω(a)ξ, πω(b)ηi ω for every a,b ∈ A and ξ, η ∈ Hπ. Next, fix an element λ = (F, V,ε) ∈ Λ. Let ε′ > 0 be the positive number obtained by applying [32, Lemma 3.5.6] to the finite subset {x, π(a)ξ : a ∈F, ξ ∈ V} FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 9 which lies in H, since H is assumed to be invariant for π(A). Let ωλ ∈ Ω be chosen large enough such that ′ |hπ(a)ξ, π(b)ηi − hπωλ (a)ξ, πωλ (b)ηi| <ε for every a,b ∈F,ξ,η ∈ V. Because H is infinite-dimensional, there is an isometry ∗ Wλ : C (πωλ (A))Hπ → H. Using [32, Lemma 3.5.6], there is a unitary operator Uλ ∈ B(H) such that kUλWλx − xk <ε and

kUλWλπωλ (a)x − π(a)xk <ε ∗ for every a ∈F, x ∈ V. Put Vλ = UλWλ : C (πωλ (A))Hπ → H. For a ∈F, x ∈ V, we note that ∗ kVλ x − xk≤kx − Vλxk <ε whence ∗ ∗ kVλπωλ (a)Vλ x − π(a)xk = kVλπωλ (a)Vλ x − Vλπωλ (a)xk + kVλπωλ (a)x − π(a)xk ∗ < kπωλ (a)kkVλ x − xk + ε ≤ (1 + kak)ε ≤ 2ε. ∗ We thus conclude that the net (Vλπωλ (a)Vλ )λ converges in the SOT to π(a)|H for ∗ ∗ H every a ∈ A. Moreover, for each λ ∈ Λ we see that C (Vλπωλ (A)Vλ ) is finite- ∗ H ∗ dimensional since it is contained in VλC (πωλ (A)) π. Thus, (Vλπωλ (a)Vλ )λ is the desired finite-dimensional approximation. (ii) The argument is similar to the one above, with some minor modifications. By assumption, there is a net

πω : A → B(Hπ), ω ∈ Ω ∗ of representations such that C (πω(A))Hπ is finite-dimensional. Moreover, for every ∗ a ∈ A, in the SOT we have that (πω(a)) converges π(a) while (πω(a)) converges π(a)∗. Thus, hπ(a)ξ, π(b)ηi = limhπω(a)ξ, πω(b)ηi ω ∗ ∗ hπ(a)ξ, π(b) ηi = limhπω(a)ξ, πω(b) ηi ω ∗ ∗ ∗ ∗ hπ(a) ξ, π(b) ηi = limhπω(a) ξ, πω(b) ηi ω for every a,b ∈ A and ξ, η ∈ Hπ. Next, fix an element λ = (F, V,ε) ∈ Λ. Let ε′ > 0 be the positive number obtained by applying [32, Lemma 3.5.6] to the finite subset {x, π(a)ξ, π(a)∗ξ : a ∈F, ξ ∈ V} which lies in H, since H is assumed to be reducing for π(A). Let ωλ ∈ Ω be chosen large enough such that ′ |hπ(a)ξ, π(b)ηi − hπωλ (a)ξ, πωλ (b)ηi| <ε , ∗ ∗ ′ |hπ(a)ξ, π(b) ηi − hπωλ (a)ξ, πωλ (b) ηi| <ε , and ∗ ∗ ∗ ∗ ′ |hπ(a) ξ, π(b) ηi − hπωλ (a) ξ, πωλ (b) ηi| <ε , for every a,b ∈F,ξ,η ∈ V. Arguing as in (i), we find an isometry ∗ Vλ : C (πωλ (A))Hπ → H 10 R. CLOUATREˆ AND A. DOR-ON with the property that kVλx − xk <ε and ∗ ∗ kVλπωλ (a)x − π(a)xk < ε, kVλπωλ (a) x − π(a) xk <ε V ∗ for every a ∈ F, x ∈ . As before, the net (Vλπωλ (a)Vλ )λ is verified to be the desired Exel–Loring approximation. An important application of the preceding sequence of lemmas is the next tool. ∗ Theorem 2.5. Let A be a separable operator algebra. Let π : C (A) → B(Hπ) and ∗ σ : C (A) → B(Hσ) be two ∗-representations. Assume that π is injective and that π|A admits an Exel–Loring approximation. Then, σ|A also admits an Exel–Loring approximation. Proof. Because π is injective and A is separable, we may find a separable closed ′ ∗ subspace Hπ ⊂ Hπ which is reducing for π(C (A)) and such that the restriction ′ ∗ ′ π : C (A) → B(Hπ) defined as

′ ′ ∗ π (t)= π(t)|Hπ , t ∈ C (A) is injective and non-degenerate. By Lemma 2.4, we see that π′ also admits an Exel–Loring approximation. Thus, we may assume that π itself is injective, non- degenerate, and that Hπ is separable. Next, we may write σ = σ′ ⊕ 0, where σ′ is a non-degenerate ∗-representation which can, in turn, be decomposed as the direct sum of cyclic ∗-representations. Lemma 2.2 implies that it suffices to establish the desired result for all these cyclic ∗-representations. Thus, we may simply assume that σ is non-degenerate and that Hσ is separable. ∗ (∞) Consider then the non-degenerate ∗-representations Π : C (A) → B(Hπ ) and ∗ (∞) Σ : C (A) → B(Hσ ) defined as Π(t)= π(t)(∞) and Σ(t)= σ(t)(∞) for every t ∈ C∗(A). In particular, we see that Π is injective, and that Π(C∗(A)) and Σ(C∗(A)) contain no compact operators. The non-degenerate ∗-representations Π and Σ⊕Π are both injective, so that by [29, Corollary II.5.6] they are approximately unitarily equivalent. By assumption, π|A admits an Exel–Loring approximation, and thus so does Π|A by Lemma 2.2. In turn, Lemma 2.3 implies that (Σ ⊕ Π)|A admits an Exel–Loring approximation. Finally, Lemma 2.4 implies that Σ|A and σ|A admit an Exel–Loring approximations as well. The final preliminary tool we require is the following, which provides Exel–Loring approximations for representations that extend to ∗-representations of some RFD C∗-cover. Lemma 2.6. Let (B, ι) be a C∗-cover of some operator algebra A such that B is RFD. Let π be a representation of A such that π ◦ ι−1 extends to a ∗-representation of B. Then, π admits an Exel–Loring approximation.

Proof. Write π : A → B(Hπ). By assumption, there is a ∗-representation σ : B → B(Hπ) such that σ ◦ ι = π. Invoking [38, Theorem 2.4], we know that σ admits an Exel–Loring approximation

σλ : B → B(Hπ), λ ∈ Λ. FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 11

This means that σλ(B)Hπ is finite-dimensional for every λ ∈ Λ, and (σλ(b)) con- verges to σ(b) in SOT for every b ∈ B. For each λ ∈ Λ, let πλ = σλ ◦ ι. Note first that ∗ C (πλ(A))Hπ = σλ(B)Hπ is finite-dimensional for every λ ∈ Λ. Next, fix a ∈ A. Then, we find ∗ ∗ πλ(a)= σλ(ι(a)) and πλ(a) = σλ(ι(a) ) ∗ for every λ ∈ Λ. We infer that (πλ(a)) converges to π(a) and (πλ(a) ) converges to ∗ π(a) in SOT. Consequently, (πλ) is an Exel–Loring approximation of π. 3. Characterizations of residual finite-dimensionality In this section, we use finite-dimensional approximations and the related tools developed in Section 2 to analyze residual finite-dimensionality of general (that is, possibly non-selfadjoint) operator algebras. We start with a characterization of residual finite-dimensionality. Theorem 3.1. Let A be an operator algebra. The following are equivalent. (i) The algebra A is RFD. (ii) Every extremal representation of A admits a finite-dimensional approxima- tion. (iii) Every extremal representation of A admits an Exel–Loring approximation. (iv) Every coextension-extremal representation of A admits a finite-dimensional approximation. Proof. (iii) ⇒ (ii): This is trivial. (ii) ⇒ (i): By virtue of [36, Theorems 1.1 and 1.2] combined with [35, Proposition 2.5], we may find a set S of extremal representations of A with the property that N M the map π∈S π is completely isometric. Let d ∈ and A ∈ d(A) with kAk = 1. Let ε> 0. There is a representation π : A → B(Hπ) in S such that L kπ(d)(A)k≥ (1 − ε). On the other hand, by assumption we see that π admits a finite-dimensional ap- proximation (πλ)λ∈Λ. Thus, there is λ ∈ Λ with the property that (d) kπλ (A)k≥ (1 − 2ε). ∗ Let Hλ = C (πλ(A))Hπ. This is a finite-dimensional reducing subspace for πλ. Define ρλ : A → B(Hλ) to be the representation

ρλ(a)= πλ(a)|Hλ , a ∈ A. We find πλ(a)= ρλ(a) ⊕ 0, a ∈ A (d) (d) whence πλ (A) is unitarily equivalent to ρλ (A) ⊕ 0, so that (d) kρλ (A)k≥ (1 − 2ε). We conclude that A is RFD. (i) ⇒ (iii): By definition of A being RFD, there is a C∗-cover (B, ι) of A such that B is an RFD C∗-algebra. Let π be an extremal representation of A. By virtue of Theorem 2.1, we may find a ∗-representation σ of B such that σ ◦ ι = π. It then follows from Lemma 2.6 that π admits an Exel–Loring approximation. (iv) ⇒ (ii): This is trivial. 12 R. CLOUATREˆ AND A. DOR-ON

(ii) ⇒ (iv): Let π be a coextension-extremal representation of A. By Theorem 2.1, we see that there is an extremal representation σ of A which is an extension of π. Since σ is assumed to admit a finite-dimensional approximation, Lemma 2.4 implies that π admits a finite-dimensional approximation as well.

Let A be an operator algebra and let π be an extremal representation of A. ∗ ∗ −1 Let (Ce (A),ε) denote the C -envelope of A. By Theorem 2.1, we see that π ◦ ε ∗ extends to a ∗-representation of Ce(A). Unfortunately, the converse statement fails ∗ in general, in the sense that there are ∗-representations of Ce(A) whose restriction to ε(A) is not extremal; in other words, there are operator algebras which are not hyperrigid [10]. We mention that this kind of pathology occurs even in the classical setting of uniform algebras on compact metric spaces [52, page 42]. In light of this discussion, for non-hyperrigid RFD operator algebras, Theorem 3.1 is inadequate to deal with general ∗-representations of the C∗-envelope. We remedy this shortcoming in the next result, at least in the separable setting.

∗ Theorem 3.2. Let A be a separable operator algebra and let (Ce(A),ε) denote its C∗-envelope. Then A is RFD if and only if π ◦ ε−1 admits an Exel–Loring ∗ approximation for every ∗-representation π of Ce(A). Proof. Assume first that π ◦ ε−1 admits an Exel-Loring approximation for every ∗ ∗-representation π of Ce(A). As explained before the theorem, this implies in particular that π ◦ ε−1, and hence π, admits an Exel–Loring approximation for every extremal representation π of A. By virtue of Theorem 3.1, we conclude that A is RFD. Conversely, assume that A is RFD. Let B denote the unitization of A (see [13, ∗ ∗ Paragraph 2.1.11]), and let (Ce (B), ι) denote the C -envelope of B. By [7, Propo- sition 2.2.4],[9, Theorem 7.1], and [10, Proposition 4.4], we may find an injective ∗ unital ∗-representation π : Ce(B) → B(Hπ) with Hπ separable such that π ◦ ι is an ∗ ∗ C extremal representation of B. Note now that Ce (B)=C (ι(A)) + I. Thus, as ∗ explained in [8, page 15], we may find two ∗-representations π0, π1 of Ce(B) such ∗ ∗ that π0(C (ι(A))) is non-degenerate, C (ι(A)) ⊂ ker π1 and π = π0 ⊕ π1. Because ∗ π is injective, we see that π0 must be injective on C (ι(A)). By [13, Paragraph ∗ ∗ 4.3.4], we know that we may choose Ce(A)=C (ι(A)) and ε = ι|A. Furthermore, [35, Proposition 2.5] implies that π0|ι(A) = π|ι(A) is extremal since π|ι(B) is. It follows from Theorem 3.1 that π0|ι(A) admits an Exel–Loring approximation. An application of Theorem 2.5 now yields the desired statement.

As mentioned above, in the case of RFD C∗-algebras, it is known that every representation admits an Exel–Loring approximation [38]. Theorems 3.1 and 3.2 make similar claims in the non-selfadjoint context, but only for certain special representations. It is then natural to wonder what can be said about the existence of Exel–Loring approximations for general representations of RFD operator algebras. As we show next, this property is in fact equivalent to a condition which is a priori stronger, namely the residual finite-dimensionality of the maximal C∗-cover.

∗ Theorem 3.3. Let A be an operator algebra and let (Cmax(A),υ) denote its max- imal C∗-cover. Then, the following statements are equivalent. ∗ ∗ (i) The C -algebra Cmax(A) is RFD. (ii) Every representation of A admits an Exel–Loring approximation. FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 13

∗ (iii) Let H be a Hilbert space and assume that Cmax(A) ⊂ B(H). Then, the repre- sentation υ : A → B(H) admits an Exel–Loring approximation. ∗ Proof. (i) ⇒ (ii): Assume that Cmax(A) is RFD. Let π : A → B(Hπ) be a represen- ∗ tation. Then, there is a ∗-representation π : Cmax(A) → B(Hπ) such that π ◦υ = π. It thus follows from Lemma 2.6 that π admits an Exel–Loring approximation. (ii) ⇒ (iii): This is trivial. b b (iii) ⇒ (i): Assume that υ has an Exel–Loring approximation

πλ : A → B(H), λ ∈ Λ. Note that linear combinations of words with letters in υ(A) ∪ υ(A)∗ are clearly ∗ ∗ ∗ dense in Cmax(A), since Cmax(A)=C (υ(A)). Therefore, to show that the latter C∗-algebra is RFD it suffices to fix such an element s 6= 0 and to construct a finite- ∗ dimensional ∗-representation ρ of Cmax(A) that does not vanish on s, since injective ∗-representations are automatically completely isometric. ∗ For each λ ∈ Λ, we obtain a ∗-representation πλ : Cmax(A) → B(H) with the property that πλ ◦ υ = πλ. Because is jointly continuous in the SOT on bounded sets, it is readily verified that (πbλ(s)) converges to s in the SOT. In particular, thereb exists λ ∈ Λ such that πλ(s) 6= 0. Note now that the space ∗ Hλ = πλ(Cmax(A))H is finite-dimensional since b ∗ ∗ ∗ πλ(Cmax(A))H =C (πλ(υ(Ab))H =C (πλ(A))H b and the latter is finite-dimensional by construction. Thus, we may define a finite- ∗ dimensional ∗-representationb ρ : Cmax(Ab ) → B(Hλ) as ∗ ρ(t)= πλ(t)|Hλ , t ∈ Cmax(A). It follows that ∗ πλ(t)=bρ(t) ⊕ 0, t ∈ Cmax(A). so in particular ρ has the required property. In light of Theorems 3.1b and 3.3, we see that the existence of finite-dimensional and Exel–Loring approximations is at the heart of Question 1. Indeed, the question is equivalent to the one asking whether all representations of an RFD operator al- gebra admit Exel–Loring approximations. Alternatively, one can ask for the weaker property (F ) defined in the introduction. Question 2. Let A be an RFD operator algebra. Does A have property (F )? In other words, do all representations of A admit a finite-dimensional approximation? To illustrate some of the difficulties associated with this question, we remark that the approximations from Theorem 3.3 can be rather complicated, even for very transparent choices of representations. For instance, let A⊂ B(H) be an operator ∗ algebra such that Cmax(A) is RFD. Let K ⊂ H be an invariant subspace for A. Consider the representation π : A → B(K) defined as π(a) = a|K for every a ∈ A. Then, by Theorem 3.3, we know that π admits an Exel–Loring approximation. But this approximation does not necessarily arise as compressions of A to finite- dimensional subspaces of K, as the following concrete example shows. Recall that a closed subspace S ⊂ H is said to be semi-invariant for A if the map

a 7→ PSa|S, a ∈ A is multiplicative. A standard verification reveals that S is semi-invariant if and only if S = M ⊖ N for some closed subspaces M, N ⊂ H that are invariant for A. 14 R. CLOUATREˆ AND A. DOR-ON

Example 3.4. Let D ⊂ C denote the open unit disc. Let A(D) be the classical disc algebra acting on the Hardy space H2. Let H∞ denote the Banach algebra of bounded holomorphic functions on D. In this example, we will freely use well- known facts about the structure of these objects; the reader may consult [41] for details. Let θ ∈ H∞ be the singular inner corresponding to the point mass at 1, i.e. z +1 θ(z) = exp , z ∈ D. z − 1   2 2 Let Kθ = H ⊖ θH and let π : A(D) → B(Kθ) be the unital representation defined as D π(a)= PKθ a|Kθ , a ∈ A( ). ∗ D Note that Cmax(A( )) is RFD by [20, Example 3]. Therefore, π admits an Exel– Loring approximation by Theorem 3.3. Nevertheless, as we show next, Kθ contains no finite-dimensional subspace which is semi-invariant for π(A(D)). To see this, assume that S ⊂ Kθ is such a subspace. We may find closed subspaces M, N ⊂ Kθ which are invariant for π(A(D)) such that S = M ⊖ N. Then, M ⊕ θH2 and N ⊕ θH2 are also closed and invariant for A(D). By Beurling’s theorem, there are distinct inner functions ϕ and ψ in H∞ such that ϕH2 = M ⊕ θH2, ψH2 = N ⊕ θH2. Because θ ∈ ϕH2 ∩ψH2, we see that ϕ and ψ must be singular inner functions with ϕH2 ⊂ ψH2 or ψH2 ⊂ ϕH2. The first possibility would imply M ⊂ N and hence S = {0} contrary to our assumption. Thus, we must have then that ψH2 ⊂ ϕH2 and S = M ⊖ N = ϕH2 ⊖ ψH2. We infer that S is infinite-dimensional (this follows for instance from [11, Exercise III.1.11]).

4. Constructing Exel–Loring approximations via semigroup coactions Motivated by Question 1, in this section we aim to identify concrete instances of ∗ RFD operator algebras A with the property that Cmax(A) is RFD. Our approach is predicated on Theorem 3.3, in that we aim to produce Exel–Loring approximations for arbitrary representations. The unifying theme throughout this section will be that of semigroups and their coactions. For more on semigroups and their C*- algebras, we recommend [28, Chapter 5] and [44–46]. Let P be a discrete semigroup. Assume that P is cancellative, so that two elements p, q ∈ P satisfy p = q whenever there is r ∈ P with rp = rq or pr = qr. For p ∈ P , define Rp = {r ∈ P : p = qr for some q ∈ P } and Lp = {q ∈ P : p = qr for some r ∈ P }.

Let r ∈ Rp. By definition, we find qr ∈ Lp such that p = qrr. Since P is cancellative, this element qr is uniquely determined. Likewise, given q ∈ Lp there is a unique rq ∈ Rp such that p = qrq. We conclude that there is a between Rp and Lp. We say that the cancellative semigroup P has the finite divisor property (or FDP) if for every p ∈ P the sets Rp and Lp are finite. Clearly the finite divisor property is inherited by subsemigroups. Semigroups with FDP include FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 15 important examples such as free monoids, left-angled Artin monoids, Baumslag– Solitar monoids, braid monoids, Thompson’s and more. Next, let λ : P → B(ℓ2(P )) denote the left regular representation. For each 2 p ∈ P , we denote by ep ∈ ℓ (P ) the characteristic function of {p}. Thus, {ep}p∈P is an orthonormal basis for ℓ2(P ). Because P is left cancellative, it is easily seen that λp is an isometry for every p ∈ P . Furthermore, for p ∈ P we see that

∗ eq if r = pq (4.1) λp(er)= (0 if r∈ / pP. 2 Let Ar(P ) ⊂ B(ℓ (P )) denote the norm closed operator algebra generated by the ∗ ∗ image of λ, and let Cr(P ) denote the C -algebra generated by the image of λ. The developments in this section hinge on the following fact. Proposition 4.1. Let P be a cancellative semigroup with FDP. For every finite 2 subset F ⊂ P , there is a representation πF : Ar(P ) → B(ℓ (P )) such that −1 P \ ∪p∈F ∪r∈Rp Lr = λ (ker πF ).

Moreover, the net (πF )F is an Exel–Loring approximation for the identity repre- sentation of Ar(P ). In particular, Ar(P ) is RFD.

Proof. For each p ∈ P , the subset Rp is finite by the assumption on the semigroup. 2 Correspondingly, let Xp ⊂ ℓ (P ) denote the finite-dimensional subspace spanned by {er : r ∈ Rp}. Note now that if r ∈ Rp, then Rr ⊂ Rp. Invoking (4.1), we find for p,s ∈ P and r ∈ Rp that ∗ (4.2) λs er ∈{0}∪{et : t ∈ Rr} ⊂ Xp and that ∗ (4.3) λs er 6= 0 if and only if s ∈ Lr.

We conclude from (4.2) that Xp is coinvariant for Ar(P ) for every p ∈ P . Next, for every finite subset F ⊂ P , we let YF = p∈F Xp. Clearly, YF is a finite-dimensional subspace which is coinvariant for A (P ). Let Q ∈ B(ℓ2(P )) rP F denote the finite-rank orthogonal projection onto YF . Thus, the map πF : Ar(P ) → B(ℓ2(P )) defined as

πF (a)= QF aQF , a ∈ Ar(P ) is a representation. It readily follows the fact that

∪p∈P Rp = P that the net (πF ) is an Exel–Loring approximation for the identity representation of Ar(P ), and that Ar(P ) is RFD. Finally, (4.3) implies that πF (λs) 6= 0 if and only if s ∈ ∪p∈F ∪r∈Rp Lr, which is equivalent to −1 P \ ∪p∈F ∪r∈Rp Lr = λ (ker πF ). 

Coactions of discrete groups on general operator algebras were introduced in [34] as a generalization of coactions of discrete groups on C*-algebras [53]. The following is a semigroup variant of [34, Definition 3.1], adapted to our context. 16 R. CLOUATREˆ AND A. DOR-ON

Definition 4.2. Let A be an operator algebra, P be a cancellative discrete semi- group, and δ : A→A⊗min Ar(P ) a completely isometric homomorphism. For each δ p ∈ P denote by Ap the p-th spectral subspace according to δ, given by δ Ap = {a ∈ A | δ(a)= a ⊗ λp}. δ We say that δ is a coaction of P on A if p∈P Ap is norm dense in A.

We remark here that in principle, theP role of Ar(P ) in the definition of a semi- group coaction can be fulfilled by a norm closed algebra generated by any contrac- tive representation of P . However, we will not need this added generality for our purposes. A useful property of semigroup coactions that we require is the following. Lemma 4.3. Let P be a cancellative discrete semigroup with the property that there is a character χ : Ar(P ) → C such that χ(λp)=1 for every p ∈ P . Let A be an operator algebra and let π : A → B(Hπ) be a representation. If δ is a coaction of P on A, then (π ⊗ χ) ◦ δ is unitarily equivalent to π. δ Proof. For p ∈ P and a ∈ Ap, we find

(π ⊗ χ) ◦ δ(a) = (π ⊗ χ)(a ⊗ λp)= π(a) ⊗ 1. The density of the sum of the spectral subspaces thus implies that (π ⊗ χ) ◦ δ is unitarily equivalent to π. Next, let F denote the directed set of finite subsets of P . Proposition 4.1 yields the existence of a net (πF )F ∈F which is an Exel–Loring approximation for the identity representation of Ar(P ). For each F ∈F, we let QF be the of A δ −1 by the norm of the generated by p∈λ (ker πF ) Ap. Definition 4.4. Let A be an operator algebra,PP be a cancellative discrete semi- group, and δ a coaction of P on A. We say that δ is residually finite–dimensional ∗ ∗ (RFD) if the C -algebra Cmax(QF ) is RFD for each F ∈F. We now take our first crucial step towards answering Question 1 for operator algebras with semigroup coactions. Proposition 4.5. Let A be an operator algebra and let P be a cancellative semi- group with FDP. Assume that there exists an RFD coaction δ of P on A. Let π : A → B(Hπ) be a representation. Then, the representation (π ⊗ id) ◦ δ admits an Exel–Loring approximation.

Proof. Let F ⊂ P be a finite subset, and let qF : A→QF denote the natural −1 δ quotient map. If p ∈ λ (ker πF ) and a ∈ Ap, then

(π ⊗ πF ) ◦ δ(a)= π(a) ⊗ πF (λp)=0. We conclude that δ Ap ⊂ ker(π ⊗ πF ) ◦ δ. −1 p∈λ X(ker πF ) 2 Thus, there is a representation ρF : QF → B(Hπ ⊗ ℓ (P ))) with the property that

(π ⊗ πF ) ◦ δ = ρF ◦ qF . ∗ Since the coaction δ is assumed to be RFD, we see that Cmax(QF ) is RFD, so by Theorem 3.3 the representation ρF admits an Exel–Loring approximation 2 ρF,µ : QF → B(Hπ ⊗ ℓ (P )), µ ∈ ΩF . FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 17

In particular, ∗ 2 ∗ 2 C ((ρF,µ ◦ qF )(A))(Hπ ⊗ ℓ (P ))=C (ρF,µ(QF ))(Hπ ⊗ ℓ (P )) is finite-dimensional for every µ ∈ ΩF . δ 2 For every p ∈ P and every a ∈ Ap, given ξ ∈ Hπ and η ∈ ℓ (P ) we obtain

lim lim(ρF,µ ◦ qF )(a)(ξ ⊗ η) = lim(ρF ◦ qF )(a)(ξ ⊗ η) F µ F

= lim((π ⊗ πF ) ◦ δ)(a)(ξ ⊗ η) F = lim π(a)ξ ⊗ πF (λp)η F = π(a)ξ ⊗ λpη

= (π(a) ⊗ λp)(ξ ⊗ η) = ((π ⊗ id) ◦ δ)(a)(ξ ⊗ η) and likewise ∗ ∗ lim lim(ρF,µ ◦ qF )(a) (ξ ⊗ η) = ((π ⊗ id) ◦ δ)(a) (ξ ⊗ η). F µ A standard density argument then reveals that there is a net in

{ρF,µ ◦ qF : F ⊂ P finite,µ ∈ ΩF } which forms an Exel–Loring approximation for (π ⊗ id) ◦ δ.

The following is the connection between coactions of semigroups on RFD oper- ator algebras, and RFD maximal C*-covers. Theorem 4.6. Let P a countable cancellative semigroup with FDP such that there is a character χ : Ar(P ) → C with χ(λp)=1 for every p ∈ P . Let A be a separable ∗ operator algebra that admits an RFD coaction of P . Then, Cmax(A) is RFD. ∗ Proof. Recall that υ : A → Cmax(A) is the completely isometric homomorphism ∗ ∗ such that (Cmax(A),υ) is the maximal C -cover of A. By Theorem 3.3, it suffices ∗ to fix an injective ∗-representation ρ : Cmax(A) → B(H) and to show that the representation ρ ◦ υ : A → B(H) admits an Exel–Loring approximation. To see this, let δ be an RFD coaction of P on A. By injectivity of the minimal tensor product, the map ∗ ∗ γ = υ ⊗ id : A⊗min Ar(P ) → Cmax(A) ⊗min Cr(P ) ∗ ∗ ∗ is a completely isometric homomorphism. Let δ : Cmax(A) → Cmax(A) ⊗min Cr(P ) be the ∗-representation such that δ ◦ υ = γ ◦ δ. Now, it follows from Lemma 4.3 that (π ⊗ χb) ◦ δ is unitarily equivalent to π. Let ∗ C χ : Cr(P ) → be a state extendingb χ. Since λp is an isometry and χ(λp) = 1, it follows that λp lies in the multiplicative of χ for every p ∈ P . Thus, χ isb a character, which in turn implies that (ρ ⊗ χ) ◦ δ is unitarily equivalent to ρ. In particular, because ρ is injective this forces δ to be an injective ∗-representationb and thus so is (ρ⊗id)◦δ. Proposition 4.5 now yieldsb anb Exel–Loring approximation for (π ⊗ id) ◦ δ. But we find b b (π ⊗ id) ◦ δ = (ρ ⊗ id) ◦ δ ◦ υ

b 18 R. CLOUATREˆ AND A. DOR-ON so that (ρ⊗id)◦δ|υ(A) admits an Exel–Loring approximation. Since P is countable, the algebra Ar(P ) is separable and we may invoke Theorem 2.5 to conclude that ρ|υ(A) admits anb Exel–Loring approximation, and hence the same is true for ρ◦υ.

The astute reader may notice that the condition that Ar(P ) admit an appro- priate character in the previous theorem is stronger than what is actually needed in the proof. Indeed, therein we only need to know that the coaction δ admits ∗ ∗ ∗ an extension to an injective ∗-representation δ : Cmax(A) → Cmax(A) ⊗min Cr(P ). Unfortunately, at present we do not know how to verify this condition without assuming the existence of a character, or whenb the coaction is trivial.

5. Algebras of semigroups and algebras of functions In this section, we exhibit several classes examples of operator algebras to which Theorem 4.6 can be applied. These include algebras arising from semigroups and algebras of functions. 5.1. Semigroup operator algebras. We recall some background material on semigroups. Let P be a cancellative semigroup. By a right ideal of P we mean a subset X ⊂ P which is closed under right multiplication. Given a subset X ⊂ P and an element p ∈ P , we define pX = {px : x ∈ X} and p−1X = {x : px ∈ X}.

Further, we denote by JP the smallest collection of right ideals that contains ∅ −1 and P , and such that X ∈ JP implies pX,p X ∈ JP . Assume now that P is embedded in a group G. The full semigroup C∗-algebra, ∗ ∗ denoted by Cs(P ), is the universal C -algebra generated by isometries {vp : p ∈ P } and self-adjoint projections {eX : X ∈ JP } such that (i) e∅ = 0; (ii) vpvq = vpq for p, q ∈ P ; −1 −1 (iii) whenever p1, q1, ..., pn, qn ∈ P satisfy p1q1 ...pn qn = e in G, then ∗ ∗ −1 −1 v v 1 ...v v = e . p1 q pn qn qn pn...q1 p1P ∗ ∗ Furthermore, we let Ds(P ) ⊂ Cs(P ) denote the C -subalgebra generated by {eX : X ∈ JP }. ∗ The defining relations above can be realized inside of Cr(P ), as we explain next. For added clarity, thoughout this section we denote the left regular representation P P 2 of P by λ . Given a subset X ⊂ P , we let 1X ∈ B(ℓ (P )) denote the operator of multiplication by the indicator function of X. Clearly, this is a self-adjoint ∗ projection. We denote by Dr(P ) the C*-subalgebra of Cr(P ) generated P {1X : X ∈ JP }. P P Now, it follows from [44, Lemma 3.1] that the families {λp : p ∈ P } and {1X : X ∈ 2 JP } inside B(ℓ (P )) satisfy the defining relations above. Hence, by universality we P ∗ ∗ obtain a surjective ∗-representation λ∗ : Cs(P ) → Cr(P ) such that P P λ∗ (p)= λp , p ∈ P and P λ∗ (eX )=1X , X ∈ JP . P ∗ Now, by [44, Lemma 3.11] there is a faithful conditional expectation Er : Cr(P ) → Dr(P ). FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 19

A cancellative semigroup P is said to be independent if whenever X ∈ JP can be written as n X = Xi i=1 [ for some X1,...,Xn ∈ JP , then necessarily we have X = Xj for some 1 ≤ j ≤ n. This property will play a crucial role below. Fortunately, many semigroups from the literature automatically satisfy independence. For instance, all right LCM semigroups are independent by [46, Lemmas 6.31 and 6.32]. Assume now that P embeds into a group and is independent. By [44, Lemma P ∗ P P P P 3.13] there is a conditional expectation Es : Cs(P ) → D so that λ∗ ◦Es = Er ◦λ∗ and P ∗ P ∗ (5.1) ker(λ∗ )= {a ∈ Cs(P ) : Es (a a)=0}. This fact will be used in the proof of the next result, which shows, roughly speaking, that certain semigroup induce ∗-representations on the reduced C∗-algebras. Proposition 5.1. Let P and Q be unital semigroups embedded into some groups G and H respectively. Suppose that P is independent and that ψ : G → H is an injective homomorphism such that ψ(P ) ⊂ Q. Then, there is a ∗-representation ∗ ∗ Ψr : Cr(P ) → Cr(Q) such that P Q Ψr(λp )= λψ(p), p ∈ P.

Q 2 Proof. For each p ∈ P , define the isometry wp = λψ(p) ∈ B(ℓ (Q)). For each Q 2 X ∈ JP , define the projection fX = 1ψ(X)Q ∈ B(ℓ (Q)). Because ψ(e) = e and ψ(P ) ⊂ Q, we see that ψ(P )Q = Q. From this, it readily follows that that the collections {wp : p ∈ P } and {fX : X ∈ JP } satisfy properties (i), (ii) and ∗ ∗ (iii) defining the full semigroup C -algebra Cs(P ). Consequently, there is a ∗- ∗ ∗ homomorphism Ψ : Cs(P ) → Cr (Q) such that

Ψ(vp)= wp, p ∈ P and Ψ(eX )= fX , X ∈ JP .

Furthermore, given p1,...,pm, q1,...,qm ∈ P , we compute using [44, Lemmas 3.1 and 3.11] that Q ∗ ∗ Q Q∗ Q Q∗ Q E ◦ Ψ(v v 1 ··· v v )= E (λ λ ··· λ λ ) r p1 q pm qm r ψ(p1) ψ(q1 ) ψ(pm) ψ(qm) Q −1 −1 1 −1 −1 if ψ(p1 q1 ··· pm qm)= e in H, = ψ(qm )ψ(pm)···ψ(q1 )ψ(p1)Q (0 otherwise Q −1 −1 1 −1 −1 if p1 q1 ··· pm qm = e in G, = ψ(qm )ψ(pm)···ψ(q1 )ψ(p1)Q (0 otherwise where the last equality follows since ψ is injective. Using that ψ(P )Q = Q again, [44, Lemma 3.13] implies that Q ∗ ∗ P ∗ ∗ Er ◦ Ψ(vp1 vq1 ··· vpm vqm )=Ψ ◦ Es (vp1 vq1 ··· vpm vqm ). We conclude from this that Q P Er ◦ Ψ=Ψ ◦ Es . 20 R. CLOUATREˆ AND A. DOR-ON

Q Combining this equality with (5.1) and recalling that Er is faithful, we infer that P ∗ ∼ ∗ P ker λ∗ ⊂ kerΨ. Since Cr(P ) = Cs(P )/ ker λ∗ , we obtain a ∗-representation Ψr : ∗ ∗ Cr(P ) → Cr(Q) with the desired property. We now give an extension [27, Lemma 7.5] that does not require the semigroups P to be Ore. Furthermore, it is not confined to the natural “diagonal” action λp 7→ P P λp ⊗ λp . Proposition 5.2. Let P and Q be unital semigroups embeddable into some groups G and H respectively. Assume that P is independent and let ϕ : G → H be a homomorphism such that ϕ(P ) ⊂ Q. Then there is a ∗-representation ∆ϕ : ∗ ∗ ∗ Cr(P ) → Cr(P ) ⊗min Cr (Q) such that P P Q ∆ϕ(λp )= λp ⊗ λϕ(p) Proof. Let ψ : G → G × H be the injective homomorphism given by ψ(g) = ∗ (g, ϕ(g)). Hence, by Proposition 5.1 we have a ∗-homomorphism Ψr : Cr(P ) → ∗ Cr(P × Q) such that P P ×Q Ψ(λp )= λ(p,ϕ(p)), p ∈ P. ∗ On the other hand, by [44, Lemma 2.16] there is a ∗- Θ : Cr(P × Q) → ∗ ∗ Cr(P ) ⊗min Cr(Q) such that P ×Q P Q Θ(λp,q )= λp ⊗ λq , p ∈ P, q ∈ Q.

The map ∆ϕ =Θ ◦ Ψr has the required properties.

Our next goal is to show that the map ∆ϕ above restricts to a coaction of Q on Ar(P ). To do this, we need the following simple observation. Lemma 5.3 (Fell’s absorption principle for semigroups). Let P be a unital cancella- tive semigroup, and suppose V : P → B(H) is a representation of P by isometries. Then, there is an isometry W ∈ B(ℓ2(P ) ⊗ H) such that

W (λp ⊗ idH) = (λp ⊗ Vp)W, p ∈ P. Proof. The map

ep ⊗ ξ 7→ ep ⊗ Vp(ξ), p ∈ P, ξ ∈ H is easily seen to extend to an isometry W : ℓ2(P ) ⊗ H → ℓ2(P ) ⊗ H. But then, for p, q ∈ P and ξ ∈ H we have

(λq ⊗ Vq)W (ep ⊗ ξ)= eqp ⊗ Vqp(ξ)= W (eqp ⊗ ξ)

= W (λq ⊗ idH)(ep ⊗ ξ).

We infer that W (λq ⊗ idH) = (λq ⊗ Vq)W as required.

We now arrive at a result which will provide us with many examples of coactions. Theorem 5.4. Let P and Q be unital semigroups embeddable into groups G and H respectively. Assume that P is independent and let ϕ : G → H be a homomorphism such that ϕ(P ) ⊂ Q. Then, there is a coaction δ of Q on Ar(P ) such that

P P Q δ(λp )= λp ⊗ λϕ(p), p ∈ P. FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 21

∗ Proof. By virtue of Proposition 5.2, there is a ∗-representation ∆ϕ : Cr(P ) → ∗ ∗ Cr(P ) ⊗min Cr(Q) such that P P Q ∆ϕ(λp )= λp ⊗ λϕ(p)

Note that the injectivity of the minimal tensor product we have Ar(P )⊗minAr(Q) ⊂ ∗ ∗ Cr(P ) ⊗min Cr(Q). Thus, we obtain a representation δ : Ar(P ) → Ar(P ) ⊗min

Ar(Q) by putting δ = ∆ϕ|Ar (P ). In particular, we conclude that δ takes values in Ar(P ) ⊗min Ar(Q), as required. By construction, it is clear that the sum of the spectral subspaces for δ is norm dense in Ar(P ). 2 Q Let V : P → B(ℓ (Q)) be given by Vp = λϕ(p), which is a representation of P by isometries. By Lemma 5.3 there is an isometry W such that P P W (λp ⊗ idH)= δ(λp )W, p ∈ P. Hence, we have W (a ⊗ idH)= δ(a)W, a ∈ Ar(P ).

Let A = [aij ] ∈ Mn(Ar(P )) for some n ∈ N. We find

kAk = k[aij ⊗ IH]k = k[Waij ⊗ IH]k = kδ(n)(A)(W ⊕ W ⊕ ... ⊕ W )k≤kδ(n)(A)k. We infer that δ is completely isometric, and thus is indeed a coaction of Q on Ar(P ). The following shows that the notion of a semigroup coaction in Definition 4.2 is deserving of its name, at least for independent semigroups embedded in groups. Corollary 5.5. Let P be a unital independent semigroup embedded in a group G. Then there is a comultiplication on Ar(P ), which is the completely isometric P P P homomorphism ∆P : Ar(P ) → Ar(P ) ⊗min Ar(P ) given by ∆P (λp )= λp ⊗ λp . If moreover P acts on some operator algebra A by δ, then we automatically have the coaction identity (δ ⊗ id) ◦ δ = (id ⊗ ∆P ) ◦ δ. Proof. The first part follows directly from Theorem 5.4 by taking ϕ to be the identity on G. For the second part, by definition of coaction it suffices to verify the δ coaction identity on each Ap, but this is easily deduced. The last piece of the puzzle is to verify that the coaction constructed in Theorem 5.4 is in fact RFD. For this purpose, the following elementary fact will be useful. Lemma 5.6. Let P and Q be semigroups. Assume that Q is cancellative. Let ϕ : P → Q be a homomorphism and let δ : Ar(P ) → Ar(P ) ⊗min Ar(Q) be a representation such that P P Q δ(λp )= λp ⊗ λϕ(p), p ∈ P. Q Then, for each q ∈ Q the set {a ∈ Ar(P ) : δ(a)= a ⊗ λq } coincides with the norm C P closure of p∈ϕ−1(q) λp . Q Proof. FixPq ∈ Q and a ∈ Ar(P ) such that δ(a)= a ⊗ λq . Let ε> 0. Then there ′ P C ′ is a finite linear combination a = p∈P cpλp with cp ∈ such that ka − a k <ε. Let h ∈ ℓ2(P ) be a unit vector and let s ∈ Q. Note that P ′ P δ(a )(h ⊗ es)= cpλp h ⊗ eϕ(p)s and δ(a)(h ⊗ es)= ah ⊗ eqs. ∈ pXP 22 R. CLOUATREˆ AND A. DOR-ON

Since Q is cancellative, we see that the set {ers : r ∈ Q} is orthonormal, whence by applying Pythagoras’ theorem we get

2 c λP − a h  p p  p∈ϕ−1(q) X   2 2

≤ c λP h ⊗ e + c λP − a h ⊗ e p p ϕ(p)s  p p  qs p/∈ϕ−1(q) p∈ϕ−1(q) X X ′ 2 2   = kδ(a − a)(h ⊗ es)k ≤ ε .

Since ε> 0 and the unit vector h ∈ ℓ2(P ) were chosen arbitrarily, this implies that C P a lies in the norm closure of p∈ϕ−1(q) λp . The reverse inclusion is obvious.

Let P and Q be semigroupsP embedded in some groups G and H respectively. A homomorphism ϕ : G → H is a (P,Q)-map if ϕ(P ) ⊂ Q and the set ϕ−1(q) is finite for every q ∈ Q. It is readily verified that in the presence of a (P,Q)-map with ϕ(P ) = Q, if P has FDP then so does Q. Surjective (P,Q)-maps with ϕ(P ) = Q are sometimes called controlled maps. These were originally introduced in [43] as generalizations of length functions. Various kinds of controlled maps were used successfully in the literature to establish amenability and nuclearity properties for semigroup C∗-algebras [4, 5, 25, 26]. We now arrive at one of the main results of the paper. Recall that a cancellative semigroup P is said to be left-amenable if it admits a left-invariant mean. More precisely, there is a state µ on ℓ∞(P ) such that for every p ∈ P , and f ∈ ℓ∞(P ) we have that µ(p · f) = µ(f) where (p · f)(q) = f(pq). For instance, all abelian semigroups are left-amenable. Theorem 5.7. Let P and Q be countable semigroups embedded in some groups G and H respectively. Assume that P and Q are independent, and that Q is left- amenable and has FDP. Assume also that there exists a (P,Q)-map ϕ : G → H. ∗ Then, there is an RFD coaction of Q on Ar(P ) and Cmax(Ar(P )) is RFD.

Proof. By Theorem 5.4, there is a coaction δ of Q on Ar(P ) such that

P P Q δ(λp )= λp ⊗ λϕ(p), p ∈ P. For q ∈ Q we let Q Bq = {a ∈ Ar(P ) : δ(a)= a ⊗ λq } be the corresponding spectral subspace for δ. For each q ∈ Q, we see by Lemma C P 5.6 that Bq is the norm closure of p∈ϕ−1(q) λp . Let (π ) be the Exel–Loring approximation for the identity representation of F P Ar(Q) given by Proposition 4.1. Thus, for each finite subset F ⊂ Q we see that

Q −1 Q \ ∪q∈F ∪r∈Rq Lr = (λ ) (ker πF ).

By definition of a coaction, we know thatAr(P ) is the norm closure of C P Bq = λp . ∈ ∈ −1 qXQ qXQ p∈ϕX(q) FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 23

We conclude that the quotient QF of Ar(P ) by the norm closure of the ideal generated by Q −1 B is spanned by the image in the quotient of q∈(λ ) (ker πF ) q

P C P λp . ∈ ∈R −1 qXF rXq p∈ϕX(Lr)

Since Q has FDP and ϕ is a (P,Q)-map, we get that QF is finite-dimensional. We ∗ infer that Cmax(QF ) is RFD by virtue of [20, Theorem 5.1]. Thus, we see that δ is an RFD coaction. Finally, since Q is left-amenable and independent, by [46, Theorem 6.42] there ∗ C Q is a character χ : Cr (Q) → . Clearly, for every q ∈ Q we have χ(λq ) = 1 since Q ∗ λq is an isometry. Thus, by Theorem 4.6 we get that Cmax(Ar(P )) is RFD.

We now extract some applications of the previous result.

Corollary 5.8. Let P be a countable, independent semigroup with FDP that is ∗ embedded in an . Then Cmax(Ar(P )) is RFD. Proof. Every abelian semigroup is left-amenable. Hence, in Theorem 5.7 we may ∗ take ϕ to be the identity, so that Cmax(Ar(P )) is RFD.

∗ Nd In paricular, the previous result shows that Cmax(Ar( )) is RFD for every d ∈ N (including d = ℵ0). We now study some classical examples of semigroups. Example 5.9 (Left-angled Artin monoids). Let Γ=(V, E) be a finite undirected graph, where two vertices are connected by at most one edge and no vertex is connected to itself. We let AΓ be the group generated by V as formal generators where, for v, w ∈ W , we have vw = wv if if there is an edge between v in w. We + then define the unital semigroup AΓ as the semigroup generated by the identity and {v : v ∈ V }. It is known that right-angled Artin groups are quasi- ordered by + their monoids (see [25]), so we get that AΓ is right LCM, and is hence independent. V V Next, consider the group Z . For each v ∈ V , let ev ∈ Z denote the indicator V function of {v}. There is a surjective ϕ : AΓ → Z such that NV ZV + NV ϕ(v)= ev. Viewing as a subsemigroup of , it is clear that ϕ(AΓ )= and −1 NV + NV ϕ (f) is finite for every f ∈ . Thus, ϕ is a (AΓ , )-map. As a consequence ∗ + of Theorem 5.7, we get that Cmax(Ar(AΓ )) is RFD.

Example 5.10 (Braid monoids). For n ≥ 3 let Bn be the group generated by elements σ1, ..., σn−1 subject to the relations

σiσi+1σi = σi+1σiσi+1 for 1 ≤ i ≤ n − 2 and

σiσj = σj σi for |i − j|≥ 2. + We let Bn be the unital semigroup generated by σ1, ..., σn−1 in Bn. + For each n ≥ 3, it follows from [14, Lemma 2.1] that Bn is right LCM, and hence is independent (see also the discussion preceding [46, Lemma 6.33]). There is a surjective group homomorphism ϕ : Bn → Z such that ϕ(σi) = 1 for every i. It is + N ∗ + readily seen that ϕ is a (Bn , )-map. By Theorem 5.7, we get that Cmax(Ar(Bn )) is RFD. 24 R. CLOUATREˆ AND A. DOR-ON

5.2. Operator algebras of functions. Let Ω ⊂ Cn be a subset and let T ⊂ C denote the unit circle. Assume that Ω is balanced, so that ζz ∈ Ω for every z ∈ Ω and every ζ ∈ T. In this case, given a function f : Ω → C and ζ ∈ T, we define fζ : Ω → C to be the function

fζ (z)= f(ζz), z ∈ Ω. Let A be an operator algebra consisting of functions on Ω. We say that A is T-symmetric if for each ζ ∈ T, the map

f 7→ fζ , f ∈ A defines a completely isometric isomorphism Φζ : A → A. It would certainly be natural here to impose some form of continuity on the map ζ 7→ Φζ . Fortunately, in our cases of interest this turns out to be automatic. Lemma 5.11. Let Ω ⊂ Cn be a balanced subset and let A be a T-symmetric operator algebra of functions on Ω. Assume that there is a collection of homogeneous polynomials spanning a dense subset of A. Then, the map

ζ 7→ Φζ (f), ζ ∈ T is norm continuous for every f ∈ A.

Proof. Given f ∈ A, we define a function δ0(f) : T → A as

δ0(f)(ζ)=Φζ (f), ζ ∈ T. For a homogeneous polynomial f ∈ A of degree n, we see that n δ0(f)(ζ)= ζ f, ζ ∈ T ε so in particular δ0(f) is continuous. A standard 3 -argument using our density assumption shows that in fact δ0(a) is continuous for every a ∈ A. Next, we show that T–symmetric operator algebras admit RFD coactions by N. Theorem 5.12. Let Ω ⊂ Cn be a balanced subset and let A be a T-symmetric operator algebra of functions on Ω. Assume that there is a collection of homogeneous polynomials spanning a dense subset of A. Then, A admits an RFD coaction by N. ∗ In particular, Cmax(A) is RFD.

Proof. Given f ∈ A, we define a function δ0(f) : T → A as

δ0(f)(ζ)=Φζ (f), ζ ∈ T.

By Lemma 5.11, we see that δ0(f) ∈ C(T; A) for every f ∈ A. Hence, we obtain a completely isometric homomorphism δ0 : A → C(T; A). Next, it is a consequence of [51, Proposition 12.5] and of the injectivity of the minimal tensor product that there is a completely isometric isomorphism Θ : A⊗min C(T) → C(T; A) such that Θ(a ⊗ g)(ζ)= g(ζ)a, ζ ∈ T for every a ∈ A and g ∈ C(T). If f ∈ A is a homogeneous polynomial of degree n, then we see that −1 n Θ (δ0(f)) = f ⊗ ζ . −1 We conclude that (Θ ◦ δ0)(A) ⊂ A⊗min A(D), where A(D) denotes the disc algebra. FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 25

∗ N T Recall now that there is a surjective ∗-representation τ : Cr( ) → C( ) such that N n N τ(λn)= ζ , n ∈ and whose restriction to Ar(N) is completely isometric. In particular, this clearly ∗ N C N N implies that there is a character χ : Cr( ) → with χ(λn) = 1 for every n ∈ . −1 D Putting ρ = (τ|Ar (N)) , we obtain a completely isometric isomorphism ρ : A( ) → Ar(N) such that n N N ρ(z )= λn, n ∈ . Define −1 δ = (id ⊗ ρ) ◦ Θ ◦ δ0 : A→A⊗min Ar(N), which is a completely isometric representation. We claim that δ is an RFD coaction of N on A. For n ∈ N, we let N An = {a ∈ A : δ(a)= a ⊗ λn} be the corresponding spectral subspace. By construction, we see that a ∈ An if and only if n Φζ (a)= δ0(a)(ζ)= ζ a, ζ ∈ T.

Hence, An is the space of all homogeneous polynomials of degree n in A, and in particular is finite-dimensional. By assumption, we must have that n∈N An is norm dense in A. This shows that δ is a coaction of N on A. It remains to show that this coaction is in fact RFD. P Let (πF ) be the Exel–Loring approximation for the identity representation of Ar(P ) from Proposition 4.1. Given a finite subset F ⊂ N and n ∈ N, we see from N N Proposition 4.1 that λn ∈ ker πF if n > max F . Thus, the quotient QF of Ar( ) by the closed ideal generated by

An −1 n∈λ X(ker πF ) has dimension at most that of An n≤max F X ∗ and in particular is finite-dimensional. Hence, Cmax(QF ) is RFD by [20, Theorem 5.1]. We conclude that δ is indeed an RFD coaction of N on A. Finally, invoking ∗ Theorem 4.6 we get that Cmax(A) is RFD. Many concrete examples of operator algebras can now be analyzed in light of the previous result. Corollary 5.13. Let Ω ⊂ Cn be a balanced bounded subset and let A(Ω) denote the ∗ closure of the polynomials inside the continuous functions C(Ω). Then, Cmax(A(Ω)) is RFD. Proof. In view of Theorem 5.12, it suffices to show that A(Ω) is a T-symmetric algebra. For f ∈ A(Ω) and ζ ∈ T, we see that

kfζk = max |f(ζz)| = max |f(w)| = kfk, f ∈ A(Ω). z∈Ω w∈Ω

This implies that the map f 7→ fζ defines an isometric isomorphism Φζ : A(Ω) → A(Ω). The fact that Φζ is in fact completely isometric follows then from [51, Theorem 3.9]. 26 R. CLOUATREˆ AND A. DOR-ON

By taking Ω above to be the open unit polydisc or the open unit ball, we conclude from Corollary 5.13 that the maximal C∗-covers of the polydisc algebra and of the ball algebra are necessarily RFD. This answers a question left open in [20]. By definition, the algebra A(Ω) considered in Corollary 5.13 is a uniform algebra, as it embeds into the commutative C∗-algebra C(Ω). Below, we deal with operator algebras of functions on Ω which do not necessarily admit such an embedding. Let n be a positive integer. Let Γ denote the set of (unitary equivalence classes of) commuting n-tuples of contractions on some separable Hilbert space. Let p ∈ C[z1,...,zn] be a polynomial in n variables. The quantity

sup kp(T1,...,Tn)k (T1,...,Tn)∈Γ is easily seen to be finite and to define a norm on C[z1,...,zn]. The corresponding completion of C[z1,...,zn] is an operator algebra [51, Chapter 18], usually referred to as the Schur–Agler class [1]. We denote it by Sn. It follows from the classical 2 von Neumann and Ando inequalities that S1 =∼ A(D) and S2 =∼ A(D ). However, for bigger n the Schur–Agler class differs from the polydisc algebra. Nevertheless, we can still prove the following.

Corollary 5.14. Let n be a positive integer and let Sn denote the corresponding ∗ Schur–Agler class. Then, Cmax(Sn) is RFD. Proof. By definition we see that Dn |p(z)|≤kpkSn , z ∈

→֒ for every polynomial p in n variables, so that there is a contractive inclusion Sn n A(D ). Next, we note that if (T1,...,Tn) is a commuting n-tuple of contractions, then so is (ζT1,ζT2,...,ζTn) for each ζ ∈ T. Therefore, there is a completely isometric isomorphism Φζ : Sn → Sn such that

Φζ (f)= fζ , f ∈ Sn. n Hence, we conclude that Sn is a T–symmetric operator algebra of functions on D , ∗ so that Cmax(Sn) is RFD by Theorem 5.12.

For our next application, we will make use of standard terminology from the theory of reproducing kernel Hilbert spaces; details can be found in [2] and [40]. Let Ω ⊂ Cn be a balanced open connected subset containing the origin. Let K : Ω × Ω → C be a kernel function normalized at the origin, so that K(z, 0)= 1 for every z ∈ Ω. We say that K is T-invariant if we have that (i) K is holomorphic in the first variable, (ii) for every z, w ∈ Ω the map ζ 7→ K(ζz,w), ζ ∈ T is continuous, (iii) for every z, w ∈ Ω and ζ ∈ T we have that K(z, w)= K(ζz,ζw). Many important classical kernels satisfy these properties, such as the Drury–Arveson kernel and the Dirichlet kernel; see [40, Example 6.1]. Let H be the reproducing kernel Hilbert space corresponding to K and let Mult(H) denote its multiplier algebra. By our assumption on K, we see that H FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 27 consists of holomorphic functions on Ω. Furthermore, because 1 ∈ H we infer Mult(H) ⊂ H. For each ζ ∈ T, the map Γζ : H → H defined as

(Γζ (f))(z)= f(ζz), f ∈ H,z ∈ Ω is easily checked to be a strongly continuous unitary path (see [40, Section 6]). For each ζ ∈ T and ϕ ∈ Mult(H), we note that Γ∗M Γ = M . ζ ϕ ζ Γζ (ϕ)

Hence, we obtain a completely isometric isomorphism Θζ : Mult(H) → Mult(H) defined as H Θζ (Mϕ)= MΓζ (ϕ), ϕ ∈ Mult( ). For each n ∈ N, we let A(H)n ⊂ Mult(H) be the subspace consisting of those n multipliers ϕ such that Θζ (Mϕ) = ζ Mϕ for every ζ ∈ T. As explained in [40, Example 6.1 (i)], for n ≥ 0 we get that A(H)n consists of homogeneous polynomials of degree n, while A(H)n = {0} for n< 0. We define A(H) to be the norm closure ∞ of n=0 A(H)n inside of Mult(H). We now arrive at another application of Theorem 5.12. P Corollary 5.15. Let Ω ⊂ Cn be a balanced open connected set containing the ori- gin. Let K be a T-invariant kernel on Ω and let H be the corresponding reproducing ∗ kernel Hilbert space. Then, Cmax(A(H)) is RFD. Proof. We make a preliminary observation. Let ϕ ∈ A(H). By definition of A(H), there is a sequence (pn) of polynomials converging to ϕ in the norm topology of Mult(H). Since the multiplier norm always dominates the supremum norm over Ω, we conclude that the sequence (pn) converges to ϕ uniformly on Ω, and thus ϕ is holomorphic on Ω. Hence, A(H) is an operator algebra of holomorphic functions on Ω. In view of Theorem 5.12, it suffices to show that A(H) is a T-symmetric algebra. For each ζ ∈ T, we let Φζ = Θζ |A(H). It is clear that Φζ (A(H)n) ⊂ A(H)n for every n ≥ 0, so that Φζ (A(H)) ⊂ A(H). Thus, Φζ : A(H) → A(H) is a completely isometric isomorphism. A special class of T-invariant kernels is that of regular unitarily invariant kernels on the ball Bn. In this case, the algebra A(H) was studied in [3] from the point of view of residual finite-dimensionality. It was shown therein (see the proof of [3, Theorem 4.2]) that the stronger requirement that A(H) be subhomogeneous imposes tremendous restrictions, and in fact implies that A(H) can be identified with the ball algebra. This provides further impetus to revealing whether A(H) enjoys other refinements of residual finite-dimensionality. Corollary 5.15 provides the first instance in this direction. References [1] Jim Agler, On the representation of certain holomorphic functions defined on a polydisc, Topics in operator theory: Ernst D. Hellinger memorial volume, 1990, pp. 47–66. [2] Jim Agler and John E. McCarthy, Pick interpolation and Hilbert function spaces, Gradu- ate Studies in Mathematics, vol. 44, American Mathematical Society, Providence, RI, 2002. MR1882259 (2003b:47001) [3] Alexandu Aleman, Michael Hartz, John McCarthy, and Stefan Richter, Multiplier tests and subhomogeneity of multiplier algebras, preprint arXiv:2008.00981 (2020). [4] Astrid an Huef, Brita Nucinkis, Camila F. Sehnem, and Dilian Yang, Nuclearity of semigroup C∗-algebras, J. Funct. Anal. 280 (2021), no. 2, 108793, 46. MR4159272 28 R. CLOUATREˆ AND A. DOR-ON

[5] Astrid an Huef, Iain Raeburn, and Ilija Tolich, HNN extensions of quasi-lattice ordered groups and their operator algebras, Doc. Math. 23 (2018), 327–351. MR3846056 [6] R. J. Archbold, On residually finite-dimensional C∗-algebras, Proc. Amer. Math. Soc. 123 (1995), no. 9, 2935–2937. [7] William Arveson, Subalgebras of C∗-algebras, Acta Math. 123 (1969), 141–224. MR0253059 (40 #6274) [8] , An invitation to C∗-algebras, Springer-Verlag, New York-Heidelberg, 1976. Graduate Texts in Mathematics, No. 39. MR0512360 [9] , The noncommutative Choquet boundary, J. Amer. Math. Soc. 21 (2008), no. 4, 1065– 1084. [10] , The noncommutative Choquet boundary II: hyperrigidity, Israel J. Math. 184 (2011), 349–385. MR2823981 [11] Hari Bercovici, Operator theory and arithmetic in H∞, Mathematical Surveys and Mono- graphs, vol. 26, American Mathematical Society, Providence, RI, 1988. MR954383 (90e:47001) [12] Kelly Bickel, Michael Hartz, and John E. McCarthy, A multiplier algebra functional calculus, Trans. Amer. Math. Soc. 370 (2018), no. 12, 8467–8482. [13] David P. Blecher and Christian Le Merdy, Operator algebras and their modules—an oper- ator space approach, London Mathematical Society Monographs. New Series, vol. 30, The Clarendon Press, Oxford University Press, Oxford, 2004. Oxford Science Publications. [14] Egbert Brieskorn and Kyoji Saito, Artin-Gruppen und Coxeter-Gruppen, Invent. Math. 17 (1972), 245–271. MR323910 [15] N.P Brown and N. Ozawa, C∗-algebras and finite-dimensional approximations, Graduate Studies in Mathematics, vol. 88, American Mathematical Society, Providence, RI, 2008. [16] Nathan Brownlowe, Nadia S. Larsen, and Nicolai Stammeier, On C∗-algebras associated to right LCM semigroups, Trans. Amer. Math. Soc. 369 (2017), no. 1, 31–68. MR3557767 [17] Toke M. Carlsen, Nadia S. Larsen, Aidan Sims, and Sean T. Vittadello, Co-universal algebras associated to product systems, and gauge-invariant uniqueness theorems, Proc. Lond. Math. Soc. (3) 103 (2011), no. 4, 563–600. [18] Rapha¨el Clouˆatre and Kenneth R. Davidson, Absolute continuity for commuting row con- tractions, J. Funct. Anal. 271 (2016), no. 3, 620–641. MR3506960 [19] Rapha¨el Clouˆatre and Michael Hartz, Multiplier algebras of complete Nevanlinna-Pick spaces: dilations, boundary representations and hyperrigidity, J. Funct. Anal. 274 (2018), no. 6, 1690– 1738. [20] Rapha¨el Clouˆatre and Christopher Ramsey, Residually finite-dimensional operator algebras, J. Funct. Anal. 277 (2019), no. 8, 2572–2616. [21] Rapha¨el Clouˆatre and Ian Thompson, Finite-dimensionality in the non-commutative Choquet boundary: peaking phenomena and C∗-liminality, preprint arXiv:2009.100226 (2020). [22] Rapha¨el Clouˆatre and Edward J. Timko, Gelfand transforms and boundary representations of complete Nevanlinna–Pick , to appears in Transactions of the AMS, preprint arXiv:1912.10331 (2019). [23] Kristin Courtney and David Sherman, Some notes on the universal C∗-algebra of a contrac- tion, Journal of Operator Theory 84 (2020), no. 1, 153–184. [24] Kristin Courtney and Tatiana Shulman, Elements of C∗-algebras attaining their norm in a finite-dimensional representation, Canad. J. Math. 71 (2019), no. 1, 93–111. [25] John Crisp and Marcelo Laca, On the Toeplitz algebras of right-angled and finite-type Artin groups, J. Aust. Math. Soc. 72 (2002), no. 2, 223–245. [26] , Boundary quotients and ideals of Toeplitz C∗-algebras of Artin groups, J. Funct. Anal. 242 (2007), no. 1, 127–156. [27] Joachim Cuntz, Siegfried Echterhoff, and Xin Li, On the K-theory of the C*-algebra generated by the left regular representation of an Ore semigroup, J. Eur. Math. Soc. (JEMS) 17 (2015), no. 3, 645–687. [28] Joachim Cuntz, Siegfried Echterhoff, Xin Li, and Guoliang Yu, K-theory for group C∗- algebras and semigroup C∗-algebras, Oberwolfach Seminars, vol. 47, Birkh¨auser/Springer, Cham, 2017. [29] Kenneth R. Davidson, C∗-algebras by example, Fields Institute Monographs, vol. 6, American Mathematical Society, Providence, RI, 1996. FINITE-DIMENSIONAL APPROXIMATIONS AND SEMIGROUP COACTIONS 29

[30] Kenneth R. Davidson and Elias G. Katsoulis, Dilation theory, commutant lifting and semi- crossed products, Doc. Math. 16 (2011), 781–868. [31] Kenneth R. Davidson, Christopher Ramsey, and Orr Moshe Shalit, The isomorphism problem for some universal operator algebras, Adv. Math. 228 (2011), no. 1, 167–218. MR2822231 (2012j:46083) [32] Jacques Dixmier, C∗-algebras, North-Holland Publishing Co., Amsterdam-New York-Oxford, 1977. Translated from the French by Francis Jellett, North-Holland Mathematical Library, Vol. 15. MR0458185 [33] Adam Dor-On, of tensor algebras arising from weighted partial systems, Trans. Amer. Math. Soc. 370 (2018), no. 5, 3507–3549. [34] Adam Dor-On, Evgenios Kakariadis, Katsoulis Elias, Marcelo Laca, and Xin Li, C∗-envelopes for tensor algebras with a coaction and co-universal C∗-algebras for product systems, arXiv preprint : 2012.12435 (2020). [35] Adam Dor-On and Guy Salomon, Full Cuntz-Krieger dilations via non-commutative bound- aries, J. Lond. Math. Soc. (2) 98 (2018), no. 2, 416–438. [36] Michael A. Dritschel and Scott A. McCullough, Boundary representations for families of representations of operator algebras and spaces, J. Operator Theory 53 (2005), no. 1, 159– 167. [37] Ruy Exel, Amenability for Fell bundles, J. Reine Angew. Math. 492 (1997), 41–73. [38] Ruy Exel and Terry A. Loring, Finite-dimensional representations of free product C∗- algebras, Internat. J. Math. 3 (1992), no. 4, 469–476. [39] Don Hadwin, A lifting characterization of RFD C∗-algebras, Math. Scand. 115 (2014), no. 1, 85–95. [40] Michael Hartz, On the isomorphism problem for multiplier algebras of Nevanlinna-Pick spaces, Canad. J. Math. 69 (2017), no. 1, 54–106. [41] Kenneth Hoffman, Banach spaces of analytic functions, Dover Publications, Inc., New York, 1988. Reprint of the 1962 original. [42] Zhengfeng Ji, Anand Natarajan, Thomas Vidick, John Wright, and Henry Yuen, MIP ∗ = RE, preprint arXiv:2001.04383 (2020). [43] Marcelo Laca and Iain Raeburn, Semigroup crossed products and the Toeplitz algebras of nonabelian groups, J. Funct. Anal. 139 (1996), no. 2, 415–440. [44] Xin Li, Semigroup C∗-algebras and amenability of semigroups, J. Funct. Anal. 262 (2012), no. 10, 4302–4340. [45] , Nuclearity of semigroup C∗-algebras and the connection to amenability, Adv. Math. 244 (2013), 626–662. [46] , Semigroup C∗-algebras, preprint arXiv:1707.0594 (2017). [47] Ralf Meyer, Adjoining a unit to an operator algebra, J. Operator Theory 46 (2001), no. 2, 281–288. [48] Meghna Mittal and Vern I. Paulsen, Operator algebras of functions, J. Funct. Anal. 258 (2010), no. 9, 3195–3225. MR2595740 (2011h:47150) [49] Paul S. Muhly and Baruch Solel, An algebraic characterization of boundary representations, Nonselfadjoint operator algebras, operator theory, and related topics, 1998, pp. 189–196. [50] A. Nica, C∗-algebras generated by isometries and Wiener-Hopf operators, J. Operator Theory 27 (1992), no. 1, 17–52. [51] Vern Paulsen, Completely bounded maps and operator algebras, Cambridge Studies in Ad- vanced Mathematics, vol. 78, Cambridge University Press, Cambridge, 2002. MR1976867 (2004c:46118) [52] Robert R. Phelps, Lectures on Choquet’s theorem, Second Ed., Lecture Notes in Mathematics, vol. 1757, Springer-Verlag, Berlin, 2001. MR1835574 [53] John C. Quigg, Discrete C∗-coactions and C∗-algebraic bundles, J. Austral. Math. Soc. Ser. A 60 (1996), no. 2, 204–221. [54] Guy Salomon, Orr M. Shalit, and Eli Shamovich, Algebras of bounded noncommutative an- alytic functions on subvarieties of the noncommutative unit ball, Trans. Amer. Math. Soc. 370 (2018), no. 12, 8639–8690. 30 R. CLOUATREˆ AND A. DOR-ON

Department of Mathematics, University of Manitoba, Winnipeg, MB, Canada Email address: [email protected]

Department of Mathematics, University of Copenhagen, Copenhagen, Denmark Email address: [email protected]